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

Frequency Permutations
for Joint Radar and Communications

Rajitha Senanayake,  Peter Smith, 
Tian Han,  Jamie Evans, 
William Moran,  and Robin Evans
Abstract

This paper presents a new joint radar and communication technique based on the classical stepped frequency radar waveform. The randomization in the waveform, which is achieved by using permutations of the sequence of frequency tones, is utilized for data transmission. A new signaling scheme is proposed in which the mapping between incoming data and waveforms is performed based on an efficient combinatorial transform called the Lehmer code. Considering the optimum maximum likelihood (ML) detection, the union bound and the nearest neighbour approximation on the communication block error probability is derived for communication in an additive white Gaussian noise (AWGN) channel. The results are further extended to incorporate the Rician fading channel model, of which the Rayleigh fading channel model is presented as a special case. Furthermore, an efficient communication receiver implementation is discussed based on the Hungarian algorithm which achieves optimum performance with much less operational complexity when compared to an exhaustive search. From the radar perspective, two key analytical tools, namely, the ambiguity function (AF) and the Fisher information matrix are derived. Furthermore, accurate approximations to the Cramer-Rao lower bounds (CRLBs) on the delay and Doppler estimation errors are derived based on which the range and velocity estimation accuracy of the waveform is analysed. Numerical examples are used to highlight the accuracy of the analysis and to illustrate the performance of the proposed waveform.

Index Terms:
Joint radar and communications, error probability, maximum likelihood, ambiguity function, Fisher information matrix, Cramer-Rao lower bound.

I Introduction

Traditionally, wireless communications and radar sensing have been designed and developed in isolation addressing domain specific challenges. Recently, there has been significant research interest in the cooperation and co-design of the two systems [1, 2, 3, 4, 5, 6, 7]. This is primarily driven by the similarities of the two systems in terms of signal processing algorithms, devices and system architecture [5]. With wireless communication systems starting to use millimeter-wave (mmWave) frequencies, which are traditionally used in radar systems, such joint radar and communication systems could also help increase the spectrum utilization. Its potential applies to a range of emerging applications including, autonomous vehicles [8], unmanned aerial vehicles [9], the Internet of Things (IoT) [10] and defence applications where communications and sensing both play important roles [11].

I-A Related work

The main challenge in developing such a joint system lies in finding a new waveform that is capable of simultaneously transmitting information and performing radar sensing. Existing literature proposes several new techniques to achieve this convergence. In [4, 5, 6], excellent reviews of joint radar and communication techniques are provided.

Several papers have proposed communication-centric designs in which traditional communication waveforms are exploited for radar sensing purposes. In [8], a new waveform suitable for long-range radar (LRR) is proposed based on the IEEE 802.11ad-based waveform which is traditionally used for wireless local area networks (WLANs). The special structure of repeated Golay sequences in the preamble of the 802.11ad-based waveform, which has good correlation properties, is exploited to estimate the radar parameters. In [12], the IEEE 802.11p waveform, which is used for dedicated short-range communication (DSRC) in vehicles, is considered for radar range estimation of the closest target. The paper also presents an alternative to forward collision detection in vehicles based on the communication waveform. In [13], new algorithms based on wireless communication standards such as IEEE 802.11ad and IEEE 802.11p have been proposed for radar range and velocity estimation.

Orthogonal frequency division multiplexing (OFDM) waveforms, which incorporate frequency diversity, have also been proposed in [14, 15, 16], where the target range and Doppler is extracted from the reflected OFDM signal. The OFDM waveform also has similarities to phase coded radar. However, due to the addition of subcarrier components it introduces a high peak-to-average power ratio (PAPR) which is problematic in typical radar operation as it requires power amplifiers with a large linear range. While some attempt has been made to improve this by randomizing the carrier phase and amplitude [17], the transmit power amplifier is not used efficiently from a radar perspective. Some papers have proposed the use of cyclic prefixed orthogonal frequency division multiplexing (CP-OFDM) which has limitations involving the length of the cyclic prefix [18]. Similar to OFDM, spread spectrum waveforms such as direct sequence spread spectrum (DSSS) have also been proposed for joint radar and communications systems [19].

From the opposite perspective, several papers have focused on using conventional radar waveforms for joint radar and communication purposes. In [20, 21], pulse position modulation and pulse amplitude modulation are considered to embed data in pulsed radar signals. The popular linear frequency modulated (LFM) waveform is considered in [22], where separate up and down chirps are used for radar and communication pulses, respectively. In this work, communication data rates are further enhanced by using π/4\pi/4-Differential Quadrature Phase Shift Keying (DQPSK) modulation. The LFM waveform is also considered in [23], resulting in a low communication symbol rate that corresponds only to the chirp rate. Recently, a chirp-based MM-ary Frequency Shift Keying (chirp-MM-FSK) modulation method has been proposed in [24], together with the frequency modulated continuous wave (FMCW) radar waveform. Embedding data into traditional radar waveforms is challenging due to the lack of randomness in the available parameters. Also, when adjusting waveform parameters it is important to make sure radar performance is not compromised. None of the above works characterise the effect of embedding data on the radar performance in-terms of key analytical tools in radar waveform design such as the ambiguity function and the Fisher information matrix. As such, analysis of the behavior of new waveforms under important phenomena such as radar clutter is lacking.

I-B Motivation and Contributions

In this paper, we focus on the stepped frequency radar waveform, which, to the best of our knowledge, has not been considered for this joint operation before, and propose a new method to randomize the frequency tones to modulate data. In order to make the paper complete and clearly readable to both the communications and radar communities we have conducted a rigorous analytical investigation on a number of important performance measures. While some of these measures are considered in the literature they are not always rigorously developed. We hope that the level of detail presented here is helpful and valuable to the reader.

Radar waveforms consisting of a sequence of step changes in frequency are popular in many radar applications and are being used extensively in the emerging automotive radar application [25, 26, 27]. In a stepped frequency waveform, the available bandwidth is divided into equally spaced frequency tones separated by a constant step Δf\Delta f Hz. The most common stepped frequency waveform employs a linear stepping pattern where the frequency of each pulse is increased by Δf\Delta f from the preceding pulse [28]. Costas codes, on the other hand, arrange the frequency tones according to a special sequence that results in an extremely good radar waveform with zero side-lobes [29]. However, the number of available Costas codes are limited and it does not provide enough randomization to be used in a joint radar communication system. Interestingly, in [25, 26], a pseudo random stepped frequency radar is considered where the transmitted waveforms consist of a train of short tones with frequencies defined according to a pseudo-random sequence. The randomization in the sequence of tones appears to be more suitable for the automotive radar application as it can significantly reduce the interference arising from a large number of radars operating in close proximity [25].

In this paper, we exploit this randomization in frequency to send data. We focus on a random stepped frequency radar waveform based on Lehmer codes, in which the transmitted waveform consists of a train of short tones with frequencies defined according to a permutation of the sequence of tones. As such, a set of MM frequency tones generates M!M! different waveforms - all suitable for typical radar applications. These waveforms also maintain a constant amplitude resulting in a PAPR of unity which is ideally suited to radar operation. Our novel contributions are detailed as follows:

  • We propose a novel joint radar and communication technique in which the incoming data is modulated based on the selection of the frequency permutation in the transmitted waveform. We provide an efficient implementation for the mapping between the incoming data and the corresponding waveform based on a combinatorial transform called the Lehmer code.

  • From a communication perspective, we consider optimum maximum likelihood (ML) detection and analyse the baseband communication in additive white Gaussian noise (AWGN) channels to provide a union bound and a nearest neighbour approximation on the block error probability of the proposed signaling scheme. Then, we extend the analysis to consider Rician fading and provide closed-form analytical expression for the same performance measures. The Rayleigh channel model is also considered as a special case.

  • We provide an efficient communication receiver implementation based on the Hungarian algorithm which always results in the optimum solution. Compared to the exhaustive search which has a time complexity of O(M!)O(M!), the Hungarian algorithm has a worst case complexity of O(M3)O(M^{3}), where MM is the number of frequency tones in the waveform.

  • From a radar perspective, we derive the ambiguity function and the Fisher information matrix of the proposed waveform. Based on the ambiguity function, we analyse the behavior of the waveform in relation to the matched filter output and provide a detailed discussion on the overall structure of the ambiguity function. Based on the Fisher information matrix we derive approximate Cramer-Rao Lower Bounds (CRLBs) on the delay and Doppler estimation error and illustrate the effect of different parameters on radar range and Doppler resolution.

Numerical examples are provided to support the discussion and also to illustrate the accuracy of the analysis. The remainder of the paper is organized as follows. In Section II, the system model and the waveform of the novel joint radar and communications system are described. The communication and radar performance of the Lehmer code based random stepped frequency radar waveform is analysed in Section III and IV, respectively. Simulation results are provided in Section V with conclusions and future research directions provided in Section VI.

II Joint Radar and Communications System

II-A System Model

Consider the system model illustrated in Fig. 1, where the transmitter sends a joint waveform for both communication and radar systems. The return signal from the radar target is received and processed by the radar receiver to estimate the range and the velocity of the target. The signal received at the communications receiver, which is equipped with multiple antennas, is processed to detect the transmitted information.

II-B Joint Radar and Communications Waveform

Refer to caption

Figure 1: Illustration on the joint radar communication system model.

Consider MM equally spaced frequency tones f0,f1,,fM1f_{0},f_{1},\ldots,f_{M-1}. Based on the fact that there are M!M! permutations of this sequence of MM frequencies we formulate M!M! stepped frequency radar waveforms. In each waveform a given frequency is used only once. As illustrated in Fig. 2, the ii-th waveform is generated using MM pulses, each TT seconds long and, the frequencies follow the ii-th permutation of the set of M!M! permutations with fmif^{i}_{m} denoting the mm-th frequency of the ii-th permutation. As such, the complex baseband representation of the ii-th waveform is given by

si(t)=EMTm=0M1sp(tmT)exp(2πfmi(tmT)),\displaystyle s_{i}(t)=\sqrt{\frac{E}{MT}}\sum_{m=0}^{M-1}s_{p}(t-mT)\exp(2\pi f^{i}_{m}(t-mT)), (1)

where

sp(t)={10tT0otherwise.\displaystyle s_{p}(t)=\left\{\begin{array}[]{ll}1~{}~{}~{}~{}~{}~{}~{}~{}0\leq t\leq T\\ 0~{}~{}~{}~{}~{}~{}~{}~{}\text{otherwise}.\end{array}\right. (4)

TT denotes the pulse duration and EE is the signal energy satisfying

0MTsi2(t)𝑑t=E,i=0,1,,M1.\displaystyle\int^{MT}_{0}s_{i}^{2}(t)dt=E,~{}~{}~{}~{}~{}~{}~{}~{}i=0,1,\ldots,M-1.

We also assume that the frequencies are orthogonal. Thus, the separation between two frequency tones Δf=n/T\Delta f=n/T, where nn is an integer. Since all frequencies are permutations of the MM frequency tones, the set of M!M! sequences are given by

ψ1={f0,,fM2,fM1},\displaystyle\psi_{1}=\{f_{0},\ldots,f_{M-2},f_{M-1}\},
ψ2={f0,,fM1,fM2},\displaystyle\psi_{2}=\{f_{0},\ldots,f_{M-1},f_{M-2}\},
\displaystyle~{}~{}~{}~{}~{}\vdots
ψM!={fM1,,f1,f0}.\displaystyle\psi_{M!}=\{f_{M-1},\ldots,f_{1},f_{0}\}.

