This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Estimating and detecting random processes on the unit circle

Changrong Liu    S. Suvorova    R. J. Evans    W. Moran    A. Melatos Electrical and Electronic Engineering Department, University of Melbourne, Parkville, Victoria 3010, Australia and OzGrav (e-mail: changrongl1@student.unimelb.edu.au) Electrical and Electronic Engineering Department, University of Melbourne, Parkville, Victoria 3010, Australia and OzGrav (e-mail: sofia.suvorova, robinje, wmoran@unimelb.edu.au). School of Physics, University of Melbourne, Parkville, Victoria 3010, Australia and OzGrav (e-mail: amelatos@unimelb.edu.au).
Abstract

Detecting random processes on a circle has been studied for many decades. The Neyman-Pearson detector, which evaluates the likelihood ratio, requires first the conditional mean estimate of the circle-valued signal given noisy measurements, which is then correlated with the measurements for detection. This is the estimator-correlator detector. However, generating the conditional mean estimate of the signal is very rarely solvable. In this paper, we propose an approximate estimator-correlator detector by estimating the truncated moments of the signal, with estimated signal substituted into the likelihood ratio. Instead of estimating the random phase, we estimate the complex circle-valued signal directly. The effectiveness of the proposed method in terms of estimation and detection is shown through numerical experiments, where the tracking accuracy and receiver operating curves, compared with the extended Kalman filter are shown under various process/measurement noise.

keywords:
Conditional Mean Estimate, Circle-Valued Signal, Estimator-Correlator, Moment Function, Random Process

1 Introduction

Detecting weak circle-valued random processes in white Gaussian noise has been analyzed by researchers for several decades. This type of signal is widely encountered in a range of fields including communication systems, where signals are modelled as either frequency or phase modulated by a Gaussian message (Snyder (1968)). Another example is optical communication with laser phase drift (Rusch and Poor (1995); Foschini et al. (1988)). Generally speaking, These signals are specific cases of non-Gaussian random processes (Kailath and Poor (1998)). In order to construct a Neyman-Pearson detector for such a process, the likelihood ratio must be evaluated, which requires computation of the causal conditional mean estimate of the random process given measurements. The estimated signal is then correlated with the measurements to produce the detection statistic (Kailath and Poor (1998); Veeravalli and Poor (1991)). This is the estimator-correlator (EC) detector (Kailath and Poor (1998)). Unfortunately, the conditional mean estimate is mostly impossible to compute except in the Gaussian case (Kailath and Poor (1998)). Instead, a near-optimal approximation of the conditional mean estimate (or equivalently, the conditional density) of the signal has to be constructed.

For circle-valued signals, various approximation techniques have been proposed. A random point on the circle is represented either by the angle Θ[π,π)\Theta\in[-\pi,\pi) or the complex numbers {eiΘ|Θ[π,π)}\{e^{i\Theta}|\Theta\in[-\pi,\pi)\}. One approach is to characterize the conditional density of Θ\Theta using a Gaussian sum approximation with the mean and variance updated, as described in Tam and Moore (1977). In Willsky (1974), the density of Θ\Theta is spanned by Fourier modes, with Fourier coefficients estimated. In Suvorova et al. (2021), the density of Θ\Theta is approximated by a von-Mises distribution with maximum entropy, and at each iteration, the density is propagated and projected back to a von-Mises distribution. Another approach is to use the extended Kalman filter (EKF) to linearize the nonlinear system dynamics and/or the measurement process to demodulate (estimate) the phase (Georghiades and Snyder (1985); Snyder (1968)). In Tretter (1985), the unwrapped phase model is assumed and fitted by a linear regression. In Scharf et al. (1980), the dynamics of the wrapped phase sequence is assumed and then estimated using dynamic programming. A similar idea has also been proposed in Suvorova et al. (2018) and Melatos et al. (2021), where the phase is modelled as a hidden Markov model and the Viterbi algorithm is used to track the random phase. A deterministic filtering approach is attempted in Zamani et al. (2011), where the process and measurement noises are both treated as unknown deterministic disturbances and a deterministic minimum-energy filter is formed by minimizing an energy cost function over phase trajectories (Coote et al. (2009)). As we can see, a large number of methods have been considered for estimating (filtering) the phase Θ\Theta, while very few papers analyze the complex signal eiΘe^{i\Theta} directly. One recent paper (Condat (2021)) minimizes a Tiknonov-regularized problem over the complex signal eiΘe^{i\Theta}, where the non-convexity introduced by the S1S^{1} constraint is relaxed using moments of measures.

