• No results found

Frequency-Domain Identification of Continuous-Time ARMA Models from Sampled Data

N/A
N/A
Protected

Academic year: 2021

Share "Frequency-Domain Identification of Continuous-Time ARMA Models from Sampled Data"

Copied!
14
0
0

Loading.... (view fulltext now)

Full text

(1)

Technical report from Automatic Control at Linköpings universitet

Frequency-Domain Identification of

Continuous-Time ARMA Models from

Sampled Data

Jonas Gillberg, Lennart Ljung

Division of Automatic Control

E-mail: gillberg@isy.liu.se, ljung@isy.liu.se

13th May 2009

Report no.: LiTH-ISY-R-2905

Accepted for publication in Automatica

Address:

Department of Electrical Engineering Linköpings universitet

SE-581 83 Linköping, Sweden

WWW: http://www.control.isy.liu.se

AUTOMATIC CONTROL REGLERTEKNIK LINKÖPINGS UNIVERSITET

Technical reports from the Automatic Control group in Linköping are available from http://www.control.isy.liu.se/publications.

(2)

Abstract

The subject of this paper is the direct identi cation of continuous-time au-toregressive moving average (CARMA) models. The topic is viewed from the frequency domain perspective which then turns the reconstruction of the continuous-time power spectral density (CT-PSD) into a key issue. The rst part of the paper therefore concerns the approximate estimation of the CT-PSD from uniformly sampled data under the assumption that the model has a certain relative degree. The approach has its point of origin in the frequency domain Whittle likelihood estimator. The discrete-or continuous-time spectral densities are estimated from equidistant samples of the output. For low sampling rates the discrete-time spectral density is modeled directly by its continuous-time spectral density using the Poisson summation for-mula. In the case of rapid sampling the continuous-time spectral density is estimated directly by modifying its discrete-time counterpart.

(3)

Frequency-Domain Identification of Continuous-Time

ARMA Models from Sampled Data

Jonas Gillberg

a

Lennart Ljung

b

aAptitude Ltd, 8 Humphries Way

Cambridge, CB24 6DL, United Kingdom Email: jonas.gillberg@aptitudegroup.com

bDepartment of Electrical Engineering, Link¨oping University

SE-581 83 Link¨oping, Sweden Email: ljung@isy.liu.se

Abstract

The subject of this paper is the direct identification of continuous-time autoregressive moving average (CARMA) models. The topic is viewed from the frequency domain perspective which then turns the reconstruction of the continuous-time power spectral density (CT-PSD) into a key issue. The first part of the paper therefore concerns the approximate estimation of the CT-PSD from uniformly sampled data under the assumption that the model has a certain relative degree. The approach has its point of origin in the frequency domain Whittle likelihood estimator. The discrete- or continuous-time spectral densities are estimated from equidistant samples of the output. For low sampling rates the discrete-time spectral density is modeled directly by its continuous-time spectral density using the Poisson summation formula. In the case of rapid sampling the

continuous-time spectral density is estimated directly by modifying its discrete-time counterpart.Copyright c°2008 IFAC

Key words: Continuous-time systems; Parameter estimation; Continuous-time ARMA;noise model; Whittle likelihood estimator

1 Introduction

The topic of this paper is a novel method for the identifi-cation of continuous-time autoregressive moving average (ARMA) models. The method works on uniformly sam-pled data and operates primarily in the frequency do-main. Its main advantage is that its direct, i.e. does not make use of the sampled version of the continuous-time system. It is also proved that the method is equivalent to interpolating an estimate of the covariance function. Parameter estimation of continuous-time systems is by no means a novel branch of system identification. See for example the survey article by Unbehauen and Rao, [28], the monograph by Sinha and Rao [22] or the PhD thesis by Mensler [13]. Research on identification of continuous-time models has primarily been concen-trated to the time-domain, with approaches such as: Poisson moment functionals, integrated sampling, or-thogonal functions. A few authors on the other hand, have tackled the problem in the frequency domain. An early reference is by Shinbrot [21] followed by the Fourier modulating function approach introduced by

Pearson et al. [16]. Frequency-domain analysis has also been the starting point for the work on input-output and noise models done by Pintelon et al. [17].

Recently there has been a renewed interest in time system identification. In particular continuous-time noise models[19],[12],[9]. See for instance the arti-cles by Larsson and S¨oderstr¨om on continuous-time AR [11] and ARMA [10] parameter estimation. The work on hybrid Box-Jenkins and ARMAX modeling by Pintelon et.al [18] and Johansson [8] also concerns this problem. The central result in this paper is a method for the esti-mation of the continuous-time power spectral density

ˆˆΦc(iω) = F`+1,Ts(iω) ˆˆΦd(e

iωTs)

by spectrally weighting the discrete-time periodogram. An optimal weighting would be

Fc

2`,Ts(iω) =

Φc(iω, θ0)

Φd(eiωTs, θ0)

(4)

where the discrete-time power spectral density is com-puted using a version of the Poisson summation formula [15] Φd(eiωTs, θ0) = X k=−∞ Φc(iω + ωsk, θ0) (1)

where ωs=Tsis the sampling frequency. Unfortunately,

the optimal weighting depended on the true parameters of the time series model, which of course are unknown during identification. The solution presented in this pa-per to the above dilemma is to replace the true model by a chain of integrations such that

Φc(iω) ≈ 1

|iω|2`,

where ` = n − m is the relative degree or pole excess. This approximation can then be used in the summation formula (1) and will yield another weighting factor

F2`,Tc s(iω) = ¯ ¯ ¯eiωTsiωT−1s ¯ ¯ ¯2` ¯ ¯ ¯Π2`−1(eiωTs) (2`−1)! ¯ ¯ ¯ such that the spectrum can be estimated as

ˆˆΦc,T(iω) =F2`,Tc s(iω) ˆˆΦd,Nt(iω).

Here, the weighting only depends on ` and the so called Euler-Frobenius polynomials Π2`−1(z).

