throbber
NIH Public Access
`Author Manuscript
`IEEE Trans Biomed Eng. Author manuscript; available in PMC 2011 June 1.
`Published in final edited form as:
`IEEE Trans Biomed Eng. 2010 June ; 57(6): 1335–1347. doi:10.1109/TBME.2010.2041002.
`
`Characterizing Nonlinear Heartbeat Dynamics within a Point
`Process Framework
`
`Zhe Chen [Member, IEEE]
`Neuroscience Statistics Research Laboratory, Harvard Medical School, Massachusetts General
`Hospital, Boston, MA 02114 USA, and also with the Department of Brain and Cognitive Sciences,
`Massachusetts Institute of Technology, Cambridge, MA 02139 USA (zhechen@neurostat.mit.edu)
`Emery N. Brown [Fellow, IEEE]
`Neuroscience Statistics Research Laboratory, Harvard Medical School, Massachusetts General
`Hospital, Boston, MA 02114 USA, and also with the Harvard-Massachusetts Institute of Technology
`(MIT) Division of Health Science and Technology, and the Department of Brain and Cognitive
`Sciences, MIT, Cambridge, MA 02139 USA (enb@neurostat.mit.edu)
`Riccardo Barbieri [Senior Member, IEEE]
`Neuroscience Statistics Research Laboratory, Harvard Medical School, Massachusetts General
`Hospital, Boston,MA02114 USA, and also with the Massachusetts Institute of Technology,
`Cambridge, MA 02139 USA (barbieri@neurostat.mit.edu)
`
`Abstract
`Human heartbeat intervals are known to have nonlinear and nonstationary dynamics. In this paper,
`we propose a model of R–R interval dynamics based on a nonlinear Volterra–Wiener expansion
`within a point process framework. Inclusion of second-order nonlinearities into the heartbeat model
`allows us to estimate instantaneous heart rate (HR) and heart rate variability (HRV) indexes, as well
`as the dynamic bispectrum characterizing higher order statistics of the nonstationary non-Gaussian
`time series. The proposed point process probability heartbeat interval model was tested with synthetic
`simulations and two experimental heartbeat interval datasets. Results show that our model is useful
`in characterizing and tracking the inherent nonlinearity of heartbeat dynamics. As a feature, the fine
`temporal resolution allows us to compute instantaneous nonlinearity indexes, thus sidestepping the
`uneven spacing problem. In comparison to other nonlinear modeling approaches, the point process
`probability model is useful in revealing nonlinear heartbeat dynamics at a fine timescale and with
`only short duration recordings.
`
`Keywords
`Adaptive filters; approximate entropy (ApEn); heart rate variability (HRV); nonlinearity test; point
`processes; scaling exponent; Volterra series expansion
`
`Correspondence to: Zhe Chen.
`This work was presented at Proceedings of the IEEE Engineering in Medicine and Biology Conference (EMBC), 2008, Vancouver, BC,
`Canada.
`Color versions of one or more of the figures in this paper are available online at http://ieeexplore.ieee.org.
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`APPLE 1034
`
`1
`
`

