physic ad 117

Upload: liezel-marasigan

Post on 14-Apr-2018

238 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/27/2019 Physic Ad 117

    1/16

    Physica D 117 (1998) 283298

    Comparisons of new nonlinear modeling techniques

    with applications to infant respiration

    Michael Small , Kevin JuddDepartment of Mathematics, University of Western Australia, Nedlands, WA 6907, Australia

    Received 1 August 1997; received in revised form 23 October 1997; accepted 2 December 1997

    Communicated by A.M. Albano

    Abstract

    This paper concerns the application of new nonlinear time-series modeling methods to recordings of infant respiratorypatterns. The techniques used combine the concept of minimum description length modeling with radial basis models. Our

    first application of the methods produced results that were not entirely satisfactory, particularly with respect to accurately

    modeling long term quantitative and qualitative features of respiration patterns. This paperdescribes a number of modifications

    of the original methods and makes a comparison of the improvements the various modifications gave. The modifications made

    were increasing the class of basis function, broadening the range of possible embedding strategies, improving the optimization

    of the likelihood of the model parameters and calculating a closer approximation to description length. The criteria used in

    the comparisons were description length, root-mean-square prediction error, model size, free-run behavior and amplitude size

    and variation.

    We use surrogate data analysis to confirm the hypothesis that the recorded data are consistent with the nonlinear models we

    construct, and not consistent with simpler models. We also investigate the free-run dynamics of the model systems to see if the

    models exhibit features consistent with physiological characteristics observed independently in the respiratory recording. This

    involved modeling breathing patterns prior to a sigh and onset of a phenomenon called periodic breathing; and comparing

    the period of cyclic amplitude modulation of the free-run dynamics of the models and the period of the subsequent periodic

    breathing observed in respiration recording. The periods were consistent in six out of seven recordings. Copyright 1998Elsevier Science B.V.

    Keywords: Radial basis modeling; Description length; Surrogate analysis; Infant respiratory patterns; Periodic breathing

    1. Introduction

    This paper describes an attempt to accurately model

    the respiratory patterns of human infants using new

    nonlinear modeling techniques. We have identified

    periodic fluctuation in regular breathing pattern of

    sleeping infants using linear modeling techniques

    Corresponding author. Tel.: +618 9380 3348; fax: +6189380 1028; e-mail: [email protected].

    [3]. An accurate, reliable and replicable method of

    building nonlinear models may further aid the iden-

    tification of such subtle periodicities and give some

    insight into the mechanisms generating them. Just as

    a differential equation model of a system can lead to

    greater understanding, so too can numerical, nonlinear

    models. In this paper it is our aim to produce accu-

    rate models of nonlinear systems (infant respiration)

    from a scalar time series measurement of that system

    (abdominal volume).

    0167-2789/98/$19.00 Copyright 1998 Elsevier Science B.V. All rights reserved

    PII S 0 1 6 7 - 2 7 8 9 ( 9 7 ) 0 0 3 1 1 - 4

  • 7/27/2019 Physic Ad 117

    2/16

    284 M. Small, K. Judd/ Physica D 117 (1998) 283298

    Initially we used a radial basis modeling algorithm

    described by Judd and Mees [1] to model recordings

    of the abdominal movements of sleeping infants. Al-

    though these radial basis models give accurate short-

    term predictions, they were not entirely satisfactory in

    the sense that simulations of the models failed to ex-

    hibit some characteristics of the original signals. Aftersome alteration of the model building algorithm, much

    better results were obtained; simulations of the models

    exhibit signals that are nearly indistinguishable from

    the original signals.

    In this paper we first describe the time series we

    will model and briefly review the nonlinear modeling

    methods of Judd and Mees [1]. We identify some fail-

    ings of simulations of models produced by this algo-

    rithm; suggest modifications that may overcome these

    problems; and finally demonstrate the improved re-

    sults we have obtained.

    1.1. Data collection

    For this study we collected measurements pro-

    portional to the cross-sectional area of the abdomen

    of infants during natural sleep (using standard non-

    invasive inductive plethysmography techniques). Such

    measurements are a gauge of lung volume.

    Nineteen healthy infants were studied at one, two,

    four and six months of age, in the sleep labora-

    tory at Princess Margaret Hospital. The study was

    approved by the Princess Margaret Hospital ethics

    committee.The unfiltered analog signal from an inductance

    plethysmograph was passed through a DC ampli-

    fier and 12 bit analog to digital converter (sampling

    at 50 Hz). The digital data were recorded in ASCII

    format and were then transferred to Unix work-

    stations at the University of Western Australia for

    analysis.

    The only practical limitation on the length of time

    for which data could be collected is the period when

    the infant remains asleep and still. The cross-sectional

    area of the lung varies with the position of the infant.

    However, in this study we are interested only in the

    variation due to the breathing and so we have been

    careful to avoid artifact due to changes in position or

    band slippage. We have made observations of up to

    2 h that are free from significant movement artifacts,

    although typically observations are in the range 5

    30 min.

    1.2. Pseudo-linear radial basis modeling

    We have previously used these data to estimate the

    correlation dimension of the respiratory patterns of

    sleeping infants [2], and to identify cyclic amplitude

    modulation (CAM) in respiration during quiet sleep

    [3]. Both these studies concluded that linear model-

    ing techniques were unable to model the dynamics

    of human respiration. 1 Furthermore, by comparing

    the correlation dimension estimates for the data and

    surrogates we were able to demonstrate that simula-

    tions from radial basis models produced dimension

    estimates that closely resembled that of the data [2].This implies that nonlinear models are more accurately

    modeling the data than are linear models. However,

    these nonlinear models appeared to have difficulty

    with some data sets, most notably those with sub-

    stantial noise contamination and data exhibiting non-

    stationarity. In this paper we attempt to improve the

    modeling techniques.