We note that all M!M! waveforms generated by the above permutations maintain a constant amplitude. This, in contrast to OFDM, results in a PAPR of unity which is suitable for typical radar operation. Moreover, these waveforms maintain a short individual pulse width TT which results in good delay resolution. A good Doppler resolution is obtained by changing the frequency from one pulse to another, thus, utilizing a total bandwidth of MΔfM\Delta f when considering the waveform as a whole. Due to the randomness in frequency, all M!M! waveforms generated by the permutations are suitable for the automotive radar application [27]. We embed data in the selection of the waveform by maintaining a direct mapping between the data symbol and the selected waveform. In other words, we take log2M!\log_{2}{M!} number of bits and depending on the symbol it represents, select the waveform to be transmitted.

Refer to caption

Figure 2: The envelope of the ii-th Lehmer code based random stepped frequency waveform.

II-C Use of Lehmer Codes

Since we have M!M! possible waveforms, one could argue that for large MM, the implementation of the mapping between incoming data and the corresponding waveform may not be practical as it requires a large lookup table. However, it is important to identify that the the use of a lookup table is unnecessary for the implementation of the proposed signaling scheme. The mapping between incoming data and the corresponding waveform can be efficiently implemented using a combinatorial transform called the Lehmer code [30]. With Lehmer codes, we first present the natural number of the incoming data symbol in the factorial number system, which is also known as the factoradic form. Then, we exploit the natural mapping between the integers 0,1,,M!10,1,\ldots,M!-1 and permutations of MM elements in lexicographic order to map incoming data to a unique permutation of frequency tones [31]. At the communications receiver, a similar decoding procedure is followed to map the detected permutation of frequency tones to the corresponding data symbol. This procedure is a straightforward mapping between integers and permutations and can be efficiently implemented as it only involves simple operations such as selection from a sequence and deletion from it. For more details on mapping and demapping using Lehmer codes please refer to [31].

III Communication Performance Analysis

In this section, we focus on the communications performance of the proposed joint radar and communications system. We consider ML detection at the communications receiver and analyse the error probability performance.

III-A Maximum Likelihood Detection

Consider a communication receiver with NN receive antennas. The N×1N\times 1 complex baseband received signal vector is given by

𝐫(t)=𝐡si(t)+𝐧(t),\displaystyle\mathbf{r}(t)=\mathbf{h}s_{i}(t)+\mathbf{n}(t), (5)

where 𝐡\mathbf{h} is the small-scale fading channel vector, si(t){s1(t),s2(t),,sM!(t)}s_{i}(t)\in\{s_{1}(t),s_{2}(t),\ldots,s_{M!}(t)\} is one of the M!M! possible waveforms bearing transmit data and 𝐧(t)\mathbf{n}(t) is an AWGN vector where each element is complex Gaussian with zero mean and variance N0N_{0}.

Assuming that the channel vector 𝐡\mathbf{h} is known at the receiver, we consider a coherent ML detector and write write the detected symbol as

s^i(t)=arg minsk(t){s1(t),s2(t),,sM!(t)}0MT|𝐡H𝐫(t)𝐡H𝐡sk(t)|2𝑑t.\displaystyle\hat{s}_{i}(t)={\underset{{s_{k}}(t)\in\{{s}_{1}(t),{s}_{2}(t),\ldots,{s}_{M!}(t)\}}{\text{arg min}}}\int_{0}^{MT}|\mathbf{h}^{H}\mathbf{r}(t)-\mathbf{h}^{H}\mathbf{h}{s_{k}}(t)|^{2}dt. (6)

Using some straight forward mathematical manipulations, the ML detection rule in (6) can be reexpressed as

s^i(t)=arg maxsk(t){s1(t),s2(t),,sM!(t)}Re(0MTsk(t)𝐡H𝐫(t)𝑑t),\displaystyle\hat{s}_{i}(t)={\underset{{s}_{k}(t)\in\{{s}_{1}(t),{s}_{2}(t),\ldots,{s}_{M!}(t)\}}{\text{arg max}}}\textrm{Re}\left(\int_{0}^{MT}{s_{k}}^{*}(t)\mathbf{h}^{H}\mathbf{r}(t)dt\right), (7)

where Re(.)\textrm{Re}(.) denotes the real part of the argument.

III-B Error Probability Performance

Consider the hypothesis that si(t)s_{i}(t), which corresponds to the the ii-th sequence, is being transmitted. The sequence is correctly detected if s^i(t)=si(t)\hat{s}_{i}(t)={s}_{i}(t). Thus, the probability of a correct decision is given by

Pc(i)=Pr[ξii=maxk{1,2,M!}ξik],\displaystyle P_{c}(i)=\textrm{Pr}\left[\xi_{ii}={\underset{k\in\{1,2,\ldots M!\}}{\textrm{max}}}\xi_{ik}\right], (8)

where

ξik=Re(0MTsk(t)𝐡H𝐫(t)𝑑t)\displaystyle\xi_{ik}=\textrm{Re}\left(\int_{0}^{MT}{s}^{*}_{k}(t)\mathbf{h}^{H}\mathbf{r}(t)dt\right)

denotes the decision variable. The transmitted data is assumed to be equally likely. Thus, the a priori probabilities of the M!M! sequences are 1/M!1/M!, and the error probability performance of the sequence can be defined based on a block error rate as

Pe=1M!i=1M![1Pc(i)].\displaystyle P_{e}=\frac{1}{M!}\sum_{i=1}^{M!}[1-P_{c}(i)]. (9)

The exact expression of PeP_{e} requires complex multi-dimensional integrals over the multivariate Gaussian density. These integrals cannot be computed in closed-form. Thus, we take a more tractable approach by considering the union bound which is given by

PePeUB=1M!i=1M!k=1,kiM!Pik,\displaystyle P_{e}\leq P_{e}^{UB}=\frac{1}{M!}\sum_{i=1}^{M!}\sum_{k=1,k\neq i}^{M!}P_{ik}, (10)

where PikP_{ik} denotes the pairwise error probability of detecting the kk-th sequence when the ii-th sequence is transmitted. Please note that while the union bound is commonly used in literature, its derivation in the present context is not straight forward as we consider an unconventional method of modulating data based on the frequency permutation. Let !x!x, given by

!x=x!i=0x(1)ii!,\displaystyle!x=x!\sum_{i=0}^{x}\frac{(-1)^{i}}{i!}, (11)

denote the number of derangments of an xx-element set [32]. By analysing the M!M! permutations of the sequence of MM frequency tones we note that we have !l(MMl)!l\binom{M}{M-l} pairwise error probabilities where the frequency tones differ by ll positions. Using this observation and considering the symmetry in pairwise error probabilities, we can further simplify (10) as

PeUB=l=2M!l(MMl)P1kl,\displaystyle P_{e}^{UB}=\sum_{l=2}^{M}!l\binom{M}{M-l}P_{1{k_{l}}}, (12)

where kl{1,2,,M}k_{l}\in\{1,2,\ldots,M\} is such that ψ1\psi_{1} and ψkl\psi_{k_{l}}, which are the frequency sequences used to generate s1(t)s_{1}(t) and skl(t)s_{k_{l}}(t), respectively, differ by ll frequency tones. Next we proceed to derive P1klP_{1k_{l}} which is the probability of detecting skl(t)s_{k_{l}}(t) when s1(t)s_{1}(t) is transmitted. Using the ML detection rule we can write P1klP_{1k_{l}} as

P1kl=Pr\displaystyle P_{1k_{l}}=\textrm{Pr} [Re(0MTs1(t)𝐡H𝐫(t)𝑑t)<Re(0MTskl(t)𝐡H𝐫(t)𝑑t)].\displaystyle\left[\textrm{Re}\left(\int_{0}^{MT}{s}^{*}_{1}(t)\mathbf{h}^{H}\mathbf{r}(t)dt\right)<\textrm{Re}\left(\int_{0}^{MT}{s}^{*}_{k_{l}}(t)\mathbf{h}^{H}\mathbf{r}(t)dt\right)\right]. (13)

Using some mathematical manipulations (13) can be reexpressed as

P1kl=Pr[ElM𝐡H𝐡<N2N1],\displaystyle P_{1k_{l}}=\textrm{Pr}\left[\frac{El}{M}\mathbf{h}^{H}\mathbf{h}<N_{2}-N_{1}\right], (14)

where N1=Re(𝐡H0MTs1(t)𝐧(t)𝑑t)N_{1}=\textrm{Re}\left(\mathbf{h}^{H}\int_{0}^{MT}{s}^{*}_{1}(t)\mathbf{n}(t)dt\right) and N2=Re(𝐡H0MTskl(t)𝐧(t)𝑑t)N_{2}=\textrm{Re}\left(\mathbf{h}^{H}\int_{0}^{MT}{s}^{*}_{k_{l}}(t)\mathbf{n}(t)dt\right). When conditioned on channel fading, N2N1N_{2}-N_{1} has a Gaussian distribution with zero mean and variance ElN0M𝐡H𝐡\frac{ElN_{0}}{M}\mathbf{h}^{H}\mathbf{h}. Thus, we can write a simple expression for P1klP_{1k_{l}} as

P1kl=Pr\displaystyle P_{1k_{l}}=\textrm{Pr} [𝐡H𝐡<αlZ],\displaystyle\left[\sqrt{\mathbf{h}^{H}\mathbf{h}}<\alpha_{l}Z\right], (15)

where αl=N0MEl\alpha_{l}=\sqrt{\frac{N_{0}M}{El}} and Z𝒩(0,1)Z\sim\mathcal{N}(0,1). Next, we proceed to solve (15) with respect to different wireless communication channel models.

III-B1 AWGN Channel

To gain fundamental insights on the error performance, first, let us focus on AWGN channels. Without loss of generality we assume a unit channel gain so that 𝐡H𝐡=N\mathbf{h}^{H}\mathbf{h}=N, which results in

P1kl=𝒬(Nαl),\displaystyle P_{1k_{l}}=\mathcal{Q}\left(\frac{\sqrt{N}}{\alpha_{l}}\right), (16)

where Q(.)Q(.) is the Gaussian Q-function. Substituting (16) into (12) the union bound on the block error rate under AWGN channels can be written as

PeUB=l=2M!l(MMl)𝒬(Nαl).\displaystyle P_{e}^{UB}=\sum_{l=2}^{M}!l\binom{M}{M-l}\mathcal{Q}\left(\frac{\sqrt{N}}{\alpha_{l}}\right). (17)

While the union bound is commonly used in communications to provide a simple upper bound on the error probability performance, it has some limitations, especially for large MM. The union bound considers all the pairwise error probabilities when calculating the error probability. Therefore, as we increase MM the bound is not very accurate, especially in the low SNR regime. To avoid such limitations, in the following, we derive another approximation based on the nearest neighbour approximation [33]. Under this approach, we only consider the pairwise error probabilities corresponding to the nearest neighbours. As the waveforms are generated using permutations of frequency tones we note that the nearest neighbours only differ by two frequency tones resulting in a new approximation given by

PeNN=M(M1)2𝒬(Nα2),\displaystyle P_{e}^{NN}=\frac{M(M-1)}{2}\mathcal{Q}\left(\frac{\sqrt{N}}{\alpha_{2}}\right), (18)

where M(M1)M(M-1) is the number of pairwise error probabilities where the frequency tones differ by two positions.