`

`Chen et al.
`
`Page 2
`
`I. Introduction
`The human heartbeat is regulated by the autonomic nervous system, and as a result, heart rate
`(HR) and heart rate variability (HRV) measurements extracted from the ECG are important
`quantitative markers of cardiovascular control [1]. A healthy heart is influenced by multiple
`neural and hormonal inputs that result in variations of the interbeat interval duration.
`Specifically, various nonlinear neural interactions and integrations occur at the neuron and
`receptor levels, and underlie the complex output of structures such as the sinoatrial (SA) node
`in response to changing levels of sympathetic and vagal activities [55]. The complex nature of
`heartbeat dynamics has been widely considered and discussed in cardiovascular literature.
`Although detailed physiology behind these complex dynamics has not been completely
`clarified, several nonlinearity measures of HRV have been pointed out as important quantifiers
`of complexity of cardiovascular control and have been proved to be of important prognostic
`value in aging and diseases [4], [25], [26],[46], [58], [60], [62].
`
`Many physiological signals are known to be nonlinear and nonstationary. In biomedical
`engineering, various nonlinear indexes, such as the Lyapunov exponent, the fractal exponent,
`or the approximate entropy (ApEn), have been proposed to characterize the nonlinear behavior
`of the underlying physiological system (e.g., [2]). It has been suggested that such nonlinearity
`indexes might provide informative indicators for diagnosing cardiovascular or brain diseases.
`Notably, some difficulties have been often encountered when validating these indexes, such
`as the presence of noise or artifact, the limited size of data samples, or the low sampling rate
`of the observed signals. All these issues shall be kept in mind when new statistical indexes are
`estimated from real signals recorded from a nonlinear system.
`
`In characterizing the nonlinear heartbeat dynamics, both linear and nonlinear system
`identification methods have been applied to R–R interval series [19], [20], [61]. Examples of
`higher order characterization for cardiovascular signals, include nonlinear autoregressive (AR)
`models, Volterra–Wiener series expansion, and Volterra–Laguerre models [2], [32], [33],
`[36]. Several authors have demonstrated the feasibility and validity of nonlinear AR models,
`suggesting that future HR dynamics studies should put greater emphasis on nonlinear analysis
`[19], [20], [31], [61]. However, none of these models have included nonlinear elements in a
`framework based on a precise statistical characterization of the heartbeat generation process,
`and all of mentioned studies used either beat series (tachograms) or discretionarily interpolated
`R–R time series instead of deriving model estimates. In this paper, we apply nonlinear modeling
`to heartbeat dynamics using a point process paradigm. The point process theory is a powerful
`statistical tool able to characterize the probabilistic generative mechanism of the heartbeat at
`each moment in time, thus allowing for estimation of instantaneous HR and HRV measures
`[7], [8]. Furthermore, inclusion of second-order nonlinear terms to the point process model
`offers an opportunity to monitor dynamic higher order spectra indexes [39], [40].
`
`The paper is organized as follows. Section II presents some background on nonlinear system
`identification by Volterra–Wiener series expansion. Section III gives a brief exposition of
`probabilistic point process model theory for heartbeat intervals, derives the instantaneous HR
`and HRV indexes, and reviews the adaptive point process filtering algorithm as well as the
`goodness-of-fit tests. Section IV is devoted to the instantaneous higher order spectral analysis
`and derivation of the dynamic bispectrum estimate, as well as the nonlinearity test for R–R
`interval series. Section V describes the synthetic data generated to test the models, as well as
`two experimental heartbeat datasets. Section VI presents the experimental results on all datasets
`using the point process models, discussing model selection, nonlinearity assessment,
`performance comparison, and irregularity characterization. Finally, discussions and conclusion
`are given in Section VII.
`
`IEEE Trans Biomed Eng. Author manuscript; available in PMC 2011 June 1.
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`2
`
`

`

`Chen et al.
`
`Page 3
`
`II. Volterra Series for Nonlinear System Identification
`The Volterra series expansion, based on the Volterra theorem, is a general method for nonlinear
`system modeling and identification [36]. In functional analysis, a Volterra series denotes a
`functional expansion of a dynamic, nonlinear, and time-invariant function. The Volterra series
`allows for representation of a wide range of nonlinear systems. Because of its generality,
`Volterra series expansion has been widely used in nonlinear modeling in engineering and
`physiology [2], [31], [36]. For instance, computational procedures based on a comparison of
`the prediction power of linear and nonlinear models of the Volterra–Wiener form have been
`applied to measure the complex dynamics of the heartbeats [6]. However, it shall be pointed
`out that all of these nonlinear models used only raw R–R intervals without modeling the point
`process nature of the heartbeats.
`
`Consider a nonlinear single-input and single-output system y = g(x). According to the Volterra
`series theory, the nonlinear system can be expanded by a (finite or infinite) set of kernel
`expansion terms
`
`(1)
`
`where M is the memory of the nonlinear system. Equation (1) only includes up to the second-
`order nonlinear term in the Volterra series expansion; however, inclusion of higher order terms
`is possible. The Volterra kernels {k0,k1,k2,…,} describe the dynamics of the system, each of
`which is associated with Volterra coefficients at different kernel orders and different time lags.
`Estimation of the Volterra coefficients is generally performed by computing the coefficients
`of an orthogonalized series, and then recomputing the coefficients of the original Volterra
`series. A common method is based on the least squares optimization [36]. In this paper, we
`apply a point process adaptive filtering approach to recursively estimate the time-varying
`Volterra coefficients.
`
`III. Heartbeat Interval Point Process Model
`A random point process is a random element whose values are “point patterns” on a set, where
`a point pattern is specified as a locally finite counting measure [23]. Specifically in the time
`domain, a simple 1-D point process consists of series of binary (0 and 1) observations, where
`the variables 1 marks the occurrence times t ∊ [0,∞) of the random events. Mathematically, we
`let N(t) define a continuous-time counting process, and let its differential dN(t) denote a
`continuous-time indicator function, where dN(t) = 1, when there is an event (such as the
`ventricular contraction) or dN(t) = 0, otherwise. Point process theory has been widely used in
`modeling various types of random events (e.g., eruptions of earthquakes, queueing of
`customers, spiking of neurons, etc.) where the timing of the events are of central interest.
`Bearing a similar spirit, the point process theory has been used for modeling human heartbeats
`[7],[8],[16]. The point process framework primarily defines the probability of having a
`heartbeat event at each moment in time. A parametric formulation of the probability function
`allows for a systematic, parsimonious estimation of the parameter vector in a recursive way
`and at any desired time resolution. Instantaneous indexes can then be derived from the
`parameters in order to quantify important features as related to cardiovascular control
`dynamics.
`
`IEEE Trans Biomed Eng. Author manuscript; available in PMC 2011 June 1.
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`3
`
`