    2. Modeling respiration

    In this section we introduce the data set that we

    will attempt to model. We use correlation dimension

    estimation and false nearest neighbor techniques to

    determine a suitable embedding dimension and exam-

    ine three alternative criteria for embedding lag to de-

    duce an appropriate value. We then apply the nonlinear

    modeling technique described by Judd and Mees [1] to

    this data set and examine the weaknesses of the result.

    1 By calculating correlation dimension dc(0) for data embed-

    ded in ,3 4 and 5 as a test statistic surrogate analysis of 27

    recordings of infant respiration from 10 infants concluded thatthe data were inconsistent with each of the linear hypothesesaddressed by Theiler et al. [4]. The details of these calculationsare beyond the scope of this discussion, see [2,5].

  • 7/27/2019 Physic Ad 117

    3/16

    M. Small, K. Judd / Physica D 117 (1998) 283298 285

    Fig. 1. Data. The data we use in our calculations. The solid line represents the data set from which we build our radial basis models.The horizontal axis is time elapsed from the start of data collection and the vertical axis is the output from the analog to digitalconvertor (proportional to cross-sectional area measured by inductance plethysmography). Note the sigh (at about 300 s) and the

    onset of periodic breathing following this.

    Fig. 2. Periodic breathing. An example of a short episode of periodic breathing after a sigh (at 580 s on the second panel). Smaller

    sighs are also present at about 275 and 470 s on the first panel. The horizontal axis is time elapsed from the start of data collectionand the vertical axis is the output from the analog to digital convertor (proportional to cross-sectional area measured by inductanceplethysmography).

    2.1. Data

    For much of the following sections we illustrate

    the calculation and comparison using just one record-

    ing, selected because it is a typical representation

    of a range of important dynamical features. The data

    set we use (see Fig. 1) is from a section of approx-

    imately 10 min of respiration of a two-month-old

    female in quiet (stages 3 and 4) sleep. These data

    exhibit a physiological phenomenon of great interest

    to respiratory specialists known as periodic breath-

    ing [6,7]. Periodic breathing is simply extreme CAM

    the minimum amplitude decreases to zero. Fig. 2

    shows an example of periodic breathing. In all other

    respects these data are typical of many of our record-

    ings. The section which we examine first is from a

    period of quiet sleep preceding the onset of periodic

    breathing.

  • 7/27/2019 Physic Ad 117

    4/16

    286 M. Small, K. Judd/ Physica D 117 (1998) 283298

    Fig. 3. False nearest neighbors. False nearest neighbor calculation for the data illustrated in Fig. 1 (1600 points sampled at 12 .5Hz)embedded with a time delay embedding,

    =5 (Rtol

    =15).

    2.2. Embedding

    Using a time delay embedding strategy and appeal-

    ing to Takens [8] embedding theorem we produce

    from our scalar time series

    y1, y2, y3, . . . ,

    a d-dimensional vector time series. Consider the

    embedding

    yt zt = (yt, yt2, . . . , ytd) t > d.To perform this transformation one must first identify

    the embedding lag and the embedding dimension d.

    We describe the selection of suitable values of these

    parameters in the following paragraphs.

    Any value of is theoretically acceptable, but the

    shape of the embedded time series will depend criti-

    cally on the choice of and it is wise to select a value

    of which separates the data as much as possible.

    General studies in nonlinear time series [9] suggest the

    mutual information criterion [10], the autocorrelation

    function [11] or one of several other criteria to choose

    . Our experience and numerical experiments sug-

    gest that selecting a lag approximately equal to one

    quarter of the quasi-period of the time series pro-

    duces comparable results to the autocorrelation func-

    tion but is more expedient. Note that the first zero

    of the autocorrelation function will be approximately

    the same as one quarter of the quasi-period if the

    data are almost periodic. Numerical experiments with

    these data have shown that either of these methods

    produce superior results to the mutual information

    criterion.

    Suitable bounds on d can be deduced by using afalse nearest neighbor analysis [12]. Numerical ex-

    periments indicate that four dimensions are sufficient

    to remove false nearest neighbors from the data, see

    Fig. 3. Furthermore, it is at approximately this em-

    bedding dimension that the correlation dimension es-

    timates appear to plateau. Takens sufficient condition

    on successful recreation of the attractor by embedding

    requires d > 2dc + 1 where dc is the correlation di-mension of the attractor. For our data with 3 < dc 4(see [2]), this would suggest that d > 8 is necessary.

    However, embedding in this dimension offers no im-

    provement to the modeling process and our false near-

    est neighbor calculations indicate that a much smaller

    value of d is sufficient.

  • 7/27/2019 Physic Ad 117

    5/16

    M. Small, K. Judd / Physica D 117 (1998) 283298 287

    Fig. 4. Initial modeling results. Free-run prediction and noise-driven simulation of a radial basis model. The plot on the left is a

    free-run prediction with no noise, on the right is a simulation driven by Gaussian noise at 10% of the root-mean-square prediction error

    (

    ti=1

    2i

    /

    N). The horizontal axis is yt for t = 1, . . . , 500, the vertical axis is the output from the analog to digital convertor(proportional to cross-sectional area measured by inductance plethysmography). From 30 trials, 27 of them exhibited fixed points.

    2.3. Modeling

    We attempt to build the best model of the form

    yt+1 = f (zt) + t,where t is the model prediction error and f :

    d is of the form

    f (zt) = 0 +n

    i=1i ytli

    +m

    j=1j+n+1

    zt cjrj

    , (1)

    where rj and j are scalar constants, 1 li < li+1 d are integers and cj are arbitrary points in

    d.