III-B2 Rician Fading Channel

In the context of vehicular communication, fading in the propagation channel is inevitable due several reasons including the low position of antennas, large numbers of reflecting metallic surfaces surrounding the transmitters and the high mobility of transmitters and receivers [34]. To gain further insights into the effect of fading, next, we focus on the independent Rician fading model in which a line-of-sight (LOS) signal is present, along with other scatterers. As such, we write the N×1N\times 1 small scale fading channel vector as

𝐡=KK+1Δ+1K+1𝐮,\displaystyle\mathbf{h}=\sqrt{\frac{K}{K+1}}\Delta+\sqrt{\frac{1}{K+1}}\mathbf{u}, (19)

where Δ\Delta denotes the complex LOS phase vector with the ii-th element having the property |Δi|2=1|\Delta_{i}|^{2}=1, 𝐮\mathbf{u} denotes the scattered component vector with the ii-th element ui𝒞𝒩(0,1)u_{i}\sim\mathcal{CN}(0,1). The strength of the LOS path is governed by the Rician factor given by KK. Based on (19) we can write 𝐡H𝐡\mathbf{h}^{H}\mathbf{h} as

𝐡H𝐡=12(K+1)χ2N2(2NK),\displaystyle\mathbf{h}^{H}\mathbf{h}=\frac{1}{2(K+1)}\chi_{2N}^{2}(2NK), (20)

where χ2N2(2NK)\chi_{2N}^{2}(2NK) denotes a non-central chi-squared distribution with 2N2N degrees of freedom and non-centrality parameter 2NK2NK. The main steps of the derivation of (20) is given in Appendix A. Substituting (20) into (15) we can write

P1kl=Pr\displaystyle P_{1k_{l}}=\textrm{Pr} [χ2N2(2NK)<clZ],\displaystyle\left[\sqrt{\chi_{2N}^{2}(2NK)}<\sqrt{c_{l}}Z\right], (21)

where cl=2αl2(K+1)c_{l}={2\alpha_{l}^{2}(K+1)} and the probability in (21) is taken both with respect to χ2N2(2NK)\chi_{2N}^{2}(2NK) and ZZ. Based on the CDF of the non-central chi-squared distribution [35], we can reexpress (21) as

P1kl=j=0((NK)jeNKj!)Pr[χ2N+2j2<clZ].\displaystyle P_{1k_{l}}=\sum_{j=0}^{\infty}\left(\frac{(NK)^{j}e^{-NK}}{j!}\right)\textrm{Pr}\left[\sqrt{\chi_{2N+2j}^{2}}<\sqrt{c_{l}}Z\right]. (22)

Noting that ZZ has a standard normal Gaussian distribution and χ2N+2j2\chi_{2N+2j}^{2} has a chi-squared distribution with 2N+2j2N+2j degrees of freedom we proceed to solve the above probability as

P1kl=j=0((NK)jeNKj!)0Q(2xcl)xN+j1exΓ(N+j)𝑑x.\displaystyle P_{1k_{l}}=\sum_{j=0}^{\infty}\left(\frac{(NK)^{j}e^{-NK}}{j!}\right)\int_{0}^{\infty}Q\left(\sqrt{\frac{2x}{c_{l}}}\right)\frac{x^{N+j-1}e^{-x}}{\Gamma(N+j)}dx. (23)

where Q(.)Q(.) is the Gaussian Q-function and Γ(.)\Gamma(.) is the Gamma function [36]. Using Craig’s formula to reexpress the Gaussian Q-function we simplify the integral in (23) as

P1kl=1πj=0((NK)jeNKj!)0π2(sin2θcl1+sin2θ)N+j𝑑θ,\displaystyle P_{1k_{l}}=\frac{1}{\pi}\sum_{j=0}^{\infty}\left(\frac{(NK)^{j}e^{-NK}}{j!}\right)\int_{0}^{\frac{\pi}{2}}\left(\frac{\sin^{2}{\theta}}{c_{l}^{-1}+\sin^{2}{\theta}}\right)^{N+j}d\theta, (24)

which can be evaluated in closed-form using [37, eq. 69] to produce

P1kl=j=0((NK)jeNKj!)(12+n=1N+j(1)n(N+jn)νn(cl)),\displaystyle P_{1k_{l}}=\sum_{j=0}^{\infty}\left(\frac{(NK)^{j}e^{-NK}}{j!}\right)\left(\frac{1}{2}+\sum_{n=1}^{N+j}(-1)^{n}\binom{N+j}{n}\nu_{n}(c_{l})\right), (25)

where νn(cl)=12(1+cl)n1/2q=0n1(n1q)(2qq)(cl4)q\nu_{n}(c_{l})=\frac{1}{2(1+c_{l})^{n-1/2}}\sum_{q=0}^{n-1}\binom{n-1}{q}\binom{2q}{q}\left(\frac{c_{l}}{4}\right)^{q}. Substituting (25) into (12) the final closed-form expression for the union bound on the block error rate can be written as

PeUB=l=2Mj=0!l(MMl)((NK)jeNKj!)(12+n=1N+j(1)n(N+jn)νn(cl)).\displaystyle P_{e}^{UB}=\sum_{l=2}^{M}\sum_{j=0}^{\infty}!l\binom{M}{M-l}\left(\frac{(NK)^{j}e^{-NK}}{j!}\right)\left(\frac{1}{2}+\sum_{n=1}^{N+j}(-1)^{n}\binom{N+j}{n}\nu_{n}(c_{l})\right). (26)

Similar to the AWGN channel, the nearest neighbour approximation under the Rician fading channel model is given by

PeNN=j=0(M(M1)(NK)jeNK2j!)(12+n=1N+j(1)n(N+jn)νn(c2)).\displaystyle P_{e}^{NN}=\sum_{j=0}^{\infty}\left(\frac{M(M-1)(NK)^{j}e^{-NK}}{2j!}\right)\left(\frac{1}{2}+\sum_{n=1}^{N+j}(-1)^{n}\binom{N+j}{n}\nu_{n}(c_{2})\right). (27)

In Section V, we illustrate the accuracy of the above analysis using numerical examples and further discuss the error probability performance of the proposed signaling scheme under Rician fading channels.

III-B3 Special Case - Rayleigh Fading Channel

Next, we focus on the independent Rayleigh fading model which is a special case of Rician fading. By setting K=0K=0 in (26), we can derive the union bound on the block error rate as

PeUB=l=2M!l(MMl)(12+n=1N(1)n(Nn)νn(cl)).\displaystyle P_{e}^{UB}=\sum_{l=2}^{M}!l\binom{M}{M-l}\left(\frac{1}{2}+\sum_{n=1}^{N}(-1)^{n}\binom{N}{n}\nu_{n}(c_{l})\right). (28)

Similarly, the nearest neighbour approximation under Rayleigh distribution can be written as

PeNN=M(M1)2(12+n=1N(1)n(Nn)νn(c2)).\displaystyle P_{e}^{NN}=\frac{M(M-1)}{2}\left(\frac{1}{2}+\sum_{n=1}^{N}(-1)^{n}\binom{N}{n}\nu_{n}(c_{2})\right). (29)

This provides an accurate approximation of the block error probability when there is no LOS signal present and the communication channel fading is Rayeigh distributed. In Section V, we illustrate the accuracy of the above analysis using numerical examples.

III-C Communications Receiver Implementation

We propose an efficient method for the implementation of the optimal communications receiver under the proposed modulation method. An optimal receiver for the fading channel implements the ML decision rule, which is given in (7), where the detected signal maximizes correlation between the transmitted signal and the received signal correlated with the channel vector. This can be implemented using a correlation receiver (or a matched filter) using the relation

rnm=Re((n1)TnT𝐡H𝐫(t)ϕm(t(n1)T)𝑑t),\displaystyle r_{nm}=\textrm{Re}\left(\int_{(n-1)T}^{nT}\mathbf{h}^{H}\mathbf{r}(t)\phi_{m}(t-(n-1)T)dt\right), (30)

where the basis function ϕm(t)\phi_{m}(t) is defined by ϕm(t)=2Tsp(t)cos(2πfm(t))\phi_{m}(t)=\sqrt{\frac{2}{T}}s_{p}(t)\cos(2\pi f_{m}(t)). Based on (30), we formulate the following matrix

R=(rnm)M×M,\displaystyle\textbf{R}=(r_{nm})\in\mathbb{R}^{M\times M}, (31)

in which the nmnm-th element, rnmr_{nm}, represents the correlation between the received signal and the basis function ϕm(t(n1)T)\phi_{m}(t-(n-1)T). Now, the ML detection rule reduces to selecting the MM elements from R in such a way that the selected MM elements are in MM different rows (because the frequencies are not repeated in a given waveform) and MM different columns (because only one frequency is used during a given pulse duration) and the sum of the MM elements is maximized.

We note that this optimization is an assignment problem which can be solved efficiently using the Hungarian algorithm [38]. More specifically, we formulate the cost matrix by negating R and follow the Hungarian algorithm to pick at most one element from each row and column in such a way that the summation is maximized. It is important to note that the Hungarian algorithm always results in the optimum solution. For the sake of completion, in Algorithm 1, we summarize the main steps of the Hungarian algorithm that we follow in order to obtain the optimum solution. A working example of the algorithm is also provided in B. For further details please refer to [38]. Compared to the exhaustive search which has a time complexity of O(M!)O(M!), the Hungarian algorithm has a worst case complexity of O(M3)O(M^{3}).

Algorithm 1 Hungarian Algorithm for the Communications Receiver
1:Negate the cost matrix R to produce (R)(-\textbf{R})
2:Subtract the row minimum from each row in (R)(-\textbf{R}) and assign the resulting matrix to R~\tilde{\textbf{R}}
3:Subtract the column minimum from each column in R~\tilde{\textbf{R}} and assign the resulting matrix to R¯\bar{\textbf{R}}
4:count=0count=0
5:while countMcount\not=M do
6:     countcount = minimum number of lines required to cover all zeros in R¯\bar{\textbf{R}}.
7:Note that covering a zero means putting a line through the row with that zero. The maximum number of lines that can be drawn through a row is one.
8:     if count==Mcount==M then
9:         break;
10:     else
11:         rmin=r_{min}= the smallest uncovered number in R¯\bar{\textbf{R}}
12:         Update R¯\bar{\textbf{R}} by subtracting rminr_{min} from all uncovered numbers and add rminr_{min} to all the elements that are covered twice
13:     end if
14:end while
15:The values of the optimal assignment are covered by the zeros in R¯\bar{\textbf{R}}

IV Radar Performance Analysis

So far, we have considered the communications performance of the Lehmer code based random stepped frequency radar waveform. In this section, we focus on the radar performance of this waveform and derive the two key analytical tools in radar waveform design, namely, the ambiguity function and the Fisher information matrix based on which we evaluate the side-lobe structure of the waveform and derive the CRLBs on velocity and range estimation. Using these analytical tools we fully characterize the delay and Doppler response of the radar waveform and discuss the effect of embedding data on the target range and velocity estimation.

IV-A Ambiguity Function

In radar waveform design, the time-frequency autocorrelation function of the transmitted waveform provides a measure of the degree of similarity between the complex envelope of the transmitted waveform and a replica of it that is shifted in time and frequency [39]. The ambiguity function is defined as the magnitude of the time-frequency autocorrelation function and it fully characterizes the behavior of a radar waveform in relation to the matched filter output [28, 40]. More specifically, it describes the output of a matched filter when the signal input to the radar receiver is delayed by τ\tau and Doppler shifted by ω\omega. The ambiguity function is useful for examining several radar parameters including, resolution, side-lobe structure and accuracy in both delay and Doppler. In fact, the local accuracy of the delay and Doppler estimates are characterized by the behavior of the ambiguity function around the origin, while the global accuracy is characterized by its broader structure [39].

