You are on page 1of 14

The Quantum Illumination Story

Jeffrey H. Shapiro1, ∗
1
Research Laboratory of Electronics
Massachusetts Institute of Technology
Cambridge, Massachusetts 02139, USA
Superposition and entanglement, the quintessential characteristics of quantum physics, have been
shown to provide communication, computation, and sensing capabilities that go beyond what clas-
sical physics will permit. It is natural, therefore, to explore their application to radar, despite the
fact that decoherence—caused by the loss and noise encountered in radar sensing—destroys these
fragile quantum properties. This paper tells the story of “quantum illumination”, an entanglement-
based approach to quantum radar, from its inception to its current understanding. Remarkably,
despite loss and noise that destroy its initial entanglement, quantum illumination does offer a target-
arXiv:1910.12277v4 [quant-ph] 23 Dec 2019

detection performance improvement over a classical radar of the same transmitted energy. A realistic
assessment of that improvement’s utility, however, shows that its value is severely limited. Nev-
ertheless, the fact that entanglement can be of value on an entanglement-breaking channel—the
meta-lesson of the quantum illumination story—should spur continued research on quantum radar.

INTRODUCTION two scenarios shown in Fig. 1. In both scenarios an op-


tical transmitter illuminates a region of space in which
Superposition and entanglement—Schrödinger’s cat a weakly-reflecting target is equally likely to be absent
being simultaneously alive and dead [1], and the “spooky or present within always-present background light. In
action at a distance” that Einstein found disturbing [2]— Fig. 1(a), the transmitter’s signal beam is a sequence
are quantum-mechanical phenomena that are moving of N high time-bandwidth product (M = T W  1),
from fundamental studies into scientific and engineering single-photon pulses. The receiver, for this single-photon
applications. Quantum computers, if realized at suffi- (SP) scenario, makes a minimum error-probability deci-
ciently large scale, will vastly outstrip the capability of sion between hypotheses H0 (target absent) and H1 (tar-
classical machines for a variety of problems in simula- get present) from observation of the light returned from
tion [3], optimization [4], and machine learning [5]. Those the interrogated region. In Fig. 1(b), the transmitter il-
large-scale quantum computers—running Shor’s quan- luminates the region of interest with a sequence of N
tum factoring algorithm [6]—will also break the public- high time-bandwidth product (M = T W  1), single-
key infrastructure on which Internet commerce currently photon signal pulses, each of which is entangled with a
relies. Quantum communication, in the form of quantum companion single-photon idler pulse; see Appendix A for
key distribution [7], however, may thwart that quantum the details. The receiver, for this quantum illumination
threat. In other work, quantum-enhanced sensing is mov- (QI) scenario, makes its minimum error-probability deci-
ing out of the laboratory and into real use, the most no- sion between H0 and H1 from observation of the retained
table example being the incorporation of squeezed-state idler light and the light returned from the interrogated
light into the laser interferometric gravitational-wave ob- region.
servatory (LIGO) [8]. A natural question to ask, there- The crucial assumptions in Lloyd’s analysis of the un-
fore, is whether quantum techniques can bring perfor- entangled and entangled scenarios are as follows.
mance gains to radar sensing. This paper will follow one
avenue of quantum radar research from its inception to • When the target is present, the roundtrip
now: quantum illumination. transmitter-to-target-to-receiver transmissivity for
the signal beam is 0 < κ  1.

• The background light’s average photon number per


LLOYD’S QUANTUM ILLUMINATION
temporal mode, NB , satisfies the low-brightness
condition NB  1.
Lloyd [9] coined the term “quantum illumination” for
his entanglement-based approach to improving an optical • For each transmitted signal pulse, at most one
radar’s capability to detect a weakly-reflecting target em- photon is returned to the receiver, regardless of
bedded in background noise that can be much stronger whether the target is absent or present, implying
than the target return. His work, which built on Sacchi’s that M NB  1.
earlier studies of quantum operation discrimination [10],
[11], compared the target-detection performance for the Under these assumptions, Lloyd identified two operating
regimes, the “good” and the “bad”, for his SP and QI
scenarios, and compared their error probabilities’ quan-
tum Chernoff bounds [12] in these regimes. In their good
∗ jhs@mit.edu regimes, SP and QI’s error probabilities have the same
2

? ?

(a) unentangled (b) entangled

FIG. 1. Scenarios presumed in Lloyd’s treatment of quantum illumination (QI). (a) Unentangled operation, in which the
transmitter illuminates the region of interest with a sequence of N single-photon signal pulses. Target absence or presence
is decided from observation of the returned light. (b) Entangled operation, in which the transmitter illuminates the region
of interest with a sequence of N single-photon signal pulses, each of which is entangled with a companion single-photon idler
pulse. Target absence or presence is decided from observation of the retained idler and the returned light.