In this paper, we focus directly on stochastic filtering of the circle-valued signal eiΘe^{i\Theta} embedded in large additive white noise, where Θ\Theta obeys different random processes. We first present the nonlinear Kushner stochastic partial differential equation for recursively computing the conditional mean estimate of the signal, or equivalently, the conditional probability density. The solution of this equation is infinite dimensional, and we approximate it by a sequence of moment functions. The choice of moment functions is determined by the fact that they form the eigenspace of the infinitesimal generator of the signal process. In doing so, solving the Kushner equation simplifies to solving a sequence of time-dependent ordinary differential equations (ODEs), which characterize the evolution of the conditional density. We then produce the approximate optimal EC detector and show its performance by simulations.

The paper is organized as follows. Section 2 describes the dynamic system on S1S^{1} and formulates the detection and estimation problems. The Kushner equation for computing the conditional density and its finite-dimensional approximation is introduced and derived in Section 3. Simulation results are included in Section 4, testing both estimation and detection performance, compared with EKF based methods. Finally, brief conclusions are drawn in Section 5.

2 Problem Formulation

2.1 System model

Consider a detection system with two hypothesis:

H0:Zt\displaystyle H_{0}:Z_{t} =σ01/2Wt, 0tT,\displaystyle=\sigma_{0}^{1/2}W_{t},\;0\leq t\leq T, (1)
H1:Zt\displaystyle H_{1}:Z_{t} =0th(Xt)𝑑s+σ01/2Wt, 0tT\displaystyle=\int_{0}^{t}h(X_{t})ds+\sigma_{0}^{1/2}W_{t},\;0\leq t\leq T (2)

where ZtZ_{t} is the noisy measurements and Xt=eiΘtS1X_{t}=e^{i\Theta_{t}}\in S^{1} is the circle-valued signal with Θt\Theta_{t} obeying an arbitrary random process. Measurement function h(Xt)h(X_{t}) in our case, is given by

h(Xt)=[Xt]h(X_{t})={\Re}[X_{t}] (3)

with []\Re[\cdot] denoting the real part. {Wt,0tT}\{W_{t},0\leq t\leq T\} is the standard Wiener process, which represents the measurement noise and is independent of XtX_{t}. σ0\sigma_{0} denotes the variance of the measurement noise and is assumed known. In this paper, we consider three specific random processes for Θt\Theta_{t}:

(i)Θt\displaystyle\text{(i)}\;\Theta_{t} =0tqΘ1/2𝑑Bs+ϕ0,\displaystyle=\int_{0}^{t}q_{\Theta}^{1/2}dB_{s}+\phi_{0}, (4)
(ii)Θt\displaystyle\text{(ii)}\;\Theta_{t} =0t(w0ds+qΘ1/2dBs)+ϕ0, and\displaystyle=\int_{0}^{t}(w_{0}ds+q_{\Theta}^{1/2}dB_{s})+\phi_{0},\text{ and} (5)
(iii)Θt=0t(wsds+qΘ1/2dBs)+ϕ0,wt=w0+0tqw1/2𝑑B~s\displaystyle\begin{split}\text{(iii)}\;\Theta_{t}&=\int_{0}^{t}(w_{s}ds+q_{\Theta}^{1/2}dB_{s})+\phi_{0},\\ w_{t}&=w_{0}+\int_{0}^{t}q_{w}^{1/2}d\tilde{B}_{s}\end{split} (6)

where qΘq_{\Theta} and qwq_{w} are known functions of phase, Θ\Theta, and frequency, ww. They represent the noise variance for phase and frequency, respectively. w0w_{0} is the known initial frequency and ϕ0\phi_{0} is the random initial phase, uniformly distributed across [0,2π)[0,2\pi). {Bt,0tT}\{B_{t},0\leq t\leq T\} and {B~t,0tT}\{\tilde{B}_{t},0\leq t\leq T\} are two independent standard Wiener processes, which are also independent of XtX_{t} and {Wt,0tT}\{W_{t},0\leq t\leq T\}.