In order to derive the ambiguity function, we re-write the expression in (1) as

si(t)=1MTm=0M1sp(tmT)exp[jωmi(tmT)],\displaystyle s_{i}(t)=\sqrt{\frac{1}{MT}}\sum_{m=0}^{M-1}s_{p}(t-mT)\exp[j\omega^{i}_{m}(t-mT)], (32)

where ωmi=2πfmi\omega^{i}_{m}=2\pi f^{i}_{m} and without loss of generality we have normalized the signal energy to one. Based on [28, eq. (4.30)], the time-frequency autocorrelation function, which is also known as the complex ambiguity function, of the above waveform can be written using si(t)s_{i}(t) and its complex conjugate si(t)s_{i}^{*}(t) as

A^(τ,ω)=si(s)si(sτ)exp(jωs)𝑑s,\displaystyle\hat{A}(\tau,\omega)=\int_{-\infty}^{\infty}s_{i}(s)s_{i}^{*}(s-\tau)\exp(j\omega s)ds, (33)

where τ\tau is the time delay relative to the expected matched filter peak output and ω\omega is the Doppler mismatch. Substituting (32) into (33) we can write

A^(τ,ω)=1MTm=0M1\displaystyle\hat{A}(\tau,\omega)=\frac{1}{MT}\sum_{m=0}^{M-1} n=0M1sp(smT)sp(sτnT)ej(ωs+ωmi(smT)ωni(sτnT))𝑑s.\displaystyle\sum_{n=0}^{M-1}\int_{-\infty}^{\infty}s_{p}(s-mT)s_{p}^{*}(s-\tau-nT)e^{j(\omega s+\omega_{m}^{i}(s-mT)-\omega_{n}^{i}(s-\tau-nT))}ds. (34)

Substituting s=smTs^{\prime}=s-mT and rearranging (34) we get

A^(τ,ω)=1MT\displaystyle\hat{A}(\tau,\omega)=\frac{1}{MT} m=0M1n=0M1ejωmTsp(s)sp(sτnT+mT)ej(ωs+ωmisωni(sτnT+mT))𝑑s.\displaystyle\sum_{m=0}^{M-1}\sum_{n=0}^{M-1}e^{j\omega mT}\int_{-\infty}^{\infty}s_{p}(s^{\prime})s_{p}^{*}(s^{\prime}-\tau-nT+mT)e^{j(\omega s^{\prime}+\omega_{m}^{i}s^{\prime}-\omega_{n}^{i}(s^{\prime}-\tau-nT+mT))}ds^{\prime}. (35)

If the complex ambiguity function of the simple pulse sp(t)s_{p}(t) is represented by A^p(τ,ω)\hat{A}_{p}(\tau,\omega), the integral in (35) can be reexpressed as

A^(τ,ω)=1MTm=0M1n=0M1\displaystyle\hat{A}(\tau,\omega)=\frac{1}{MT}\sum_{m=0}^{M-1}\sum_{n=0}^{M-1} A^p(τ+nTmT,ωωni+ωmi)ej(ωmTωni((mn)Tτ)),\displaystyle\hat{A}_{p}(\tau+nT-mT,\omega-\omega_{n}^{i}+\omega_{m}^{i})e^{j(\omega mT-\omega_{n}^{i}((m-n)T-\tau))}, (36)

where A^p(τ,ω)\hat{A}_{p}(\tau,\omega) is given by

Ap^(τ,ω)={ejωτ(ejωTejωτjω)Tτ<0ejωTejωτjω0τT0otherwise.\displaystyle\hat{A_{p}}(\tau,\omega)=\left\{\begin{array}[]{ll}e^{j\omega\tau}\left(\frac{e^{j\omega T}-e^{-j\omega\tau}}{j\omega}\right)~{}~{}~{}~{}~{}~{}~{}~{}-T\leq\tau<0\\ \frac{e^{j\omega T}-e^{j\omega\tau}}{j\omega}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}0\leq\tau\leq T\\ 0~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\text{otherwise}.\end{array}\right. (40)

Then, the ambiguity function of the waveform is defined as the magnitude of A^(τ,ω)\hat{A}(\tau,\omega)

A(τ,ω)=|A^(τ,ω)|.\displaystyle{A}(\tau,\omega)=|\hat{A}(\tau,\omega)|. (41)

The expression in (41) succinctly characterizes the behavior of the Lehmer code based stepped frequency radar waveform in the whole delay-Doppler plane.

The matched filter output when there is no Doppler mismatch is given by the zero-Doppler response A(τ,0){A}(\tau,0), which can be found by setting ω=0\omega=0 in (41) which results in

A(τ,0)\displaystyle{A}(\tau,0) =|1MTm=0M1n=0M1A^p(τ+(nm)T,ωni+ωmi)ej(ωni((mn)Tτ))|.\displaystyle=\left|\frac{1}{MT}\sum_{m=0}^{M-1}\sum_{n=0}^{M-1}\hat{A}_{p}(\tau+(n-m)T,-\omega_{n}^{i}+\omega_{m}^{i})e^{j(-\omega_{n}^{i}((m-n)T-\tau))}\right|. (42)

Note that A(τ,0){A}(\tau,0) is the autocorrelation function of si(t)s_{i}(t). Similarly, the matched filter output when there is no delay is given by the zero-delay response A(0,ω){A}(0,\omega), which can be found by setting τ=0\tau=0 in (41) which results in

A(0,ω)\displaystyle{A}(0,\omega) =|1MTm=0M1n=0M1A^p((nm)T,ωωni+ωmi)ej(ωmTωni((mn)T))|.\displaystyle=\left|\frac{1}{MT}\sum_{m=0}^{M-1}\sum_{n=0}^{M-1}\hat{A}_{p}((n-m)T,\omega-\omega_{n}^{i}+\omega_{m}^{i})e^{j(\omega mT-\omega_{n}^{i}((m-n)T))}\right|. (43)

Note that A(0,ω){A}(0,\omega) is the Fourier transform of the magnitude squared of si(t)s_{i}(t). In radar, the zero-Doppler response and the zero-delay response are used to quantify the range estimation accuracy and the velocity estimation accuracy, respectively [28]. This is further discussed using numerical examples in Section V.

IV-B Fisher Information Matrix

For the ii-th transmitted waveform, we define the complex envelope of the received radar waveform as given in [39, eq. 3]

r(t)=b~si(tτ)ejωt+n(t),\displaystyle r(t)=\tilde{b}s_{i}(t-\tau)e^{j\omega t}+n(t), (44)

where τ\tau and ω\omega are the unknown delay and Doppler to be estimated, the multiplier b~\tilde{b} represents fading induced by multiple reflecting surfaces on the target and is modelled by a zero mean and unit variance complex Gaussian random variable and n(t)𝒞𝒩(0,N0)n(t)\sim\mathcal{CN}(0,N_{0}) represents the AWGN. Then, the 2×22\times 2 Fisher information matrix for τ\tau and ω\omega is defined as [39]

𝐉=[J11J12J21J22,]\displaystyle{\bf{J}}=\begin{bmatrix}J_{11}&J_{12}\\ J_{21}&J_{22},\end{bmatrix} (45)

where we identify subscript 1 with τ\tau and subscript 2 with ω\omega. Based on [39, eq. (63) - (65)] we can write the elements of 𝐉{\bf{J}} as

J11=C[ω2¯(ω¯)2],\displaystyle J_{11}=C[\overline{\omega^{2}}-(\bar{\omega})^{2}], (46)
J12=J21=C[ωτ¯ω¯τ¯],\displaystyle J_{12}=J_{21}=C[\overline{\omega\tau}-\bar{\omega}\bar{\tau}], (47)
J22=C[τ2¯(τ¯)2],\displaystyle J_{22}=C[\overline{\tau^{2}}-(\bar{\tau})^{2}], (48)

where C=2N0(11+N0)C=\frac{2}{N_{0}}\left(\frac{1}{1+N_{0}}\right) and ω2¯\overline{\omega^{2}}, ωτ¯\overline{\omega\tau} and τ2¯\overline{\tau^{2}} are given in [39, eq. (67), (68) and (69)], respectively. The term ω¯\overline{\omega} can be found by replacing ω2\omega^{2} by ω\omega in [39, eq. (67)] and similarly τ¯\overline{\tau} can be found by replacing u2u^{2} by uu in [39, eq. (69)]. Using lengthy mathematical manipulations, we solve (46) - (48) and find closed-form expressions for the Fisher information matrix elements as

J112BCT(1(M1M)cos(ω0T)),\displaystyle J_{11}\approx\frac{2BC}{T}\left(1-\left(\frac{M-1}{M}\right)\cos{\left(\omega_{0}T\right)}\right), (49)
J12=J21CT22m=0M1(2m+1)ωmi,\displaystyle J_{12}=J_{21}\approx-\frac{CT^{2}}{2}\sum_{m=0}^{M-1}{(2m+1)\omega_{m}^{i}}, (50)
J22=CM2T212,\displaystyle J_{22}=\frac{CM^{2}T^{2}}{12}, (51)

where BB is the finite filter bandwidth of the radar receiver and ω0=2πf0\omega_{0}=2\pi f_{0}. As the derivations of the J11J_{11}, J12J_{12} and J22J_{22} terms involve lengthy calculations, we have presented them separately in Appendices C, D and E, respectively. Note that the pulse width TT in the above expressions are that of the perfectly rectangular pulse before bandwidth limiting. It is a good approximation to the width of the bandwidth-limited pulse for large BTBT. Also, the approximation in (49) comes from neglecting the small terms in the resulting integrals which vanish as we set BTBT\rightarrow\infty. The term ω0T\omega_{0}T represents the phase shift between adjacent frequency tones and is further explained in Appendix C.

Based on the Fisher information matrix we can bound the variance of delay and Doppler estimation errors which directly corresponds to the target range and velocity estimation accuracy in radar. More specifically, the variance of any unbiased estimate is bounded by the diagonal elements of 𝐉1{\bf{J}}^{-1} [39]. Thus, the CRLBs on the delay estimation error and the Doppler estimation error, which we denote by CRLBτ\textrm{CRLB}_{\tau} and CRLBω\textrm{CRLB}_{\omega}, respectively, can be approximated as

CRLBτC1(M2T2M2B(1(M1M)cos(ω0T))3T3(m=0M1(2m+1)ωmi)2),\displaystyle\textrm{CRLB}_{\tau}\approx C^{-1}\left(\frac{M^{2}T}{2M^{2}B\left(1-\left(\frac{M-1}{M}\right)\cos{(\omega_{0}T)}\right)-3T^{3}\left(\sum_{m=0}^{M-1}(2m+1)\omega_{m}^{i}\right)^{2}}\right), (52)
CRLBωC1(2BT(1(M1M)cos(ω0T))M2BT6(1(M1M)cos(ω0T))T44(m=0M1(2m+1)ωmi)2).\displaystyle\textrm{CRLB}_{\omega}\approx C^{-1}\left(\frac{\frac{2B}{T}\left(1-\left(\frac{M-1}{M}\right)\cos{(\omega_{0}T)}\right)}{\frac{M^{2}BT}{6}\left(1-\left(\frac{M-1}{M}\right)\cos{(\omega_{0}T)}\right)-\frac{T^{4}}{4}\left(\sum_{m=0}^{M-1}(2m+1)\omega_{m}^{i}\right)^{2}}\right). (53)

We can further simplify (52) and (53) based on [39, eq. (94), (95)] by ignoring the effect of the ωτ¯\overline{\omega\tau} term on the CRLB. This simplification allows us to draw more insights on the effect of different parameters on the estimation errors as follows

CRLBτC1(MT2B(M(M1)cos(ω0T))),\displaystyle\textrm{CRLB}_{\tau}\approx C^{-1}\left(\frac{MT}{2B\left(M-(M-1)\cos{(\omega_{0}T)}\right)}\right), (54)
CRLBωC1(12M2T2).\displaystyle\textrm{CRLB}_{\omega}\approx C^{-1}\left(\frac{12}{M^{2}T^{2}}\right). (55)

Note that when we set M=1M=1, the right hand sides of (54) and (55) reduce to C1T/2BC^{-1}T/2B and 12C1/T212C^{-1}/T^{2}, respectively, which are the CRLBs on the delay and Doppler estimation errors for a simple pulse [28, 41]. Based on (54) we note that the bound on the delay estimation error is proportional to the simple pulse length TT. In other words, the delay estimation error is inversely proportional to the effective bandwidth of the simple pulse (which is approximately 1/T1/T). Based on (55) we note that the bound on the Doppler estimation accuracy is determined by the total pulse width. More specifically, the Doppler estimation error decreases as we increase MM and TT. If we keep the energy of the signal fixed, and increase TT the accuracy in Doppler will increase and the accuracy in delay will decrease. These observations are further discussed in Section V using numerical examples.