At the end of the paper, it will be shown that this type of spectral weighting is equivalent to interpolating an es-timate of the covariance function {ˆr(kTs)}Nk=0t−1by

car-dinal B-splines. It will also be proved that this setup can be interpreted as time-domain construction of the continuous-time measurements {y(t)}T

t=0 with a spline based kernel Gc `,Ts(t) such that ˆ yc(t) = NXt−1 k=0 y(kTs)Gc`,Ts(t − kTs).

After the continuous-time periodogram is calculated, the parameters can then be estimated using the continuous-time Whittle likelihood approach

ˆ θ = arg min θ X k=1 ˆˆΦc,T(iωk)

Φc(iωk, θ)+ log Φc(iωk, θ).

where there is no need to use the discrete-time power spectral density. All the ideas in this paper should be

seen as a direct continuation of the groundbreaking work on CARMA estimation initiated by Erik K. Larsson and Torsten S¨oderstr¨om[11,9,10].

2 Outline

First, the model structure for continuous-time series models used in the paper is introduced in Section 3. Then, in Section 4, a method for the indirect frequency-domain estimation of a continuous-time ARMA model is presented, in order to familiarize the reader with the subject. Then, the procedure for the direct estimation of the continuous-time model is introduced in Section 5. As explained before, this approach requires the estima-tion of the continuous-spectrum, and a method for this is explained in Section 6. Section 8 then contains a brief introduction to cardinal B-splines and their time- and frequency domain properties. Finally, in Section 9 it is shown that for uniformly sampled data this method is equivalent to interpolation of the covariance function by polynomial splines.

3 Model and Representations

The content of this paper will focus on frequency-domain identification of continuous-time autoregressive moving average (CARMA) models. In this context, observed data {y(t)}T

t=0 or {y(kTs)}Nk=0 is modeled as Gaussian

noise passed through a continuous-time dynamical sys-tem. The system is parameterized as

e(t)

- Hc

y(t)

¡¡

-y(kTs)

Fig. 1. Input and sampling setup for a continuous-time sys-tem.

y(t) =Hc(s, θ)e(t, θ)

Hc(s, θ) =B(s, θ)

A(s, θ) (2)

where e(t) is used informally to denote continuous-time white noise such that

Ee(t, θ) = 0

E{e(t, θ)e(s)} = σ2δ(t − s).

The numerator and the denominator in the transfer func-tion are

A(s, θ) = sn+ a1sn−1+ a2sn−2+ · · · + an

B(s, θ) = sm+ b

(5)

and the vector of parameters is defined as θ = [a1 a2 . . . an b1 b2 . . . bm λ]T where λ = σ2 is the

variance of the driving noise. Here m < n, so the system is strictly proper and Ey2(t) < ∞. A schematic view

of the setup is illustrated in Figure 1.This type of con-figuration could also be interpreted such that the true output would have the spectrum

Φ(iω, θ) = λ ¯ ¯ ¯ ¯B(iω, θ)A(iω, θ) ¯ ¯ ¯ ¯ 2 (3) It should be noted that securing identifiability for continuous-time models which are inferred from discrete-time data is a delicate matter. The transformation from continuous- to discrete-time could possibly destroy in-formation and make two different continuous-time mod-els seem identical. In this paper the authors will assume identifiability of the models and the model structure used for identification is the same as that of the true system. However, for the interested reader we refer to discussion in [5],[25] and [14], together with the method for determining identifiability which is found in [26]. 4 Indirect estimation

A straightforward way to do frequency domain identifi-cation of a continuous time series model from uniformly sampled data would be to compute the discrete-time pe-riodogram from sampled data such that

ˆˆΦd(eiωTs) = ¯ ¯Yd(eiωTs) ¯ ¯2 where Yd ¡ eiωkTs¢=1 Nt Nt X k=1 y(kTs)e−iωkTs.

The parameters, would then be estimated using the discrete-time Whittle likelihood procedure

ˆ θ = arg min θ X k=1 ˆˆΦd(eiωkTs) Φd(eiωkTs, θ)+ log Φd(e iωkTs, θ) (4)

where the discrete-time spectrum is expressed in terms of continuous-time parameters by the relationship

Φd(eiωTs, θ) =

X

k=−∞

Φc(iω + iωsk, θ). (5)

A drawback connected to using the formula in (5) is of course the infinite sum. An approximations can however be achieved with a limited number of terms

Φd(eiωTs) = Nf

X

k=−Nf

Φc(iω + iωsk) (6)

when the continuous-time system is strictly proper. The idea of truncating the Poisson summation formula can now be used to illustrate the detrimental effect that comes with the exclusion of higher order terms. For more efficient ways of approximating the summation we refer the readers to [2] and [27].

4.1 Numerical Example

In Figure 2 and Table 4.1, we have estimated the second-order continuous-time autoregressive (AR) model

y(t) =Hc(s)e(t) Hc(s) = 1 s2+ a 1s + a2 (7) E{e(s)e(t)} =δ(s − t)σ2

with the true parameters σ = 1, a1= 2 and a2= 2. The

duration of the data set was T = 1000s with the sam-pling interval Ts = 1s. Estimates were produced using

NM C = 250 Monte-Carlo simulations and the

informa-tion in the table and figure is based on the mean of those values. The method that has been employed is that of (4) where the discrete time spectrum has been approxi-mated using (6). −100 −90 −80 −70 −60 −50 −40 −30 −20 −10 0 Magnitude (dB) 10−2 10−1 100 101 102 −180 −135 −90 −45 0 Phase (deg) Bode Diagram Frequency (rad/sec)

Fig. 2. Bode diagram comparing the Whittle likelihood

es-timator with Nf = 0 (dashdot) and Nf = 5 (dashed) in (6)

to the true system (solid). The system is Hc(s) =s2+aσ 1s+a2

where σ = 1, a1 = 2 and a2 = 2. The sampling interval

is Ts = 0.8 and frequencies up to the Nyquist frequencies

have been used. When Nf = 0 the estimated system will

be biased. For the case Nf = 5 the dashed curve is almost

identical to the solid curve of the true system.