    The integer parameters n and m are selected to min-imize the description length [10] as described in [1].

    Here () represents the class of radial basis functionfrom which the model will be built. We choose to use

    Gaussian basis functions because they appear to be

    capable of modeling a wide variety of phenomena.

    The data set consists of 20 000 points sampled at

    50 Hz. This is oversampled for our purposes and we

    thin the data set to one in four points and truncate it

    to a length of 1600 (see Fig. 1). We set d = 4 andchoose = 5.

    Trials with the modeling algorithm as described in

    [1] produced some problems with the model simula-

    tions (see Fig. 4). None of the simulations look like

    the data. When periodic orbits are evident they are still

    unlike the data; the waveform is symmetric, whereas

    the data have a definite asymmetry. Moreover the free-

    run models often exhibit stable fixed points. This is

    extremely undesirable as it is evidently not an accu-

    rate representation of the dynamics of respiration

    breathing does not tend to a fixed point, usually.

    The remainder of this paper shall be concerned with

    addressing these problems. These problems are the

    result of three main deficiencies in the initial mod-

    eling algorithm: (i) it over fits the data; (ii) it does

    not produce appropriate simulations; and (iii) models

    are not consistent or reproducible. We will attempt to

    improve upon these problems whilst considering the

    many competing criteria for a good model.

    3. Improvements

    Before we can attempt to improve our modeling

    procedure we must be clear on what we mean by im-

    provement. There are several criteria that might be

    imposed to achieve a good model.

    Modeling criteria measure quantities such as the

    number of parameters in the model, its prediction error

    and description length. It is desirable to have a model

    with few parameters, a small description length and a

    small root mean square prediction error.

    Algorithmic criteria are concerned with optimizing

    the modeling algorithm, to ensure that it searches the

    broadest possible range of basis functions as efficiently

  • 7/27/2019 Physic Ad 117

    6/16

    288 M. Small, K. Judd/ Physica D 117 (1998) 283298

    as possible. Unfortunately a larger search space comes

    at the expense of more computation.

    Qualitative criteria consider properties of the dy-

    namics of models; for example, the behavior observed

    in the simulations of the model. In modeling breath-

    ing, for example, we expect something like stable

    periodic (or quasi-periodic) solutions; divergence orstable fixed points seem unlikely. Furthermore we ex-

    pect the shape of the periodic solution to closely match

    the shape of the data and to occupy the same region

    of phase space.

    Modeling results should also be reproducible and

    representative. It does not seem unreasonable to ex-

    pect consistent, repeatable results from a modeling

    algorithm, both qualitatively and quantitatively. Re-

    producibility can be examined by repeatedly modeling

    a single data set. Furthermore, the model should be

    representative in that when making many simulations

    of the model, we ought to obtain time series of whichthe original data are representative. Representativity

    can be measured with the assistance of surrogate tests

    using a statistic such as the correlation dimension es-

    timates or cyclic amplitude modulation.

    In the following sections we consider improvements

    of the basic modeling procedure by: (i) broadening the

    class of basis functions; (ii) using a more targeted se-

    lection algorithm; (iii) making more accurate estimates

    of description length; (iv) local optimization of nonlin-

    ear parameters; (v) using reduced linear modeling to

    determine embedding strategies; and (vi) simplifying

    the embedding strategies using a form of sensitivityanalysis.

    3.1. Basis functions

    In this section we introduce a broader class of ba-

    sis functions. This will produce an algorithm that is

    capable of modeling a wider range of phenomena.

    First we expand the embedding strategy so that

    instead of radial (spherical) basis functions we in-

    troduce cylindrical basis functions. Detailed argu-

    ments about the advantages of these basis functions aredescribed elsewhere [13]. Generalize the functional

    form (1) to

    f (zt) = 0 +n

    i=1i ytli

    +m

    j=1j+n+1

    Pj(zt cj)rj

    , (2)

    where li , rj, j, cj, n, m are as described previously

    and Pj d: dj (dj < d) are projections ontoarbitrary subsets of coordinate axes.

    The functions Pj can be thought of as a local em-

    bedding strategy. Each basis function has a different

    projection Pj and so each Pj(zt cj) is dependenton a different set of coordinate axes.

    We actually generalize the choice of embed-

    ding strategy further by selecting the best lags

    from the set {0, 1, 2, . . ., d}, not only subsets of{0, , 2 , . . . , d }. It seems that by allowing the se-lection of different embedding strategies in different

    parts of phase space the model gives better free-run

    behavior. This indicates that, naturally enough, the

    optimal embedding strategy is not uniform over phase

    space. Selecting from this larger set of embedding

    lags is equivalent to embedding with a time lag of 1 in

    d. However, the modeling algorithm rarely selects

    more than a d-dimensional local embedding. There-

    fore, these improved results are not contrary to our

    previous estimates of optimal embedding dimension.

    They do allow for an embedding in more than 2 dc +1dimensions (satisfying Takens sufficient condition) if

    necessary. As noted earlier the choice of embedding

    lag is largely arbitrary.Furthermore, to increase the curvature of the basis

    functions we replace the choice of

    (x) = exp

    x2

    2

    by

    (x,) = exp

    (1 ) x

    ,

    where 1 < < R is the curvature 2 and (1 )/ is acorrection factor so that

    (x,) dx = 1. Hence,

    maintaining consistent notation

    2 To prevent large values of the second derivative of f it isnecessary to provide an upper bound R on .

  • 7/27/2019 Physic Ad 117

    7/16

    M. Small, K. Judd / Physica D 117 (1998) 283298 289

    (x,) =

    2(1 )

    x/2

    ,

    and the basis functions become functions of the form

    2(1 j)

    j

    Pj(zt cj)rj

    j/2

    ,

    where

    (x) = exp x2

    2.