`

`Chen et al.
`
`A. Heartbeat Interval
`
`Page 4
`
` detected from the ECG, let RRj = uj −
`Suppose we are given a set of R-wave events
`uj−1 > 0 denote the jth R-R interval, or equivalently, the waiting time until the next R-wave
`event. By treating the R-wave as discrete events, we may develop a point process probability
`model in the continuous time domain [7].
`
`Assuming history dependence, the probability distribution of the waiting time t−uj until the
`next R-wave event follows an inverse Gaussian model:
`
`where uj denotes the previous R-wave event occurred before time t, μRR(t) represents the first-
`moment statistic (mean) of the distribution, and θ > 0 denotes the shape parameter of the inverse
`Gaussian distribution, whose role is to model the tail shape of the distribution (when θ → ∞,
`the inverse Gaussian distribution becomes more like a Gaussian distribution). As p(t) indicates
`the probability of having a beat at time t given that a previous beat has occurred at uj and
`μRR(t) can be interpreted as signifying the most probable moment when the next beat could
`occur. By definition, p(t) is characterized at each moment in time, at the beat as well as in-
`between beats. We can also estimate the second-moment statistic (variance) of the inverse
`Gaussian distribution as
`. The use of an inverse Gaussian distribution to
`characterize the R–R intervals' occurrences is motivated by the fact that if the rise of the
`membrane potential to a threshold initiating the cardiac contraction is modeled as a Gaussian
`random walk with drift, then the probability density of the times between threshold crossings
`(the R–R intervals) is indeed the inverse Gaussian distribution [7]. In [16], we have compared
`heartbeat interval fitting point process models using different probability distributions, and
`found that the inverse Gaussian model achieved the overall best fitting results. The parameter
`μRR(t) denotes the instantaneous R–R mean that can be modeled as a generic function of the
`past (finite) R–R values μRR(t)=g(RRt−1,RRt−2,…,RRt−h), where RRt−j denotes the previous
`jth R–R interval occurred prior to the present time t. In our previous work [8], [14], [16], the
`history dependence is defined by expressing the instantaneous mean μRR(t) as a linear
`combination of present and past R–R intervals (in terms of an AR model), i.e., function g is
`linear. Here, we propose to include the nonlinear terms of past R–R intervals by defining the
`instantaneous RR mean as follows:
`
`(2)
`
`. Here the coefficients a0(t),{ai(t)}, and {bkl(t)} correspond to
`where
`the time-varying zero-, first-, and second-order Volterra kernel coefficients. The zero order
`coefficient a0 accounts for the nonzero mean of the R–R series. Equation (2) can be interpreted
`as a discrete Volterra–Wiener series with degree of nonlinearity d = 2 and memory h = max
`{p, q} [6]. As μRR(t) is defined in a continuous time fashion, we can obtain an instantaneous
`R–R mean estimate at a very fine timescale (with an arbitrarily small bin size Δ), which requires
`no interpolation between the arrival times of two beats. Given the proposed parametric model,
`the nonlinear indexes of the HR and HRV will be defined as a time-varying function of the
`parameters ξ(t)=[a0(t),a1(t),…,ap(t),b11(t),…,bqq(t),θ(t)].
`
`IEEE Trans Biomed Eng. Author manuscript; available in PMC 2011 June 1.
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`4
`
`

