Wednesday, April 24, 2024

Super Radar: Pioneering Research Overcomes Historic Trade-Offs Between Distance and Detail

Advanced Radar Technology Art Concept
Breakthrough radar technology improves distance resolution,
impacting multiple industries and enabling advanced applications
like detailed surface characterization.

Summary of Paper

The article discusses a breakthrough in radar technology that significantly improves range resolution, allowing for much more detailed characterization of surfaces and distinction between closely spaced objects. The key points are:

  1. Researchers from Chapman University and other institutions have developed new radar technology using interference functions that overcomes the traditional trade-off between observation distance and detail.
  2. The new method can resolve distances between objects over 100 times better than the long-believed limit of one-quarter wavelength. It can measure distances between objects to about 20 μm, which is 35,000 times smaller than the previous Rayleigh-resolved limit, even when using low frequency radio waves.
  3. This breakthrough could have important applications in fields like military, construction, archaeology, and medical sensing. It opens up new possibilities for efficient humanitarian demining and high-resolution non-invasive medical imaging.
  4. The technology relies on purposely designed self-referencing waveforms that generate unique interference patterns when reflected off of closely spaced surfaces. Parameter estimation techniques are then used to precisely determine the distances.
  5. This amplitude-based ranging paradigm is fundamentally different from traditional temporally-resolved methods. It allows radar to maintain high spatial resolution even at very long wavelengths that penetrate well through various media.

In summary, this novel radar technology provides orders of magnitude improvement in range resolution compared to current limits, promising to significantly enhance capabilities across multiple industries and enable new high-resolution remote sensing applications. 

Principle of Operation

The key innovation in this new radar technology is the use of specially designed "self-referencing" waveforms that create unique interference patterns when reflected off surfaces at different depths. This allows tiny differences in reflection times to be discerned from the amplitude variations in the combined reflected signal.

Here's a step-by-step explanation of how it works:

  1. A purposely crafted "interference-class" waveform is transmitted toward the target. This waveform has specific properties:
    1.    a) It contains regions with steep gradients that are very sensitive to interference effects.
    2.    b) It also has flat "reference" regions that are insensitive to interference, providing a baseline.
  2. When this waveform reflects off objects at different depths (e.g. two surfaces of an object), the reflections will have slight time delays between them.
  3. These time-delayed reflections overlap and interfere with each other, creating an interference pattern in the combined reflected signal that returns to the radar receiver.
  4. The interference pattern will have its peaks and troughs shifted very slightly based on the tiny time delay between the reflections. This shift is much smaller than the actual waveform features.
  5. By comparing the amplitudes of the interference region to the flat reference regions, and applying parameter estimation techniques, these minuscule time delays can be precisely quantified.
  6. From the time delays, the distance between the reflecting surfaces can be calculated with micrometer-scale precision, even though the radar wavelength is much larger (e.g. low frequency radio waves).

The key is that the interference pattern contains information at a much finer scale than the waveform itself. By deliberately designing waveforms to create these telltale interference patterns and having a reliable reference, tiny path length differences can be discerned without having to resolve the actual time delays, which would be impossible at long wavelengths.

This allows radar to achieve extremely high range resolution while still enjoying the benefits of low frequency waves, such as better penetration through many materials. The technique is robust because the self-referencing design makes it insensitive to signal attenuation or amplification that affects the entire waveform.

Validation of Principle

The researchers have experimentally validated their technique through a series of proof-of-principle experiments:
  1. They tested the ranging resolution using a setup where they sent the crafted waveforms through cables of different lengths (introducing controlled time delays) and observed the resulting interference patterns.
  2. In one experiment, they used a 72 cm cable to introduce a 7.6 nanosecond round-trip delay. By varying the bandwidth of their "triangle function" waveform, they showed they could distinguish the signal with and without the delay cable down to about 1 MHz bandwidth. This corresponds to a resolution over 100 times better than the inverse bandwidth limit.
  3. In another experiment, they used just the tiny path length difference introduced by a BNC T-junction (about 1.3 cm). With a high sample rate and bandwidth, they could clearly distinguish the signal with and without the T-junction connected, demonstrating sub-millimeter resolution.
  4. They also quantified the ability to measure the distance between two reflectors when the reflections are overlapped in time (temporally unresolved). Using their specially designed waveform and varying the bandwidth slightly, they showed they could measure changes in the reflection delay corresponding to about 20 μm, which is around 35,000 times smaller than the nominal resolution given by the waveform's bandwidth.
  5. The experimental results matched well with theoretical predictions, validating the underlying principles.
These experiments demonstrate the technique's ability to achieve super-resolution ranging in controlled conditions. The researchers note that the principles apply across the electromagnetic spectrum, so this technique could be adapted to different frequency ranges for different applications.