bound, Despite the idealized nature of Lloyd’s QI analysis—


it presumes an on-demand source of high-T W entan-
Pr(e)SP ≤ e−N κ /2 and Pr(e)QI ≤ e−N κ /2, (1) gled signal-idler photon pairs, lossless idler storage, and
perfect realization of an optimum quantum receiver—
which, because N κ equals the average number of signal its enormous predicted performance enhancement in the
photons returned when the target is present, whereas 0 bad regime motivated a great deal of follow-on research.
is the average number of signal photons returned when Some of that research, unfortunately, took much of the
the object is absent, equals the signal shot-noise limit for air out of QI’s balloon, as we will describe below. Before
laser communication with on-off-keying (OOK) modula- doing so, however, a brief preface about classical versus
tion. Despite (1)’s applying to SP and QI’s good regimes, quantum radar is in order.
QI still enjoys a substantial good-regime performance ad- We were careful, earlier in this section, not to describe
vantage over SP, because SP’s good regime is limited to the comparison between SP and QI operation as one be-
κ  NB , whereas QI’s extends to the much larger pa- tween a classical radar (SP) and a quantum radar (QI).
rameter region in which κ  NB /M . In quantum optics, see, e.g., [15], [16], it is conventional
The comparison between SP and QI operation is rad- to reserve the appellation “quantum” for those systems
ically different in their bad regimes. Here Lloyd found whose performance analysis requires the quantum theory
that of photodetection. In particular, their performance can-
2 not be correctly quantified from the semiclassical theory
Pr(e)SP ≤ e−N κ /8NB
/2, for κ  NB , (2)
of photodetection, in which light is treated as a classical
and (possibly stochastic) electromagnetic wave and the dis-
creteness (quantization) of the electron charge gives rise
2
Pr(e)QI ≤ e−N κ M/8NB
/2, for κ  NB /M . (3) to shot noise. Furthermore, measurements of light beams
that are in coherent states [17], or classically-random
These bounds mimic the background-limited error prob- mixtures thereof, in any of the three basic photodetec-
ability of OOK laser communication, with NB being the tion paradigms—direct detection, homodyne detection,
background’s brightness for SP operation, and NB /M or heterodyne detection—do not require quantum pho-
being that brightness for QI operation. The bad-regime todetection theory to obtain correct measurement statis-
results reveal a double advantage for QI: first, its bad tics [15]. Hence such states are called classical states.
regime applies in a smaller κ range than that of SP; Systems that employ nonclassical states—states other
and second, when both systems are in their bad regimes, than coherent states or their random mixtures—require
QI enjoys a factor-of-M greater error-probability expo- the use of quantum photodetection theory. In this re-
nent (effective signal-to-noise ratio). At optical frequen- gard, we note that single-photon states and entangled
cies, high time-bandwidth product is easy to obtain, e.g., states are not classical states, and hence Lloyd’s perfor-
a 1 µs pulse at 300 THz center frequency (1 µm wave- mance comparison is between two quantum radars, one
length) with 1 THz (1/3-percent fractional) bandwidth of which employs entanglement (QI) while the other (SP)
yields M = 106 , in which case Lloyd’s bad-regime QI has does not.
a 60 dB higher effective signal-to-noise ratio than its SP A coherent state—other than the zero-photon (vac-
competitor. Moreover, and remarkably, this advantage uum) state—contains a random number of photons. So,
is afforded despite the background noise’s destroying the Shapiro and Lloyd [18] compared Lloyd’s QI to a classi-
initial entanglement, i.e., the retained and returned light cal radar by replacing Lloyd’s SP transmitter in Fig. 1(a)
are not entangled. with a coherent-state transmitter (an ideal laser) that
3

produces a sequence of N pulses, each of which has unity term this operating regime the “bad” regime for Tan et
average photon number. The quantum Chernoff bound al.’s CS and QI systems. Lloyd’s bad-regime QI system
for their coherent-state radar, offered an M -fold error-probability exponent improve-
√ √ ment over its SP counterpart, where a time-bandwidth
1+NB − NB )2 product M ≥ 106 is easily attainable. In contrast,
Pr(e)CS ≤ e−N κ( /2, (4)
Tan et al.’s QI system only affords a factor-of-four (6 dB)
which applies for all 0 ≤ κ ≤ 1 and for all NB ≥ 0, improvement in error-probability exponent over its CS
reduces to counterpart of the same transmitted energy, regardless
of an M ≥ 106 time-bandwidth product. Moreover, be-
Pr(e)CS ≤ e−N κ /2, (5) cause Tan et al. do not restrict their M -mode signal and
idler state spaces to the span of their M -mode vacuum
for the low-brightness (NB  1) background that Lloyd’s and single-photon states, their work does not fall prey to
QI analysis assumed. Shapiro and Lloyd thus showed issues identified by Shapiro and Lloyd.
that a coherent-state radar matched the performance of Figure 3 illustrates the behaviors of Tan et al.’s quan-
Lloyd’s QI radar in the latter’s good regime, and was sub- tum Chernoff bounds for κ = 0.01, NS = 0.01, and
stantially better when QI operated in its bad regime. For- NB = 20. Also included in this figure is the Bhat-
tunately for quantum illumination, Shapiro and Lloyd’s tacharyya lower bound on Pr(e)CS from Ref. [19]. It is
work was not the end of the story; Tan et al. [19] had important to remember that quantum Chernoff bounds
already analyzed a Gaussian-state QI system that out- are known to be exponentially-tight upper bounds, while
performed all classical radars of the same transmitted it is known that the Bhattacharyya lower bound is always
energy. loose. Furthermore, Tan et al. showed that the CS radar
affords the lowest error probability of any classical-state
radar of the same average transmitted energy. Thus, be-
TAN ET AL.’S QUANTUM ILLUMINATION cause Fig. 3 shows Pr(e)UB LB
QI becoming lower than Pr(e)CS
for sufficiently high M values, it provides definitive proof
Tan et al. compared the classical and quantum radar that a QI radar can, in principle, outperform all classical
scenarios, shown in Fig. 2, for detecting a weakly- radars of the same transmitted energy for the target-
reflecting (0 < κ  1) target that is equally likely detection scenario addressed by Tan et al. We will say
to be absent or present within always-present, high- more later about why we refer to Fig. 3’s QI advantage
brightness (NB  1) background light. The classical over CS operation as being “in principle”. For now, it
radar, in Fig. 2(a), illuminates the region of interest with suffices to point out that when Ref. [19] appeared there
a coherent-state (laser) pulse of average photon num- was no known receiver that provided any QI performance
ber M NS , where NS  1 and M = T W  1. Tar- advantage over CS operation. Why that was so and how
get absence or presence is then decided from observa- it was overcome require some understanding of quantum
tion of the returned light. The QI radar, in Fig. 2(b), photodetection theory and the quantized electromagnetic
carves duration-T , entangled signal and idler pulses from field’s Gaussian states. So those topics are next on our
the outputs of a continuous-wave spontaneous parametric agenda.
downconversion (SPDC) source, whose phase-matching The mth temporal modes of Tan et al.’s QI signal and
bandwidth is W , and whose signal and idler have av- idler from (B1) and (B2) have associated photon anni-
erage photon number per temporal mode NS  1; see hilation operators, âSm and âIm , whose adjoints are the
Appendix B for the details. Target absence or presence photon creation operators, â†Sm and â†Im . These names
is then decided from observation of the retained idler and originate from the operators’ actions on their mode’s
the returned light. number states, which, for K = S, I, obey
Pirandola and Lloyd [20] developed a formula for com-
puting the quantum Chernoff bound for the task of dis- √
n |n − 1iKm , for n = 1, 2, . . . ,
tinguishing between two arbitrary multi-mode Gaussian âKm |niKm = (8)
0, for n = 0,
states. Exploiting this tool, Tan et al. showed that the
quantum Chernoff bounds for their coherent-state (CS) and
and QI radars behave rather differently than those for