    Broadening the class of basis functions has in-

    creased the complexity of the search algorithm.

    Hopefully, it will also have broadened the search

    space sufficiently to encompass functions which can

    more accurately model the data. To overcome this

    increased search space we consider a more efficient

    search algorithm.

    3.2. Directed basis selection

    The method of Judd and Mees [1] involves ran-

    domly generating a large set of basis functions

    {(z cj/rj)}j = {j}j and evaluating them ateach point of the embedded time series z to give the

    matrix V = [1|2| |M]. Following an iterativescheme they repeatedly select columns from this ma-

    trix (and the corresponding candidate basis function)

    to add to the optimal model.

    Instead, we select a new set of candidate basis func-

    tions {j}j (and a new matrix V) at each expansion ofthe optimal model. We then identify the column k ofV

    that best fits the residuals (orthogonal to the previously

    selected basis functions) and select the corresponding

    basis function k . All the other candidate basis func-

    tions {j}j=1,...,M;j=k are ignored and forgotten at thenext iteration. Because a new set of basis functions

    is selected at each expansion, all the candidate basis

    functions are much more appropriately placed. 3

    The improvement in modeling achieved by this will

    require a greater computation time. Furthermore, the

    selection of basis functions that more closely fit the

    3 Basis functions are selected according to either a uniformdistribution or the probability distribution induced by the mag-nitude of the modeling prediction error.

    data may, possibly, increase the number of basis func-

    tions allowed by the description length criterion. To

    alleviate this problem we introduce a harsher more

    precise version of description length.

    3.3. Description length

    The minimum description length criterion, sug-

    gested by Rissanen [10], is used by Judd and Mees

    [1] to prevent over fitting. However, the original im-

    plementation of minimum description length used by

    Judd and Mees only provides a description length

    penalty for the coefficient j of each of the radial ba-

    sis functions (and linear terms). Each basis function

    also has a radius rj and coordinates cj which must

    also be specified to some precision, and hence should

    also be included in the description length calculation.

    In [1] j is j truncated to some finite precision j,

    then the description length is expressed as

    L(z, ) = L(z|) + L(), (3)where

    L(z|) = ln P (z|)is the description length of the model prediction errors

    (the negative log likelihood of the errors) and

    L() m+n+1

    j=1ln

    j

    is the description length of the truncated parameters,

    is an inconsequential constant.

    We generalize Eq. (3) and include the finite preci-

    sions ofrj and cj. Let represent the vector of all the

    model parameters (j, cj, and rj) and the truncation

    of those parameters to precision . Then

    L(z, ) = L(z|) + L(), (4)where

    L() (d+2)m+n+1

    j=1

    ln

    j.

    Now the problem becomes one of choosing to

    minimize (4). By assuming that is not far from the

  • 7/27/2019 Physic Ad 117

    8/16

    290 M. Small, K. Judd/ Physica D 117 (1998) 283298

    maximum likelihood solution (see Section 3.4), one

    can deduce that

    L(z, ) L(z|) + 12 TQ

    + k ln k

    j=1ln j, (5)

    where k = ((d+ 2)m + n + 1). Minimizing (5) gives(as in [1])

    (Q)j = 1/j,

    where Q = DL(z|) is the second derivative ofthe negative log likelihood with respect to all the

    parameters.

    Although algebraically complicated, this expression

    can be solved relatively efficiently by numerical meth-

    ods. However, by assuming that the precision of the

    radii and the position of the basis function must be

    approximately the same, 4 one can circumvent a great

    deal of the computational difficulty, and simply cal-

    culate the precision of j assuming the same values

    for the corresponding precisions of the coordinates cj.

    Much of the computational complexity of calculat-

    ing description length could be avoided by utilizing

    the Schwarz criterion [14]. Indeed, from experience

    it appears that the Schwarz criterion gives compara-

    ble size models. However, Schwarz criterion does not

    take into account the relative accuracy of different ba-

    sis functions an important feature of minimum de-

    scription length.

    3.4. Local optimization of nonlinear model

    parameters

    Once the best (according to sensitivity analysis) ba-

    sis function has been selected we improve on its place-

    ment by attempting to maximize the likelihood

    P (z|, 2)

    = 1(2 2)N/2

    exp(y V )T(y V )

    22,

    4 Since a slight change in radii will affect the evaluation of abasis function over phase space in the same way as an equalsmall change in the position of the basis function.

    where y V = is the model prediction error, and2 is the standard deviation of the (assumed to be)

    Gaussian error. By setting 2 =ti=1 2i /N and tak-ing logarithms one gets that

    ln P (z

    |)

    =N

    2 +ln2

    N

    N/2

    + ln

    ti=1

    2i

    N/2. (6)

    To maximize the likelihood we optimize Eq. (6) by

    differentiating ln (t

    i=1 2i )

    N/2 with respect to rj, cj,

    and j. This calculation is algebraically messy, but

    computationally straightforward provided a good op-

    timization package is used.

    3.5. Reduced linear modeling for embeddingselection

    Allowing different embedding strategies from such

    a wide class (due to the expansion of the class of basis

    functions in Section 3.1) increases the computational

    complexity of the modeling process. However, to cir-

    cumvent this we note that for Gaussian basis functions

    the first-order Taylor Series expansion gives

    Pj(zt cj)rj

    = d

    i=1 pi (zt cj)2rj

    d

    i=1

    |pi (zt cj)|rj

    , (7)

    where pi :d is the coordinate projection onto

    the ith coordinate. We then build a minimum descrip-

    tion length model of the residual of the form (7). From

    this we deduce that the basis functions selected are

    a good indication of an appropriate embedding strat-

    egy. Although this method is approximate it is hoped

    that this will provide a useful and efficient innovation

    within the modeling algorithm.