However, it's important to note that these are initial proof-of-principle experiments in idealized setups. Further research will be needed to develop practical systems that can handle real-world complexities, such as more complex target geometries, variable reflectivities, signal distortions, and noise. Nonetheless, these initial laboratory demonstrations provide strong evidence for the fundamental validity and potential of the technique.

Artifacts and Reproducibility

The paper does provide some experimental data to support their claims, but you're right that having access to more detailed test artifacts and datasets would be valuable for independent verification and to assess the robustness and limitations of the technique.

The experimental results presented in the paper include:

1. A plot (Figure 3) showing the signal amplitude vs. bandwidth for their "triangle function" waveform, with and without a delay cable, demonstrating the ability to resolve the delay at low bandwidths.

2. Numerical results for the experiment with the BNC T-junction, giving the signal values with and without the junction, showing a statistically significant difference.

3. A plot (Figure 4) of the signal vs. time shift for their specially designed waveform, showing the experimental data points matching well with the theoretical curve.

4. An inset in Figure 4 showing the change in signal for small incremental frequency shifts, with error bars, which is used to estimate the achievable distance resolution.

While these results support their claims, the paper doesn't seem to include or link to complete datasets that would allow others to fully reproduce and verify the experiments. Ideally, for independent verification, one would want access to:
  • - The exact waveform definitions used in the experiments
  • - Raw data from the oscilloscope readings
  • - Detailed experimental setup parameters
  • - Analysis scripts used to process the data and generate the results
It's not uncommon for initial proof-of-principle papers to focus on presenting the key results rather than providing comprehensive data and code for replication. However, in fields like physics and engineering, there is a growing movement towards more open and reproducible research practices.

It would be reasonable and potentially valuable for the research community to request that the authors make their full experimental data and analysis scripts available, either as supplementary information with the paper or in a public data repository. This would allow other researchers to independently verify the results, test the techniques under different conditions, and build upon the work.

Ultimately, the true test of a new technique like this is whether other research groups are able to replicate and extend the results. If this technique proves to be robust and widely applicable, we would expect to see follow-up papers from both the original authors and other groups, providing further experimental validation and exploring practical applications. The additional data and details provided in such follow-up work would help to solidify the evidence for the technique's capabilities and limitations.

The paper does include some analytical and simulation results to validate the underlying principles of the technique.
  • 1. Analytical expressions: The authors introduce analytical expressions for their specially designed waveforms. These include:
   - Equation 2, which defines their custom waveform as a product of a "canvas" function and a polynomial.
   - Equation 7, which defines a signal metric S that quantifies the interference effect based on the peak and trough amplitudes in different regions of the waveform.
   These expressions provide a mathematical foundation for understanding how the waveforms are constructed and how the interference effects are quantified.
  • 2. Theoretical plots: The paper includes plots of the theoretical waveforms and their Fourier transforms (Figure 2). These plots demonstrate the key properties of the waveforms, such as the steep gradients in the interference region and the flat reference regions.
  • 3. Theoretical signal curve: In Figure 4, the experimental data points for the signal S as a function of time shift are plotted alongside a theoretical curve. The close match between the theoretical curve and the experimental points supports the validity of the underlying analytical model.
  • 4. Scaling arguments: The authors use scaling arguments to estimate the achievable distance resolution based on the slope of the signal curve and the bandwidth resolution. These theoretical estimates are consistent with the experimental results.
These analytical and theoretical components of the paper provide a conceptual and quantitative framework for understanding how the technique works and what its fundamental capabilities and limitations are.

However, the paper does not appear to include detailed numerical simulations that test the technique's performance under a wide range of conditions. Such simulations could be valuable for exploring factors like:
  • - The effects of noise and signal distortions
  • - The impact of target geometry and reflectivity variations
  • - The performance limits with different waveform designs
  • - The scalability to different frequency ranges and applications
Rigorous numerical simulations could help to further validate the technique and guide its optimization and practical implementation. They could also help to identify potential challenges or limitations that may not be apparent from the simplified analytical models or idealized experimental setups.