`

`Chen et al.
`
`Page 5
`
`B. Instantaneous Indices of HR and HRV
`HR is defined as the reciprocal of the R–R interval. For t measured in seconds, a new variable
`r=c(t−uj)−1 (where c = 60 s/min) can be defined in beats per minute (bpm). By the change-of-
`variables formula, the HR probability p(r)=p(c(t−ut)−1) is given by
`
`and the mean and the standard deviation of HR r can be derived [7], [8], as given by μHR and
`standard deviation σHR, respectively
`
`(3)
`
`(4)
`
`(5)
`
`where
`
` and
`
`.
`
`It is known from point process theory [7], [8], [13] that the conditional intensity function (CIF)
`λ(t) is related to the interevent probability p(t) with a one-to-one relationship
`
`The estimated CIF can be used to evaluate the goodness-of-fit of the proposed probability
`model for the heartbeat interval point process probability model. The quantity λ(t)Δ yields
`approximately the probability of observing a beat during the [t,t + Δ) interval in the sense that
`[23]
`
`(6)
`
`where Ht denotes all of available history information (subject to causality) up to time t.
`
`C. Adaptive Point Process Filtering
`In order to track the unknown parameters of vector ξ in a nonstationary environment, we can
`recursively estimate them via adaptive point process filtering [8]. Upon time discretization, we
`have the following equation updates at discrete-time index k:
`
`(7)
`
`(8)
`
`IEEE Trans Biomed Eng. Author manuscript; available in PMC 2011 June 1.
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`5
`
`

`

`Chen et al.
`
`Page 6
`
`(9)
`
`(10)
`
`where P and W denote the parameter and noise covariance matrices, respectively; Δ =0.005s
` denote the first- and second-
`denotes the time bin size; Δλk = ∂λk/∂ξk and
`order partial derivatives of the CIF with respect to ξ at time t = kΔ, respectively. The indicator
`variable nk = 1 if a heart beat occurs in time ((k−1)Δ,kΔ] and 0 otherwise. The point process
`filtering described in (7) through (10) can be viewed as a “point process analog” of the Kalman
`filtering (for continuous-valued observations). In (7) and (8), the a priori estimates ξk|k−1 and
`Pk|k−1 are computed, while in (9) and (10), the a posteriori estimates ξk|k and Pk|k are computed.
`In (9), [nk − λkΔ] can be viewed as the innovations term computed from the point process filter,
`and (10) is derived based on a Gaussian approximation of the log posterior. Clearly, as the
`innovations term is likely to be nonzero even in the absence of a beat, the parameters are always
`updated at each step.
`
`Once the vector ξ has been estimated within (0, T], one can compute the probability density
`function (pdf) p(t) as well as the CIF estimate λ(t) [from (6)] in time interval (0, T]. Furthermore,
`we can compute the cumulative log-likelihood (denoted by ) of the point process observations
`[13], [23]
`
`(11)
`
`where the indicator variable dN(τ) = 1 (or nk = 1), if a beat occurs at time τ (or within the time
`interval ((k − 1)Δ, kΔ]) and dN(τ) = 0 (or nk = 0), otherwise. The log-likelihood function (11)
`defines the logarithm of the joint probability of all random events (i.e., beats), and the second
`equality holds when the bin size Δ is sufficiently small, and ℓk = nk logλk−λk approximates the
`instantaneous log-likelihood function ℓ(t) = logλ(t)dN(t)/dt−λ(t).
`
`D. Model Selection and Goodness-of-fit Tests
`Our method requires the user to predetermine a proper model order {p, q} for the Volterra series
`expansion. In general, a tradeoff between model complexity and goodness-of-fit arises when
`a point process model is considered. In practice, the order of the model (2) may be determined
`based on the akaike information criterion (AIC) (by prefitting a subset of the data using either
`point process filter or local likelihood method [7], [35]) as well as the Kolmogorov–Smirnov
`(KS) statistic in the post hoc analysis. For different values p and q, we can compare the AIC
`and choose the parameter setup with the minimum AIC value
`
`where dim(ξ) = p + q2 + 2 denotes the dimensionality of parameter vector ξ in the nonlinear
`model. Once the order {p, q} is determined, the initial Volterra coefficients will be estimated
`by the method of least squares [59]: specifically, the coefficients {ai} are optimized by solving
`a Yule–Walker equation for the linear part using the first 200 sample points, and the coefficients
`
`IEEE Trans Biomed Eng. Author manuscript; available in PMC 2011 June 1.
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`6
`
`