  • 7/27/2019 Physic Ad 117

    9/16

    M. Small, K. Judd / Physica D 117 (1998) 283298 291

    Fig. 5. Improved modeling results. Free-run prediction and noise-driven simulation of a radial basis model. The plot on the left is afree-run prediction with no noise, on the right is a simulation driven by Gaussian noise at 10% of the root-mean-square prediction

    error (

    ti=1

    2i

    /

    N). The horizontal axis is yt for t = 0, . . . , 500, the vertical axis is the output from the analog to digitalconvertor (proportional to cross-sectional area measured by inductance plethysmography).

    3.6. Simplifying embedding strategies

    Our final, very rudimentary alteration is designed

    to account for some of the approximation required in

    the reduced linear modeling of the embedding strate-gies. Given an embedding strategy suggested by the

    method of Section 3.5, we generate additional can-

    didate basis functions by using embedding strategies

    whose coordinates are subsets of the coordinates of the

    embedding strategy suggested by the linear modeling

    methods.

    4. Results

    After implementing the alterations described in the

    preceding section, we again apply our methods tothe same data set. This section describes the results

    of these calculations and examines some of the im-

    provements in the final model. We also examine the

    individual effect of each modification and the effec-

    tiveness of this modeling procedure in seven different

    data sets (from six infants). Because of its physiologi-

    cal significance, all the data sets selected for this anal-

    ysis exhibit CAM suggestive of periodic breathing. We

    compare dimension estimates for the original data sets

    and simulations from the models. Finally, we apply

    a linear modeling technique discussed elsewhere [3]

    to detect CAM within the respiratory traces of sleep-

    ing human infants, and present some results. That is,

    we compare the CAM present in the data following a

    sigh to that present in the models built from the data

    preceding the sigh.

    4.1. Improved modeling

    Fig. 5 shows a section of free-run prediction, and

    noisy simulation for a representative model. Using

    an interactive three-dimensional viewer (see Fig. 6),

    it is possible to determine that these models also have

    many more common structural characteristics than

    those created in Section 2.3. The size, placement,

    shape and local embedding dimensions of the basis

    functions of the models have many similarities.

    Importantly, all of these models have a similar free-

    run behavior. The free-run predictions are as large (in

    amplitude) as the data; this was a substantial problem

    with the original modeling procedure. Moreover, thefree-run behavior with noise appears more realistic

    and the shape of the simulations mimics very closely

    that of the data. Fig. 7 shows a short segment of a

    simulation, along with the data. Note the similarities

    in the shape of the prediction and the data. Finally,

    the simulations exhibit a measurable cyclic amplitude

    modulation which we use in Section 4.3.2 to infer

    the presence of cyclic amplitude modulation in the

    original time series.

    4.2. Effect of individual alterations

    Table 1 lists some characteristics of models built

    from the data in Fig. 1 using various methods. The

  • 7/27/2019 Physic Ad 117

    10/16

    292 M. Small, K. Judd/ Physica D 117 (1998) 283298

    Fig. 6. Cylindrical basis model. A pictorial representation of the interactive three-dimensional viewer we used. The axes range from

    1.715415 to 3.079051, the same range of values as the data. The point (1.7, 1.7, 1.7) is in the front center, foreground.The cylinders, prisms and sphere represent the placement (cj) and size (rj) of different basis functions with different embeddingstrategies. the X, Y, and Z coordinates shown correspond to yt, yt5, and yt15, respectively. The coloring of the basis functionsrepresents the value of the coefficients (j).

    different modeling strategies are: (A) the initial

    method (described in Section 2.3); (B) extended ba-

    sis functions and embedding strategies (Section 3.1);

    (C) directed basis selection (Section 3.2); (D)

    more accurate approximation to description length

    (Section 3.3); (E) local optimization of nonlin-

    ear model parameters (Section 3.4); (F) reduced

    linear modeling to select embedding strategies

    (Section 3.5); and (G) simplifying embedding strate-

    gies (Section 3.6). These alterations to the algorithm

    were progressively added in various combinations

    and characteristics of the observed models measured.

    The initial procedure (A) produced very bad free-

    run predictions; 27 out of 30 trials produced simula-

    tions with fixed points. Extending the class of basis

    functions and adding cylindrical basis functions (B)

    vastly improved this (only eight out of 30 simula-

    tions did not have periodic (or quasi-periodic) orbits).

    Most of the periodic orbits in these simulations were

    smaller than the data (did not occupy the same part

    of phase space) and one divergent simulation was ob-

    served (hence the large standard deviation in Table 1).

    This approach decreased the prediction error without

    affecting either the model size or description length

    (clearly, the required precision of the parameters was

    greater).

    Directed basis selection (C) greatly increased the

    size of the model and decreased error whilst improv-

    ing free-run behavior not only in amplitude but also

    shape. The increase in computational time could al-

    most entirely be due to the greater model size. Improv-

    ing the description length calculation (D) decreased

    the model size whilst, predictably increasing predic-

    tion error. This also caused a surprising increase in

  • 7/27/2019 Physic Ad 117

    11/16

    M. Small, K. Judd / Physica D 117 (1998) 283298 293

    Fig. 7. Short-term behavior. Comparison of simulation and data. The solid line is the data, the dotdashed is a free-run prediction,

    the dashed is a simulation driven by noise (20% of

    ti=1 2i /N). The initial conditions for the artificial simulations are identical

    and are taken from the data. The vertical axis is the output from the analog to digital convertor (proportional to cross-sectional area

    measured by inductance plethysmography).