Figure 2 illustrates the frequency-domain bias which could occur in the transfer function if only the central term (Nf = 0) in (6) have been used, which means that

(6)

we have assumed that

Φd(eiωTs, θ) = Φc(iω, θ).

This will produce a biased estimate which is illustrated by the dashed-dotted line in the Bode diagram for the system. A set of systems with virtually no bias has also been estimated using Nf = 5 in (6). The mean value of

these are also illustrated in the figure as a dotted line which is almost identical to the true system which is represented by the solid line.

In Table 1 the mean parameter values are also illustrated for Nf = 0 and Nf = 5 in (6). From the table we see

that assuming that the discrete-time spectrum is equal to the continuous-time spectrum will also produce bias in the parameters. However, including Nf = 5 terms in

(6) will almost entirely remove this.

Table 1

Mean values of NM C = 250 parameter estimates of the

sys-tem is Hc(s) = s2+aσ1s+a2 where σ = 1, a1= 2 and a2= 2

using the Whittle likelihood estimator with (6).The

sam-pling interval is Ts= 0.8 and frequencies up to the Nyquist

frequencies have been used. When Nf = 0, the parameters

will carry bias. For the case Nf = 5 the parameters are close

those of the true system.

Method a1= 3 a2= 2 σ = 1

Nf = 0 5.090122 3.217543 1.675111

Nf = 5 3.077730 2.044676 1.020823

5 Direct Estimation

Assume for a moment that it would be possible to mea-sure the true continuous-time output {y(t)}T

t=0 of the

system in (3). Then it would be possible to compute the Fourier transform of this signal

ˆ Y (iω) = 1 T Z T 0 y(t)e−iωtdt,

and its periodogram

ˆˆΦ(iω) = ¯¯ ¯ ˆY (iω) ¯ ¯ ¯2= 1 T ¯ ¯ ¯ ¯ ¯ Z T 0 ˆ y(t)e−iωtdt ¯ ¯ ¯ ¯ ¯ 2 would be distributed as

ˆˆΦ(iωk) = |YT(iωk)|2∼ AsExp Φ(iωk, θ0).

The periodogram would also be asymptotically indepen-dent at the frequencies {ωk}Nk=1ω where

ωk=

T k

if T is the time of observation of the output. Here, it is assumed that T is so large so that the asymptotic expression can be considered valid.

When an estimate of the continuous-time power spec-trum is available, a model can be identified by the fol-lowing Maximum-Likelihood (ML) procedure described in the papers by Whittle[34] or Dzhaparidze [6]

L(θ, ˆˆΦ) =

X

k=1

ˆˆΦ(iωk)

Φ(iωk, θ)+ log Φ(iωk, θ)

ˆ

θ = arg min

θ L(θ, ˆˆΦ).

The main problem here is that the discrete-time pe-riodogram can be readily found from the uniformly sampled data while we need the continuous-time peri-odogram to use the method described above. A natural question is then to ask if it is possible to estimate the latter from the former? The answer to this question is yes, and the next section will explain how.

6 Estimating the Continuous-Time Spectrum In this section, we will be looking for ways to approxi-mate the continuous-time power spectral density. In par-ticular expression such as

ˆˆΦc,T(iω) = F2`,Tc s(iω) ˆˆΦd,Nt(e

iωTs) (8)

with a suitable function Fc

2`,Ts(iω). In order to find such

a function, we could argue as follows: if the true param-eters θ0 of (3) were known, we would have

E ˆˆΦd,Nt(e

iωTs) = Φ

d(eiωTs, θ0)

and

E ˆˆΦc,T(iω) = Φc(iω, θ0).

This means that the ideal spectral weighting in (8) would be characterized by

Fc

2`,Ts(iω) =

Φc(iω, θ0)

Φd(eiωTs, θ0). (9)

Since θ0 is unknown, we cannot construct F2`,Tc s(iω) in

this fashion, but the point is that as Ts→ 0, F2`,Tc s(iω)

in (9) will approach Fc

2`,Ts(iω), which is defined in (10).

It turns out that Fc

2`,Ts(iω) is real, positive, and does

not depend on the signal parameters θ0, but only on the

relative degree (pole excess) ` = n − m of the time series model. This is what we will show now.

(7)

Let the model be strictly proper, stable and ` = n − m be its relative degree (or pole excess), i.e. the difference between the number of poles and zeros of the system. Then, at high frequencies, we can approximate the sys-tem as a chain of ` integrators with the continuous-time spectrum

Φc(iω) =

1

|iω|2`.

The discrete-time spectrum for this model is then well known as [32] Φd(eiωTs) = 1 |iω|2` + X k6=0 1 |iω + iωsk|2` = T 2` s (−1)` (2` − 1)! eiωTsΠ 2`−1(eiωTs) (eiωTs− 1)2`

which means that we get the spectral weighting

Fc 2`,Ts(iω) = Φc(iω) Φd(eiωTs) = 1 |iω|2` 1 |iω|2` + P k6=0|iω+iω1sk|2` = ³ eiωTs−1 iωTs ´2` eiωTsΠ2`−1(eiωTs) (2`−1)! , (10)

where Π2`−1(z) are the Euler-Frobenius polynomials,

[33] Π`(z) = b`1z`−1+ b`2z`−2+ · · · + b`` where b`k = k X m=1 (−1)k−mm` µ n + 1 k − m, k = 1, . . . , `.