Lloyd’s SP and QI radars, i.e., Tan et al. found â†Km |niKm = n + 1 |n + 1iKm , for n = 0, 1, 2, . . . , (9)
Pr(e)CS ≤ e−M κNS /4NB /2, (6) where the ket vectors {|niKm : n = 0, 1, 2, . . . , } rep-
resent states containing exactly n photons. Ideal pho-
and ton counting on the mth signal and idler modes mea-
Pr(e)QI ≤ e−M κNS /NB /2, (7) sures the photon-number operators N̂Sm ≡ â†Sm âSm and
N̂Im ≡ â†Im âIm , respectively. It then follows that ideal
where, in both cases, 0 < κ  1, NS  1, and NB  1 photon counting on the coherent state given in (B4)
are assumed. By analogy with Lloyd’s work, we might yields a Poisson-distributed output with mean NS , as
4

LASER
? SPDC
?

(a) coherent state (b) entangled state

FIG. 2. Scenarios presumed in Tan et al.’s treatment of quantum illumination (QI). (a) Coherent-state operation, in which a
laser transmitter illuminates the region of interest with a coherent-state signal pulse of average photon number M NS . Target
absence or presence is decided from observation of the returned light. (b) Entangled operation, in which the transmitter
illuminates the region of interest with a duration-T signal pulse carved from the low-brightness (average photon number per
temporal mode NS  1) output of a continuous-wave spontaneous parametric downconversion (SPDC) source whose phase-
matching bandwidth W satisfies M = T W  1. Target absence or presence is decided from observation of the retained idler
and the returned light.

the Heisenberg uncertainty principle forces this measure-


Pr(e)UB
CS ment to incur extra noise on each quadrature that is not
present in homodyne measurement of a single quadra-
ture [15].
Pr(e)UB
QI
Gaussian states of the mth signal and idler modes, âSm
and âIm , are the quantum analogs of classical, complex-
Pr(e)LB
CS
valued Gaussian random variables, aSm and aIm . Thus,
Gaussian states of these two modes are completely char-
acterized by knowledge of their first and second mo-
ments [21], hâKm i, h∆â†Km ∆âJm i, and h∆âKm ∆âJm i,
where K = S, I, J = S, I, h·i denotes ensemble aver-
age, and ∆âKm ≡ âKm − hâKm i. The coherent √state
from (B4) is a Gaussian state with hâSm i = NS ,
and h∆â†Sm ∆âSm i = h∆â2Sm i = 0. The entangled
signal-idler state from (B3) is also Gaussian. It is a
two-mode squeezed vacuum (TMSV) state [21], whose
FIG. 3. Error-probability bounds for Tan et al.’s CS and QI measurement statistics are completely characterized by
radars [19]: Pr(e)UB
CS is the quantum Chernoff bound for the hâSm i = hâIm i = 0, hâ†Sm âSm i = hâ†Im âIm i = NS ,
CS radar; Pr(e)UB
QI is the quantum Chernoff bound for the QI hâ2Sm i = hâ2Im i = 0, hâ†Sm âIm i = 0, and hâSm âIm i =
radar; and Pr(e)LB
CS is the Bhattacharyya lower bound for the
p
CS radar. NS (NS + 1).
The preceding brief introduction to Gaussian states
provides enough information to understand the origin of
expected from the coherent state’s photodetection statis- Tan et al.’s QI advantage and why conventional opti-
tics being obtainable from semiclassical (shot-noise) the- cal receivers—direct detection, homodyne detection, and
ory [15]. heterodyne detection—do not realize any of that advan-
The signal and idler’s mth temporal modes have tage. The cross correlations of all zero-mean classical
operator-valued quadrature components, signal-idler states must obey [22]
q
âKm + â†Km âKm − â†Km

|hâSm âIm i| ≤ hâ†Sm âSm ihâ†Im âIm i, (11)
Re(âKm ) = and Im(âKm ) = ,
2 2j and
(10) q
for K = S, I, and ideal (quantum-limited) optical homo- |hâSm âIm i| ≤ hâ†Sm âSm ihâ†Im âIm i. (12)
dyne detection with the appropriate local oscillator fields For arbitrary, zero-mean quantum states (11) applies, but
measures these operators [15]. The positive-operator- (12) becomes the less restrictive condition [23]
valued measurements (POVMs) associated with âSm and r
âIm [21] can be realized by ideal (quantum-limited) op- |hâSm âIm i| ≤ max (hâ†Km âKm i) min (hâ†Km âKm i + 1).
tical heterodyne detection [15]. Although heterodyne K=S,I K=S,I
detection provides information about both quadratures, (13)
5