    Table 1

    Algorithmic performance: Comparison of the modeling algorithm with various improvements

    Modeling method Nonlinear parameters RMS error MDL Free-run amplitude CPU time (s)ti=1

    2i

    /

    N

    A 12.52.4 0.1350.016 1086157 0.000.91 283103A + B 12.52.4 0.1130.011 1090155 1.2231.90 292113A + B + C 24.54.3 0.1040.015 1123198 1.581.04 567167A + B + D 10.72.3 0.1220.016 975191 0.3424.91 744504A + B + C + D 14.53.5 0.1230.018 909210 1.501.09 16261203A + B + C + D + E 9.52.9 0.1410.012 735131 1.591.31 20471253A + B + C + D + F 13.73.6 0.1260.009 87081 1.3117.48 43481959A + B + C + D + E + F 11.03.1 0.1170.013 990119 1.1717.94 58422973A + B + C + D + E + F + G 11.43.2 0.1170.011 980110 1.871.00 57532360

    The seven different modeling procedures are the initial routine described by Judd and Mees, and six alterations described in

    Section 3. Modeling methods are: (A) the initial method; (B) extended basis functions and embedding strategies; (C) directed basisselection; (D) exact description length; (E) local optimization of nonlinear model parameters; (F) reduced linear modeling to select

    embedding strategies; and (G) simplifying embedding strategies. Results are from 30 attempts at modeling the data described inSection 2.1 and Fig. 1. The numbers quoted are (mean value)(standard deviation). Calculations were performed on a SiliconGraphics Indy running at 133MHz with 16 Mbytes of RAM. CPU time is measured in seconds using MATLABs cputime

    command.

    calculation time an indication of the computational

    difficulty solving (Q)j = 1/j, where Q is thesecond derivative with respect to all the model para-

    meters (or at least and r). Because there is a harsh

    penalty these models are far less likely to be over

    fitting the data. Combining the improved description

  • 7/27/2019 Physic Ad 117

    12/16

    294 M. Small, K. Judd/ Physica D 117 (1998) 283298

    Fig. 8. Periodic breathing. An example of periodic behavior in one of our data sets. The solid region was used to build a nonlinear

    radial basis model. Note that periodic breathing begins immediately after the sigh. The vertical axis is the output from the analog todigital convertor (proportional to cross-sectional area measured by inductance plethysmography).

    length calculation and directed basis selection pro-

    duced models comparable in both size and fitting error

    to before either alteration was implemented (A + B).However, free-run behavior had an amplitude closer

    to the mean amplitude of the data and exhibited anasymmetric waveform similar to the data.

    Addition of the nonlinear optimization (E) and local

    linear modeling (F) routines caused the greatest in-

    crease to computational time. Individually these meth-

    ods did not offer any considerable improvement to

    the other model characteristics. However, many of the

    statistics indicate a decrease in the variation between

    trials. Combined, these modifications gave a slight im-

    provement in prediction error and description length

    whilst making the model smaller. They produced more

    realistic simulations although the amplitude was

    smaller than that of the data.Finally, the simple procedure of checking that sim-

    pler embedding strategies would not produce better

    (or equally good) results (G) caused a substantial im-

    provement. This is perhaps due in part to the previous

    optimization and local linear methods, particularly the

    approximate nature of the local linear modeling. Re-

    moving the coordinates helped produce some appre-

    ciable improvement in suitability of the embedding

    strategies suggested by the approximate local linear

    methods. The local linear methods often produce a

    high-dimensional local embedding (many significant

    coordinates), eliminating some of these will usually

    only slightly increase the prediction error. This sim-

    ple addition increases the amplitude to a realistic level

    (approximately 1.9 whilst the mean breath size for the

    data is about 2.3 5) whilst decreasing the proportion

    of fixed point and divergent trajectories to the lowest

    level (8 and 0 of the 30 models, respectively) with-

    out appreciably changing the description length, pre-diction error, or model size whilst decreasing slightly

    the calculation time (and variance in calculation time).

    Furthermore, these models have far more structural

    similarities (in the size and placement of basis func-

    tions) than the previous models have, indicating that

    these models are far more consistent.

    The remainder of this section is devoted to some

    applications of these modeling methods and tests of

    their representability.

    4.3. Modeling results

    From over 200 recordings of 19 infants, we identi-

    fied seven data sets from six infants for a more careful

    analysis. All seven of these data sets include a sigh

    followed by a period of breathing exhibiting cyclic

    amplitude modulation (CAM) [3]. Our present discus-

    sion examines the analysis of these data sets.

    In this section we examine the free-run behavior

    of data sets created from seven models of seven data

    sets from six sleeping infants. We compare the corre-

    lation dimension of the data and simulations from mo-

    dels. Following this we compare the period of CAM

    5 Note, however, that the data are slightly nonstationary whilstthe model is not.

  • 7/27/2019 Physic Ad 117

    13/16

    M. Small, K. Judd / Physica D 117 (1998) 283298 295

    Fig. 9. Surrogate calculations. Comparison of dimension estimates for data and surrogates. The three figures on the left are dimension

    estimates (for embedding dimension from 3 to 5, shown from top to bottom) for a model of Bs2t8. The left three plots aresimilar results for a model of Ms1t6. All surrogates are simulation driven by Gaussian noise with a standard deviation of half theroot-mean-square one step prediction error. Each picture contains one dimension estimate for the data (solid line), and 30 surrogates

    (dotted). The two data sets used in these calculations are shown in Figs. 1 and 8, respectively.

    detected in the free-run predictions from the models to

    that visually evident after a sigh. Fig. 8 illustrates oneof the data sets used in our analysis. This is the only

    set of data to exhibit periodic breathing, the others

    merely exhibited strong amplitude modulation after

    the sigh for 2560s ( 1530 breaths). Neverthelessthe change that the respiratory system undergoes after

    a large sigh is of great interest to respiratory physio-

    logists. We examine the system before and after a sigh

    to determine evident physiological similarities in the

    mechanics of breathing.