Finally, observe that since Fc

2`,Ts(iω) ≥ 0 this actually

means that F2`,Tc s(iω) = ³ eiωTs−1 iωTs ´2` eiωTsΠ2`−1(eiωTs) (2`−1)! = ¯ ¯ ¯eiωTsiωT−1s ¯ ¯ ¯2` ¯ ¯ ¯Π2`−1(eiωTs) (2`−1)! ¯ ¯ ¯ In Figure 3 we see that there is a very good correspon-dence between Φc(iω, θ0)/Φd(iω, θ0) and F2`,Tc s(iω) for

the system in (7). This observation is verified by the fol-lowing theoretical result.

Theorem 1 Let the model in (2) be strictly proper,

sta-ble and ` = n−m be its relative degree. Assume Fc

2`,Ts(iω)

is defined as in (10). Then, for each ω

lim Ts→0 Fc 2`,Ts(iω)Φd(e iωTs, θ 0) = Φc(iω, θ0) at the rate of T2` s .

PROOF. Let Φc(iω, θ0) = Φc(iω) and Φd(eiωTs, θ0) =

Φd(eiωTs). Then for each ω choose Tssuch that −ωN <

ω < ωN where ωN = TΠs is the Nyquist frequency. As a

consequence, Fc

2`,Ts(iω) > 0 since by Lemma 3.2 in [32]

Fc 2`,Ts(iω) = 1 |iω|2` 1 |iω|2` + P k6=0|iω+iω1sk|2` . and 1 |iω + iωsk|2` (11)

will only have singularities at ω = ωsk.This will in turn

mean that ¯ ¯Φc(iω) − F2`,Tc s(iω)Φd(e iωTs)¯¯ = ¯ ¯Fc 2`,Ts(iω) ¯ ¯ ¯ ¯ ¯ ¯ ¯ 1 Fc 2`,Ts(iω) Φc(iω) − Φd(eiωTs) ¯ ¯ ¯ ¯ ¯ ¯ ¯ ¯ ¯ ¯ 1 Fc 2`,Ts(iω) Φc(iω) − Φd(eiωTs) ¯ ¯ ¯ ¯ ¯ since Fc

2`,Ts(iω) < 1. Using Lemma 3.2 in [32] again with

the aid of the Poisson summation formula it possible to show that ¯ ¯ ¯ ¯ ¯ 1 Fc 2`,Ts(iω) Φc(iω) − Φd(eiωTs) ¯ ¯ ¯ ¯ ¯= ¯ ¯ ¯ ¯ ¯Φc(iω)|iω| 2` X k=−∞ 1 |iω + iωsk|2` X k=−∞ Φc(iω + iωsk) ¯ ¯ ¯ ¯ ¯ = ¯ ¯ ¯ ¯ ¯ ¯ X k6=0

Φc(iω)|iω|2`− Φc(iω + iωsk)|iω + iωsk|2`

|iω + iωsk|2` ¯ ¯ ¯ ¯ ¯ ¯

since the first terms in the sums cancel. Because the model in (2) is proper and stable with relative degree ` it is possible to find a C > 0 such that Φc(iω)|iω|2`≤ C

(8)

−100 −8 −6 −4 −2 0 2 4 6 8 10 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 ω

Fig. 3. Comparison of Φc(iω, θ0)/Φd(eiωTs) (solid) to

Fc

2`,Ts(iω) (dashdot) for the system Hc(s) =

σ s2+a1s+a2

where σ = 1, a1 = 2 and a2 = 2. In the inner plot Ts = 2

and in the outer plot Ts= 0.5. As can be seen from the

fig-ure, the discrepancy is small at these sampling intervals and

decreases as Ts decrease.

for all ω. Hence

= ¯ ¯ ¯ ¯ ¯ ¯ X k6=0

Φc(iω)|iω|2`− Φc(iω + iωsk)|iω + iωsk|2`

|iω + iωsk|2` ¯ ¯ ¯ ¯ ¯ ¯ X k6=0 ¯

¯Φc(iω)|iω|2`− Φc(iω + iωsk)|iω + iωsk|2`

¯ ¯

|iω + iωsk|2`

X

k6=0

Φc(iω)|iω|2`+ Φc(iω + iωsk)|iω + iωsk|2`

|iω + iωsk|2` X k6=0 2C |iω + iωsk|2` X k6=0 2C |iωsk|2` ≤ 4C µ Ts 2` ∞X k=1 1 k2`

where the last step comes from the definition of the sam-pling time ωs. Since the system is strictly proper and

` ≥ 1 the sum Pk=1 1

k2` will be convergent [1]. This

means that ¯ ¯Φc(iω) − F2`,Tc s(iω)Φd(e iωTs)¯¯ ≤ 4C µ Ts 2` ∞X k=1 1 k2`

and the result will follow. Q.E.D.

From the reasoning above, a reasonable estimate of the

continuous-time spectrum would therefore be

ˆˆΦc(iω) = ¯ ¯ ¯eiωTsiωT−1s ¯ ¯ ¯2` ¯ ¯ ¯Π2`−1(eiωTs) (2`−1)! ¯ ¯ ¯ ˆˆΦd(eiωTs). (12)

It should be noted that theorem above only concern the relationship between the true continuous- and discrete-time power spectral densities as Ts → 0. When the

method in (12) is used and the duration of the dataset

T is limited or short, spectral leakage will occur in the

estimated spectrum ˆˆΦd(eiωTs) . This will in turn cause

bias in the parameter estimates obtained from the Whit-tle method. However, this is a problem concerning small sample sizes and the Whittle estimator as such. The au-thors have thought about this issue, but feels that a rig-orous treatment is outside the scope of this paper. For a more detailed empirical analysis of this difficult topic we refer the interested reader to [3].

7 Numerical Illustration

In Table 2 we have compared the performance of the approach in the section above, which we call Method 2, to using the discrete-time Whittle likelihood approach with the exact discrete time spectrum in (5), which we call Method 1. We have used different sampling intervals, and the mean parameter values have been estimated by

NM C = 250 Monte-Carlo simulations. The system is the

one in (7) and the correspondence between the mean parameter estimates is good.