2.2 Neyman-Pearson detector

The Neyman-Pearson (EC) detector is the optimal detector in the sense that it maximizes the detection probability (Pd) with given false alarm probability (Pf). In order to form this type of detector, the log likelihood ratio logΛt\log\Lambda_{t} has to be evaluated, which is given by

logΛt\displaystyle\log\Lambda_{t} =0tX^s𝑑Zs120tX^s2𝑑s,\displaystyle=\int_{0}^{t}\hat{X}_{s}dZ_{s}-\frac{1}{2}\int_{0}^{t}\hat{X}_{s}^{2}ds, (7)
X^t\displaystyle\hat{X}_{t} =𝐄[Xt|t;H1]\displaystyle=\mathbf{E}[X_{t}|\mathcal{F}_{t};H_{1}] (8)

with t\mathcal{F}_{t} the filtration of σ\sigma-algebras generated by {Zτ,0τt}\{Z_{\tau},0\leq\tau\leq t\}, given in (2) with st\mathcal{F}_{s}\subseteq\mathcal{F}_{t}, st\forall s\leq t.

From (7) and (8), the EC detector is formed in two steps: 1.) calculate the conditional mean estimate X^t\hat{X}_{t}; 2.) form the log likelihood ratio logΛT\log\Lambda_{T} at the termination time TT and compare it with a pre-specified threshold to claim H0H_{0} or H1H_{1} correspondingly.

3 Nonlinear Filtering: compute conditional mean estimate

We now consider computation of the conditional mean estimate of the signal using the Kushner equation.

3.1 Kushner equation: recursive filtering of the conditional density

Given the filtered probability space (Ω,,;{t})(\Omega,\mathcal{F},\mathbb{P};\{\mathcal{F}_{t}\}) with Ω\Omega the underlying sample space and \mathbb{P} the probability measure defined on \mathcal{F}, suppose the signal and measurement processes are given by

dxt=b(xt)dt+σ(xt)dBt,x(0)=x0dZt=h(xt)dt+σ01/2dWt,Z0=0\begin{split}dx_{t}&=b(x_{t})dt+\sigma(x_{t})dB_{t},\>x(0)=x_{0}\\ dZ_{t}&=h(x_{t})dt+\sigma_{0}^{1/2}dW_{t},\>Z_{0}=0\end{split} (9)

with xtdx_{t}\in\mathbb{R}^{d} and b:ddb:\mathbb{R}^{d}\to\mathbb{R}^{d}, σ:dd×p\sigma:\mathbb{R}^{d}\to\mathbb{R}^{d\times p}, h:dh:\mathbb{R}^{d}\to\mathbb{R}, a=12σσTa\overset{\triangle}{=}\frac{1}{2}\sigma\sigma^{\rm{T}} with superscript T\rm{T} being the transpose of a matrix. The filtering problem can be formulated as: for a general measurable function g(xt)g(x_{t}), calculate g^(xt)\hat{g}(x_{t}) by

g^(xt)=𝐄[g(xt)|t]=dg(xt)(xtdx|t)=dg(x)p(x,t|t)𝑑x\begin{split}\hat{g}(x_{t})&\overset{\triangle}{=}\mathbf{E}[g(x_{t})|\mathcal{F}_{t}]\\ &=\int_{\mathbb{R}^{d}}g(x_{t})\mathbb{P}(x_{t}\in dx|\mathcal{F}_{t})\\ &=\int_{\mathbb{R}^{d}}g(x)p(x,t|\mathcal{F}_{t})dx\end{split} (10)

where p(x,t|t)p(x,t|\mathcal{F}_{t}) denotes the conditional density of xx at time tt. The recursive formula for evolving p(x,t|t)p(x,t|\mathcal{F}_{t}) is described by the Kushner equation (Kushner (1967))

dp(x,t|t)=p(x,t|t)dt+(h(xt)h^(xt))σ01[dZth^(xt)dt]p(x,t|t)\begin{split}dp(x,t|\mathcal{F}_{t})&=\mathcal{L}^{*}p(x,t|\mathcal{F}_{t})dt\\ &+(h(x_{t})-\hat{h}(x_{t}))\cdot\sigma_{0}^{-1}[dZ_{t}-\hat{h}(x_{t})dt]p(x,t|\mathcal{F}_{t})\end{split} (11)

with \mathcal{L}^{*} the adjoint of the infinitesimal generator \mathcal{L}, given respectively by