    For each of our seven data sets, we identify the lo-

    cation of the sigh, and extract data sets of 1501 points

    spanning 120 s preceding the sigh. From these data

    sets the respiratory rate of each recording was estab-

    lished and the period of respiration deduced. Each data

    set was embedded in 4 with a lag equivalent to the

    integer closest to one quarter of the period. We thenapplied our modeling algorithm.

    4.3.1. Surrogate analysis

    To determine exactly how similar data and model

    simulations are, we employ an obvious generalization

    of the surrogate data analysis used by Theiler et al. [4].

    The principle of surrogate data is the following. One

    first assumes that the data come from some specific

    class of dynamical system, possibly fitting a paramet-

    ric model to the data. One then generates surrogate

    data from this hypothetical system (producing simu-

    lations from the parametric model) and calculates var-

    ious statistics of the surrogate and original data. The

    surrogate data will give the expected distribution of

  • 7/27/2019 Physic Ad 117

    14/16

    296 M. Small, K. Judd/ Physica D 117 (1998) 283298

    Table 2

    Periodic behavior: Comparison of CAM after apnea (apparent to visual inspection), the second set of results, and CAM detected in

    the models limit cycle, the first set of results

    Subject Sex Age Model CAM in free-run CAM after sigh

    (months) size (breaths) (s) (breaths) (s)

    A(As4t2) Male 6 8 (7) 56 a 14 a 5 25Bb(Bs2t8) Female 2 7 (6) 6 9 6 9Bb(Bs3t1) 6 (5) 5 10 5 10

    G(Gs2t4) Female 2 4 (3) 5 11 5 9H(Hs1t2) Male 1 5 (3) 89 a 11 a 9 13

    M(Ms1t6) Female 1 6 (4) None None 5 14.5R(Rs2t4) Male 2 8 (6) 9 18 8 16

    Data sets Ms1t6 and Bs2t8 exhibited periodic breathing. For each data set marked cyclic amplitude modulation (CAM) occurred

    after a sigh and was measured by inspection. Radial basis models were built on a section of quiet sleep preceding the sigh, noisefree limit cycles exhibited periodicities that were measured in both time and breaths from the simulation.

    a Not strictly periodic but rather exhibited a chaotic behavior. Model size is m + n(m), see Eq. (2).

    statistic values and one can check that the original

    have atypical value. If the original data have atypical

    statistics, then we reject the hypothesis that the sys-

    tem that generated the original data is of the assumedclass. One always begins with simple assumptions and

    progresses to more sophisticated models if the data

    are inconsistent with the surrogate data. Theilers [4]

    original paper frames surrogate analysis in the con-

    text of noise-driven linear systems; a recent paper [17]

    presents surrogate analysis in a more general context,

    similar to the above.

    In our case, however, we are not interested in

    determining what type of system generated the data

    at least not at present (we have considered this else-

    where [2]). A simpler null hypothesis (for example

    [15,16]) consistent with the data does not concern ushere. What is of greater interest to us is determin-

    ing if the models really do behave like the data. By

    calculating models and generating free-run predic-

    tions from those models, we are in fact generating

    surrogate data. The similarity of the value of various

    statistics applied to data and surrogate can be used to

    gauge the accuracy of the model. Fig. 9 shows calcu-

    lations of correlation dimension estimates (following

    the methods of Judd [18,19]) for data and surrogate.

    Our calculations indicate a very close agreement be-

    tween the correlation dimension of the data and that of

    the simulations. In six of the seven data sets the corre-

    lation dimension estimate dc(0) for the data is within

    two standard deviations of the mean value of dc(0)

    estimated from the ensemble of surrogates for all val-

    ues of0 for which both converged. In the remaining

    data set the value of correlation dimension differed by

    more than two standard deviations only at the small-est values of 0 (the finest detail in the data). In all

    calculations, dc(0) for the data is within three stan-

    dard deviations of the mean value of dc(0) estimated

    from the ensemble of surrogates. With respect to cor-

    relation dimension, our models are producing results

    virtually indistinguishable from the data.

    4.3.2. Application

    Previously [3], we have used a form of reduced

    autoregressive modeling (RARM) to detect CAM in

    the regular breathing of infants during quiet sleep. We

    apply nonlinear modeling methods here with two aimsin mind: to demonstrate the accuracy of our modeling

    methods; and to further demonstrate that CAM evident

    during periodic breathing and in response to apnea or

    sigh is also present during quiet, regular breathing.

    We have built nonlinear models following the meth-

    ods outlined in this paper of the regular respiration

    of six sleeping infants immediately preceding seven

    sighs and the consequential onset of periodic or CAM

    respiration. For each of these models we produce sim-

    ulations both driven by Gaussian noise, and without

    noise. The noiseless simulations approach a stable pe-

    riodic (or chaotic, quasi-periodic) orbit which may ex-

    hibit slight CAM. Table 2 summarizes the results of

    these calculations.