As this is a novel technique, it's likely that future work by both the original authors and other researchers will include more extensive simulations and analytical modeling to further explore and validate the approach. Such theoretical and computational studies, combined with additional experimental data, will be important for establishing the robustness, versatility, and practical potential of this super-resolution radar technique.


Researchers have developed new radar technology that improves distance resolution significantly, promising enhancements in various fields such as military and archaeology. Credit: SciTechDaily.com

New interference radar functions employed by a team of researchers from Chapman University and other institutions improve the distance resolution between objects using radar waves. The results may have important ramifications in military, construction, archaeology, mineralogy, and many other domains of radar applications.

This first proof-of-principle experiment opens a new area of research with many possible applications that can be disruptive to the multi-billion dollar radar industry. There are many new avenues to pursue both in theory and experiment.

Resolving a Decades-Old Challenge

The discovery addresses a nine-decade-old problem that requires scientists and engineers to sacrifice detail and resolution for observation distance — underwater, underground, and in the air. The previous bound limited the distance estimated between objects to be one quarter of the wavelength of radio waves; this technology improves the distance resolution between objects using radar waves.

“We believe this work will open a host of new applications as well as improve existing technologies,” says John Howell, the lead author of the article published in Physical Review Letters and highlighted as an Editors’ Suggestion paper (see Radar Resolution Gets a Boost). “The possibility of efficient humanitarian demining or performing high-resolution, non-invasive medical sensing is very motivating,” Howell adds.

Enhancements in Radar Resolution

Howell and a team of researchers from the Institute for Quantum Studies at Chapman University, the Hebrew University of Jerusalem, the University of Rochester, the Perimeter Institute and the University of Waterloo have demonstrated range resolution more than 100 times better than the long-believed limit. This result breaks the trade-off between resolution and wavelength, allowing operators to use long wavelengths and now have high spatial resolution.

By employing functions with both steep and zero-time gradients, the researchers showed that it was possible to measure extremely small changes in the waveform to precisely predict the distance between two objects while still being robust to absorption losses. To an archaeologist, this creates the ability to distinguish between a coin deep underground from a pottery shard.

The breakthrough idea relies on the superposition of specially-crafted waveforms. When a radio wave reflects from two different surfaces, the reflected radio waves add to form a new radio wave. The research team uses purpose-designed pulses to generate a new kind of superposed pulse. The composite wave has unique sub-wavelength features that can be used to predict the distance between the objects.

Transforming Radar Sensing

“In radio engineering, interference is a dirty word and thought of as a deleterious effect. Here, we turn this attitude on its head, and use wave interference effects to break the long-standing bound on radar ranging by orders of magnitude,” says Andrew Jordan, director of Quantum Studies at Chapman University. “In remote radar sensing, only a small amount of the electromagnetic radiation is returned to the detector. The tailored waveforms that we designed have the important property of being self-referencing, so properties of the target can be distinguished from loss of signal.”

Howell adds, “We are now working to demonstrate that it is possible to not only measure the distance between two objects, but many objects or perform detailed characterization of surfaces.”

Reference: “Super Interferometric Range Resolution” by John C. Howell, Andrew N. Jordan, Barbara Šoda and Achim Kempf, 2 August 2023, Physical Review Letters.
DOI: 10.1103/PhysRevLett.131.053803

“Super Interferometric Range Resolution” 

by John C. Howell, Andrew N. Jordan, Barbara Šoda and Achim Kempf

Abstract

We probe the fundamental underpinnings of range resolution in coherent remote sensing. We use a novel class of self-referential interference functions to show that we can greatly improve upon currently accepted bounds for range resolution. We consider the range resolution problem from the perspective of single-parameter estimation of amplitude versus the traditional temporally resolved paradigm. We define two figures of merit: (i) the minimum resolvable distance between two depths and (ii) for temporally subresolved peaks, the depth resolution between the objects. We experimentally demonstrate that our system can resolve two depths greater than 100× the inverse bandwidth and measure the distance between two objects to approximately 20μm (35 000 times smaller than the Rayleigh-resolved limit) for temporally subresolved objects using frequencies less than 120 MHz radio waves.

  • Received 7 March 2023
  • Accepted 13 June 2023

DOI:https://doi.org/10.1103/PhysRevLett.131.053803

Published by the American Physical Society under the terms of the Creative Commons Attribution 4.0 International license. Further distribution of this work must maintain attribution to the author(s) and the published article’s title, journal citation, and DOI.