The TMSV is a zero-mean Gaussian p state that vio- hâRm âIm iH1 for those beams’ mth modes. The returned
lates (12), because |hâSm
â Im
i| = N S (N S + 1) > NS = and retained light’s target-present phase-insensitive cross
correlation, hâ†Rm âIm iH1 , can be measured in second-
q
† †
hâSm âSm ihâIm âIm i. Thus it is a nonclassical state, as
order interference between the two beams. But Tan et
we already knew from (B3), and in fact maximally en- al.’s QI system has hâ†Rm âIm iHj = 0 for j = 0, 1, i.e., re-
tangled, because it saturates √ the √bound in (13). Fur- gardless of whether the target is absent or present, and its
thermore, with âRm = κ âSm + 1 − κ âBm being the hâRm âIm iHj cannot be measured in second-order inter-
photon annihilation operator for the returned light’s mth ference; see Appendix C for the details. Guha and Erk-
mode when the target is present, where âBm is the photon men [26] recognized this problem, and offered a partial
annihilation operator of the relevant background-light solution with their optical parametric amplifier (OPA)
mode [19], Tan et al.’s QI system has the conditional receiver. That receiver, its experimental realization, and
cross-correlation other aspects of obtaining QI’s performance advantage
p are treated in the next section.
hâRm âIm iH1 = κNS (NS + 1), (14)

given the target-present hypothesis H1 . When NS  1,


this conditional cross-correlation, which is the signature QI RECEIVERS AND EXPERIMENTS
of target presence in Tan et al.’s QI √system because
hâRm âIm iH0 = 0, greatly exceeds the κ NS classical- To understand Guha and Erkmen’s OPA receiver re-
state limit on this cross correlation that applies when quires some results from the quantum theory of nonlin-
hâ†Sm âSm i = hâ†Im âIm i = NS . When NS  1, however, ear optics. Because that theory also underlies the SPDC
QI’s target-presence signature is only slightly better than behavior we have already been employing, a brief intro-
the classical limit [24]. So, it should be no surprise that duction encompassing both SPDC and OPA operation is
the preferred operating regime for Tan et al.’s QI sys- germane.
tem has low signal brightness, NS  1. That this pre- In crystals that have a second-order nonlinear suscepti-
ferred operating regime has high background brightness, bility, such as lithium niobate (LiNbO3 ) or potassium ti-
NB  1, follows from Nair’s no-go theorem [25], which tanyl phosphate (KTiOPO4 ), a strong pump beam at fre-
shows that QI offers only an inconsequential performance quency ωP can interact with weak signal and idler beams
improvement over CS operation’s Pr(e)CS ≤ e−M κNS /2 at lower frequencies ωS and ωI satisfying ωS + ωI = ωP .
when background light can be neglected. Continuous-wave SPDC has no inputs at the signal and
It might seem it would be easy to build a receiver idler frequencies. Nevertheless, it produces outputs at
capable of reaping the benefit associated with QI’s en- those frequencies. These outputs can be regarded as
hanced cross-correlation signature in the 0 < κ  1, arising from a photon-fission process, in which a sin-
NS  1, NB  1 regime. After all, âRm âIm can be gle pump photon splits into a signal-idler photon pair.
expanded into four terms that are products of the re- Energy conservation at the single-photon level requires
turn and idler modes’ quadrature components, and these that h̄ωS + h̄ωI = h̄ωP . Momentum conservation at
quadrature components can be measured individually the single-photon level requires that h̄kS + h̄kI = h̄kP ,
by means of homodyne detection. Unfortunately, this where kJ , for J = S, I, P , are the propagation vectors
approach does not work, because we need to measure for the signal, idler, and pump photons. Even with
both quadratures of the return and idler modes, and these conservation conditions obeyed—which can only be
the Heisenberg uncertainty principle precludes that being achieved within the crystal’s phase-matching bandwidth,
done, e.g., by heterodyne detection, without incurring |ωK − ωKo | ≤ πW for K = S, I, about the signal and
additional noise. The net effect is that neither homo- idler’s center frequencies, ωSo and ωIo —continuous-wave
dyne detection nor heterodyne detection can be used to SPDC produces only pW of signal and idler per mW of
provide any QI target-detection performance advantage pump power, see, e.g., [27].
over its CS counterpart. Direct detection (time-resolved For the QI application, we want the signal and idler
photon counting) is also unable to provide a QI advan- to be single-spatial-mode fields, so the former can be
tage. Although SPDC produces signal and idler pho- formed into a tight transmitter beam, which is why
tons in time-coincident photon pairs, photon-coincidence our description of Tan et al.’s QI system—and, simi-
counting—as routinely used, e.g., in SPDC ghost imag- larly, of Lloyd’s QI system—only considered temporal
ing [16]—does not provide a usable QI signature for tar- modes. The Gaussian-state treatment of single-spatial-
get presence. This failure is because the presence of high- mode SPDC [28] shows that the signal’s frequency com-
brightness background light in the return implies that ponent at ωSo + ω is only correlated with the idler’s
every detection of an idler photon will have a coincident frequency component at ωIo − ω, which explains Ap-
detection from the returned light, regardless of target ab- pendix B’s choice of temporal modes, and justifies our
sence or presence. only paying attention to correlations between the mth
The signature of target presence in Tan et al.’s QI sys- modes of the signal and idler in our treatment of Tan et
tem is the conditional phase-sensitive cross correlation al.’s QI system.
between its returned light and its retained idler, namely Guha and Erkmen’s receiver for Tan et al.’s QI system
6

uses a continuous-wave OPA operating at extremely low Zhang et al. [29] reported the first experimental demon-
gain, which means that it can be realized with a crystal stration of Tan et al.’s QI system using an OPA receiver
identical to the one used for that system’s SPDC source. in a table-top setup. Owing to a variety of experimental
In the receiver, however, the returned light and the re- nonidealities, that experiment yielded only a 20% signal-
tained idler light are applied as the signal and idler inputs to-noise ratio (SNR) improvement—equivalent to a 20%
to the OPA crystal,
√ and the √ idler output’s mth mode is error-probability exponent advantage—over a CS system