IV-C Discussion

As is the case with any radar waveform with a frequency sweep across the waveform, such as the LFM waveform, the Lehmer code based random stepped frequency radar waveform introduces range-Doppler coupling. Thus, when both the delay and Doppler are unknown, there is an ambiguous region in the delay-Doppler plane. In radar, the classical method of resolving this ambiguity is by transmitting another waveform with the opposite frequency sweep [39]. In our joint radar communication system, this is automatically taken care of by the randomization of the incoming data, because, the transmitted radar waveform will differ based on the communication data, and hence, the effect of the overall range-Doppler coupling will be negligible in average. More specifically, the off-diagonal terms in the Fisher information matrix, i.e., the J12J_{12} and J21J_{21} terms in (50) change with ωmi\omega_{m}^{i}, and hence on the particular frequency permutation of the waveform. They are maximized when ωmi\omega_{m}^{i} values are set in ascending order and the resulting ambiguous region lies in the first and third quadrants of the delay-Doppler plane. Similarly, J12J_{12} and J21J_{21} terms are minimized when ωmi\omega_{m}^{i} values are set in descending order and the resulting ambiguous region lies in the second and fourth quadrants of the delay-Doppler plane.

The overall structure of the AF resulting from the Lehmer code based random stepped frequency radar waveform has a narrow main-lobe and small side-lobes. Such an AF is important to achieve good radar performance specially with clutter. There is a useful averaging effect on the AF side-lobes which occurs because the communications tends to randomly select among all possible frequency combinations which means that clutter enters via the average AF side-lobe level and this is reasonably small for our waveform. The positioning of these side-lobes in the range-Doppler plane changes with the transmitted radar waveform. In situations where AF side-lobe clutter must be minimized, a more advanced coding scheme can be used in which the incoming data is mapped to radar waveforms in such a way that the overall effect of clutter is minimized in average. This could be a topic of interest for future.

V Numerical Examples

In this section we present numerical examples illustrating the performance of the Lehmer code based random stepped frequency radar waveform and the proposed baseband signaling model. Fig. 3 - Fig. 6 focus on the communication functionality while Fig. 4 - Fig. 10 focus on the radar functionality.

Refer to caption

Figure 3: The block error rate versus average received SNR for N=2N=2 and M=4,8,16M=4,8,16. The results are for the proposed baseband signaling model in AWGN.

Fig. 3 plots the block error rate versus the average received signal-to-noise ratio (SNR) in an AWGN channel with two diversity branches, i.e., N=2N=2, and M=4,8M=4,8 and 16. The frequencies are chosen to maintain orthogonality, i.e., the frequency separation Δf=1/T\Delta f=1/T. Note that the waveform energy is assumed to be one and is kept constant as we increase MM. The simulation results are generated using Monte-Carlo simulations while the analytical approximations are generated using the union bound in (17) and the nearest neighbour approximation in (18). Both the union bound and the nearest neighbour approximation sit above the simulation curves and accurately approximate the block error rate at the high SNR regime. Since the waveform energy is kept constant the block error rate increases as we increase MM.

Refer to caption

Figure 4: The block error rate versus average received SNR for N=4N=4, M=2,4M=2,4 and K=0.5,4K=0.5,4. The results are for the proposed baseband signaling model in a Rician fading channel.

Fig. 4 plots the block error rate versus the average received SNR under the Rician fading channel model with a Rician KK factor 0.5 and 1. We set the number of antennas N=4N=4 and change M=2,4M=2,4. Similar to Fig. 3, the simulation results are generated using Monte-Carlo simulations while the analytical approximations are generated using the union bound in (26) and the nearest neighbour approximation in (27). As we increase the Rician KK factor from 0.5 to 4, the communication channel experiences a stronger LOS signal, as a result of which, the block error probability performance improves. For M=2M=2, the union bound and the nearest neighbour approximation reduces to the exact block error probability, and hence coincide with the simulation results. We note that the nearest neighbour approximation accurately follows the simulated results generated using the Monte-Carlo simulations, especially in the high SNR regime. As we increase MM the union bound becomes quite loose. This is expected as the union bound takes into account all the pairwise error probabilities which is a significant number as we increase MM. Therefore, in the next examples we only illustrate the nearest neighbour approximation which provides a better approximation of the block error probability than the union bound.

Refer to caption

Figure 5: The block error rate versus average received SNR for N=4N=4 and M=2,4,8M=2,4,8. The results are for the proposed baseband signaling model in a Rayleigh fading channel.

Fig. 5 plots the block error rate versus the average received SNR under the Rayleigh fading channel model (K=0K=0). We set the number of antennas N=4N=4 and change M=2,4,8M=2,4,8 while keeping the energy of the waveform at one. The analytical approximations are generated using the the nearest neighbour approximation in (29). We observe that (29) provides an accurate approximation of the block error probability under Rayleigh fading channels. Similar to Fig. 3, since the waveform energy is kept constant the block error rate increases as we increase MM.

Refer to caption

Figure 6: The block error rate versus average received SNR for K=1K=1, M=8M=8 and N=2,3,4N=2,3,4. The results are for the proposed baseband signaling model in Rician fading channel.

Fig. 6 plots the block error rate versus the average received SNR by changing the number of antennas as N=2,3N=2,3 and 44. We consider the Rician fading channel model and set K=1K=1 and M=8M=8. The analytical results are generated using the nearest neighbour approximation in (27). As expected, when we increase the number of receive antennas the block error rate decreases due to the receive diversity gain. We also observe that the nearest neighbour approximation becomes more accurate with increasing NN.

Refer to caption

Figure 7: Ambiguity function of a Lehmer code based random stepped frequency radar waveform for M = 8. The permutation of the frequency sequence is arranged in ascending order.

Next we focus on the radar performance of the waveform. Fig. 7 and Fig. 8 plot the ambiguity function of the proposed Lehmer code based random stepped frequency radar waveform for two different permutations of M=8M=8 frequency tones. In Fig. 7 the frequency tones are in ascending order while in Fig. 8 they are in descending order. Subplots (a) and (b) present the three-dimensional surface plot and the contour plot, respectively. The energy of the signal is normalized so that the ambiguity function has a maximum value of one. Note that, as with any waveform the maximum of the ambiguity function occurs at t=ω=0t=\omega=0. Around the origin, the ambiguity function rarely changes as we change the permutation of the frequency sequence, thus the local accuracy remains the same irrespective of the transmitted waveform. The broader structure of the ambiguity function, however, can change depending on the permutation. This can be clearly observed by the contour plots given in Fig. 7 and Fig. 8. Similar to the traditional LFM waveform, the ambiguity function of the Lehmer code based random stepped frequency radar waveform is skewed in the delay-Doppler plane giving rise to range-Doppler coupling. As discussed in Section IV-C, this behaviour is expected as our waveform is sweeping across a bandwidth of MΔfM\Delta f. As shown in subplot (b) of Fig. 7, when the permutation of the frequency sequence is ordered in ascending order, the ambiguous region lies in the first and third quadrants. When the permutation of the frequency sequence is ordered in descending order, the ambiguous region lies in the second and fourth quadrants.

Refer to caption

Figure 8: Ambiguity function of a Lehmer code based random stepped frequency radar waveform for M = 8. The permutation of the frequency sequence is arranged in descending order.

More insight into the ambiguity function can be obtained from the one-dimensional cuts through the delay and the Doppler axis. Fig. 9 plots the zero-delay cut of the ambiguity function for M=4,6M=4,6 and 8 while keeping the energy of the waveform constant at one. As we increase MM, more frequency tones are included in the waveform which results in different side-lobe structures. Overall the waveform achieves reasonable side-lobe performance. Around the origin, the curvature of the main-lobe of the zero-delay cut increases as we increase MM, thus improving the accuracy of the Doppler estimation. This also agrees with the analytical relationship we observed in (55) using the approximation on the CRLB of the Doppler estimation error, which decreases as we increase the total pulse width.

Fig. 10 plots the zero-Doppler cut of the ambiguity function by fixing M=4M=4 and changing the pulse width as T,T/2T,T/2 and T/4T/4 while keeping the energy of the waveform constant at one. As we change the pulse width, Δf\Delta f was changed accordingly to maintain the orthogonality between frequency tones such that TΔf=1T\Delta f=1. From the plot we note that around the origin, the curvature of the main-lobe of the zero-Doppler cut increases as we decrease the pulse width, thus improving the accuracy of the range estimation. This also agrees with the analytical relationship we observed in (54) using the approximation on the CRLB on the delay estimation error, which decreases as we decrease the pulse width.

Refer to caption

Figure 9: The zero-delay cut of the ambiguity function for M = 4, 6, 8.

Refer to caption

Figure 10: The zero-Doppler cut of the ambiguity function for simple pulse width = T,T2T,\frac{T}{2} and T4\frac{T}{4}.

Refer to caption


Figure 11: Behavior of the approximated delay CRLB in (54).

Refer to caption

Figure 12: Behavior of the approximated Doppler CRLB in (55).

Finally, Figs. 11 and 12 illustrate the behaviour of the CRLBs of delay and Doppler estimation errors derived in (54) and (55), respectively. Similar to Fig. 10, in Fig. 11 we set M=4M=4 and change the pulse width as T,T/2T,T/2 and T/4T/4 while keeping the energy of the waveform constant at one. The figure illustrates the behaviour of the approximated CRLB on the delay estimation error, versus the received SNR. As we decrease the pulse width the delay CRLB decreases making a more accurate delay estimation. Similar to Fig. 9, in 12 we consider M=4,6M=4,6 and 8 while keeping the energy of the waveform constant at one. As we increase MM the Doppler CRLB decreases making a more accurate Doppler estimation.

VI Conclusion and Future Work

A Lehmer code based random stepped frequency radar waveform is presented as a suitable candidate for simultaneous data transmission and radar sensing. A signaling scheme is proposed in which data is embedded based on the selection of the frequency permutation in the transmitted waveform. From the data communication perspective, the union bound and the nearest neighbour approximation on the block error probability is analysed based on the optimum ML detector. Closed-form expressions are derived for the block error probability under the baseline case of AWGN channels and the more general case of Rician fading channels. Based on the Hungarian algorithm an efficient implementation of the communications receiver is also presented. From a radar sensing perspective, the performance of the radar waveform is fully characterized based on the ambiguity function and Fisher information matrix. CRLBs on delay and Doppler estimation errors are also derived to analyse the accuracy of range and velocity resolution. Simulation results demonstrate the accuracy of the analytical derivations and illustrate the communication and radar performance of the proposed waveform. In the following, we outline several possible directions for further research.

  • The communication performance in this paper focused on the block error rate. It would be desirable to formalise the extension to bit and symbol error rates which will require the selection of a subset of waveforms from the set of M!M! waveforms. This will bring in important aspects relating to linear block codes and numbering permutations.