`

`Chen et al.
`
`Page 7
`
`{bij} are estimated by fitting the residual error via least squares. Here, we use a separate
`estimation instead of a joint estimation procedure for the Volterra coefficients because we like
`to preserve the interpretation of the linear AR coefficients (such as the stability, which is
`assured by keeping the roots inside the unit circle). A joint estimation procedure is possible
`based on orthogonal projection, cross correlation, or least squares [36],[59], but it may destroy
`the structure described by the linear AR coefficients {ai}, which will be used to estimate the
`parametric AR spectrum defined later.
`
`The goodness-of-fit of the point process model is based on the KS test [13]. Given a point
`process specified by J discrete events: 0 <u1 <…<uJ <T, the random variables
` are defined for j = 1, 2, …, J − 1. If the model is correct, then the variables
`vj = 1−exp(−zj) are independent, uniformly distributed within the region [0, 1], and the variables
`gj = Φ−1(vj) (where Φ(•) denotes the cumulative distribution function (cdf) of the standard
`Gaussian distribution) are sampled from an independent standard Gaussian distribution. To
`compute the KS test, the vj s are sorted from smallest to largest, and plotted against the cdf of
`the uniform density defined as j − 0.5/J. If the model is correct, the points should lie on the 45°
`line. The 95% confidence interval lines are defined as y=x±1.36/(J−1)1/2. The KS distance,
`defined as the maximum distance between the KS plot and the 45° line, is used to measure the
`lack-of-fit between the model and the data. The autocorrelation function of the gj s:
`
`, can also be computed. If the gjs are independent, ACF(m)
`shall be small (around 0 and within the 95% confidence interval 1.96/(J−1)1/2) for any lag m.
`
`V. Quantitative Tools: Spectral Analysis and Nonlinearity Test
`A. Instantaneous Higher Order Spectral Analysis
`Given the Volterra–Wiener expansion for the instantaneous R–R interval mean {μRR(t)}, we
`may compute the time-varying parametric (linear) autospectrum
`
`(12)
`
`By integrating (12) in each frequency band, we may compute the index within the very low
`frequency (VLF) (0.01–0.05 Hz), LF (0.05–0.15 Hz), or HF (0.15–0.5 Hz) ranges. In addition,
` denote the Fourier transform of the second-
`let
`order kernel coefficients {bkl(t)} (all of which together are viewed as discrete samples from a
`2-D impulse response function). From (2), it is known that [39], [40]
`
`(13)
`
`where C(f1,f2t) denotes the bispectrum (Fourier transform of the third-order moment). Note
`that we use the approximation “≈” instead of equality “=” in (13), since the equality only strictly
`holds when the input variables are jointly Gaussian, which is not necessarily true in our case.
`The bispectrum is an important tool for evaluating the presence of nonlinearity in stationary
`time series [9], [39], [40]. From (13), we then can estimate the dynamic bispectrum C(f1,f2,t).
`From the Parseval theorem, we also know that that the sum (or integral) of the square of a
`function is equal to the sum (or integral) of the square of its transform,
` (the second equality follows from
`
`IEEE Trans Biomed Eng. Author manuscript; available in PMC 2011 June 1.
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`7
`
`