In Figure 4 we have compared the mean values of

NM C = 250 Monte-Carlo simulations of parameter

estimates versus the sample time Ts. The dotted line

represents the method where we have assumed that ˆˆΦc,T(iω) = ˆˆΦd,Nt(eiωTs). The solid line shows the result

when ˆˆΦc,T(iω) = F2`,Tc s(e

iωTs) ˆˆΦd,N

t(eiωTs) as in (10).

Frequencies up to 2π rad/s have been used and the time of observation is T = 1000 s. For low sampling rates the difference between the two estimates is significant. This is due to the bias that occurs if no spectral weighting by Fc

2`,Ts(iω) is used.

8 Introduction to Cardinal B-Splines

In this section there will be a short introduction to some aspects of B-spline interpolation. The purpose is that of preparing the reader for the results in the remaining part of the paper. Most of the material is based on the two excellent papers by Unser [30,31] on B-splines and signal processing.

Assume that we by interpolation want to approximate some 2` times differentiable function r(t) by the

(9)

B-Table 2

Comparison of mean values for NM C = 250 Monte-Carlo

simulations of parameter estimates versus the sample time

Ts. Method 1 is using the discrete-time Whittle likelihood

ap-proach with the exact discrete time spectrum. Method 2 em-ploys the continuous-time Whittle likelihood approach with the spectral estimator (10). The system is Hc(s) =s2+aσ1s+a2

where σ = 1, a1= 3 and a2= 2 and the performance of the

methods are similar.

Ts Method a1= 3 a2= 2 σ = 1 0.6 1 3.04 ± 0.22 2.02 ± 0.13 1.00 ± 0.047 0.6 2 2.99 ± 0.21 2.02 ± 0.13 1.01 ± 0.047 0.5 1 3.02 ± 0.16 2.02 ± 0.11 1.01 ± 0.034 0.5 2 3.05 ± 0.18 2.04 ± 0.12 1.02 ± 0.036 0.4 1 3.03 ± 0.18 2.02 ± 0.11 1.01 ± 0.038 0.4 2 3.04 ± 0.18 2.02 ± 0.11 1.01 ± 0.039 0.3 1 3.03 ± 0.17 2.01 ± 0.11 1.01 ± 0.035 0.3 2 3.03 ± 0.17 2.02 ± 0.12 1.01 ± 0.035 0.2 1 3.00 ± 0.19 2.01 ± 0.12 1.00 ± 0.038 0.2 2 3.00 ± 0.16 2.01 ± 0.12 1.00 ± 0.031 0.1 1 3.02 ± 0.18 2.02 ± 0.11 1.01 ± 0.035 0.1 2 3.02 ± 0.18 2.02 ± 0.12 1.01 ± 0.036 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0 5 10 a1 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0 2 4 6 a2 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0 1 2 3 σ Ts

Fig. 4. Comparison of mean values for NM C = 250 Monte–

Carlo simulations of parameter estimates versus the sample

time Ts. The dotted line represents the method where we

have assumed that ˆˆΦc,T(iω) = ˆˆΦd,Nt(e

iωTs). The solid line

shows the result when ˆˆΦc,T(iω) = F2`,Tc s(e

iωTs) ˆˆΦd,N t(e

iωTs)

as in (10). The system is Hc(s) = s2+aσ1s+a2 where

σ = 1, a1= 2 and a2 = 2. Frequencies up to 2π rad/s have

been used and the time of observation is T = 1000 s. For low sampling rates the difference between the two estimates is significant.

splines such that

ˆ r(t) = X k=−∞ c(k)Bc 2`,Ts(t − kTs) (13) where Bc

2`,Tsis the B-spline interpolation kernel of order

2` which will be defined more precisely later on. Also, we define the sampled version of spline kernel function as

B2`,Td s(k) = B

c

2`,Ts(kTs).

Then, the interpolation conditions for the bi-infinite se-quence {r(kTs)}∞k=−∞of equidistantly distributed data

points will be r(lTs) = ˆr(lTs), l = −∞ . . . ∞ (14) = X k=−∞ c(k)Bd 2`,Ts(l − k), (15)

Taking the z-transform of this expression will then give

R(z) = X k=−∞ r(kTs)z−k= C(z)B2`,Td s(z), where C(z) = X k=−∞ c(k)z−k, Bd2`,Ts(z) = X k=−∞ B2`,Td s(k)z −k.

Extracting the z-transform of the coefficients is then just a matter of inversion, and we will get

C(z) =£Bd

2`,Ts(z)

¤−1

R(z).

Hence, it remains is to compute the Z-transform

Bd

2`,Ts(z) in order to calculate C(z).

8.1 Z-transform of a B-Spline

The B-spline basis function of order 2` is defined as [29]

Bc 2`,Ts(t) =(` + 1)Ts2` Ts(· − t) 2`−1 + (` + 1)! = 2` X l=0 (−1)l (2` − 1)!Ts2`−1 µ 2` l((2` − l)Ts− t)2`−1+ .

where ∆Ts is the delta operator [7]

Tsf (t) =

f (t + Ts) − f (t)

Ts

(10)

and

f+2`−1=

½

f2`−1 if f ≥ 0

0 if f < 0

In turn, this means that the transformed version of the spline basis function can be computed as [30]

Bd2`,Ts(z) = X k=−∞ B2`,Td s(k)z −k =z−(2`) 2` X l=0 (−1)l (2` − 1)! µ 2` lz2`−1 X k=0 k2`−1zk =z−(2`)(1 − z)2` (2` − 1)! X k=0 k2`−1zk. (16)

Because of the relationship [20]

X

k=0

k2`−1zk= zΠ2`−1(z)

(1 − z)2`, |z| ≤ 1

between the Euler-Frobenius polynomials Π`(z) and the

summation in (16), the following explicit expression,

Bd

2`,Ts(z) =

z−2`+1Π

2`−1(z)

(2` − 1)! . is true for the sampled version of a B-spline.