  • Related to the above point, it would also be desirable to understand the tradeoffs between the achievable communication data rates and the error probability, which directly relates to the minimum distance between selected waveforms.

  • The analysis could be extended to randomize the phase of each frequency tone, on top of the frequency permutation. This will increase the communication data rate and allow the incorporation of traditional modulation methods such as MM-ary Phase Shift Keying (MPSK).

  • Another desirable extension would be to relax the constraint on frequency permutations and allow the frequencies to repeat within the waveform. This will also increase the communication data rate, but the analysis of its effect on the radar performance is nontrivial.

Appendix A Derivation of 𝐡H𝐡\mathbf{h}^{H}\mathbf{h} in (20)

Using (19) we can express 𝐡H𝐡\mathbf{h}^{H}\mathbf{h} as a summation given by

𝐡H𝐡=i=1N|KK+1Δi+1K+1ui|2.\displaystyle\mathbf{h}^{H}\mathbf{h}=\sum_{i=1}^{N}\left|\sqrt{\frac{K}{K+1}}\Delta_{i}+\sqrt{\frac{1}{K+1}}u_{i}\right|^{2}. (56)

Note that |Δi|2=1|\Delta_{i}|^{2}=1. We can write an expression that is statistically identical to 𝐡H𝐡\mathbf{h}^{H}\mathbf{h} as

𝐡H𝐡=𝑑i=12N|δi+γiK+1|2,\displaystyle\mathbf{h}^{H}\mathbf{h}\overset{d}{=}\sum_{i=1}^{2N}\left|\delta_{i}+\frac{\gamma_{i}}{\sqrt{K+1}}\right|^{2}, (57)

where X=𝑑YX\overset{d}{=}Y implies the statistical equivalence when XX has the same distribution as YY, γi=Re(ui)\gamma_{i}=\textrm{Re}(u_{i}) for all i{1,2,,N}i\in\{1,2,\ldots,N\}, γi=Im(ui)\gamma_{i}=\textrm{Im}(u_{i}) for all i{N+1,N+2,,2N}i\in\{N+1,N+2,\ldots,2N\}, δi=KK+1\delta_{i}=\sqrt{\frac{K}{K+1}} for all i{1,2,,N}i\in\{1,2,\ldots,N\} and δi=0\delta_{i}=0 for all i{N+1,N+2,,2N}i\in\{N+1,N+2,\ldots,2N\}. Using some mathematical manipulations (57) can be written as

𝐡H𝐡=𝑑12(K+1)i=12N|δi2(K+1)+γ~i|2,\displaystyle\mathbf{h}^{H}\mathbf{h}\overset{d}{=}\frac{1}{2(K+1)}\sum_{i=1}^{2N}\left|\delta_{i}\sqrt{2(K+1)}+{\tilde{\gamma}_{i}}\right|^{2}, (58)

where γ~i=2γi𝒩(0,1).\tilde{\gamma}_{i}=\sqrt{2}\gamma_{i}\sim\mathcal{N}(0,1). Thus, we have a sum of squares of 2N2N Gaussian random variables resulting in a non-central chi-squared distribution given by

𝐡H𝐡=𝑑12(K+1)χ2N2(2NK).\displaystyle\mathbf{h}^{H}\mathbf{h}\overset{d}{=}\frac{1}{2(K+1)}\chi_{2N}^{2}(2NK). (59)

Appendix B Working example of the Hungarian Algorithm

In this appendix we have provide a working example of the Hungarian algorithm used in the communications receiver. Suppose the number of receiving antennas N=4N=4. We consider the following example in which the matrix 𝑹\boldsymbol{R} is given by

𝑹=[4326210442535443].\boldsymbol{R}=\left[\begin{array}[]{cccc}-4&-3&-2&-6\\ -2&1&0&-4\\ 4&-2&5&-3\\ 5&4&-4&3\end{array}\right]. (60)

To select 44 elements that are in 44 different rows and 44 different columns and the sum of the 44 elements is maximised, we apply the Hungarian algorithm to 𝑹-\boldsymbol{R}. The steps of the algorithm are illustrated in Fig. 13 and explained as follows.

Refer to caption


Figure 13: An example of the Hungarian algorithm.

In the first step we subtract the row minimum from each row, which results in the second matrix illustrated in Fig. 13. In the second step we subtract the column minimum from each column, which results in the third matrix illustrated in Fig. 13. In the third step we determine the minimum number of lines that are required to cover all zeros in the matrix. In this example, all zeros can be covered using 33 lines drawn across the third column, the second row and the fourth row as indicated in grey shading in the fourth matrix of Fig. 13. Since the number of lines required 33 is lower than the size of the matrix 44, we continue with the fourth step. We subtract the smallest uncovered number 11 from all uncovered elements and add it to all elements that are covered twice, which results in the fifth matrix illustrated in Fig. 13. Then again we return to the third step. The minimum number of lines that are required to cover all zeros in the matrix is 4, which is equivalent to the size of the matrix. Therefore, an optimal assignment exists among the zeros in the matrix. The zeros cover an optimal assignment and the resulting frequency sequence is {f2,f1,f0,f3}\{f_{2},\,f_{1},\,f_{0},\,f_{3}\}.

Appendix C Derivation of the J11J_{11} term

Let us first focus on ω2¯\overline{\omega^{2}} which is defined as

ω2¯=ω2|Si(jω)|2dω2π,\displaystyle\overline{\omega^{2}}=\int_{-\infty}^{\infty}\omega^{2}|S_{i}(j\omega)|^{2}\frac{d\omega}{2\pi}, (61)

where Si(jω)S_{i}(j\omega) is the Fourier transform of si(u)s_{i}(u) and can be derived as

Si(jω)=TMm=0M1sinc((ωmiω)T2)ej(ωmT(ωmiω)T2).\displaystyle S_{i}(j\omega)=\sqrt{\frac{T}{M}}\sum_{m=0}^{M-1}\textrm{sinc}\left((\omega_{m}^{i}-\omega)\frac{T}{2}\right)e^{-j\left(\omega mT-(\omega_{m}^{i}-\omega)\frac{T}{2}\right)}. (62)

Substituting (62) into (61) we write

ω2¯=T2πMm=0M1n=0M1ej(ωmiωni)T2ω2sinc((ωωmi)T2)sinc((ωωni)T2)ejωT(mn)𝑑ω.\displaystyle\overline{\omega^{2}}=\frac{T}{2\pi M}\sum_{m=0}^{M-1}\sum_{n=0}^{M-1}e^{j(\omega_{m}^{i}-\omega_{n}^{i})\frac{T}{2}}\int_{-\infty}^{\infty}\omega^{2}\textrm{sinc}\left((\omega-\omega_{m}^{i})\frac{T}{2}\right)\textrm{sinc}\left((\omega-\omega_{n}^{i})\frac{T}{2}\right)e^{j\omega T(m-n)}d\omega. (63)

Note that ω2¯\overline{\omega^{2}} does not converge. This is because we have assumed a perfect rectangular pulse, i.e, one with zero rise and zero fall time, for sp(u)s_{p}(u) in si(u)s_{i}(u). In practice, however, pulses are not perfectly rectangular because a zero rise time or a zero fall time requires an infinite bandwidth. Therefore, similar to [41] we assume that the spectrum of sp(u)s_{p}(u) has the form of a sinc function but the bandwidth is limited to a finite value of BB such that the integral in (63) becomes

ω2¯=T2πMm=0M1n=0M1ej(ωmiωni)T2ωniBπωmi+Bπω2sinc((ωωmi)T2)sinc((ωωni)T2)ejωT(mn)𝑑ω.\displaystyle\overline{\omega^{2}}=\frac{T}{2\pi M}\sum_{m=0}^{M-1}\sum_{n=0}^{M-1}e^{j(\omega_{m}^{i}-\omega_{n}^{i})\frac{T}{2}}\int_{\omega_{n}^{i}-B\pi}^{\omega_{m}^{i}+B\pi}\omega^{2}\textrm{sinc}\left((\omega-\omega_{m}^{i})\frac{T}{2}\right)\textrm{sinc}\left((\omega-\omega_{n}^{i})\frac{T}{2}\right)e^{j\omega T(m-n)}d\omega. (64)

This is equivalent to passing a perfectly rectangular pulse through a filter of width BB. Substituting v=ωT2v=\frac{\omega T}{2}, vm=ωmiT2v_{m}=\frac{\omega_{m}^{i}T}{2} and vn=ωniT2v_{n}=\frac{\omega_{n}^{i}T}{2}, where we have dropped the superscript ii for notational simplicity, we can reexpress (64) as

ω2¯=4πMT2m=0M1n=0M1ej(vmvn)vnBπT2vm+BπT2v2sinc(vvm)sinc(vvn)ej2vk𝑑v,\displaystyle\overline{\omega^{2}}=\frac{4}{\pi MT^{2}}\sum_{m=0}^{M-1}\sum_{n=0}^{M-1}e^{j(v_{m}-v_{n})}\int_{v_{n}-\frac{B\pi T}{2}}^{v_{m}+\frac{B\pi T}{2}}v^{2}\textrm{sinc}\left(v-v_{m}\right)\textrm{sinc}\left(v-v_{n}\right)e^{j2vk}dv, (65)

where k=mnk=m-n. Next we use the identities

v2(vvm)(vvn)=1+(vm+vn)vvm+vn2(vvm)(vvn),\displaystyle\frac{v^{2}}{(v-v_{m})(v-v_{n})}=1+\frac{(v_{m}+v_{n})}{v-v_{m}}+\frac{v_{n}^{2}}{(v-v_{m})(v-v_{n})}, (66)

and

sin(vvm)sin(vvn)ej2vk=12cos(vnvm)ej2vk14ej(2(k1)v+vm+vn)14ej(2(k+1)vvmvn),\displaystyle\sin{(v-v_{m})}\sin{(v-v_{n})}e^{j2vk}=\frac{1}{2}\cos{(v_{n}-v_{m})e^{j2vk}}-\frac{1}{4}e^{j(2(k-1)v+v_{m}+v_{n})}-\frac{1}{4}e^{j(2(k+1)v-v_{m}-v_{n})}, (67)

to reexpress (65) as

ω2¯=4πMT2m=0M1n=0M1ej(vnvm)Jn,m,\displaystyle\overline{\omega^{2}}=\frac{4}{\pi MT^{2}}\sum_{m=0}^{M-1}\sum_{n=0}^{M-1}e^{j(v_{n}-v_{m})}J_{n,m}, (68)

where

Jn,m\displaystyle J_{n,m} =vnBπT2vm+BπT2(1+(vm+vn)vvm+vn2(vvm)(vvn))(12cos(vnvm)ej2vk\displaystyle=\int_{v_{n}-\frac{B\pi T}{2}}^{v_{m}+\frac{B\pi T}{2}}\left(1+\frac{(v_{m}+v_{n})}{v-v_{m}}+\frac{v_{n}^{2}}{(v-v_{m})(v-v_{n})}\right)\left(\frac{1}{2}\cos{(v_{n}-v_{m})e^{j2vk}}\right.
14ej(2(k1)v+vm+vn)14ej(2(k+1)vvmvn))dv.\displaystyle\left.-\frac{1}{4}e^{j(2(k-1)v+v_{m}+v_{n})}-\frac{1}{4}e^{j(2(k+1)v-v_{m}-v_{n})}\right)dv. (69)