`

`Chen et al.
`
`Page 8
`
`. Let b(t) denote a vector that contains
`the conjugate symmetry property), or
`all of coefficients {bkl(t)}, in light of (13), we may compute an index that quantifies the
`fractional contribution of the linear terms on the total power as follows:
`
`(14)
`
`where | · | denotes either the norm of a vector or the modulus of a complex variable. The
`spectrum norm defines the area integrated over the frequency range under the spectral density
`curve. Since the norm units of spectral and bispectral density are the same, their ratio ρ(t) is
`dimensionless (note that the unit of {bkl}{bkl} is in 1/second, and the unit of norm |Q(f, t)| is
`in second, thus their product is unitless). As a function of the estimated parameters, this index
`can also be estimated at each moment in time and is updated at the beat as well as in-between
`beats. The ratio ρ defined in (14) can be viewed as a dynamic counterpart of the following
`static power ratio [footnote: As one reviewer pointed out, the ratio defined in (15) hardly
`reaches close to 0, and the reviewer also suggested to define an alternative ratio index (power
`spectrum – bispectrum)/power spectrum, which is bounded between 0 and 1]:
`
`(15)
`
`where the power spectrum and bispectrum will be calculated by Fourier transform using the
`observed (nonequally spaced) R–R interval time series (with a stationarity assumption). A
`lower value of the ratio (15) implies that the fraction of the bispectral power is higher, thus
`pointing at a more significant nonlinear component in the time series.
`
` are both cycles/
`It shall be noted that the frequency units appearing in
` and
`beat, since the autoregression of μRR (t) is conducted on previous beats instead of previous
`instantaneous {μRR (t–i)}. This alternative modeling strategy, however, would require a large
`number of tags in the linear AR model to compensate for the use of fine timescale. Another
`way to compute the spectra of interest in the unit of cycles/second is to consider the estimated
`{μRR(t)}or {μHR(t)} series and compute the power spectrum or bispectrum using a direct
`method. However, this would require a windowing technique and would not allow for
`instantaneous estimates. As a consequence of the change from cycles/second to cycles/beat,
`certain spectral distortion between the spectrum of counts (SOC) and the spectrum of interval
`(SOI) might be expected [11], [24], [37], [48], especially when the beat-to-beat intervals have
`a large variance. As this issue could become critical when precise estimate of specific
`oscillatory frequencies are needed, its effects are less noticeable in total power computations.
`
`B. Nonlinearity Test
`In the literature, there are many nonlinearity indexes being proposed for time series analysis,
`such as the correlation dimension [29], the Lyapunov exponent [2], [29], [51], the time
`reversibility index [29], [50], and the prediction error [6]. Common methods require the
`computation of surrogate time series in order to construct a hypothesis test. The standard
`procedure is to assume Gaussianity and stationarity, and to perform a Fourier transform
`followed by phase randomization and inverse Fourier transform (such a procedure preserves
`the first and second-order moment statistics while discarding the phase information). In this
`paper, we consider a specific established time-domain method [9] as applied to the R–R time
`series for testing the presence of nonlinearity in the heartbeat intervals.
`
`IEEE Trans Biomed Eng. Author manuscript; available in PMC 2011 June 1.
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`8
`
`

`

`Chen et al.
`
`Page 9
`
`The test developed in [9] uses a phase scrambled bootstrap technique for testing the presence
`of nonlinearity of a time series based upon the third-order moment statistics. The basic idea of
`this method is to compare the estimated third-order moment of the tested series with a set of
`limits generated from linear stationary phase scrambled bootstrap data: large differences shall
`indicate nonlinearity or possibly nonstationarity [9]. The null hypothesis assumes that the given
`time series is linear and stationary. The result of the hypothesis test is either H = 0 (which
`indicates that the null hypothesis is accepted and P >0.05) or H = 1 (which indicates that the
`null hypothesis is rejected with 95% confidence). In the considered simulated series and real
`recordings, we restricted the test to short-term dependence by setting the number of laps M =
`8, and a total of 500 bootstrap replications were simulated for every test.
`
`In order to test the tracking ability of the nonlinear point process model, and to compare its
`performance with the standard filter with only linear dynamics, we generate a synthetic
`heartbeat dataset. Specifically, without postulating a second-order nonlinear system as
`assumed in our model, we used the chaotic Rössler time series governed by the following
`differential equations [49] [Footnote: Of note, the heartbeat dynamics reflected in the synthetic
`set are not directly associated with real physiological generation mechanisms, and it is neither
`implied that the heartbeat dynamics be chaotic].
`
`V. Data
`
`The time series were simulated by the Runge–Kutta integration using conditions a = 0.15, b =
`0.20, and c = 10.0, with step size of 0.01. A total of 3000 data points were generated, one for
`every three x-axis values was chosen, and 1000 data points (representing the generated R–R
`intervals) were finally selected. The simulated deterministic time series is illustrated in Fig. 1
`(first panel). The nonlinearity test described earlier indicates the synthetic time series are
`significantly nonlinear (H = 1 and P < 1e-6).
`
`In order to validate the proposed algorithms' performance as related to real physiological
`dynamics, we have considered two experimental datasets. The first heartbeat dataset was
`recorded under the “autonomic blockade assessment of the sympathovagal balance and
`respiratory sinus arrhythmia” protocol. Detailed description of the experimental data was
`given in [53]. The recorded R–R interval time series last about 5 min for each epoch. In the
`drug administered state, after a control recording stage in rest condition, either atropine (ATR,
`0.04 mg/kg iv over 5 min, parasympathetic blockade) or propranolol (PROP, 0.2 mg/kg iv over
`5 min, sympathetic blockade) was delivered to the subject. In the double blockade (DB) epoch,
`the inputs from both sympathetic and parasympathetic branches of the autonomic nervous
`system were suppressed [53]. A total of 17 healthy volunteers participated in the study. Due
`to space limit, the results of four representative subjects (two from the ATR group and two
`from the PROP group, both were randomly selected) are listed in Table I. These four subjects
`have been tested and reported on a previous analysis with a linear predictive model [14],
`[16]. In Fig. 2, we show the R–R interval series of one representative subject in the control
`supine condition.
`
`The second heartbeat dataset, which was retrieved from a public source: Physionet
`(http://www.physionet.org/) [28], consists of R–R time series recorded from 12 congestive
`heart failure (CHF) patients (from BIDMC-CHF Database) and 16 healthy subjects (from MIT-
`BIH Normal Sinus Rhythm Database). Each R–R time series was artifact-free (upon human's
`visual inspection and artifact rejection) and lasted about 50 min (small segments of the original
`over 20 h recordings). In Fig. 3, we show the R–R interval series from one representative
`
`IEEE Trans Biomed Eng. Author manuscript; available in PMC 2011 June 1.
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`NIH-PA Author Manuscript
`
`9
`
`