given by âout
Im = G âIm + G − 1 âRm , where the gain, of the same transmitted energy.
G > 1, is chosen to optimize target-detection perfor- Prior to Zhang et al.’s work, Lopaeva et al. [30] re-
mance. Ideal photon counting is done on all idler-output ported a QI-like experiment in which an SPDC source
modes, which measures their total photon-number oper- and a photon-counting correlations were used to ob-
ator, tain an SNR advantage over a correlated-thermal-state
(correlated-noise radar) probe. That experiment, how-
(M −1)/2
ever, did not exploit entanglement, and its QI-like system
âout†
X
out
N̂T ≡ Im âIm . (15) only outperformed a correlated-thermal-state system of
m=−(M −1)/2 the same transmitted energy, not a CS system of that
energy. A more QI recent experiment, by England et
Although written as a sum of modal photon-number op- al. [31], used an SPDC source and photon-coincidence
erators, the N̂T measurement is easily realized by pho- counting for QI, but it operated in the low-brightness,
ton counting on the OPA’s duration-T idler output in re- NB  1, background regime, wherein QI offers no ad-
sponse to the duration-T returned light and retained idler vantage over a coherent-state radar.
inputs. The conditional means—given target absence or The failure of OPA reception to achieve QI’s full 6 dB
presence—for the N̂T measurement are as follows: advantage in error-probability exponent is a consequence
(M −1)/2
of Tan et al.’s QI scenario being one of mixed-state hy-
pothesis testing and OPA reception’s using mode-pair

Ghâ†Im âIm i + (G − 1)hâRm â†Rm iHj
X
hN̂T iHj =
measurements. Such an arrangement falls into the class
m=−(M −1)/2
of local operations plus classical communication (LOCC)
processing, which is known [32], [33] to be suboptimal for
p 
+2 G(G − 1) Re[hâRm âIm iHj ] . (16)
mixed-state hypothesis testing. There is an in-principle
approach to realizing QI’s full performance advantage
Here, the phase conjugation of the returned light’s con- over CS operation—a Schur transform on a quantum
tribution to the OPA’s idler output makes QI’s phase- computer [34]—but Zhuang et al. [35] have proposed a
sensitive cross-correlation signature of target presence receiver structure for this purpose that does not require
observable in a photon-counting measurement, cf. Ap- a full-blown quantum computer. They did so by exploit-
pendix C, where without phase conjugation it is only the ing SPDC’s inverse operation, sum-frequency generation,
phase-insensitive cross correlation that can be measured to go beyond the bounds set by LOCC processing.
in a photon-counting interferometer.
A signal-idler photon pair produced by continuous-
It is the high brightness of the returned light’s back-
wave SPDC with phase-matching bandwidth W is in a
ground component, in comparison the low brightness of
quantum state satisfying
the retained idler, that leads to the OPA receiver’s opti-
mum gain satisfying G − 1  1. At this optimum gain πW

Z
value, Guha and Erkmen found their receiver’s quantum |ψiSI ∝ |ωSo + ωiS |ωIo − ωiI , (18)
Chernoff bound to be −πW 2π

Pr(e)OPA ≤ e−M κNS /2NB /2, (17) where |ωSo +ωiS (|ωIo −ωiI ) denotes a single-photon sig-
nal (idler) of frequency ωSo +ω (ωIo −ω) and, as has been
in QI’s usual 0 < κ  1, NS  1, NB  1, operating implicitly assumed earlier, the signal (idler) brightness
regime, implying a 3 dB improvement in error-probability is taken to be constant over the phase-matching band-
exponent over Tan et al.’s CS system of the same trans- width. If this photon pair illuminates a crystal identical
mitted energy. to the one used for its SPDC generation, then, with ex-
Implicit in (17) are two important assumptions made tremely low probability, sum-frequency generation (SFG)
by Guha and Erkmen, both of which arise from the in- can occur, viz., a photon-fusion process in which the
terferometric nature of applying the returned light and signal-idler photon pair is converted to a single photon
the retained idler to the signal and idler inputs of a low- at the pump frequency, ωP = ωSo + ωIo . Zhuang et
gain OPA. The first assumption is that the idler is stored al. realized that SFG’s being a coherent process involv-
losslessly. The second is that the idler storage is matched ing all of QI’s mode pairs offered a path to QI recep-
in time delay and phase to those of the light returned tion that was not bound by the limitations of LOCC
from the target (when it is present). We will revisit operation. Nevertheless, their work made the rather sig-
these assumptions later in discussing QI’s utility for re- nificant assumption that SFG could be done with 100%
alistic radar scenarios. For now, it suffices to note that efficiency at the photon-pair level, something that is far
7