8.2 Fourier Transform of a Spline

Altogether, the discussion above means that when a function is represented as a spline in (13), its continuous-time Fourier transform will be

ˆ R(iω) = Z −∞ ˆ r(t)e−iωtdt = Z −∞ Ã X k=−∞ c(k)B2`,Tc s(t − Tsk) ! e−iωtdt = X k=−∞ c(k) Z −∞ Bc2`,Ts(t − Tsk)e −iωtdt = Z −∞ Bc 2`,Ts(t)e −iωtdt X k=−∞ c(k)e−iωTsk. (17)

If we use the relationship [20]

Bc 2`,Ts(iω) = Z −∞ Bc 2`,Ts(t)e −iωtdt = µ 1 − e−iωTs 2` ,

and insert this expression into (17) we get ˆ Rc(iω) =F2`,Tc s(iω)Rd(e iωTs) = B c 2`,Ts(iω) Bd 2`,Ts(e iωTs)Rd(e iωTs), where Fc 2`,Ts(iω) = ³ eiωTs−1 ´2` eiωTsΠ`(z) (2`−1)! ,

8.3 Fundamental Polynomial B-Spline Function

An interesting consequence of line of reasoning in the subsection above is that if we interpret Fc

2`,Ts(iω) as

the continuous-time Fourier transform of some function

Fc 2`,Ts(t) such that Fc 2`,Ts(iω) = Bc 2`,Ts(iω) Bd 2`,Ts(e iωTs) = Z −∞ Fc 2`,Ts(t)e −iωtdt. Then, Fc

2`,Ts(t) is the so called fundamental spline

func-tion of order 2` (see for instance Lecture 4 in [20]) which

corresponds to the solution of the interpolation problem

δ(l) = X k=−∞ c(k)B2`,Tc s(Tsl − Tsk) l = −∞, . . . , ∞ where δ(l) = ½ 1 l = 0 0 l 6= 0

is the Kronecker delta function. This means that we can also write our interpolation function in (13) as

ˆ rc(t) = X k=−∞ r(Tsk)F2`,Tc s(t − Tsk)

which is called the Lagrange form [4] of the spline repre-sentation. The functions Fc

2`,Ts(t − kTs) can be thought

of as an orthogonal basis for the linear mapping from the interpolation points to the spline of order 2` with equidistantly distributed knots.

8.4 Interpolation of the Auto-Correlation Function

It is a well known fact that the auto-correlation function

ˆ rd(kTs) = 1 Nt Nt X l=k+1 y(lTs)y((k + l)Ts), 0 ≤ k ≤ Nt

(11)

is related to the discrete-periodogram ˆˆΦd,Nt(e iωTs) such that [24] ˆˆΦd,Nt(e iωTs) =¯¯Y d,Nt(e iωTs)¯¯2 = NXt−1 k=−(Nt−1) ˆ rd(kTs)e−iωTs

This implies that estimating the continuous-time peri-odogram from the discrete-time periperi-odogram using the spectral weighting Fc

2`,Ts(iω) will yield

ˆˆΦc,T(iω) = F2`,Tc s(iω) ˆˆΦd,Nt(e iωTs) = F2`,Tc s(iω) NXt−1 k=−(Nt−1) ˆ rd(kTs)e−iωkTs = T Z −T ˆ rc(t)e−iωtdt.

Hence, the estimate of the continuous-time autocorrela-tion funcautocorrela-tion can be represented as

ˆ rc(t) = NXt−1 k=−(Nt−1) ˆ rd(kTs)F2`,Tc s(t − kTs) where Fc

2`,Ts(t) can also be interpreted as the

fundamen-tal spline basis function of order 2`. In the case of con-tinuous time series models and uniformly sampled data, one is therefore interpolating the covariance function in-stead of the output as in the case of input-output mod-els.

9 Interpretations of Spectral Factorization Imagine that it is possible to compute the spectral factor

Gc

`,Ts(iω) of F

c

2`,Ts(iω) such that

Fc 2`,Ts(iω) = ¯ ¯Gc `,Ts(iω) ¯ ¯2 .

Then, you could write the estimate of the continuous-time spectrum as ˆˆΦc,T(iω) =F2`,Tc s(iω) ˆˆΦd,Nt(e iωTs) =Fc 2`,Ts(iω) ¯ ¯Yd,Nt(e iωTs)¯¯2 =¯¯Gc `,Ts(iω)Yd,Nt(e iωTs)¯¯2

This in turn would mean that we could define a function

ˆ yc(t) = NXt−1 k=0 y(kTs)Gc`,Ts(t − kTs). such that ˆ Yc(iω) = Gc`,Ts(iω)Yd,Nt(e iωTs) = NXt−1 k=0

y(kTs)Gc(iω)e−iωTsk

= NXt−1 k=0 y(kTs) Z −∞ Gc `,Ts(t − kTs)e −iωtdt = Z −∞ ˆ yc(t)e−iωtdt

Now, as the reader might have noticed, the entity

Fc

2`,Ts(iω) is just the Fourier transform of the

funda-mental spline function Fc

2`,Ts(iω) introduced in Section

8.3 and its factorization would be

Fc 2`,Ts(iω) = ¯ ¯ ¯eiωTs−1iωTs ¯ ¯ ¯2` ¯ ¯ ¯B2`−1(eiωTs) (2`−1)! ¯ ¯ ¯ = ¯ ¯ ¯ ¯ ¯C(e iωTs) µ eiωTs−1 iωTs`¯¯ ¯ ¯ ¯ 2 where |C(z)|2= ¯ ¯ ¯ ¯B(2` − 1)!2`−1(z) ¯ ¯ ¯ ¯ .

This is possible, since if z is a root of B2`−1(z) then so

is 1/z [23]. Hence we have Gc `,Ts(iω) = C(e iωTs) µ eiωTs−1 iωTs` ,

and since the Fourier transform of a traditional B-spline of order ` is Bc `,Ts(iω) = µ eiωTs−1 iωTs` the function Gc `,Ts can be expressed as Gc`,Ts(t) = X k=0 c(k)B`,Tc s(t − kTs) (18)

with the convolution property

Fc 2`,Ts(t) = Z −∞ Gc `,Ts(t − τ )G c `,Ts(τ ).