It is important to note that all of the integrals in (C) are small and of order one when k0,±1k\neq 0,\pm 1. This is because for q0q\neq 0, the integrals ejqv𝑑v\int e^{jqv}dv, ejqvvvm𝑑v\int\frac{e^{jqv}}{v-v_{m}}dv and ejqv(vvm)(vvn)𝑑v\int\frac{e^{jqv}}{(v-v_{m})(v-v_{n})}dv for mnm\neq n can all be computed using trigonometric integrals [36, eq. (2.641.1)-(2.641.2)] giving order one integrals. Hence, the dominant terms in Jn,mJ_{n,m} are k=0k=0 (i.e., n=mn=m) and k=±1k=\pm 1 (i.e., n=m±1n=m\pm 1) which can be written as

Jn,n=14vnBπT2vm+BπT2(1+2vnvvn+vn2(vvn)2)(2ej(2v+2vn)ej(2v2vn))𝑑v\displaystyle J_{n,n}=\frac{1}{4}\int_{v_{n}-\frac{B\pi T}{2}}^{v_{m}+\frac{B\pi T}{2}}\left(1+\frac{2v_{n}}{v-v_{n}}+\frac{v_{n}^{2}}{(v-v_{n})^{2}}\right)\left(2-e^{j(-2v+2v_{n})}-e^{j(2v-2v_{n})}\right)dv (70)
Jn,n+1=14ej(vn+vn+1)vnBπT2vm+BπT2(1+vn+vn+1vvn+1+vn2(vvn)(vvn+1))𝑑v+𝒪(1)\displaystyle J_{n,n+1}=-\frac{1}{4}e^{j(v_{n}+v_{n+1})}\int_{v_{n}-\frac{B\pi T}{2}}^{v_{m}+\frac{B\pi T}{2}}\left(1+\frac{v_{n}+v_{n+1}}{v-v_{n+1}}+\frac{v_{n}^{2}}{(v-v_{n})(v-v_{n+1})}\right)dv+\mathcal{O}(1) (71)
Jn,n1=14ej(vn+vn1)vnBπT2vm+BπT2(1+vn+vn1vvn1+vn2(vvn)(vvn1))𝑑v+𝒪(1)\displaystyle J_{n,n-1}=-\frac{1}{4}e^{j(v_{n}+v_{n-1})}\int_{v_{n}-\frac{B\pi T}{2}}^{v_{m}+\frac{B\pi T}{2}}\left(1+\frac{v_{n}+v_{n-1}}{v-v_{n-1}}+\frac{v_{n}^{2}}{(v-v_{n})(v-v_{n-1})}\right)dv+\mathcal{O}(1) (72)

For large BB, we can ignore the terms involving inverse powers of vvnv-v_{n}, vnvn1v_{n}-v_{n-1} and vvn+1v-v_{n+1} as these are also order one. Hence, we approximate Jn,nBπT2J_{n,n}\approx\frac{B\pi T}{2}, Jn,n+1BπT4ej(vn+vn+1)J_{n,n+1}\approx-\frac{B\pi T}{4}e^{j(v_{n}+v_{n+1})} and Jn,n1BπT4ej(vn+vn1)J_{n,n-1}\approx-\frac{B\pi T}{4}e^{j(v_{n}+v_{n-1})} which results in

ω2¯\displaystyle\overline{\omega^{2}} 2BMT(Mn=0M212ej(vnvn+1)ej(vn+vn+1)n=1M112ej(vnvn1)ej(vn+vn1)).\displaystyle\approx\frac{2B}{MT}\left({M}-\sum_{n=0}^{M-2}\frac{1}{2}e^{j(v_{n}-v_{n+1})}e^{j(v_{n}+v_{n+1})}-\sum_{n=1}^{M-1}\frac{1}{2}e^{j(v_{n}-v_{n-1})}e^{-j(v_{n}+v_{n-1})}\right). (73)

After some straightforward mathematical manipulations (73) can be reexpressed as

ω2¯2BMT(Mn=0M2cos(ωniT)).\displaystyle\overline{\omega^{2}}\approx\frac{2B}{MT}\left(M-\sum_{n=0}^{M-2}\cos{\left(\omega_{n}^{i}T\right)}\right). (74)

Note that the argument inside the cosine function in (74), i.e., ωniT\omega_{n}^{i}T simply corresponds to the phase change of the waveform at T,2T,,(M1)TT,2T,\ldots,(M-1)T. Since we consider orthogonal frequencies such that the separation between two frequency tones Δf=n/T\Delta f=n/T, where nn is an integer, we can write ωni=2π(f0+qniΔf)\omega_{n}^{i}=2\pi(f_{0}+q_{n}^{i}\Delta f) where qniq_{n}^{i} is an integer and further reduce (74) to derive

ω2¯2BT(1(M1M)cos(ω0T)).\displaystyle\overline{\omega^{2}}\approx\frac{2B}{T}\left(1-\left(\frac{M-1}{M}\right)\cos{\left(\omega_{0}T\right)}\right). (75)

From (75) we learn that, if we set ω0T=0\omega_{0}T=0, the cosine function takes its maximum value of one and (75) reduces to

ω2¯2BTM.\displaystyle\overline{\omega^{2}}\approx\frac{2B}{TM}. (76)

Similarly, if we set ω0T=π\omega_{0}T=\pi, the cosine function takes its minimum value of negative one and (75) reduces to

ω2¯2BT(2M1M).\displaystyle\overline{\omega^{2}}\approx\frac{2B}{T}\left(\frac{2M-1}{M}\right). (77)

Next, we focus on ω¯\overline{\omega} which is defined as

ω¯=ω|Si(jω)|2dω2π.\displaystyle\overline{\omega}=\int_{-\infty}^{\infty}\omega|S_{i}(j\omega)|^{2}\frac{d\omega}{2\pi}. (78)

Following similar steps to ω2¯\overline{\omega^{2}} we write (78) as

ω¯=2πMTm=0M1n=0M1ej(vmvn)vnBπT2vm+BπT2vsinc(vvm)sinc(vvn)ej2vk𝑑v,\displaystyle\overline{\omega}=\frac{2}{\pi MT}\sum_{m=0}^{M-1}\sum_{n=0}^{M-1}e^{j(v_{m}-v_{n})}\int_{v_{n}-\frac{B\pi T}{2}}^{v_{m}+\frac{B\pi T}{2}}v\textrm{sinc}\left(v-v_{m}\right)\textrm{sinc}\left(v-v_{n}\right)e^{j2vk}dv, (79)

which can be reexpressed as

ω¯=2πMTm=0M1n=0M1ej(vnvm)Jn,m,\displaystyle\overline{\omega}=\frac{2}{\pi MT}\sum_{m=0}^{M-1}\sum_{n=0}^{M-1}e^{j(v_{n}-v_{m})}J_{n,m}, (80)

where

Jn,m\displaystyle J_{n,m} =vnBπT2vm+BπT2(vm(vvm)(vmvn)+vn(vvn)(vnvm))(12cos(vnvm)ej2vk\displaystyle=\int_{v_{n}-\frac{B\pi T}{2}}^{v_{m}+\frac{B\pi T}{2}}\left(\frac{v_{m}}{(v-v_{m})(v_{m}-v_{n})}+\frac{v_{n}}{(v-v_{n})(v_{n}-v_{m})}\right)\left(\frac{1}{2}\cos{(v_{n}-v_{m})e^{j2vk}}\right.
14ej(2(k1)v+vm+vn)14ej(2(k+1)vvmvn))dv.\displaystyle\left.-\frac{1}{4}e^{j(2(k-1)v+v_{m}+v_{n})}-\frac{1}{4}e^{j(2(k+1)v-v_{m}-v_{n})}\right)dv. (81)

Noting that all the integrals in (C) are order one, for large BB we can approximate

ω¯0.\displaystyle\overline{\omega}\approx 0. (82)

Based on (75) and (82) we can approximate the J11J_{11} term as given in (49).

Appendix D Derivation of the J12J_{12} term

Based on (82) we note that ω¯0\bar{\omega}\approx 0. Hence, we first approximate the J12J_{12} term in (47) by

J12Cωτ¯.\displaystyle J_{12}\approx C\overline{\omega\tau}. (83)

Next, we focus on ωτ¯\overline{\omega\tau} which is defined in [39, eq. (68)] as

ωτ¯=Im(usi(u)si(u)u𝑑u),\displaystyle\overline{\omega\tau}=\textrm{Im}\left(\int_{-\infty}^{\infty}us_{i}(u)\frac{\partial s_{i}^{*}(u)}{\partial u}du\right), (84)

where Im(.)\textrm{Im}(.) denotes the imaginary part of the argument, si(u)s_{i}(u) is given in (32) and si(u)s_{i}^{*}(u) denotes the complex conjugate of si(u)s_{i}(u). We reexpress si(u)s_{i}(u) as

si(u)=1MTm=0M1sp(umT)e2πjfmi(umT),\displaystyle s_{i}(u)=\sqrt{\frac{1}{MT}}\sum_{m=0}^{M-1}s_{p}(u-mT)e^{2\pi jf_{m}^{i}(u-mT)}, (85)

Next, we approximate sp(u)s_{p}(u) by an increasing sequence of smooth functions, sp(n)(u)s_{p}^{(n)}(u), vanishing and with vanishing derivatives at the endpoints 0,T0,T. Specifically, we require

sp(n)(u)sp(n)(umT)=0,s_{p}^{(n)}(u)s_{p}^{(n)}(u-mT)=0, (86)

and

sp(n)(u)sp(n)(umT)=0form0,u.s_{p}^{(n)}(u){s^{\prime}}_{p}^{(n)}(u-mT)=0form\neq 0,\ \forall u. (87)

Let us denote the integral within the brackets in (84) by Φ\Phi and approximate Φ\Phi by Φ(n)\Phi^{(n)} where we are replacing sp(u)s_{p}(u) by sp(n)(u)s_{p}^{(n)}(u) in the formula for si(u)s_{i}(u) and write

Φ(n)=\displaystyle\Phi^{(n)}=\int_{-\infty}^{\infty} u(m=0M1sp(n)(umT)e2πjfmiu)\displaystyle u\Bigl{(}\sum_{m=0}^{M-1}s_{p}^{(n)}(u-mT)e^{2\pi jf_{m}^{i}u}\Bigr{)}
×(m=0M1sp(n)(umT)e2πjfmium=0M12πjfmisp(n)(umT)e2πjfmiu)du.\displaystyle\times\Bigl{(}\sum_{m^{\prime}=0}^{M-1}{s^{\prime}}_{p}^{(n)}(u-m^{\prime}T)e^{-2\pi jf_{m^{\prime}}^{i}u}-\sum_{m^{\prime}=0}^{M-1}2\pi jf_{m^{\prime}}^{i}s_{p}^{(n)}(u-m^{\prime}T)e^{-2\pi jf_{m^{\prime}}^{i}u}\Bigr{)}\,du. (88)

In view of sp(n)(u)sp(n)(umT)=0s_{p}^{(n)}(u)s_{p}^{(n)}(u-mT)=0, the expression in (D) becomes

Φ(n)=m=0M1(usp(n)(umT)(sp(n)(umT)du2πjfmiusp(n)(umT)2du).\displaystyle\Phi^{(n)}=\sum_{m=0}^{M-1}\Bigl{(}\int_{-\infty}^{\infty}us_{p}^{(n)}(u-mT)\bigl{(}{s^{\prime}}_{p}^{(n)}(u-mT)\,du-2\pi jf_{m}^{i}\int_{-\infty}^{\infty}us_{p}^{(n)}(u-mT)^{2}\,du\Bigr{)}. (89)

Focusing just on the imaginary part of Φ(n)\Phi^{(n)} we can write an approximation for ωτ¯\overline{\omega\tau} as

ωτ¯\displaystyle\overline{\omega\tau} 2πm=0M1fmiusp(n)(umT)2𝑑u\displaystyle\approx-2\pi\sum_{m=0}^{M-1}f_{m}^{i}\int_{-\infty}^{\infty}us_{p}^{(n)}(u-mT)^{2}\,du
=πm=0M1fmi(2m+1)T2.\displaystyle=-\pi\sum_{m=0}^{M-1}f_{m}^{i}{(2m+1)T^{2}}. (90)

Substituting (D) into (83) we can derive an approximation for J12J_{12} as

J12πCT2m=0M1(2m+1)fmi,J_{12}\approx-\pi CT^{2}\sum_{m=0}^{M-1}(2m+1)f_{m}^{i}, (91)

which can be simplified to produce (50).

Appendix E Derivation of the J22J_{22} term

Let us first focus on τ¯\bar{\tau} which is defined as

τ¯=u|si(u)|2𝑑u.\displaystyle\bar{\tau}=\int_{-\infty}^{\infty}u|s_{i}(u)|^{2}du. (92)

Substituting (32) into (92) we can write

τ¯=m=0M1mT(M+1)uMT𝑑u.\displaystyle\bar{\tau}=\sum_{m=0}^{M-1}\int_{mT}^{(M+1)}\frac{u}{MT}du. (93)

After some straight-forward linear algebra the above summation can be reduced to

τ¯=MT2.\displaystyle\bar{\tau}=\frac{MT}{2}. (94)

Similarly, we proceed to solve τ2¯\overline{\tau^{2}} which is defined as

τ2¯=u2|si(u)|2𝑑u.\displaystyle\overline{\tau^{2}}=\int_{\infty}^{\infty}u^{2}|s_{i}(u)|^{2}du. (95)

Substituting (32) into (95) we can write

τ2¯=m=0M1mT(M+1)u2MT𝑑u,\displaystyle\overline{\tau^{2}}=\sum_{m=0}^{M-1}\int_{mT}^{(M+1)}\frac{u^{2}}{MT}du, (96)

which can be solved using some straight-forward linear algebra to derive

τ2¯=M2T23.\displaystyle\overline{\tau^{2}}=\frac{M^{2}T^{2}}{3}. (97)

Substituting (94) and (97) into (48) the final expression of J22J_{22} term can be derived as given in (51).

References

  • [1] C. Sturm and W. Wiesbeck, “Waveform design and signal processing aspects for fusion of wireless communications and radar sensing,” Proc. IEEE, vol. 99, pp. 1236––1259.
  • [2] A. Hassanien, M. G. Amin, Y. D. Zhang, and F. Ahmad, “Signaling strategies for dual-function radar communications: an overview,” IEEE Aerosp. Electron. Syst. Mag., vol. 31, pp. 36––45, Oct. 2016.
  • [3] A. R. Chiriyath, B. Paul, and D. W. Bliss, “Radar-communications convergence: Coexistence, cooperation, and co-design,” IEEE Trans. Cognitive Commun. Netw., vol. 3, pp. 1–12, Mar. 2017.
  • [4] L. Zheng, M. Lops, Y. C. Eldar, and X. Wang, “Radar and communication coexistence: An overview: A review of recent methods,” IEEE Signal Process. Mag., vol. 36, pp. 85–99, Sep. 2019.
  • [5] J. A. Zhang, F. Liu, C. Masouros, R. W. H. Jr., Z. Feng, L. Zheng, and A. Petropulu, “An overview of signal processing techniques for joint communication and radar sensing,” Available Online: [arXiv:2102.12780], pp. 1–18, Feb. 2021.
  • [6] L. Han and K. Wu, “Joint wireless communication and radar sensing systems—state of the art and future prospects,” IET Microwaves, Antennas Propag., vol. 7, pp. 876––885.
  • [7] F. Liu, C. Masouros, A. P. Petropulu, H. Griffiths, and L. Hanzo, “Joint radar and communication design: Applications, state-of-the-art, and the road ahead,” IEEE Trans. Commun., vol. 68, pp. 3834––3862, Feb. 2020.
  • [8] P. Kumari, J. Choi, N. González-Prelcic, and R. W. Heath, “IEEE 802.11ad-based radar: An approach to joint vehicular communication-radar system,” IEEE Trans. Veh. Technol., vol. 67, pp. 3012–3027, Apr. 2018.
  • [9] X. Chen, Z. Feng, Z. Wei, and X. Y. Feifei Gao and, “Performance of joint sensing-communication cooperative sensing UAV network,” IEEE Trans. Veh. Technol., vol. 69, pp. 15 545–15 556, Dec. 2020.
  • [10] E. Cianca, M. D. Sanctis, and S. D. Domenico, “Radios as sensors,” IEEE Internet of Things Journal, vol. 4, pp. 363––373, Apr. 2017.
  • [11] D. Ma, N. Shlezinger, T. Huang, Y. Liu, and Y. C. Eldar, “Joint radar-communication strategies for autonomous vehicles: Combining two key automotive technologies,” IEEE Signal. Process. Mag., vol. 37, pp. 85–97, Jul. 2020.
  • [12] R. C. Daniels, E. R. Yeh, and R. W. Heath, “Forward collision vehicular radar with IEEE 802.11: Feasibility demonstration through measurements,” vol. 67, pp. 1404––1416, Feb. 2018.
  • [13] Y. Zeng, Y. Ma, and S. Sun, “Joint radar-communication: Low complexity algorithm and self-interference cancellation,” in IEEE Global Communications Conference (GLOBECOM), vol. 1, Abu Dhabi, United Arab Emirates, Feb. 2019, pp. 1–7.
  • [14] D. Garmatyuk, J. Schuerger, K. Kauffman, and S. Spalding, “Wideband OFDM system for radar and communications,” in IEEE Radar Conf., vol. 1, pp. 1–6.
  • [15] T. Guo and R. Qiu, “OFDM waveform design compromising spectral nulling, side-lobe suppression and range resolution,” in IEEE Radar Conf., Cincinnati, OH, USA, May 2014, pp. 1424––1429.
  • [16] A. Turlapaty, Y. Jin, and Y. Xu, “Range and velocity estimation of radar targets by weighted OFDM modulation,” in IEEE Radar Conf., Cincinnati, OH, USA, May 2014, pp. 1358––1362.
  • [17] E. Mozeson and N. Levanon, “Multicarrier radar signals with low peak-to-mean envelope power ratio,” in IEE Radar Sonar and Navigation, May 2003, pp. 71–77.
  • [18] M. Braun, C. Sturm, and F. K. Jondral, “Maximum likelihood speed and distance estimation for OFDM radar,” in IEEE Radar Conference, Jun. 2010, pp. 256–261.
  • [19] X. Shaojian, C. Bing, and Z. Ping, “Radar-communication integration based on DSSS techniques,” in 8th Int. Conf. Signal Process., Beijing, China, Nov. 2006, pp. 1–4.
  • [20] M. Bocquet, C. Loyez, C. Lethien, N. Deparis, M. Heddebaut, A. Rivenq, and N. Rolland, “A multifunctional 60-GHz system for automotive applications with communication and positioning abilities based on time reversal,” in IEEE The 7th European Radar Conf., pp. 1–7.
  • [21] C. D. Richmond, P. Basu, R. E. Learned, J. Vian, A. Worthen, and M. Lockard, “Performance bounds on cooperative radar and communication systems operation,” in 2016 IEEE Radar Conference (RadarConf), 2016, pp. 1–6.
  • [22] M. Roberton and E. R. Brown, “Integrated radar and communications based on chirped spread-spectrum techniques,” in IEEE MTT-S International Microwave Symposium Digest, vol. 1, Jun. 2003, pp. 611–614.
  • [23] S. D. Blunt, P. Yatham, and J. Stiles, “Intrapulse radar-embedded communications,” IEEE Trans. Aerosp. Electron. Syst., vol. 46, pp. 1185––1200, Jul. 2010.
  • [24] S. Dwivedi, A. N. Barret, P. Sen, and G. Fettweis, “Target detection in joint frequency modulated continuous wave (FMCW) radar-communication system,” in 2019 16th International Symposium on Wireless Communication Systems (ISWCS), pp. 277–282.
  • [25] S. Skaria, A. Al-Hourani, R. J. Evans, K. Sithamparanathan, and U. Parampalli, “Interference mitigation in automotive radars using pseudo-random cyclic orthogonal sequences,” Sensors 2019, 4459, Oct. 2019.
  • [26] A. Al-Hourani, R. J. Evans, P. M. Farrell, B. Moran, M. Martorella, S. Kandeepan, S. Skafidas, and U. Parampalli, Millimeter-wave Integrated Radar Systems and Techniques.   Amsterdam, The Netherlands: Academic Press Library in Signal Processing, 2018.
  • [27] A. Al-Hourani, R. J. Evans, B. Moran, S. Kandeepan, and U. Parampall, “Efficient range-Doppler processing for random stepped frequency radar in automotive application,” in 2017 IEEE 85th Vehicular Technology Conference (VTC Spring), Sydney, NSW, Nov. 2017, pp. 1–7.
  • [28] M. Richards, Fundamentals of Radar Signal Processing, 2nd ed.   Mc GrawHill Education, 2014.
  • [29] N. Levanon and E. Mozeson, Radar Signals, 1st ed.   John Wiley and Sons, Inc., 2004.
  • [30] D. H. Lehmer, “Teaching combinatorial tricks to a computer,” in In Proceedings of Symposia in Applied Mathematics; Bellman, R.; Hall, M., Jr., Eds., vol. 10, American Mathematical Society: Providence, RI, USA, May 1960, pp. 179––193.
  • [31] D. Knuth, Sorting and Searching, The Art of Computer Programming, 2nd ed.   Addison-Wesley Professional, 1973.
  • [32] M. Hassani, “Derangements and applications,” Journal of Integer Sequences, vol. 6, pp. 1–6, Aug. 2003.
  • [33] J. G. Proakis and M. Salehi, Digital Communications.   NY, USA: McGraw-Hill, 2008.
  • [34] L. Bernadó, T. Zemen, F. Tufvesson, A. F. Molisch, and C. F. Mecklenbräuker, “Time-and frequency-varying KK-factor of non-stationary vehicular channels for safety-relevant scenarios,” IEEE Transactions on Intelligent Transportation Systems, vol. 16, pp. 1007–1017, Apr. 2015.
  • [35] N. L. Johnson and S.Kotz, Continuous Univariate Distributions, 1st ed.   A Wiley Series in Probability and Statistics, 1970.
  • [36] I. Grashteyn and I. Ryzhik, Table of Integrals, Series, and Products, 7th ed.   30 Corporate Drive, Suite 400, Burlington, MA 01803, USA: Academic Press, 2007.
  • [37] R. K. Mallik and M. Z. Win, “Analysis of hybrid selection/maximal ratio combining in correlated Nakagami fading,” IEEE Trans. Commun., vol. 50, pp. 1372––1383, Aug. 2002.
  • [38] H. W. Kuhn, The Hungarian Method for the Assignment Problem.   Wiley Online Library, 1955.
  • [39] H. L. VanTrees, Detection, Estimation, and Modulation Theory: Radar-Sonar Processing and Gaussian Signals in Noise.   NY, USA: John Wiley & sons, INC, 2001.
  • [40] A. Rihaczek, Principles of High Resolution Radar.   685, Canton Street, Norwood, MA: Artech House, INC, 1996.
  • [41] M. I. Skolnik, “Theoretical accuracy of radar measurements,” IRE Transactions on Aeronautical and Navigational Electronics, vol. 7, pp. 1960––1383, Dec. 1960.