Published by the American Physical Society

Physics Subject Headings (PhySH)

Energy Science & TechnologyInterdisciplinary PhysicsPhysics of Living SystemsAtomic, Molecular & Optical

Focus

Key Image

Radar Resolution Gets a Boost

Published 2 August 2023

A low-frequency radar method with improved resolution could aid in the detection of landmines and archeological objects.

See more in Physics

Authors & Affiliations

John C. Howell1,2,*, Andrew N. Jordan1,3,4, Barbara Šoda5, and Achim Kempf5

  • 1Institute for Quantum Studies, Chapman University, Orange, California 92866, USA
  • 2Racah Institute of Physics, The Hebrew University of Jerusalem, Jerusalem, Israel, 91904
  • 3Center for Coherence and Quantum Optics, University of Rochester, Rochester, New York 14627, USA
  • 4Department of Physics and Astronomy, University of Rochester, Rochester, New York 14627, USA
  • 5Department of Applied Mathematics and Department of Physics, University of Waterloo, and Perimeter Institute for Theoretical Physics, Waterloo, Ontario, Canada

  • *johhowell@chapman.edu

Article Text

Range resolution [1–6] is the ability to determine the distance between two objects along the same line-of-sight when performing remote sensing. The prevailing thought is that radar range resolution is inextricably linked to the inverse bandwidth of a pulse [5,6] or to the wavelength of the electromagnetic wave owing to the coherent nature of the interfering wavefronts. We quote, “Wave theory indicates that the best vertical resolution that can be achieved is one-quarter of the dominant wavelength [7]. Within that vertical distance any reflections will interfere in a constructive manner and result in a single, observed reflection” [8]. The desire for better range resolution has driven scientists and engineers to ever-higher frequencies for radar and lidar [9–12]. However, this comes at a severe cost because transmission through and reflection from various material media is critically tied to frequency [8,13–16]. We dramatically improve upon these widely accepted limits of range resolution using a novel class of self-referenced functions to demonstrate several orders of magnitude improvement in range resolution beyond known limits.

For transform-limited pulses, two radar targets are considered range resolved when the range resolution distance dr obeys the inequality

drcτ2,
(1)

where c is the speed of light, τ is the pulse width, and 2 comes from the round-trip of the pulse [5,6]. Phase or frequency encoding are commonly employed to realize high time-bandwidth product pulses, which when combined with match-filtered pulse compression, lead to high temporal resolution. Therefore, a more general range resolution for a pulse is set by the inverse bandwidth [5,6]. Going beyond these limits has been historically difficult. Wong and Elefthiardes used superoscillations to reduce ranging uncertainty by 36% [17]. Recently, Komissarov et al. used partially coherent radar in an attempt to decouple range resolution from the signal bandwidth [18] achieving a factor of 10 improvement. We classify the current state of the art as a “temporally resolved” paradigm.

Here, we introduce a new “amplitude resolved” paradigm using self-referenced interference-class functions as shown in Fig. 1. An interference-class pulse is sent to a remote object. Multiple scattering depths along the same line of sight result in the interference of temporally shifted versions of the waveform. The resultant waveform is measured by the receiver. In theory, by using parameter estimation, we can determine range, distance between objects, relative scattering amplitudes, etc. We expect that amplitude-resolution ranging methods have been unexplored because of the inability to distinguish between loss and subresolved interference peaks.

FIG. 1.

Free-space (a) and guided-wave (b) and (c) ranging experimental schematics. Using parameter estimation techniques on interference-class waveforms, we estimate the distance between two scattering depths from a target which would otherwise be temporally subresolved.

Our system relies on parameter estimation from the interference between coherent pulses. A fundamentally different parameter estimation has also been used to overcome the spatial Rayleigh resolution limit of incoherent sources using mode sorting [19–23], allowing for fundamental definitions of spatial resolution [24–26]. Of note, Ansari et al used two incoherent optical pulses and mode decomposition to achieve supertemporal resolution [27].

In designing our functions, we employ two behaviors. First, a region of the function must be very sensitive to interference requiring extended and steep temporal gradients. Second, a zero-gradient region within the function, which is insensitive to the interference, is used as an amplitude reference. In this manner, as long as all portions of the pulse experience the same attenuation (or amplification) and there is a flat spectral response in a medium or upon reflection, the range resolution properties of the pulse are preserved.

We consider three types of pulse functions as shown in Fig. 2(a). The dashed orange line shows a specially designed interference-class, bandlimited pulse. The second interference class function is shown in solid green line, which we denoted as a “triangle” function. The dotted blue line shows a standard sinc2 pulse, which is not an interference-class function, but is only used as a means of comparison of a Rayleigh criterion for temporal pulse resolution. The bandlimit for the sinc2 and the custom function, in the image is the same. The corresponding fast Fourier transforms are shown in Fig. 2(b).

FIG. 2.

Pulses and fast Fourier transforms. Panel (a) shows the functions used in the Letter. Panel (b) shows their Fourier transforms. Panel (c) shows the interfered waveforms from the coherent interference of two equal-amplitude pulses.

For the first type of self-referenced pulses, we use bandlimited function theory (e.g., used in superoscillations [28–32]) to generate our specially designed functions. While it is likely that similar behavior could be achieved with many different types bandlimited functions, we use the formalism of Šoda and Kempf [32]. We generate our tailored and bandlimited functions by exploiting the product of a bandlimited “canvas” c(t) function and arbitrary (e.g., Taylor) polynomials fn(t):

g(t)fn(t)cm(t),
(2)

where we use

cm(t)=sinc(Ωtm)m
(3)

and

fn(t)=n1k=0aktk.
(4)

For m>n this is a square-integrable function with bandlimit set by Ω. See Supplemental Material for more information [33]. We note that the specific polynomial function used herein is not unique—more generalized, optimized functions that have the features we describe here will be pursued in future work.

For the second type of self-referenced functions, we introduce an idealized line-segment function we call a “triangle pulse” Tr(t), shown in solid green in Fig. 2(a). The triangle pulse is not bandlimited, so we consider approximations to the bandwidth based on the Fourier transform. Nevertheless, this function is better for measuring the minimum distance between two objects (range resolved objects) owing to the linearity of S over the full interference range.

Our third type of function is not an interference-class function, but is used simply to define a temporally resolved function. Similar to the Rayleigh criterion [34], when the peak of one pulse is separated by a distance greater than first minimum of a second pulse, the pulses are considered to be resolved. A classic example is the sinc2,

c2(t)=sinc(Ωt2)2.
(5)

For a bandlimited pulse, the minimally Rayleigh-resolved temporal shift tR (the time analog of resolvability in space) is given by tR=2π/Ω. How deeply we can superresolve the targets is quantified by the ratio

rs=td/tR,
(6)

where td is the temporal delay between the two returning pulses. Since rs is both a function of the delay td and the bandwidth tR=2π/Ω, for the work herein it is more precise to change the bandlimit to test the fundamental properties of the relative shift in the system rather than changing the relative pulse delay.

We define a signal S akin to balanced interferometric detection used to measure transverse deflections [35], namely,

S=AcmaxAcminAlmaxAlmin,
(7)

where Acmax, Acmin are the maximum and minimum amplitude of the function in the steep center region, respectively and Almax, Almin are the maximum and minimum of the flat temporal lobes.

Consider the ranging experiments shown in Fig. 1. In Fig. 1(a), waveforms are sent to a two-depth target. Upon reflection, the two reflected waveforms interfere and are sent to the receiver. The resultant waveform is measured and processed to estimate the distance between the two depths of the target.

To better test the limits of this technique, we use low-noise guided-wave experiments as shown in panels 1(b) and 1(c). Figure 1(b) shows a guided-wave equivalent to the free-space radar. However, there are unequal amplitudes from the scattered waves measured at the receiver based on the splitting ratios of the tee junction. To create equal-amplitude interference, we used the experiment shown in Fig. 2(c), which is the radio wave equivalent of the Michelson interferometer. In Fig. 1(c), the delay cable is connected to both channels 1 and 2. Channel 1 measures the interfered waveform and channel 2 measures the input (non-interfered) waveform. Both input channels on the scope are set at 1MΩ to achieve the desired reflections and measurement. The experiments in Figs. 2(b) and 2(c) utilize high precision frequency tunability to achieve ultrasmall relative pulse shifts rs.

The functions are generated numerically, consisting of 4000 points and a duration of 40tR units. These signals were uploaded to an arbitrary waveform generator (AWG). The bandlimit of the system is set by the repetition rate of the arbitrary waveform generator. For example, a repetition rate of 1 MHz results in a 40 MHz bandlimit for the bandlimited pulses.