()\displaystyle\mathcal{L}^{*}(\cdot) =i,j=1d2xixj(aij())i=1dxi(bi()),\displaystyle=\sum_{i,j=1}^{d}\frac{\partial^{2}}{\partial x_{i}\partial x_{j}}(a_{ij}(\cdot))-\sum_{i=1}^{d}\frac{\partial}{\partial x_{i}}(b_{i}(\cdot)), (12)
()\displaystyle\mathcal{L}(\cdot) =i,j=1daij2xixj+i=1dbixi.\displaystyle=\sum_{i,j=1}^{d}a_{ij}\frac{\partial^{2}}{\partial x_{i}x_{j}}+\sum_{i=1}^{d}b_{i}\frac{\partial}{\partial x_{i}}. (13)

To uniquely characterize the conditional density, we pair (11) with an arbitrary test function g(x)g(x) with compact support, where the duality pairing is denoted by ,\langle\cdot,\cdot\rangle. We then have

dp(x,t|t),g(x)=\displaystyle d\Big{\langle}p(x,t|\mathcal{F}_{t}),g(x)\Big{\rangle}= p(t,x|t),g(x)dt+(hh^)\displaystyle\Big{\langle}\mathcal{L}^{*}p(t,x|\mathcal{F}_{t}),g(x)\Big{\rangle}dt+\Big{\langle}(h-\hat{h})
σ01[dZth^dt]p(x,t|t),g(x)\displaystyle\cdot\sigma_{0}^{-1}[dZ_{t}-\hat{h}dt]p(x,t|\mathcal{F}_{t}),g(x)\Big{\rangle} (14)

which further implies

dg^(xt)=g^(xt)dt+g(xt)h(xt)g^(xt)h^(xt)σ01[dZth^(xt)dt],\begin{split}d\hat{g}(x_{t})=&\mathcal{L}\hat{g}(x_{t})dt\\ &+\langle g(x_{t})h(x_{t})\rangle-\hat{g}(x_{t})\hat{h}(x_{t})\cdot\sigma_{0}^{-1}[dZ_{t}-\hat{h}(x_{t})dt],\end{split} (15)

here \mathcal{L} is defined in (13). Both ^\hat{} and \langle\cdot\rangle denote the conditional mean with respect to the conditional density p(x,t|t)p(x,t|\mathcal{F}_{t}). g^(xt)=p(x,t|t),g(x)\hat{g}(x_{t})\overset{\triangle}{=}\Big{\langle}p(x,t|\mathcal{F}_{t}),g(x)\Big{\rangle} is defined in (10).

3.2 Choice of the test function

From the previous section, we observe the duality relationship between the conditional density and the estimated test function. In order to fully characterize p(x,t|t)p(x,t|\mathcal{F}_{t}), an infinite number of statistics (or test functions g()g(\cdot)) are required. In this paper, we set the test functions to be the moment functions of the signal XtX_{t}, given as Xn(t)=einΘtX_{n}(t)\overset{\triangle}{=}e^{in\Theta_{t}} for n=0,N1n=0\cdots,N-1, considering firstly, Xn(t)X_{n}(t) can be regarded as the characteristic function of Θ\Theta, which completely determines the properties of the probability distribution of Θ\Theta. Secondly, Xn(t)X_{n}(t) spans the eigenspace of the generator \mathcal{L}, as discussed in the following section. In doing so, solving an infinite dimensional Kushner equation boils down to solving NN’s ODEs.

As we can see from (15), the filtering process is composed of two terms: the one-step prediction (the first term) and the update (the second term). We will derive the solutions for X^n\hat{X}_{n} in terms of these two properties.

3.3 Prediction: compute X^n(t)\mathcal{L}\hat{X}_{n}(t)

In order to obtain the infinitesimal generator of the signal process, we need to write the signal dynamics in Ito differential form, for three scenarios. In particular, we have for scenario (i) and (ii)

(i)dXn\displaystyle{\text{(i)}}\;dX_{n} =qΘn22Xndt+inqΘ1/2XndBt,\displaystyle=-\frac{q_{\Theta}n^{2}}{2}X_{n}dt+inq_{\Theta}^{1/2}X_{n}dB_{t}, (16)
(ii)dXn\displaystyle{\text{(ii)}}\;dX_{n} =(inw0qΘn22)Xndt+inqΘ1/2XndBt.\displaystyle=(inw_{0}-\frac{q_{\Theta}n^{2}}{2})X_{n}dt+inq_{\Theta}^{1/2}X_{n}dB_{t}. (17)