beyond the current capability of nonlinear optics. Fur- in which that occurred, 0 < κ  1, NS  1, and
thermore, the presence of high-brightness background NB  1, is not realistic for optical wavelengths. Weakly-
light drove them to using multiple cycles of SFG and reflecting (low radar cross-section) targets are of great
photon-counting measurements, but ideal realization of interest, and SPDC sources naturally produce NS  1
these steps resulted in a receiver whose quantum Chernoff emissions, but NB  1 does not prevail at optical
bound for QI matched that found by Tan et al. Zhuang et wavelengths. For example, a typical value for the sky’s
al. augmented their SFG receiver with an appropriate daytime spectral radiance at the eyesafe 1.55 µm wave-
feed-forward (FF) circuit—inspired by the Dolinar re- length is Nλ ∼ 10 W/m2 SR µm [39], from which we find
ceiver [36] for optimum quantum reception of coherent- that [40]
state signals—and showed that the resulting FF-SFG re-
ceiver provided minimum error-probability performance NB = π106 λ3 Nλ /h̄ω 2 ∼ 10−6 , (19)
in QI target detection. Despite this receiver’s being well
and nighttime Nλ (and hence NB ) values are several or-
beyond the reach of available technology, it is neverthe-
ders of magnitude lower. So, the advantage expected
less important because it allowed determination of QI’s
from ideal QI operation at optical frequencies would only
receiver operating characteristic (ROC), i.e., its detection
accrue were there bright-light jamming. Consequently,
probability, PD , versus false-alarm probability, PF , when
a significant amount of interest in QI from the radar
optimum quantum reception is employed [37]. QI’s ROC
community only arose after Barzanjeh et al. [41] pro-
is crucial because it is a far better target-detection per-
posed an approach for doing QI in the microwave region,
formance metric than error probability, as radar targets
where the naturally-occurring background does satisfy
should not be presumed equally likely to be absent or
NB  1 [42], and most target-detection radars operate.
present. Not surprisingly, QI’s ROC improvement over
Figure 4 shows Barzanjeh et al.’s transmitter and
CS operation of the same transmitted energy turned out
receiver concepts. Their transmitter uses an electro-
to be equivalent to a 6 dB increase in effective SNR [37].
optomechanical (EOM) converter to create a microwave
QI’s aforementioned performance advantages—in er-
signal that is entangled with an optical idler, and it trans-
ror probability or ROC—assume that the target return,
mits the microwave signal to irradiate the region in which
when present, has known amplitude and phase, a situa-
the target may be present. Then, it uses another EOM
tion that seldom occurs in light detection and ranging (li-
converter to upconvert the returned microwave signal
dar) applications. At lidar wavelengths, most target sur-
to the optical region for a phase-conjugate (PC) joint
faces are sufficiently rough that their returns are speck-
measurement with the retained idler. Guha and Erk-
led, i.e., they have Rayleigh-distributed amplitudes and
men [26] had previously shown that a PC receiver’s per-
uniformly-distributed phases. QI’s OPA receiver—which
formance advantage over a conventional radar is equiva-
affords a 3 dB-better-than-classical error-probability ex-
lent to that of an OPA receiver, i.e., a 3 dB advantage in
ponent for a return with known amplitude and phase—
error-probability exponent under ideal conditions. Even
fails to offer any performance gain for Rayleigh-fading
though this advantage could easily fall prey to system
targets. The SFG receiver from Ref. [35]—whose error-
nonidealities, the fact that QI was now predicted to offer
probability exponent for a nonfading target achieves QI’s
an advantage at microwave frequencies ignited a great
full 6 dB advantage over optimum classical operation—
deal of attention from the radar community, prompted
outperforms the classical system for Rayleigh-fading tar-
in part by some inaccurate reporting [43]. It is now time,
gets [38]. In this case, however, QI’s advantage is subex-
therefore, to turn our attention to those nonidealities,
ponential under ideal operating conditions, so that its
some of which were discussed in Refs. [41], [44], and con-
benefit is far more vulnerable to nonidealities such as
front the realities of seeking a QI advantage for target
were encountered in Zhang et al.’s OPA receiver for the
detection. We will begin that assessment with the issue
non-fading scenario.
of idler-storage loss.
The OPA, PC, SFG, and FF-SFG receivers for QI all
require that the idler be stored for a roundtrip, radar-
MICROWAVE QUANTUM ILLUMINATION: to-target-to-radar, propagation time. If the idler-storage
CONCEPT AND REALITIES transmissivity is κI (0 < κI ≤ 1), then the quantum
Chernoff bounds of interest for non-fading targets be-
Lloyd’s QI presumed operation at optical wavelengths, come
because high-sensitivity photodetection systems have
long been limited by noise of quantum-mechanical ori- Pr(e)QI ≤ e−M κκI NS /NB /2, (20)
gin. Hence the optical region was the natural setting
for FF-SFG or SFG receivers, and
in which to seek a quantum advantage. Tan et al. con-
tinued to focus on optical wavelengths, because that is Pr(e)QI ≤ e−M κκI NS /2NB /2, (21)
where SPDC sources provide the entanglement needed
for Gaussian-state QI. Unfortunately, although Tan et for OPA or PC receivers, in QI’s preferred 0 < κ  1,
al. found QI’s performance to exceed that of all classi- NS  1, NB  1 operating regime. Comparing these re-
cal radars of the same transmitted energy, the regime sults with (6)’s quantum Chernoff bound for a CS radar
8

(a) QI radar configuration (b) EOM converter

FIG. 4. Schematic for Barzanjeh et al.’s microwave quantum illumination [41]. (a) QI radar configuration. (b) Electro-
optomechanical (EOM) converter configuration.

shows that 6 dB of idler loss will eliminate QI’s perfor- surements, leads to a 10 log10 (KB ) dB performance loss
mance advantage for FF-SFG or SFG reception, and 3 dB for each bin. So, assuming ideal equipment, simulta-
of idler loss suffices for that purpose for OPA or PC re- neous interrogation of two resolution bins sacrifices the
ception. Barzanjeh et al. [41] was the first QI target- entire performance advantage of an OPA receiver, and
detection paper to explicitly comment on the adverse im- simultaneous interrogation of four resolution bins gives
pact of idler-storage loss, noting that the idler-storage up the entire performance advantage of an FF-SFG re-
loss incurred in the best presently-available technique ceiver. In this regard it is important to note that op-
for optical idler storage—an optical-fiber delay line— tical amplification cannot be used, prior to idler split-
precludes their system’s offering any performance advan- ting, to mitigate the preceding problem. This failure
tage for targets more than 11.25 km away. is because the amplifier’s unavoidable amplified spon-
Even were the preceding gloomy assessment of long- taneous emission noise will be much stronger than the
range, microwave-QI target detection obviated by a amplified idler, making the latter useless for QI. There
breakthrough in idler-storage technology, microwave QI is another resolution-related problem with QI: the target
would still suffer from its difficulty in obtaining a suf- should lie entirely within a single polarization-azimuth-
ficiently high time-bandwidth product with pulse dura- elevation-range-Doppler resolution bin throughout the
tions that make sense for microwave radars. High time- signal pulse’s full duration. If such is not the case, there
bandwidth product is critical for QI because—unlike will be a mismatch between the temporal behavior of the
a conventional radar, whose performance improves at returned radiation’s target-present component and the
constant pulse duration and bandwidth with increas- temporal behavior of the stored idler. This mismatch,
ing transmitter power in the absence of clutter—QI which will degrade the performance of the QI receivers
systems’ performance advantage degrades at constant we have considered, is quantified by a normalized overlap
pulse duration and bandwidth with increasing transmit- integral, 0 < κm ≤ 1, that reduces QI’s error-probability
ter power [44]. So, even though Barzanjeh et al.’s EOM exponent by a factor of κm . The mismatch problem is
converters—which are intrinsically narrowband—could especially significant for optical QI, where a 1 THz band-
be replaced with broadband microwave-entanglement width implies that <1 mm target range extent is needed
generation in a traveling-wave parametric amplifier [45], to ensure κm ≈ 1.
the available time-bandwidth products in the microwave Some additional points worth noting are as follows:
region are still dwarfed by what is easily achieved at op- (1) Las Heras et al. [47] have claimed that QI target de-
tical wavelengths. For example, the 1/3-percent frac- tection can unveil an electromagnetically-cloaked target
tional bandwidth at 1 µm wavelength that gives a 1 µs from the phase shift it creates, hence potentially reviving
pulse duration a 106 time-bandwidth product only gives the radar community’s interest in QI. But Las Heras et
that pulse duration a 102 time-bandwidth product at al. assume an OPA receiver and neglect idler-storage loss,
1 cm wavelength. Pushing to mm-wave operation with so their concept’s utility is limited by the same consid-
higher fractional bandwidth and longer pulse duration erations cited above for Barzanjeh et al.’s system. (2)
will afford considerably higher time-bandwidth product, De Palma and Borregaard [48] have shown that the state
but that will run afoul of another consideration: QI’s produced by SPDC is the optimum transmitter state for
single-bin-per-pulse interrogation limitation [44]. Tan et al.’s QI scenario if the objective is to obtain the
Unlike a conventional radar, QI can only interrogate a fastest decay of the miss probability, PM ≡ 1 − PD , for a
single polarization-azimuth-elevation-range-Doppler res- given false-alarm probability, PF , as the transmitted en-
olution bin at a time [46], if it is derive its full perfor- ergy is increased. In effect, their result implies that the
mance advantage over a conventional radar. Interroga- SPDC state optimizes the ROC for Tan et al.’s QI sce-
tion of KB bins simultaneously, by splitting the stored nario. So, because QI target detection for equally-likely
idler into KB equal-strength pieces to make said mea- target absence or presence has an error probability sat-
9

isfying Pr(e) = [PF + (1 − PD )]/2, where (PF , PD ) lies distributions with covariance matrices
on the system’s ROC, De Palma and Borregaard’s result
GA (NB + NF )I2 02
 
provides strong evidence that a 6 dB advantage for QI’s QCN
ΛH 0 = , (23)
error-probability exponent is the best that can be done 2 02 (NS + NF )I2
relative to a coherent-state transmitter of the same en-
ergy in this equally-likely situation. (3) Initial microwave and
QI experiments have been performed by Luong et al. [49], "
(NR + NF )I2 Cq (θ)
#
GA
[50], and by Barzanjeh et al. [51]. Both used Joseph- ΛQCN
H1 ,κ,θ = . (24)
son junction parametric amplifiers to produce entangled 2 CTq (θ) (NS + NF )I2
signal and idler in the GHz region, and both performed
pre-amplified heterodyne detection at the signal and idler In these equations: GA is the pre-amplifiers’ gain; I2 is
frequencies. Furthermore, both found substantial perfor- the 2 × 2 identity matrix; 02 is thep
2 × 2 matrix of zeros;
mance gains for their QI systems over their chosen clas- NR ≡ κNS + NB ,; and Cq (θ) = κNS (NS + 1) Rq (θ)
sical comparison cases. In Luong et al.’s experiments the with
comparison case was a classically-correlated-noise radar, 
cos(θ) sin(θ)

whereas in Barzanjeh et al.’s work comparisons were Rq (θ) ≡ . (25)
made with both a classically-correlated-noise radar and sin(θ) − cos(θ)
a coherent-state radar with incoherent post-heterodyne
The CCN radar we will consider uses a high-brightness
processing. As the next section will show, their reception
classical noise source whose output is divided—with an
techniques preclude their QI experiments from outper-
appropriate splitter—into a low-brightness signal and a
forming an optimized classically-correlated-noise (CCN)
high-brightness idler. Like the QCN radar, the CCN
radar. For that reason, we will refer to their QI setups
radar will employ pre-amplified heterodyne detection of
as quantum-correlated-noise (QCN) radars, as opposed
its retained idler and its return from the region interro-
to a true QI radar, viz., the Tan et al. scenario with an
gated by its signal. Conditioned on the true hypoth-
OPA receiver, which outperforms all classical radars of
esis and the κ, θ values, the CCN radar’s M mode-
the same transmitted energy [52].
pair outputs are also a set of independent, identically-
distributed, complex-valued, 2D random column vectors
whose quadrature components have zero-mean Gaussian
distributions, but now with covariance matrices
CORRELATED-NOISE RADARS: QUANTUM
GA (NB + NF )I2 02
 