`

`Chen et al.
`
`Page 10
`
`healthy subject. Since these recordings have longer durations, they have been deemed as
`particularly suitable for studying complex heartbeat interval dynamics [42], [46].
`
`V1. Results
`A. Model Selection and Goodness-of-Fit Tests
`Using the synthetic dataset, we have conducted several analyses to assess model order selection
`for both linear and nonlinear models. The results of AIC and KS statistics are shown in Table
`I, which are computed from fitting all simulated data points. As seen from Table I, according
`to AIC, the best fit is given by the nonlinear model (8,4), followed by (10,4), whereas according
`to the KS statistic (smaller KS distance), the best fit is given by the nonlinear model (10,4).
`Overall, it is important to notice that for the same level of model complexity, the nonlinear
`model generally achieves a better KS statistic than the linear model, but only when the
`predictive power from the linear part is sufficient—this can be seen from the relatively poorer
`performances of nonlinear models (4,4) and (5,9) in Table I. In this analysis, we selected the
`nonlinear model (10,4) as the optimal nonlinear model (with estimated instantaneous indexes
`shown in Fig. 1), for which the KS plot and autocorrelation plot for fitting the synthetic
`heartbeat data are shown in Fig. 4. As a comparison, the KS plot from the linear AR(14) model
`is also shown (see Fig. 4, left panel). It is worth noting that we have also simulated a linear
`Gaussian AR model for the R–R time series and have compared the performance between the
`linear and nonlinear predictive models—it was found that goodness-of- fit performance by the
`linear model is generally better than the one by a nonlinear model with the same model
`complexity (data not shown).
`
`For the two experimental datasets, we also conducted a preliminary model selection analysis
`(based on the AIC using the first 5-min recordings). Specifically, for testing the linear model
`alone, AIC analysis indicated p = 8 as the optimal linear order in almost all cases. In order to
`keep the number of unknown parameters relatively small, while the size of parameters from
`both linear and nonlinear models remain approximately the same (for fair comparison), we set
`p = 8 for linear modeling and empirically set p = 4 ~ 6 and q = 2 for nonlinear modeling. The
`fitting results are summarized in Tables II and III. Some representative tracking results are
`shown in Figs. 2 and 3, and their respective KS plots are illustrated in Figs. 5 and 6. In general,
`the results related to model fitting improvement vary among subjects in both datasets, and in
`some subjects we also found that the nonlinear model did not improve the goodness-of-fit
`compared with the linear model with equal model complexity.
`
`B. Nonlinearity
`Model selection and goodness-of-fit tests on synthetic data validate the nonlinear quantification
`as evaluated by the point process framework, demonstrating that indexes, such as the ρ and
`ratio (defined in (

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket