Optimal Power Allocation for OFDM-based Ranging Using Random Communication Signals
Abstract
High-precision ranging plays a crucial role in future 6G Integrated Sensing and Communication (ISAC) systems. To improve the ranging performance while maximizing the resource utilization efficiency, future 6G ISAC networks have to reuse data payload signals for both communication and sensing, whose inherent randomness may deteriorate the ranging performance. To address this issue, this paper investigates the power allocation (PA) design for an OFDM-based ISAC system under random signaling, aiming to reduce the ranging sidelobe level of both periodic and aperiodic auto-correlation functions (P-ACF and A-ACF) of the ISAC signal. Towards that end, we first derive the closed-form expressions of the average squared P-ACF and A-ACF, and then propose to minimize the expectation of the integrated sidelobe level (EISL) under arbitrary constellation mapping. We then rigorously prove that the uniform PA scheme achieves the global minimum of the EISL for both P-ACF and A-ACF. As a step further, we show that this scheme also minimizes the P-ACF sidelobe level at every lag. Moreover, we extend our analysis to the P-ACF case with frequency-domain zero-padding, which is a typical approach to improve the ranging resolution. We reveal that there exists a tradeoff between sidelobe level and mainlobe width, and propose a project gradient descent algorithm to seek a locally optimal PA scheme that reduces the EISL. Finally, we validate our theoretical findings through extensive simulation results, confirming the effectiveness of the proposed PA methods in reducing the ranging sidelobe level for random OFDM signals.
Index Terms:
Integrated Sensing and Communication, OFDM, ranging sidelobe, power allocation, frequency zero-paddingI Introduction
The International Telecommunication Union (ITU) has recently endorsed a global vision for 6G, identifying Integrated Sensing and Communication (ISAC) technology as one of six critical use cases for next-generation networks [1]. To date, ISAC has been widely recognized as a key enabler for a number of emerging 6G applications, including autonomous vehicles, Internet of Things (IoT) networks, unmanned aerial vehicle (UAV) networks, and the low-altitude economy, which require simultaneous communication and sensing functionalities[2, 3, 4].
The ISAC system is a unified framework that provides both wireless communication and radar sensing functions. ISAC significantly enhances spectrum utilization and mitigates spectrum conflicts between radar and communication systems. Furthermore, it reduces the size and energy consumption of the equipment. Due to its high spectrum efficiency and low hardware costs, ISAC technology has garnered substantial attention from both academia and industry [5, 6, 7, 8, 9, 10]. ISAC waveforms, as the basis of ISAC technology, have been studied extensively [11, 12, 13, 14]. In general, waveform design methodologies for ISAC can be categorized into sensing-centric design, communication-centric design, and joint design approaches [3]. For the sensing-centric design, the priority of sensing functionality is higher than that of the communication counterpart. In such cases, one typically focuses on maximizing sensing performance or implementing basic communication capabilities over existing sensing waveforms or infrastructures[15],[16]. Communication-centric schemes, on the other hand, rely on the existing communication waveforms and standard-compatible protocols, such as single-carrier (SC) signals, orthogonal time frequency space (OTFS) signals and orthogonal frequency division multiplexing (OFDM) signals [17],[18]. Different from the above two methods, the joint design strategy creates new ISAC waveforms, aiming to achieve an optimized balance between sensing and communication capabilities[19],[20].
The communication-centric approach is expected to be more advantageous for ISAC networks due to its lower implementation complexity[5],[6]. Towards that end, numerous researches have been centering on communication-centric ISAC transmission, with a particular focus on waveform design and performance analysis [21, 22, 23, 24]. However, a critical challenge arises because communication data symbols must inherently exhibit randomness to carry meaningful information, which can adversely affect sensing accuracy. This phenomenon has recently been characterized as the deterministic-random tradeoff (DRT), a fundamental limitation in ISAC system design [25]. As a result, developing optimal communication-centric waveforms that mitigate sensing performance degradation becomes a crucial research objective. Against this backdrop, our recent work [26] took an initial step towards answering this fundamental question: What is the optimal communication-centric ISAC waveform under random signaling? In [26], we established the superiority of OFDM modulation in achieving the lowest ranging sidelobe. Specifically, we demonstrated that among all communication waveforms with a cyclic prefix (CP), OFDM is the only globally optimal waveform that achieves the lowest ranging sidelobe for both quadrature amplitude modulation (QAM) and phase shift keying (PSK) constellations. This holds true in terms of both the expectation of the integrated sidelobe level (EISL) and the expected sidelobe level at each lag of the periodic auto-correlation function (P-ACF). Furthermore, we proved that for communication waveforms without a CP, OFDM serves as a locally optimal waveform for QAM/PSK constellations, achieving a local minimum of the EISL of the aperiodic auto-correlation function (A-ACF).
Although the superiority of OFDM modulation over other waveforms in reducing ranging sidelobe level has been demonstrated in [26], it generally assumed a uniform power distribution across transmitted symbols, leaving the optimal power allocation (PA) strategies for random ISAC signaling unexplored. Numerous studies have investigated PA schemes for OFDM-based ISAC systems [27, 28, 29, 30, 31]. In [27], the authors proposed three distinct PA schemes aimed at maximizing the mutual information (MI) of radar sensing under different scenarios. In [28], the authors designed an optimization algorithm to obtain the PA design over subcarriers, thus the sensing MI of the OFDM-based ISAC waveform is improved. The authors in [29] conceived a power minimization-based joint subcarrier assignment and power allocation (PM-JSAPA) approach for ISAC systems. Similarly, a joint subcarrier and PA method for the integrated OFDM with interleaved subcarriers (OFDM-IS) waveform was proposed in [30]. The proposed method minimizes the autocorrelation sidelobe level of the integrated OFDM-IS waveform, using a cyclic minimization algorithm to solve the optimization problem. In [31], the authors proposed two designs for a radar-communication spectrum sharing problem by maximizing the output signal-to-interference-plus-noise ratio (SINR) at the radar receiver while maintaining certain communication throughput and power constraints. Although extensive studies on OFDM PA have been conducted for deterministic signaling, none have rigorously addressed the issue of PA design for random signaling.
In this paper, we investigate the PA design for random OFDM communication signals, in the context of target range estimation under a monostatic ISAC setup. In this setup, the ISAC transmitter (Tx) sends out an OFDM communication signal modulated with random symbols for dual purposes of delivering information to communication users and multi-target ranging. The signal is received by communication users, and simultaneously reflected from distant targets back to a sensing receiver (Rx) that is colocated with the ISAC Tx. As a result, despite its inherent randomness, the sensing Rx has full knowledge of the ISAC signal. To evaluate the ranging performance under matched filtering algorithms, we adopt the auto-correlation function (ACF) of the ISAC signal as a performance indicator, considering both periodic and aperiodic convolutions for signals with and without CP, respectively. For clarity, we summarize our main contributions as follows:
-
•
We develop a generic framework to analyze the ranging performance of OFDM-based random ISAC signals. Specifically, we analyze P-ACF and A-ACF for the considered signal by deriving their closed-form expressions.
-
•
We prove that the uniform PA scheme achieves the lowest EISL for both P-ACF and A-ACF. In particular, it also achieves the lowest average sidelobe level at every lag of the P-ACF.
-
•
We derive a closed-form expression for P-ACF with PA and frequency zero-padding, where the latter is a classical method to improve the range resolution of the matched filtering output. We demonstrate that in the case of frequency zero-padding, uniform PA does not yield the lowest possible ranging sidelobe. To address this, we formulate an optimization problem and develop the Projected Gradient Descent (PGD) algorithm to conceive a PA scheme that minimizes the ranging sidelobe.
-
•
By comparing ranging performances, we observe that while the PA scheme obtained via the PGD algorithm achieves a lower sidelobe level, it results in a slight increase in the mainlobe width compared to the uniform PA scheme. Building upon this observation, we introduce a new constraint to limit the mainlobe width and employ the Successive Convex Approximation (SCA) algorithm to solve the corresponding optimization problem. Our results indicate the existence of a tradeoff between the mainlobe width and the sidelobe level.
The remainder of this paper is organized as follows. Section II introduces the system model of the ISAC system under consideration and the corresponding performance metrics. Section III derives the closed-form expressions for both P-ACF and A-ACF of random ISAC signals using OFDM modulation and PA schemes. The optimal PA strategies of P-ACF and A-ACF with OFDM modulation for random ISAC signals are discussed in Section IV. Section V conceives optimal PA strategies for P-ACF with frequency zero-padding and OFDM modulation in random ISAC signals. Section VI presents simulation results to validate the theoretical analysis. Finally, Section VII concludes the paper.
Notations: Matrices are denoted by bold uppercase letters (e.g., ), vectors are represented by bold lowercase letters (e.g., ), and scalars are denoted by normal font (e.g., ); The th entry of a vector , and the -th entry of a matrix are denoted as and , respectively; and denote the Kronecker product and the vectorization, , , and stand for transpose, Hermitian transpose, and the complex conjugate of the matrices; norm is written as , and represents the expectation operation; The notation Diag() denotes the diagonal matrix obtained by placing the entries of on its main diagonal.
II System Model
II-A ISAC Signal Model
This paper considers a monostatic ISAC system. The ISAC Tx emits a signal modulated with random communication symbols, which is received by a communication Rx and simultaneously reflected back to a sensing Rx by one or more targets at varying ranges. The sensing Rx, collocated with the ISAC Tx, performs matched filtering to estimate target delay parameters using the known random ISAC signal.
Let represent the communication symbols to be transmitted. When the OFDM modulation is adapted and the power allocation is considered, the discrete-time domain signal is given by
(1) |
where is the normalized discrete Fourier transform (DFT) matrix of size , and the matrix is the PA matrix, in the form of
(2) |
where represents the power allocated to th subcarrier, with the constraints and , for all .
Assumption 1 (Unit Power and Rotational Symmetry).
We focus on constellations with a unit power, zero mean, and zero pseudo-variance, defined as
(3) |
The unit-power normalization in Assumption 1 enables fair sensing and communication performance under varying constellation formats. Moreover, the properties of zero mean and zero pseudo-variance hold for most practical constellations (e.g., PSK and QAM families, excluding BPSK and 8-QAM).
To proceed, let us further define the kurtosis of a constellation as
(4) |
Under unit power and zero mean conditions, kurtosis reduces to the fourth-order moment . Using the Power Mean Inequality, we arrive at
(5) |
Specifically, all PSK constellations exhibit , while all QAM constellations have .
II-B Sensing Performance Metric
In radar systems, the Ambiguity Function (AF) of the signal is an important performance indicator for sensing [32].
(6) |
where is the radar signal, is the time delay and is Doppler shift. For the sake of simplicity, this paper considers only the zero-Doppler slice and sets aside the characterization of the complete AF as our future work. By letting , (6) can be rewritten as
(7) |
which is known as the ACF of the signal , and is well-recognized as a key performance indicator for ranging. In a discrete form, the ACF can be defined as the linear or periodic self-convolution of the discrete signal , depending on whether a CP is added.
II-B1 Aperiodic ACF (A-ACF)
(8) |
where is the th shift matrix in the form of
(9) |
II-B2 Periodic ACF (P-ACF)
In both cases, the sidelobe level of the ACF plays a critical role. Let us take the A-ACF as an example. The sidelobe of is defined as
(12) |
where == = is the mainlobe of the ACF. The integrated sidelobe level (ISL) can be expressed as [34], [35]
(13) |
Given the random nature of the signal , we define the average sidelobe level as the primary sensing performance metric. This metric is expressed as
(14) |
When is zero, denotes the average mainlobe level. Correspondingly, the EISL is given by
(15) |
Consequently, seeking the optimal PA strategy is equivalent to solving the following stochastic optimization problem:
(16) |
which is the EISL normalized by the expected mainlobe level.
III Statistical Characterizations of P-ACF and A-ACF
In [26], the closed-form expressions of P-ACF and A-ACF without PA have been theoretically analyzed. In this section, we systematically investigate the closed-form expressions of both P-ACF and A-ACF with PA in OFDM systems.
III-A The P-ACF Case
In this section, we present the main result of the P-ACF case with PA, which is generally easier to tackle than its A-ACF counterpart. Let us first demonstrate the basic structure of the periodic shift matrix in Lemma 1.
Lemma 1.
The periodic shift matrix can be decomposed as
(17) |
where is the th column of .
The P-ACF with PA can be expressed as
(18) |
where and is
(19) |
Using Lemma 1, (18) can be recast as
(20) |
Proposition 1.
The average squared P-ACF with PA is
(21) |
Proof.
See Appendix A. ∎
As a consequence, the average mainlobe of the P-ACF with PA is
(22) |
Corollary 1.
The EISL of the P-ACF with PA is
(23) |
Proof.
See Appendix B. ∎
III-B The A-ACF Case
In this section, we evaluate the ranging performance of random ISAC signals with PA by investigating the sidelobe level of A-ACF.
Similarly to (18), the expression of A-ACF with PA is given by
(24) |
where is a shift matrix and the size is . To apply Lemma 1, (24) can be rewritten as
(25) |
where is a periodic shift matrix with size and is a all-zero vector of length . According to Lemma 1, it can be decomposed as
(26) |
where is the th column of , which is the normalized DFT matrix of size .
Accordingly, (25) can be rewritten as
(27) |
where contains the first columns of . By denoting , it follows that
(28) |
The average mainlobe level and the EISL of the A-ACF are and respectively.
Proposition 2.
The average mainlobe and the EISL of the A-ACF with PA are respectively
(29) | ||||
(30) |
Proof.
See Appendix C. ∎
IV Optimal Power Allocation schemes
In this section, we analyze the optimal PA based on OFDM modulation for both the P-ACF and A-ACF cases.
IV-A The P-ACF Case
Theorem 1 (Global Optimality of the Uniform PA for Ranging).
When the OFDM signaling scheme is employed, the uniform PA is the only scheme that achieves the lowest normalized EISL of the P-ACF across all constellations.
Proof.
It is evident that both the average mainlobe and the EISL vary with the PA scheme. Therefore, we normalize the EISL with respect to the average mainlobe, yielding the expression
(31) |
indicating that the normalized EISL is minimized if and only if is minimized. Moreover, note that
(32) |
due to the Cauchy-Schwarz Inequality. The equality is attained if and only if . This suggests that the normalized EISL is minimized under uniform PA, completing the proof. ∎
Theorem 2 (The Uniform PA achieves the Lowest Sidelobe at Every Lag).
When the OFDM signaling scheme is used, the uniform PA is the only scheme that achieves the lowest sidelobe level at every delay index of the P-ACF for all constellations.
Proof.
According to (21) and (22), the normalized is shown as follows:
(33) |
When ,
(34) |
which maximizes the numerator of the second term of the right-hand side (RHS) in (33). Moreover, by applying an argument similar to the proof of Theorem 1, the denominator is minimized, which, in turn, maximizes the second term of the RHS of (33). Therefore, only when , the normalized reaches to its minimum, completing the proof. ∎
When the uniform PA scheme is employed, the average mainlobe level is
(35) |
the average sidelobe level is
(36) |
and the normalized EISL is expressed as
(37) |
An interesting observation is that for PSK constellations (), both the individual and the integrated sidelobe levels are zero under OFDM signaling with uniform power. This implies that the P-ACF of an OFDM-PSK signal with uniform power is always a unit impulse function. This observation is consistent with the results introduced in [26].
IV-B The A-ACF Case
Theorem 3 (Uniform PA is Globally Optimal for All Constellations).
When the OFDM signaling scheme is used, the uniform PA is the only PA scheme that achieves the lowest normalized EISL of the A-ACF for all constellations.
Proof.
See Appendix D ∎
At this time, the average mainlobe of A-ACF is
(38) |
and the normalized EISL of A-ACF is
(39) |
V CP-OFDM With power allocation and frequency zero-padding
In this section, we analyze the P-ACF for the OFDM scheme incorporating PA and frequency zero-padding. Zero-padding refers to the process of appending zeros to the end of a signal or data sequence to extend its length. Specifically, zero-padding in the frequency domain involves adding zero values to the high-frequency portion of the signal’s spectrum, thereby increasing the overall length of the spectrum. By doing this, the resolution of the time-domain signal is enhanced through increasing the number of sampling points. Consequently, the accuracy of the range estimation may be improved.
According to (20), when the OFDM modulation and PA are applied, the P-ACF is the inverse discrete Fourier transform (IDFT) of the sequence up to a factor . Through frequency zero-padding with null subcarriers, where represents the padding factor, we generate an augmented sequence with an extended length of samples, yielding
(40) |
where represents the all-zero vector with length .
Analogous to the derivation in (20), we construct an enhanced P-ACF incorporating both PA and frequency-domain zero-padding. The modified P-ACF, denoted as , is expressed as
(41) |
which is the IDFT of the sequence up to a factor .
Proposition 3.
The expectation of the integrated sidelobe level EISLZP and the average mainlobe level of the P-ACF with PA and frequency zero-padding can be expressed as
(42) | ||||
(43) |
where
(44) | ||||
(45) | ||||
(46) |
Proof.
See Appendix E. ∎
The normalized can be denoted as
(47) |
Through comprehensive theoretical analysis and numerical simulations, we found that the uniform PA scheme ceases to be optimal for minimizing the normalized when frequency zero-padding is implemented. To address this limitation, we formulate the following constrained optimization problem to derive a suitable PA strategy that can minimize the normalized .
(48) | ||||
Notably, the optimization problem in (48) requires minimizing a fractional objective function over the simplex. To solve this constrained problem, we adopt the PGD algorithm to obtain a locally optimal solution. As outlined in Algorithm 1, the PGD method enhances the conventional Gradient Descent (GD) by introducing a projection step, which ensures that each iteration remains within the feasible region [36], [37]. The gradient of the objective function at an arbitrary point is given by
(49) |
where
(50) | ||||
(51) |
Moreover, the projector over the simplex can be expressed as
(52) |
where represents the all-one vector with length and denotes the feasible region of the optimization variable .
The average sidelobe levels achieved by the proposed PA scheme, obtained via Algorithm 1, are illustrated in Fig. 5 and Fig. 6. In these figures, “optimal PA” denotes the PA scheme derived from Algorithm 1. The results indicate that the optimal PA offers considerable sidelobe suppression compared to the “uniform PA”. However, this improvement is accompanied by a slight broadening of the mainlobe, which in turn leads to a degradation in range resolution performance.
To mitigate the adverse effects of PA on mainlobe width, we propose to add a new constraint in the original optimization problem (48). The enhanced formulation is given by:
(53) | ||||
The newly added constraint ensures that the average sidelobe level at delay remains within 3 dB of the average mainlobe level . This condition is equivalent to .
Compared to the original formulation in (48), the introduction of the third constraint in (53) makes it challenging to project onto the feasible region. As a result, the PGD algorithm is no longer directly applicable for solving the modified optimization problem. Upon analyzing the problem in (53), we observe that although all three constraints are convex, the objective function remains non-convex. Given these conditions, the Successive Convex Approximation (SCA) algorithm becomes a suitable alternative. SCA is a powerful framework for handling non-convex optimization problems, as it addresses non-convex optimization problems through iterative minimization of convex surrogate functions [36, 38, 39]. The core procedure of the SCA algorithm lies in constructing an appropriate surrogate function and determining a suitable step size . In Algorithm 2, the surrogate function is defined as
(54) |
which is the first-order Taylor expansion of the objective function near the th iteration point. Furthermore, the step size at each iteration is determined via the exact line search method.
VI Simulation Results
In this section, we further validate our conclusions and theoretical analysis by presenting numerical results. All the simulation results are attained by averaging over random realizations.
First, we focus on the P-ACF with PA of 16-QAM and 64-QAM. In Fig. 1, the sidelobe performances of 16-QAM and 64-QAM constellations for CP-OFDM with uniform PA are shown. It can be observed that the theoretical values align closely with the simulation results, confirming the accuracy of the derivations in Section III-A. Additionally, a 3 dB performance gain in the peak-to-sidelobe level ratio is achieved by doubling . In Fig. 2, the sidelobe performances of the P-ACF for 64-QAM with two PA schemes, uniform PA and random PA, are compared. It is clear that the uniform PA scheme achieves a much lower sidelobe level than the random PA scheme.


In addition to the P-ACF, the sidelobe level performance of the A-ACF with PA is shown in Fig. 3 for the 16-QAM and 16-PSK constellations. We examine the normalized EISL without CP under different PA schemes. We observe that the theoretical results match perfectly with the numerical simulations for both the average and random PA schemes, confirming the accuracy of the theoretical derivations in Section III-B. Additionally, the normalized EISL for the uniform PA scheme is significantly lower than that of the random PA scheme in both the 16-QAM and 16-PSK constellations.


To verify the theoretical formulation of the normalized in (47), we present comparative simulations of computed values for different constellation schemes and PA methods in Fig. 4, where the proposed “optimal PA” scheme utilizes the power vector obtained through Algorithm 1. The close agreement between theoretical predictions and simulation measurements validates our analytical model. Notably, under the “optimal PA” scheme, the normalized for 16-PSK is significantly lower than that of the “uniform PA” case. For 16-QAM, the reduction is more modest, but the “optimal PA” still achieves a lower normalized compared to the uniform counterpart.
Next, we examine the sidelobe performances of 16-QAM and 16-PSK constellations with frequency zero-padding and different PA schemes. Fig. 5 and Fig. 6 display the P-ACF of 16-QAM and 16-PSK under OFDM signaling across different PA schemes and values of . For both 16-QAM and 16-PSK, the “optimal PA” configuration achieves a lower sidelobe level than that of the “uniform PA”. Analyzing these figures further, while the “optimal PA” reduces the sidelobe level, it also slightly increases the mainlobe width compared to the “uniform PA” scheme.


In Fig. 7, we compare the convergence results of Algorithm 2 for 16-QAM and 16-PSK under different 3 dB mainlobe width constraints, where the limiting positions are set to , and , respectively. The “uniform PA” scheme and the “unconstrained mainlobe width” scheme serve as baseline references, where the latter refers to the PA vector obtained from Algorithm 1 without any constraint on mainlobe width. It is observed that as the 3 dB constraint becomes more stringent (i.e., smaller ), the convergence result increasingly resembles that of the “uniform PA” scheme. Conversely, when the constraint is relaxed (i.e., larger ), the convergence result tends to align more closely with the “unconstrained mainlobe width” case.




Combining Figs. 7, 8, and 9, we observe that as the PA scheme approaches the “uniform PA” scheme, the mainlobe width becomes narrower, yet the sidelobe level increases. This trend holds for both 16-QAM and 16-PSK constellations. These results suggest a tradeoff between mainlobe width and sidelobe level, allowing for a flexible selection of PA schemes based on specific performance requirements.
VII Conclusion
In this paper, we investigated optimal PA for OFDM-based random ISAC signals. Our findings demonstrated that the uniform PA achieves the global minimum EISL for both periodic and aperiodic autocorrelation functions for all constellations. Additionally, we proved that this scheme achieves the lowest sidelobe at every P-ACF lag. For frequency zero-padding scenarios, where uniform PA is suboptimal, we employed efficient optimization algorithms to design tailored PA schemes that balance the trade-off between sidelobe suppression and mainlobe width. The effectiveness of our theoretical framework was further validated through comprehensive simulations. Future research should explore the optimal signal basis and PA schemes for two-dimensional ambiguity function (2D AF) design.
Appendix A Proof of Proposition 1
By referring to (20), the squared P-ACF with power allocation can be expressed as
(55) |
Correspondingly, the average squared P-ACF with PA is
(56) |
Appendix B Proof of Corollary1
Appendix C Proof of proposition 2
According to (III-B)
(59) |
Expanding yields
(60) |
Therefore
(61) |
where . To simplify (61), can be decomposed [26] as
(62) |
where
(63) | ||||
(64) |
with being the all-zero matrix of size , and
(65) |
Plugging (62) into (61), we can get
(66) | |||
(67) | |||
(68) |
According to (61) - (68), the average mainlobe is
(69) |
(70) |
Appendix D Proof of theorem 3
By observing the structure of , we can extract all the odd rows and reorganize them into a matrix , while all the even rows can be reorganized as .
(81) |
where is a diagonal matrix with its n-th diagonal entry begin .
We set , so . Because of the specificity of the , we can find that all odd-numbered columns of can form an identity matrix, and all even-numbered columns can form a cyclic matrix.
(82) | |||
(83) |
Similarly, the first term of the RHS of (80) can be divided according to odd and even columns. Accordingly, the normalized EISL of A-ACF (80) can be rewritten as (84).
(84) |
According to (2) and (82), we can get
(85) |
Plugging (85) into the first term of (84), this term can be reduced to (86). When the value of is minimized, the value of (86) is also minimized. According to (31) and (32), only when , (86) reaches the minimum.
(86) |
Based on the conditions that and , the Cauchy-Schwarz Inequality can be used in the second term of (84), as is shown in (D). Only when every is equal, the (D) reaches the minimum. According to (83), when , the conditions for equality are met.
In conclusion, when , the normalized EISL achieves the global minimum.
(87) |
Appendix E Proof of proposition 3
According to (41), the sidelobe of is
(88) |
The average sidelobe of P-ACF with PA and frequency zero-padding is
(89) |
Taking into the above equation, the average mainlobe of the P-ACF with PA and frequency zero-padding is
(90) |
References
- [1] ITU-R WP5D, “Draft New Recommendation ITU-R M. [IMT. Framework for 2030 and Beyond],” 2023.
- [2] M. Chafii, L. Bariah, S. Muhaidat, and M. Debbah, “Twelve scientific challenges for 6G: Rethinking the foundations of communications theory,” IEEE Communications Surveys & Tutorials, vol. 25, no. 2, pp. 868–904, 2023.
- [3] F. Liu, Y. Cui, C. Masouros, J. Xu, T. X. Han, Y. C. Eldar, and S. Buzzi, “Integrated sensing and communications: Toward Dual-Functional Wireless Networks for 6G and Beyond,” IEEE Journal on Selected Areas in Communications, vol. 40, no. 6, pp. 1728–1767, 2022.
- [4] 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 Transactions on Communications, vol. 68, no. 6, pp. 3834–3862, 2020.
- [5] Z. Wei, H. Qu, Y. Wang, X. Yuan, H. Wu, Y. Du, K. Han, N. Zhang, and Z. Feng, “Integrated sensing and communication signals toward 5G-A and 6G: A survey,” IEEE Internet of Things Journal, no. 13, pp. 11 068–11 092, 2023.
- [6] Z. Wei, Y. Wang, L. Ma, S. Yang, Z. Feng, C. Pan, Q. Zhang, Y. Wang, H. Wu, and P. Zhang, “5G PRS-Based Sensing: A sensing reference signal approach for joint sensing and communication system,” IEEE Transactions on Vehicular Technology, vol. 72, no. 3, pp. 3250–3263, 2023.
- [7] A. R. Chiriyath, B. Paul, and D. W. Bliss, “Radar-communications convergence: Coexistence, cooperation, and co-design,” IEEE Transactions on Cognitive Communications and Networking, vol. 3, no. 1, pp. 1–12, 2017.
- [8] A. Hassanien, M. G. Amin, E. Aboutanios, and B. Himed, “Dual-function radar communication systems: A solution to the spectrum congestion problem,” IEEE Signal Processing Magazine, vol. 36, no. 5, pp. 115–126, 2019.
- [9] J. A. Zhang, F. Liu, C. Masouros, R. W. Heath, Z. Feng, L. Zheng, and A. Petropulu, “An overview of signal processing techniques for joint communication and radar sensing,” IEEE Journal of Selected Topics in Signal Processing, vol. 15, no. 6, pp. 1295–1315, 2021.
- [10] S. Li, F. Zhang, T. Mao, R. Na, Z. Wang, and G. K. Karagiannidis, “Transmit beamforming design for ISAC with stacked intelligent metasurfaces,” IEEE Transactions on Vehicular Technology, pp. 1–6, 2024.
- [11] W. Zhou, R. Zhang, G. Chen, and W. Wu, “Integrated sensing and communication waveform design: A survey,” IEEE Open Journal of the Communications Society, vol. 3, pp. 1930–1949, 2022.
- [12] Z. Xiao and Y. Zeng, “Waveform design and performance analysis for full-duplex integrated sensing and communication,” IEEE Journal on Selected Areas in Communications, vol. 40, no. 6, pp. 1823–1837, 2022.
- [13] F. Zhang, T. Mao, R. Liu, Z. Han, S. Chen, and Z. Wang, “Cross-domain dual-functional ofdm waveform design for accurate sensing/positioning,” IEEE Journal on Selected Areas in Communications, vol. 42, no. 9, pp. 2259–2274, 2024.
- [14] T. Mao, J. Chen, Q. Wang, C. Han, Z. Wang, and G. K. Karagiannidis, “Waveform design for joint sensing and communications in millimeter-wave and low terahertz bands,” IEEE Transactions on Communications, vol. 70, no. 10, pp. 7023–7039, 2022.
- [15] R. M. Mealey, “A method for calculating error probabilities in a radar communication system,” IEEE Transactions on Space Electronics and Telemetry, vol. 9, no. 2, pp. 37–42, 1963.
- [16] M. Roberton and E. Brown, “Integrated radar and communications based on chirped spread-spectrum techniques,” in IEEE MTT-S International Microwave Symposium Digest, 2003, vol. 1, 2003, pp. 611–614.
- [17] P. Raviteja, K. T. Phan, Y. Hong, and E. Viterbo, “Interference cancellation and iterative detection for orthogonal time frequency space modulation,” IEEE Transactions on Wireless Communications, vol. 17, no. 10, pp. 6501–6515, 2018.
- [18] C. Sturm and W. Wiesbeck, “Waveform design and signal processing aspects for fusion of wireless communications and radar sensing,” Proceedings of the IEEE, vol. 99, no. 7, pp. 1236–1259, 2011.
- [19] F. Liu, L. Zhou, C. Masouros, A. Li, W. Luo, and A. Petropulu, “Toward dual-functional radar-communication systems: Optimal waveform design,” IEEE Transactions on Signal Processing, vol. 66, no. 16, pp. 4264–4279, 2018.
- [20] F. Liu, Y.-F. Liu, A. Li, C. Masouros, and Y. C. Eldar, “Cramér-Rao bound optimization for joint radar-communication beamforming,” IEEE Trans. Signal Process., vol. 70, pp. 240–253, Dec. 2021.
- [21] C. Sturm and W. Wiesbeck, “Waveform design and signal processing aspects for fusion of wireless communications and radar sensing,” Proceedings of the IEEE, vol. 99, no. 7, pp. 1236–1259, 2011.
- [22] X. Chen, Z. Feng, Z. Wei, P. Zhang, and X. Yuan, “Code-division ofdm joint communication and sensing system for 6G machine-type communication,” IEEE Internet of Things Journal, vol. 8, no. 15, pp. 12 093–12 105, 2021.
- [23] Y. Zeng, Y. Ma, and S. Sun, “Joint radar-communication with cyclic prefixed single carrier waveforms,” IEEE Transactions on Vehicular Technology, vol. 69, no. 4, pp. 4069–4079, 2020.
- [24] L. Gaudio, M. Kobayashi, G. Caire, and G. Colavolpe, “On the effectiveness of OTFS for joint radar parameter estimation and communication,” IEEE Transactions on Wireless Communications, vol. 19, no. 9, pp. 5951–5965, 2020.
- [25] Y. Xiong, F. Liu, Y. Cui, W. Yuan, T. X. Han, and G. Caire, “On the fundamental tradeoff of integrated sensing and communications under Gaussian channels,” IEEE Trans. Inf. Theory, vol. 69, no. 9, pp. 5723–5751, Jun. 2023.
- [26] F. Liu, Y. Zhang, Y. Xiong, S. Li, W. Yuan, F. Gao, S. Jin, and G. Caire, “OFDM achieves the lowest ranging sidelobe under random ISAC signaling,” 2024. [Online]. Available: https://arxiv.org/abs/2407.06691
- [27] J. Zhu, J. Mu, Y. Cui, and X. Jing, “Mutual information maximization via joint power allocation in integrated sensing and communications system,” China Communications, vol. 21, no. 2, pp. 129–142, 2024.
- [28] R. Yao, Z. Wei, L. Su, L. Wang, and Z. Feng, “Low-PAPR integrated sensing and communication waveform design,” in 2023 IEEE Wireless Communications and Networking Conference (WCNC), 2023, pp. 1–6.
- [29] C. Shi, F. Wang, S. Salous, and J. Zhou, “Joint subcarrier assignment and power allocation strategy for integrated radar and communications system based on power minimization,” IEEE Sensors Journal, vol. 19, no. 23, pp. 11 167–11 179, 2019.
- [30] Y. Chen, G. Liao, Y. Liu, H. Li, and X. Liu, “Joint subcarrier and power allocation for integrated ofdm waveform in radcom systems,” IEEE Communications Letters, vol. 27, no. 1, pp. 253–257, 2023.
- [31] F. Wang and H. Li, “Power allocation for coexisting multicarrier radar and communication systems in cluttered environments,” IEEE Transactions on Signal Processing, vol. 69, pp. 1603–1613, 2021.
- [32] N. Levanon, E. Mozeson, R. Signals, and C. Levanon, “Radar signals,” 2013. [Online]. Available: https://api.semanticscholar.org/CorpusID:9503789
- [33] P. Stoica, H. He, and J. Li, “On designing sequences with impulse-like periodic correlation,” IEEE Signal Processing Letters, vol. 16, no. 8, pp. 703–706, 2009.
- [34] J. Song, P. Babu, and D. P. Palomar, “Optimization methods for designing sequences with low autocorrelation sidelobes,” IEEE Transactions on Signal Processing, vol. 63, no. 15, pp. 3998–4009, 2015.
- [35] P. Stoica, H. He, and J. Li, “New algorithms for designing unimodular sequences with good correlation properties,” IEEE Transactions on Signal Processing, vol. 57, no. 4, pp. 1415–1425, 2009.
- [36] S. Boyd and L. Vandenberghe, Convex Optimization. Cambridge University Press, 2004.
- [37] S. Lu, F. Liu, F. Dong, Y. Xiong, J. Xu, Y.-F. Liu, and S. Jin, “Random ISAC signals deserve dedicated precoding,” IEEE Transactions on Signal Processing, vol. 72, pp. 3453–3469, 2024.
- [38] A. Liu, V. K. N. Lau, and B. Kananian, “Stochastic successive convex approximation for non-convex constrained stochastic optimization,” IEEE Transactions on Signal Processing, vol. 67, no. 16, pp. 4189–4203, 2019.
- [39] M. Kojic, Z. Luzanin, and I. Stojkovska, “Adaptive stochastic approximation algorithm,” Numerical Algorithms, vol. 76, pp. 917–937, 12 2017.