Comparing (17) with (16), a constant rotation term inw0inw_{0} is added to the drift coefficient. Considering the martingale process in both circumstances: inqΘ1/2XndBtinq_{\Theta}^{1/2}X_{n}dB_{t}, with the property 𝐄[stinqΘ1/2Xn(τ)𝑑Bτ|s]=0\mathbf{E}[\int_{s}^{t}inq_{\Theta}^{1/2}X_{n}(\tau)dB_{\tau}|\mathcal{F}_{s}]=0, we have a separate infinitesimal generator for each XnX_{n}, where n=inw0qΘn22\mathcal{L}_{n}=inw_{0}-\frac{q_{\Theta}n^{2}}{2}, with w0=0w_{0}=0 in scenario (i).

The differential equation for XnX_{n} in scenario (iii) can be obtained in the same manner

dXn=(inwtqΘn22)Xndt+inqΘ1/2XndBt.dX_{n}=(inw_{t}-\frac{q_{\Theta}n^{2}}{2})X_{n}dt+inq_{\Theta}^{1/2}X_{n}dB_{t}. (18)

However, as shown here, the randomness appears in the drift coefficient. Now it is difficult to interpret the generator by simply writing n(w)=inw(t)qΘn22\mathcal{L}_{n}(w)=inw(t)-\frac{q_{\Theta}n^{2}}{2} with w(t)w(t) a random process. Instead, n(w)\mathcal{L}_{n}(w) should be interpreted in the mean sense by marginalizing out the random variable w(t)w(t). In other words, we have the infinitesimal generator computed by n=𝐄w[n(w)|t]=wn(w)p(w,t|t)𝑑w\mathcal{L}_{n}=\mathbf{E}_{w}[\mathcal{L}_{n}(w)|\mathcal{F}_{t}]=\int_{w\in\mathbb{R}}\mathcal{L}_{n}(w)p(w,t|\mathcal{F}_{t})dw, with dX^nnX^ndtd\hat{X}_{n}-\mathcal{L}_{n}\hat{X}_{n}dt still a martingale process. Alternatively, we introduce the new state variables as Xmn=wmeinΘX_{mn}\overset{\triangle}{=}w^{m}e^{in\Theta}, with m=0,,M1m=0,\dots,M-1 and n=0,N1n=0,\dots N-1, where superscript mm and nn denote the mmth and nnth power, respectively. This is an “augmented” state space by noting that Xn=Xmn|m=0X_{n}=X_{mn}|_{m=0}. For predicting XmnX_{mn}, likewise, we write down the differential equation for XmnX_{mn} as

(iii)dXmn=(qwm(m1)2Xm2,n+inXm+1,nqΘn22Xmn)dt+qw1/2mXm1,ndB~t+inqΘ1/2XmndBt\begin{split}{\text{(iii)}}\;dX_{mn}=&\Big{(}\frac{q_{w}m(m-1)}{2}X_{m-2,n}+inX_{m+1,n}\\ &-\frac{q_{\Theta}n^{2}}{2}X_{mn}\Big{)}dt+q_{w}^{1/2}mX_{m-1,n}d\tilde{B}_{t}\\ &+inq_{\Theta}^{1/2}X_{mn}dB_{t}\end{split} (19)

with mn(Xmn)=qwm(m1)2Xm2,n+inXm+1,nqΘn22Xmn\mathcal{L}_{mn}(X_{mn})=\frac{q_{w}m(m-1)}{2}X_{m-2,n}+inX_{m+1,n}-\frac{q_{\Theta}n^{2}}{2}X_{mn} for each XmnX_{mn}. Here by augmenting the state space from XnX_{n} to XmnX_{mn}, the time-variant generator is replaced by a time-invariant generator at the cost of higher computational complexity.