The following example will illustrate the above line of reasoning for the case when ` = 2.

(12)

9.1 Example

Let ` = 2, and we will have

Fc 2`,Ts(iω) = F c 4,Ts(iω) = ³ eiωTs−1 iωTs ´4 B3(eiωTs) 6 where B3(z) = z2+ 4z + 1.

Now, we can write

B3(z) = z2+ 4z + 1

= (z + 2 −√3)(z + 2 +√3)

= z

2 −√3(1 + z

−1(2 −3))(1 + z(2 −3))

which means that one can choose

C(z) =√6 q

2 −√3 1

1 + z−1(2 −3).

Hence one will have

Gc`,Ts(iω) = 6 q 2 −√3 ³ eiωTs−1 iωTs ´2 eiωTs eiωTs+2−3 and Gc2,Ts(iω) = 6 q 2 −√3 X k=0 (−1)k(2 −√3)kBc2,Ts(t − kTs).

This function is illustrated in Figure 5.

10 Conclusions

In this paper, the problem of frequency domain estima-tion of continuous time series models was presented. The central result was a method for the estimation of the continuous-time spectrum from uniformly sampled data. It was proved that the method would approach the true spectrum when the sampling time approaches zero. It could also be interpreted as a way of reconstructing the continuous-time covariance function by means of poly-nomial splines. The method would manifest itself as way of spectrally weighting the discrete-time power spectral density in order suppress the folding effects due to sam-pling. Finally, the spectral factor of the spectral weight-ing was found to correspond to a construction of the mea-sured signals by means of polynomial spline functions.

0 1 2 3 4 5 6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 −5 −4 −3 −2 −1 0 1 2 3 4 5 −0.2 0 0.2 0.4 0.6 0.8 1

Fig. 5. The function Gc

2,1(t) (upper figure) defined in (18)

and its convolution Fc

4,1(τ ) = R −∞ Gc 2,1(τ − t)Gc2,1(t)dt (lower figure). References

[1] M. Abramowitz and I.A. Stegun. Handbook of Mathematical Functions. Dover, New York, NY, 1972.

[2] M.J. Chambers. The esimation of continuous parameter long-memory time series models. Econometric Theory, (12):374– 90, 1996.

[3] A. Contreras-Cristian, E. Gutierrez-Pena, and S. Walker. A note on whittle’s likelihood. Communications in Statistics, Simulation and Computation, 35(4):857–875, 2006.

[4] C. de Boor. A Practical Guide to Splines. Springer-Verlag, New York, 1978.

[5] J. Durbin. Efficient fitting of linear models for continuous stationary time series from discrete data. Bulletin of Institute of International Statistics, (38):273–82, 1961.

[6] K.O. Dzhaparidze. On the estimation of the spectral parameters of a Gaussian stationary process with rational spectral density. Theory of Probability and Its Applications, 15(3):531–538, 1970.

(13)

[7] G. C. Goodwin, J. I. Yuz, and H. Garnier. Robustness issues in continuous-time system identification from sampled data. In Proc. of the 16th IFAC World Congress, Prague, July 2005.

[8] R. Johansson. Identification of continuous-time models. IEEE Trans. on Signal Processing, 42(4):887–897, 1994. [9] E.K. Larsson. Identification of Stochastic Continuous-time

Systems. PhD thesis, Department of Information Technology, Division of Systems and Control, Uppsala Sweden, 2003. [10] E.K. Larsson and M. Mossberg. On possibilities for

estimating continuous-time ARMA parameters. In Proc. of the IFAC Symposium on System Identification 2003, Rotterdam, August 2003.

[11] E.K. Larsson and T. S¨oderstr¨om. Continuous-time ar parameter estimation by using properties of sampled systems. In Proc. of the 15th IFAC World Congress, Barcelona, July 2002.

[12] L. Ljung. Initialization aspects for subspace and output-error identification methods. In Proc. 5th European Control Conference (ECC), Cambridge, UK, September 2003. [13] M. Mensler. Analyse et etude comparative de methodes

d’identification des systemes a representation continue. PhD thesis, Centre de Recherche en Automatique de Nancy, Universite Henri Poincare, Nancy, 1999.

[14] S.M. Pandit and S.M. Wu. Unique estimates of the parameters of a continuous-time stochastic process. Biometrika, 1997.

[15] A. Papoulis. Probability, Random Variables and Stochastic Processes. McGraw-Hill, New York, NY, 1965.

[16] A.E. Pearson and Y. Shen. Weighted least squares/mft algorithms for linear differential system identification. In CDC 1993, December 1993.

[17] R. Pintelon and J. Schoukens. Identification of continuous-time systems using arbitrary signals. Automatica, 33(5):991– 994, 1997.

[18] R. Pintelon and J. Schoukens. Box-jenkins continuous-time modeling. Automatica, 36:983–991, 2000.

[19] G. P. Rao and H. Garnier. Numerical illustrations of the relevance of direct continuous-time model identification. In Proc. of the 15th IFAC World Congress, Barcelona, July 2002.

[20] I. J. Schoenberg. Cardinal Spline Interpolation. SIAM, Philadelphia, PA, 1973.

[21] M. Shinbrot. On the analysis of linear and nonlinear systems. Trans. ASME, 79:547–552, 1957.

[22] N.K. Sihna and G.P. Rao. Identification of continuous-time systems : methodology and computer implementation. Kluwer, Dordrecht, 1991.

[23] S.L. Sobolev. On the roots of Euler polynomials. Soviet Mathematics Doklady, 18(4):935–938, 1977.

[24] P. Stoica and R.L. Moses. Introduction to to spectral analysis. Prentice-Hall, Upper Saddle River, 1996.