  • 7/27/2019 Physic Ad 117

    15/16

    M. Small, K. Judd / Physica D 117 (1998) 283298 297

    In all but one data set, CAM was present in the free-

    run prediction of the nonlinear model. The absence of

    CAM in one model may either indicate a lack of mea-

    surable CAM in the data or a poor model (the data

    are illustrated in Fig. 8). All other data sets produced

    nonlinear models that exhibited CAM, the period of

    which matched that observed after a sigh during visu-ally apparent CAM.

    5. Conclusion

    We have successfully modified and applied pseudo-

    linear modeling techniques suggested by Judd and

    Mees [1] to respiratory data from human infants. We

    found that the initial modeling procedure had some

    difficulties capturing all the anticipated features of res-

    piratory motion (it was not periodic). Some alteration

    to the algorithm and a considerable increase to compu-tational time provided results which display dynamics

    very similar to those observed during respiration of

    infants in quiet sleep (not only did the models exhibit

    a periodic limit cycle, but its shape was very similar

    to the data).

    Correlation dimension and the methods of surro-

    gate data demonstrated that the models did indeed pro-

    duce simulations with qualitative dynamical features

    indistinguishable from the data. Short term free-run

    predictions appeared to behave similarly to the data;

    and, most significantly, we were able to deduce the

    presence of CAM in sections of quiet sleep precedingsighs by observing this behavior in free-run predic-

    tions of models built from that data. This confirms our

    earlier observations from linear models of tidal vol-

    ume [3] and the observation of a (greater than) two-

    dimensional attractor in reconstructions from data [2].

    Based on the results of Section 4, we are able to

    deduce that some of the alterations (specifically ex-

    tending the class of basis functions, and directed basis

    selection) improved short-term prediction. Other alter-

    ations reduced the size of the model (accurate approx-

    imation to description length) and improved free-run

    dynamics (extending the class of basis function, local

    optimization and linear modeling methods to predict

    embedding strategies). A combination of these meth-

    ods is required to produce an accurate model of the

    dynamics.

    We conclude that the modeling methods presented

    here and in [1] are capable of accurate modeling

    breathing dynamics (along with a wide variety of other

    phenomena, see for example [20]). Furthermore, we

    have presented some evidence that the CAM presentduring periods of periodic breathing (when tonic

    drive is reduced) is also present, but more difficult to

    observe, during eupnea (normal respiration).

    Acknowledgements

    We wish to thank Madeleine Lowe and Stephen

    Stick of Princess Margaret Hospital for Children for

    supplying the infant respiratory data, and for physio-

    logical guidance.

    References

    [1] K. Judd, A. Mees, On selecting models for nonlinear timeseries, Physica D 82 (1995) 426444.

    [2] M. Small, K. Judd, M. Lowe, S. Stick, Is breathingin infants chaotic? Dimension estimates for respiratory

    patterns during quiet sleep, J. Appl. Physiol., to appear.[3] M. Small, K. Judd, S. Stick, Linear modelling techniques

    detect periodic respiratory behaviour in infants duringregular breathing in quiet sleep, Am. J. Resp. Crit. CareMed., 153 (1996) A79.

    [4] J. Theiler, S. Eubank, A. Longtin, B. Galdrikian, J.D.Farmer, Testing for nonlinearity in time series: The method

    of surrogate data, Physica D 58 (1992) 7794.[5] M. Small, K. Judd, Detecting nonlinearity in experimental

    data, Int. J. Bifurc. & Chaos 8 (1998), in press.[6] D. H. Kelly, D.C. Shannon, Periodic breathing in infants

    with near-miss sudden infant death syndrome, Pediatrics63 (1979) 355360.

    [7] W.B. Mendelson, Human Sleep: Research and ClinicalCare, Plenum Medical Book Company, 1987.

    [8] F. Takens, Detecting strange attractors in turbulence,Lecture Notes in Mathematics, vol. 898, 1981, pp. 366381.

    [9] H.D.I. Abarbanel, R. Brown, J.J. Sidorowich, L.S.Tsimring, The analysis of observed chaotic data in physical

    systems, Rev. Mod. Phys. 65 (4) (1993) 13311392.[10] J. Rissanen, Stochastic Complexity in Statistical Inquiry,

    World Scientific, Singapore, 1989.

    [11] M.B. Priestly, Non-linear and Non-Stationary Time SeriesAnalysis, Academic Press, New York, 1989.

    [12] M.B. Kennel, R. Brown, H.D.I. Abarbanel, Determining

    embedding dimension for phase-space reconstruction using

  • 7/27/2019 Physic Ad 117

    16/16

    298 M. Small, K. Judd/ Physica D 117 (1998) 283298

    a geometric construction, Phys. Rev. A 45 (1992) 34033411.

    [13] K. Judd, A. Mees, Embedding as a modeling problem,Physica D, to appear.

    [14] G. Schwarz, Estimating the dimension of a model, Ann.Stat. 6 (1978) 461464.

    [15] J. Theiler, On the evidence for low-dimensional chaos in an

    epileptic electroencephalogram, Phys. Lett. A 196 (1995)

    335341.[16] J. Theiler, P. Rapp, Re-examination of the evidence

    for low-dimensional, nonlinear structure in the humanelectroencephalogram, Electroencephalogr. Clin. Neuro-

    physiol. 98 (1996) 213222.

    [17] J. Theiler, D. Prichard, Constrained-realization Monte-Carlo method for hypothesis testing, Physica D 94 (1996)221235.

    [18] K. Judd, An improved estimator of dimension and somecomments on providing confidence intervals, Physica D

    56 (1992) 216228.[19] K. Judd, Estimating dimension from small samples,

    Physica D 71 (1994) 421429.

    [20] K. Judd, A. Mees, Modeling chaotic motions of a stringfrom experimental data, Physica D 92 (1996) 221236.