To perform the ranging experiments as shown in Fig. 1(c), the arbitrary waveform generator sends the signal down the cable. The g(t) pulse is sent to channel 1 of the oscilloscope and is teed to another cable which then adds a one-way temporal measured delay of tc=4.4ns (the measured delay is 3.8 ns when not connected to channel 2 and the measured cable length is lc=72cm) implying that the pulse delay is td=2tc=8.8ns. The reflected signal from channel 2 interferes in channel 1 with the original displaced signal the g(t) function. When using the bandlimited function g(t), for example, channel 1 then measures g(t)+g(t+2tc) and similarly for the other functions.

Figure 2(c) shows the resultant interference waveform in channel 1 for the three different types of pulses used herein for ts=0.5tR. There are several important features of this graph that should be noted. First, the c2(t) interference pattern is not resolved, as expected, since ts=0.5tR. Second, the interference regions in the center of the g(t) (peaks have changed value) and triangle functions Tr(t) (interference plateaus) have changed dramatically. Third, the heights of the side lobes are still roughly constant for both interference-class functions.

We consider two important figures of merit for this range resolution system: (i) the minimum distance to amplitude-resolve two objects and (ii) the distance resolution between objects when the objects are amplitude resolved but still temporally subresolved (i.e., rs<1).

The triangle function is designed to amplitude resolve two objects with depths that are closely spaced along the same line of sight. Typically, in radar, the spectral bandwidth is given by the width of the spectrum at the 3 dB down point. However, owing to the irregular spectrum of the triangle pulse we use a conservative estimate based on where most of the power is found. Using the bandwidth from the Fourier transform as shown in Fig. 2(b), we can see that most of the power lies below 20% of the bandlimit for the bandlimited functions using the same pulse repetition rate. As noted above, the bandlimited frequency was 40× the repetition rate. From this observation, we define a conservative bandwidth of the triangle function to be 8× the pulse repetition rate.

The resolving power of the triangle function can be determined from Fig. 3. Here, we used the 72 cm delay cable disconnected from channel 2 and only connected to channel 1 yielding a 7.6 ns round-trip delay time. We change the bandwidth of the triangle pulse until we can no longer distinguish between the S curve with and without a cable. Here, we see that at approximately 1 MHz bandwidth, it is possible to distinguish the signal with and without the delay cable. This corresponds to rs0.008 or better than 100 times the inverse bandwidth.

FIG. 3.

The signal S is plotted against bandwidth of the triangle function with (interference) and without (no interference) a delay cable.

To further demonstrate the power of this technique, we removed the 72 cm cable and used only the extra path length of a bnc T junction, which has about 1.3 cm path length. We used a 2.5GSa/s triangle function with 10 samples per time unit tR (401 total points as opposed to 4000 described above) and measured the interference signal on an 800 MHz oscilloscope. The estimated bandwidth of the signal, based on the Fourier transform, was approximately 100 MHz, but with some small amount of frequency content up to 500 MHz. With and without the T junction, the value of S was 0.5573±0.0002 and 0.5511±0.0004, respectively, or a separation of 15 standard deviations. This implies sub-mm resolution. We then added a male-to-male bnc connector that added another couple centimeters and the value dropped to 0.5285±0.0001.

The second important figure of merit for range resolution is the ability to determine the distance between two objects for temporally subresolved pulses. Figure 4 shows the signal S vs rs (the ratio of the round-trip cable delay to the Rayleigh criterion) using the bandlimited function g(t). It can be seen in Fig. 4 that the theory and experiment agree well.

FIG. 4.

Signal S vs time shift. The signal S from Eq. (7) is computed and measured as a function of the relative shift of two g(t) functions.

The slope of S tells us the sensitivity of the system to changes in the relative temporal shift of the two functions. The S function for the bandlimited pulse is roughly quadratic in the region of zero shift and becomes linear after about rs=0.5 or half the Rayleigh resolved time. We want to know how well this system can determine the relative time between two shifts rs1 and rs2. For a fixed cable length,

Δrs=2tcΔΩ2π,
(8)

where Δrs=rs2rs1 and ΔΩ/2π is the smallest resolvable change in the bandwidth of the pulse.

In the inset of Fig. 4, it shows the change in the signal of a small incremental frequency shift. Using the 72 cm cable with tc=4.4ns delay and a 2 MHz repetition rate ( Ω/2π=80MHz bandlimit) resulted in approximately ts=0.69 shift. The bandwidth of the pulses was changed from 79.960 to 80 MHz in increments of 8.0 kHz. Each trace was averaged 512 times and each setting was measured 10 times with 15 to 30 sec between each measurement. Assuming a signal to noise ratio of 1, we can resolve 80 MHz bandwidth changes down to about 3.2 kHz. These data imply Δrs=2.8×105. The inferred depth resolution of our system is then Δx=vctRΔrs/220μm. Thus, for a target with two equal-amplitude, temporally subresolved reflected pulses, we can measure the relative distance between them down to 35 000 times below the Rayleigh limit and several orders of magnitude below the timing resolution of the oscilloscope.

While the primary emphasis of this paper concerns the simplest case of equal-amplitude reflections like those obtained using the setup in Fig. 1(c), in realistic applications, target reflection amplitude will typically be unequal, like those generated in Fig. 1(b). In the Supplemental Material [33], we show a two-parameter signal that involves both a temporal shift and disparate amplitudes and disparate amplitudes based.

We note several important comments. (i) These experiments were done using low frequency radio wavelengths. However, they are equally valid in all parts of the electromagnetic spectrum. (ii) As demonstrated preliminarily, the system can be generalized to account for disparate reflection amplitudes and multiple layers by creating more exotic functions and signal analysis. (iii) It is straightforward to convert time resolution to space resolution by transversely scanning the receiver in Fig. 1(a) and solving the inverse problem opening up high resolution imaging.

In summary, we have demonstrated both theoretically and experimentally that it is possible to obtain range resolution far better than the Rayleigh criterion or the inverse bandwidth. We employed the coherent aspects of radio wave transmission and detection to measure sensitive interference patterns. In the future work, we will explore the fundamental limits of this technique, as well as apply this method to more realistic ranging tasks in the field.

Supplemental Material

Click to Expand

References

  1. Donald R. Wehner, High Resolution Radar, (Artech House, Inc., Norwood, MA, (1987), p. 484.
  2. Jian Li, Guoqing Liu, Nanzhi Jiang, and Petre Stoica, Moving target feature extraction for airborne high-range resolution phased-array radar, IEEE Trans. Signal Process. 49, 277 (2001).
  3. K. B. Cooper, R. J. Dengler, G. Chattopadhyay, E. Schlecht, J. Gill, A. Skalare, I. Mehdi, and P. H. Siegel, A high-resolution imaging radar at 580 GHz, IEEE Microwave Wireless Compon. Lett. 18, 64 (2008).
  4. Fangzheng Zhang, Qingshui Guo, and Shilong Pan, Photonics-based real-time ultra-high-range-resolution radar with broadband signal generation and processing, Sci. Rep. 7, 13848 (2017).
  5. Merrill Ivan Skolnik, Introduction to Radar Systems, (McGraw Hill Book Co., New York, (1980), p. 590.
  6. Nadav Levanon and Eli Mozeson, Radar Signals (John Wiley & Sons, New York, 2004).
  7. Robert E. Sheriff, Seismic Stratigraphy—Applications to Hydrocarbon Exploration, American Association of Petroleum Geologists, ), 10.1306/M26490C1, 1977 http://dx.doi.org/10.1306/M26490C1.
  8. Adrian Neal, Ground-penetrating radar and its use in sedimentology: Principles, problems and progress, Earth-Sci. Rev. 66, 261 (2004).
  9. R. T. H. Collis, Lidar, Appl. Opt. 9, 1782 (1970).
  10. Ulla Wandinger, Introduction to Lidar (Springer, New York, 2005).
  11. Xiaoye Liu, Airborne lidar for dem generation: Some critical issues, Prog. Phys. Geogr. 32, 31 (2008).
  12. Behnam Behroozpour, Phillip A. M. Sandborn, Ming C. Wu, and Bernhard E. Boser, Lidar system architectures and circuits, IEEE Commun. Mag. 55, 135 (2017).
  13. Tomohiro Oguchi, Electromagnetic wave propagation and scattering in rain and other hydrometeors, Proc. IEEE 71, 1029 (1983).
  14. Ahmed I. Al-Shamma’a, Andrew Shaw, and Saher Saman, Propagation of electromagnetic waves at MHz frequencies through seawater, IEEE Trans. Antennas Propag. 52, 2843 (2004).
  15. Andrew Shaw, A. I. Al-Shamma’a, S. R. Wylie, and D. Toal, in 2006 European Microwave Conference, (IEEE, 2006), pp. 572–575.
  16. Shan Jiang and Stavros Georgakopoulos, Electromagnetic wave propagation into fresh water, J. Electromagn. Anal. Appl. 2011, 261 (2011).