3.4 Update: incorporate the innovation dZth^dtdZ_{t}-\hat{h}dt

In all three scenarios, we have measurement function h(Xt)=[Xt]h(X_{t})=\Re[X_{t}] from (3). This is a considerable advantage since hh can be regarded as a linear operator in both the subspace {Xn}n=0N1\{X_{n}\}_{n=0}^{N-1} and {Xmn}m=0,n=1M1,N1\{X_{mn}\}_{m=0,n=1}^{M-1,N-1} by noting

(i)&(ii)hXn=Xn+1+Xn12h^=X^1+X^12,\displaystyle\begin{split}{\text{(i)}}\&\text{(ii)}\;\langle hX_{n}\rangle&=\Big{\langle}\frac{X_{n+1}+X_{n-1}}{2}\Big{\rangle}\\ \hat{h}&=\frac{\hat{X}_{1}+\hat{X}_{1}^{*}}{2},\end{split} (20)
(iii)hXmn=Xm,n+1+Xm,n12h^=X^01+X^012\displaystyle\begin{split}{\text{(iii)}}\;\langle hX_{mn}\rangle&=\Big{\langle}\frac{X_{m,n+1}+X_{m,n-1}}{2}\Big{\rangle}\\ \hat{h}&=\frac{\hat{X}_{01}+\hat{X}_{01}^{*}}{2}\end{split} (21)

with ()()^{*} denoting the complex conjugation. Combining (16), (17), (19) with (20), (21) and substituting into (15), we can compute X^n(t)\hat{X}_{n}(t) for (i)&(ii) and X^mn(t)\hat{X}_{mn}(t) for (iii) recursively by solving a sequence of stochastic ODEs.

The log-likelihood ratio, defined in (7) can then be updated recursively as well by substituting in X^t\hat{X}_{t} from (15) without too much effort, considering X^t=X^n(t)|n=1\hat{X}_{t}=\hat{X}_{n}(t)|_{n=1} for (i)&(ii) and X^t=X^mn(t)|m=0,n=1\hat{X}_{t}=\hat{X}_{mn}(t)|_{m=0,\;n=1} for (iii).

4 Simulations and results

4.1 Implementation of the Neyman-Pearson (EC) detector

Since scenario (i) is analogous to (ii) with w0=0w_{0}=0, we only do experiments for (ii) and (iii). The Neyman-Pearson (EC) detector is composed of a conditional mean estimator and a log-likelihood ratio detector, hence we need to check the performance of both estimation and detection. For estimation, we generate the synthetic signal Xsyn=AsyneiΘsynX_{\rm syn}=A_{\rm syn}e^{i\Theta_{\rm syn}} with Θsyn\Theta_{\rm syn} given in (5) and (6). Noisy measurements are generated from (2). For detection, we randomly generate measurements from (1) or (2), where signal Xsyn=AsyneiΘsynX_{\rm syn}=A_{\rm syn}e^{i\Theta_{\rm syn}} is generated as above. We fix termination time TT and compare logΛT\log\Lambda_{T} with a pre-specified threshold and claim detection if logΛT\log\Lambda_{T} is greater than the threshold, and vice versa. We simulate both the signal process and the measurement process numerically using the Euler-Maruyama method with sampling interval Δt=0.1\Delta t=0.1s. The parameters are listed in Table 1.

Table 1: Simulation parameters
parameter description value
ϕ0\phi_{0} initial phase distU(0,2π)\overset{\rm dist}{\sim}U(0,2\pi)
qΘq_{\Theta} phase noise variance 101,10210^{-1},10^{-2}
qwq_{w} frequency noise variance 10810^{-8}
σ0\sigma_{0} observation noise variance 1,,801,\dots,80
w0w_{0} initial frequency 0.012, 0.0320.012,\;0.032Hz
|Asyn||A_{\rm{syn}}| absolute amplitude of synthetic signal 11
SNR =|Asyn|Δtσ0=\sqrt{\frac{|A_{\rm{syn}}|\Delta t}{\sigma_{0}}} signal to noise ratio 0.0354,0.31620.0354,\ldots 0.3162
N1N-1 highest moment of XtX_{t} 1111
M1M-1 highest moment of wtw_{t} 33
Δt\Delta t sampling time interval 10110^{-1} sec
LL length of synthetic data 10410^{4}
T=LΔtT=L\Delta t termination time 10310^{3} sec

4.2 Estimation performance

Two realizations of EC estimated [X^t]\Re[\hat{X}_{t}], with phase dynamics described in (ii) and (iii) are shown in Fig. 1 and Fig. 2, compared with EKF estimated [X^t]\Re[\hat{X}_{t}]. From both plots, with SNR = 0.1, the signal is submerged in large noise. However, the EC estimated signal (blue) is closer to the synthetic signal (yellow) compared with EKF estimated signal (red). When frequency experiences small wandering in scenario (iii), as displayed in Fig. 2, by estimating the product XmnX_{mn}, we can still extract X^t\hat{X}_{t} with higher accuracy than EKF estimated signal.