[25] L.G. Telser. Discrete smaples and moving sums in stationary stochastic processes. Journal of the American Statistical Association, (62):484–99, 1967.

[26] H. Tsai. Quasi-maximum likelihood estimation for a class of continuous-time long-memory processes. Journal of Time Series Analysis, (26):691–713, 2005.

[27] H. Tsai. Quasi-maximum likelihood estimation of long-memory limiting aggregate processes. Statistica Sinica, (16):691–713, 2005.

[28] H. Unbehauen and G. P. Rao. Continuous-time approaches to system identification - a survey. Automatica, 26(1):23–35, 1990.

[29] M. Unser, A. Aldroubi, and M. Eden. Fast B-spline transforms for continuous image representation and interpolation. IEEE Trans. Pattern Analysis and Machine Intelligence, 13(3):277–285, 1991.

[30] M. Unser, A. Aldroubi, and M. Eden. B-spline signal processing: Part I - theory. IEEE Trans. Signal Processing, 41(2):821–833, 1993.

[31] M. Unser, A. Aldroubi, and M. Eden. B-spline signal processing: Part II - efficient design and application. IEEE Trans. Signal Processing, 41(2):834–847, 1993.

[32] B. Wahlberg. Limit results for sampled systems. International Journal of Control, 48(3):1267–1283, 1988. [33] S. R. Weller, W. Moran, B. Ninnes, and A. D. Pollinton.

Sampling zeros and the Euler-Frobenius polynomials. IEEE Transactions on Automatic Control, 46(2):340–343, 2001. [34] P. Whittle. Gaussian estimation in stationary time series.

Bulletin of the International Statistical Institute, 39:105–130, 1961.

Jonas Gillberg was born 1975 in V¨anersborg, Sweden. He received his Bachelor of Science degree in Business Administration and Master of Science degree in Elec-trical Engineering in 2000, both at Link¨oping University. In 2002 he joined the Automatic Control group at Link¨oping University. He earned his Ph.D.degree in Auto-matic Control in 2006. He currently works as an advanced process control consultant with Aptitude Ltd, serving the oil, gas and petrochemicals industry.

Lennart Ljung received his

Ph.D. in Automatic Control from the Lund Institute of Technology in 1974. Since 1976 he is Professor of the chair of Automatic Control in Linkoping, Sweden, and is cur-rently Director of the Competence Center ”Information Systems for Industrial Control and Supervi-sion” (ISIS). He has held visiting positions at Stanford and MIT and has written several books on Sys-tem Identification and Estimation. He is an IEEE Fellow, an IFAC Fellow and an IFAC Advi-sor as well as a Member of the Royal Swedish Academy of Sciences (KVA), a Member of the Royal Swedish Academy of Engineering Sciences (IVA), an Honorary Member of the Hungarian Academy of Engineering and a Foreign Associate of the US National Academy of Engineering (NAE). He has received honorary doctorates from the Baltic State Technical University in St. Petersburg, from the Uppsala University, Sweden, from the Technical University of Troyes, France, and from the Catholic University of Leuven, Belgium. In 2002 he received the Quazza Medal from IFAC, in 2003 the Hendryk W. Bode Lecture Prize from the IEEE Control Systems So-ciety and he is the recipient of the IEEE Control Systems Award for 2007.

(14)

Avdelning, Institution Division, Department

Division of Automatic Control Department of Electrical Engineering

Datum Date 2009-05-13 Språk Language  Svenska/Swedish  Engelska/English   Rapporttyp Report category  Licentiatavhandling  Examensarbete  C-uppsats  D-uppsats  Övrig rapport  

URL för elektronisk version http://www.control.isy.liu.se

ISBN  ISRN



Serietitel och serienummer

Title of series, numbering ISSN1400-3902

LiTH-ISY-R-2905

Titel

Title Frequency-Domain Identication of Continuous-Time ARMA Models from Sampled Data

Författare

Author Jonas Gillberg, Lennart Ljung Sammanfattning

Abstract

The subject of this paper is the direct identi cation of continuous-time autoregressive moving average (CARMA) models. The topic is viewed from the frequency domain perspective which then turns the reconstruction of the continuous-time power spectral density (CT-PSD) into a key issue. The rst part of the paper therefore concerns the approximate estimation of the CT-PSD from uniformly sampled data under the assumption that the model has a certain relative degree. The approach has its point of origin in the frequency domain Whittle likelihood estimator. The discrete-or continuous-time spectral densities are estimated from equidistant samples of the output. For low sampling rates the discrete-time spectral density is modeled directly by its continuous-time spectral density using the Poisson summation formula. In the case of rapid sampling the continuous-time spectral density is estimated directly by modifying its discrete-time counterpart.

References

Related documents

When rendering the high detail model the program is limited by the GPU it does not matter if you split up the CPU overhead in multiple threads or if you remove the overhead

Det här för att bidra med värdefull kunskap till näringslivet och forskare om sambandsexponeringar kan öka försäljningen på både komplementprodukten och

The implementation of the variable node processing unit is relatively straight- forward, as shown in Fig. As the extrinsic information is stored in signed magnitude format, the data

Goldie som kvinnan med guldhåret heter, använder sin sexualitet för att bli skyddad från att bli mördad utan att hon berättar vilka problem hon är i, till skillnad från äldre

Microscopic changes due to steam explosion were seen to increase diffusion of the smaller 3-kDa dextran dif- fusion probe in the earlywood, while the latewood structure was

- Ett samlingsbegrepp för om jag spelar korta eller långa fraser, snabbt eller långsamt, varierar min time och mer övergripande rytmiskt?. Vilka notvärden använder jag mig av

Other tentative explanations for the association between smoking prevalence and cigarette production include access to cigarettes through retailing practices, cul- tural and

Figure 4.5 shows how the existing variables browser is used to expose interactive simulation variables red from the embedded server to the user.. The motivation behind is that the