  17. Alex M. H. Wong and George V. Eleftheriades, Superoscillatory radar imaging: Improving radar range resolution beyond fundamental bandwidth limitations, IEEE Microwave Wireless Compon. Lett. 22, 147 (2012).
  18. Rony Komissarov, Vitali Kozlov, Dmitry Filonov, and Pavel Ginzburg, Partially coherent radar unties range resolution from bandwidth limitations, Nat. Commun. 10, 1423 (2019).
  19. Mankei Tsang, Quantum Imaging beyond the Diffraction Limit by Optical Centroid Measurements, Phys. Rev. Lett. 102, 253601 (2009).
  20. H. Shin, Kam Wai Clifford Chan, H. J. Chang, and R. W. Boyd, Quantum Spatial Superresolution by Optical Centroid Measurements, Phys. Rev. Lett. 107, 083603 (2011).
  21. Mankei Tsang, Subdiffraction incoherent optical imaging via spatial-mode demultiplexing, New J. Phys. 19, 023054 (2017).
  22. Weng-Kian Tham, Hugo Ferretti, and Aephraim M. Steinberg, Beating Rayleigh’s Curse by Imaging Using Phase Information, Phys. Rev. Lett. 118, 070801 (2017).
  23. Yiyu Zhou, Jing Yang, Jeremy D. Hassett, Seyed Mohammad Hashemi Rafsanjani, Mohammad Mirhosseini, A. Nick Vamivakas, Andrew N. Jordan, Zhimin Shi, and Robert W. Boyd, Quantum-limited estimation of the axial separation of two incoherent point sources, Optica 6, 534 (2019).
  24. Mikael P. Backlund, Yoav Shechtman, and Ronald L. Walsworth, Fundamental Precision Bounds for Three-Dimensional Optical Localization Microscopy with Poisson Statistics, Phys. Rev. Lett. 121, 023904 (2018).
  25. Carmine Napoli, Samanta Piano, Richard Leach, Gerardo Adesso, and Tommaso Tufarelli, Towards Superresolution Surface Metrology: Quantum Estimation of Angular and Axial Separations, Phys. Rev. Lett. 122, 140505 (2019).
  26. J. Řeháček, M. Paúr, B. Stoklasa, D. Koutný, Z. Hradil, and L. L. Sánchez-Soto, Intensity-Based Axial Localization at the Quantum Limit, Phys. Rev. Lett. 123, 193601 (2019).
  27. V. Ansari, B. Brecht, J. Gil-Lopez, J. M. Donohue, J. Řeháček, Z. Hradil, L. L. Sánchez-Soto, and C. Silberhorn, Achieving the ultimate quantum timing resolution, PRX Quantum 2, 010301 (2021).
  28. Y. Aharonov, S. Popescu, and D. Rohrlich, How can an infra-red photon behave as a gamma ray, Tel-Aviv University Preprint TAUP, 1847–1890, 1990.
  29. Michael V. Berry, Optical currents, J. Opt. A 11, 094001 (2009).
  30. Sacha Kocsis, Boris Braverman, Sylvain Ravets, Martin J. Stevens, Richard P. Mirin, L. Krister Shalm, and Aephraim M. Steinberg, Observing the average trajectories of single photons in a two-slit interferometer, Science 332, 1170 (2011).
  31. Konstantin Y. Bliokh, Aleksandr Y. Bekshaev, Abraham G. Kofman, and Franco Nori, Photon trajectories, anomalous velocities and weak measurements: A classical interpretation, New J. Phys. 15, 073022 (2013).
  32. Barbara Šoda and Achim Kempf, Efficient method to create superoscillations with generic target behavior, Quantum Stud. Math. Found. 7, 347 (2020).
  33. See Supplemental Material at http://link.aps.org/supplemental/10.1103/PhysRevLett.131.053803 for additional information and calculations about bandlimited functions, the Rayleigh criterion, signal analysis and multi-parameter estimation.
  34. Max Born and Emil Wolf, Principles of Optics: Electromagnetic Theory of Propagation, Interference and Diffraction of Light (Elsevier, New York, 2013).
  35. P. B. Dixon, D. J. Starling, A. N. Jordan, and J. C. Howell, Ultrasensitive Beam Deflection Measurement via Interferometric Weak Value Amplification, Phys. Rev. Lett. 102, 173601 (2009).

No comments:

Post a Comment

New Fusion Record Achieved in Tungsten-Encased Reactor

New Fusion Record Achieved in Tungsten-Encased Reactor Isaac Schultz Summary The WEST (Tungsten Environm...