Refer to caption
Figure 1: Conditional mean estimate under scenario (ii): Plot of [X^t]\Re[\hat{X}_{t}] estimated by EC (blue) and EKF (red) at SNR = 0.1 with Θt\Theta_{t} defined in (5), where w0=0.012w_{0}=0.012Hz and qΘ=102q_{\Theta}=10^{-2}. The bottom panel is a magnified version of the top.
Refer to caption
Figure 2: Conditional mean estimate under scenario (iii): Plot of [X^t]\Re[\hat{X}_{t}] estimated by EC (blue) and EKF (red) at SNR = 0.1 with Θt\Theta_{t} defined in (6), where w0=0.032w_{0}=0.032Hz, qΘ=102q_{\Theta}=10^{-2} and qw=108q_{w}=10^{-8}. The bottom panel is a magnified version of the top.

4.3 Detection performance

To quantify the performance of the approximate EC detector, receiver operating characteristic (ROC) curves with different SNRs under scenario (ii) with qΘ=101q_{\Theta}=10^{-1} and w0=0.012w_{0}=0.012Hz are displayed in Fig. 3, compared with the EKF detector as well. In Fig. 4, ROCs of the EC and EKF detector under scenario (iii) with qΘ=101q_{\Theta}=10^{-1}, qw=108q_{w}=10^{-8} and w0=0.012w_{0}=0.012Hz are plotted. From Fig. 3, as SNR drops from 0.1 (top) to 0.0816 (middle) and then to 0.0577 (bottom), Pd of the EC detector (blue) descends from 0.9 to 0.6, then to 0.4 at Pf =102=10^{-2}. Fig. 4 exhibits a similar trend, with higher Pd as SNR gets larger. Comparing two top panels between Fig. 3 and Fig. 4, when the frequency is slowly wandering, with qw=108q_{w}=10^{-8} in Fig. 4, EC experiences a slight degradation, with Pd drops from 0.9 to 0.8 at SNR = 0.1 and Pf = 10210^{-2}. Throughout all six panels, the EC detector advantages over the EKF detector with higher Pd across the Pf range, especially when Pf <102<10^{-2}.

Refer to caption
Figure 3: ROC curves under scenario (ii): false alarm probability vs detection probability as SNR drops from 0.1 (top) to 0.0816 (middle) then to 0.0577 (bottom), with qΘ=101q_{\Theta}=10^{-1} and w0=0.012w_{0}=0.012Hz.
Refer to caption
Figure 4: ROC curves under scenario (iii): false alarm probability vs detection probability as SNR drops from 0.1 (top) to 0.0816 (middle) then to 0.0577 (bottom), with qΘ=101q_{\Theta}=10^{-1}, qw=108q_{w}=10^{-8} and w0=0.012w_{0}=0.012Hz.

5 Conclusion

In this paper, we focus on estimating and detecting random processes on the circle. The structure of the Neyman-Pearson (or EC) detector is split into two parts: a conditional mean estimate and a log-likelihood ratio detector. The conditional mean estimate, which is described by the nonlinear Kushner stochastic partial differential equation is approximated by filtering the first NN moments of the circle-valued signal. Instead of approximating the conditional density of the phase, we characterize the conditional density of the circle-valued signal directly, by pairing it with the moments of the signal. When the frequency is also a random process, in addition to filtering the moment function of the signal, we estimate the product of moment functions for frequency and signal, resulting in a deterministic generator.

We perform Monte Carlo simulations to verify the estimation performance by displaying realizations of the EC and EKF estimated signals and plotting the ROC curves for the detection, comparing also with the EKF detector. Overall, EC estimated signal has better tracking (estimating) capacity and higher detection probability than EKF based methods.

{ack}

The authors acknowledge support from the Australian Research Council (ARC) through the Centre of Excellence for Gravitational Wave Discovery (OzGrav) (grant number CE170100004) and an ARC Discovery Project (grant number DP170103625).

References

  • Condat (2021) Condat, L. (2021). Tikhonov regularization of circle-valued signals. 10.48550/ARXIV.2108.02602. URL https://arxiv.org/abs/2108.02602.
  • Coote et al. (2009) Coote, P., Trumpf, J., Mahony, R., and Willems, J.C. (2009). Near-optimal deterministic filtering on the unit circle. In Proceedings of the 48h IEEE Conference on Decision and Control (CDC) held jointly with 2009 28th Chinese Control Conference, 5490–5495. 10.1109/CDC.2009.5399999.
  • Foschini et al. (1988) Foschini, G.J., Greenstein, L.J., and Vannucci, G. (1988). Noncoherent detection of coherent lightwave signals corrupted by phase noise. IEEE Transactions on Communications, 36(3), 306–314.
  • Georghiades and Snyder (1985) Georghiades, C.N. and Snyder, D.L. (1985). A proposed receiver structure for optical communication systems that employ heterodyne detection and a semiconductor laser as a local oscillator. IEEE Trans. Commun., 33(4), 382–384. 10.1109/TCOM.1985.1096303. URL https://doi.org/10.1109/TCOM.1985.1096303.
  • Kailath and Poor (1998) Kailath, T. and Poor, H.V. (1998). Detection of stochastic processes. IEEE Transactions on Information Theory, 44(6), 2230–2231.
  • Kushner (1967) Kushner, H. (1967). Nonlinear filtering: The exact dynamical equations satisfied by the conditional mode. IEEE Transactions on Automatic Control, 12(3), 262–267.
  • Melatos et al. (2021) Melatos, A., Clearwater, P., Suvorova, S., Sun, L., Moran, W., and Evans, R.J. (2021). Hidden markov model tracking of continuous gravitational waves from a binary neutron star with wandering spin. iii. rotational phase tracking. Phys. Rev. D, 104, 042003. 10.1103/PhysRevD.104.042003. URL https://link.aps.org/doi/10.1103/PhysRevD.104.042003.
  • Rusch and Poor (1995) Rusch, L.A. and Poor, H.V. (1995). Effects of laser phase drift on coherent optical cdma. IEEE journal on selected areas in communications, 13(3), 577–591.
  • Scharf et al. (1980) Scharf, L., Cox, D., and Masreliez, C. (1980). Modulo-2π\pi phase sequence estimation (corresp.). IEEE Transactions on Information Theory, 26(5), 615–620. 10.1109/TIT.1980.1056241.
  • Snyder (1968) Snyder, D. (1968). The state-variable approach to analog communication theory. IEEE Transactions on Information Theory, 14(1), 94–104. 10.1109/TIT.1968.1054093.
  • Suvorova et al. (2021) Suvorova, S., Howard, S.D., and Moran, B. (2021). Tracking rotations using maximum entropy distributions. IEEE Transactions on Aerospace and Electronic Systems, 57(5), 2953–2968. 10.1109/TAES.2021.3067614.
  • Suvorova et al. (2018) Suvorova, S., Melatos, A., Evans, R.J., Moran, W., Clearwater, P., and Sun, L. (2018). Phase-continuous frequency line track-before-detect of a tone with slow frequency variation. IEEE Transactions on Signal Processing, 66(24), 6434–6442. 10.1109/TSP.2018.2877176.
  • Tam and Moore (1977) Tam, P. and Moore, J. (1977). A gaussian sum approach to phase and frequency estimation. IEEE Transactions on Communications, 25(9), 935–942. 10.1109/TCOM.1977.1093926.
  • Tretter (1985) Tretter, S. (1985). Estimating the frequency of a noisy sinusoid by linear regression (corresp.). IEEE Transactions on Information Theory, 31(6), 832–835. 10.1109/TIT.1985.1057115.
  • Veeravalli and Poor (1991) Veeravalli, V.V. and Poor, H.V. (1991). Quadratic detection of signals with drifting phase. The Journal of the Acoustical Society of America, 89(2), 811–819.
  • Willsky (1974) Willsky, A. (1974). Fourier series and estimation on the circle with applications to synchronous communication–i: Analysis. IEEE Transactions on Information Theory, 20(5), 577–583. 10.1109/TIT.1974.1055280.
  • Zamani et al. (2011) Zamani, M., Trumpf, J., and Mahony, R. (2011). Minimum-energy filtering on the unit circle.