VERSUS CLASSICAL
ΛCCN
H0 = , (26)
2 02 (NI + NF )I2
Here, for greater generality in our consideraton of
and
correlated-noise radars, we will use " #
GA (NR + NF )I2 Cc (θ)
√ √ ΛCCN = , (27)
âRm = κ ejθ âSm +1 − κ âBm , for |m| ≤ (M − 1)/2, H1 ,κ,θ
2 CTc (θ) (NI + NF )I2
(22)

to model the returned light’s M modal annihilation op- where Cc (θ) = κNS NI Rc (θ) with
erators when the target is present, where 0 < κ  1
and 0 ≤ θ ≤ 2π may be deterministic and known, as cos(θ) − sin(θ)
 
in Tan et al. [19] (κ  1 known and θ = 0), or ran- Rc (θ) ≡ . (28)
sin(θ) cos(θ)
dom with a known √ joint probability distribution, as in
Zhuang et al. [37] ( κ and √ θ statistically independent, Note that SPDC forces the signal and idler bright-
with a Rayleigh-distributed κ having hκi  1, and a nesses at the QCN radar’s transmitter to be identi-
uniformly-distributed θ). We will also assume that these cal, whereas, by starting with a high-brightness clas-
noise radars use pre-amplified heterodyne detection, with sical noise source and using a highly-asymmetric split-
noise figure NF ≥ 1, where 1 is the ideal (quantum- ting ratio, the CCN radar’s transmitter can have a low-
limited) value [53]. brightness (NS  1) signal that matches that of the
The QCN radar we shall consider uses an SPDC source QCN radar while retaining a high-brightness (NI  1)
and pre-amplified heterodyne detection at the signal idler. As we will soon see, operating the CCN radar with
and idler frequencies. Conditioned on target absence NI  1  NS is a crucial point that was missed in the
(hypothesis H0 ) or presence (hypothesis H1 ) and the microwave QI experiments to date [49]–[51].
κ, θ values, the M mode-pair outputs from the QCN The relative target-detection performance of the
radar’s heterodyne detectors are a set of independent, QCN and CCN radars is most conveniently under-
0
identically-distributed, complex-valued, 2D random col- stood by transforming
√ their
√ mode-pair data to {am ≡
∗ T
umn vectors, {am ≡ [ aRm aIm ]T : |m| ≤ (M − 1)/2}, [ aRm aIm / NS + 1 ] / GA } for the QCN radar, and
√ √
whose quadrature components have zero-mean Gaussian {a0m ≡ [ aRm aIm / NI ]T / GA } for the CCN radar.
10

Given the true hypothesis and the κ, θ values, the third (QI-OPA) uses the QCN radar’s transmitter
√ plus
QCN radar’s transformed mode pairs are independent OPA reception with gain value G = 1+NS / NB [26] and
and identically distributed with zero-mean Gaussian- an ideal photon counter. The fourth (CS-Het) and fifth
distributed quadrature components having covariance (CS-Hom) radars each transmit a coherent-state pulse
matrices of average photon number Ntot , and perform quantum-
limited heterodyne detection and quantum-limited ho-
(NB + NF )I2 02
" #
0 1 modyne detection, respectively.
ΛHQCN = , (29)
1+ N F −1 I
 
0
2 02 Figure 5 shows the ROCs for these radars, plotted
N +1 2 S as log10 (PM ) versus log10 (PF ) for PF , PM ≤ 0.5. The
and QCN, CCN, and QI-OPA ROCs were computed with
 √  Van Trees’s ROC approximation technique [54], which
0 1  (NR + NF )I2  κNS Rc (θ) is known to be accurate for M  1 and especially so [55]
ΛHQCN = √   . (30) for Gaussian signals in Gaussian noise, as we have for the
1 ,κ,θ
2 κNS RTc (θ) 1 + NF −1 I
N +1 2 QCN and CCN radars. The CS-Het ROC [54],
S
p p
Similarly, given the true hypothesis and the κ, θ val- PD = Q( 2M κNS /(NB + 1), −2 ln(PF )), (33)
ues, the CCN radar’s transformed mode pairs are in-
dependent and identically distributed with zero-mean where
Gaussian-distributed quadrature components having co- Z ∞
2
+α2 )/2
variance matrices Q(α, β) ≡ du ue−(u I0 (uα), (34)
β
1 (NB + NF )I2 02
 
0
ΛHCCN = , (31) with I0 (x) being the zeroth-order modified Bessel func-
0
2 02 (1 + NF /NI )I2
tion of the first kind, was computed exactly, as was the
and CS-Hom ROC [54],
√ p
PD = Q(Q−1 (PF ) − 4M κNS /(2NB + 1)),
" #
0 1 (NR + NF )I2 κNS Rc (θ) (35)
ΛHCCN
1 ,κ,θ
= √ . (32)
2 κNS RTc (θ) (1 + NF /NI )I2 where
∞ 2
It is now clear that the QCN and CCN radars have e−u /2
Z
identical conditional statistics in the limit NI → ∞ when Q(x) ≡ du √ , (36)
x 2π
their detectors are quantum limited (NF = 1). Hence
we expect that their quantum-limited performance will with Q−1 (·) being its inverse function, which obeys
be virtually the same for NI  1. Moreover, for non- Q−1 [Q(x)] = x for −∞ < x < ∞.
ideal detectors (NF > 1), the CCN radar outperforms Figure 5 shows two expected behaviors: (1) the QCN
the QCN radar when NI > NF (NS + 1)/(NF − 1), be- and CCN ROCs are indistinguishable; and (2) the QI-
cause the CCN radar’s transformed idler modes then have OPA radar outperforms the CS-Hom, CCN, and CS-Het
lower noise than those of the QCN radar, while all the radars. It also shows one unexpected behavior: the CS-
other second moments of the two radars are identical. Hom and CCN ROCs are also indistinguishable, see Ap-
Furthermore, these conclusions apply for κ, θ determin- pendix D for an explanation of why this occurs. Note
istic and known, as well as for κ, θ random with a known that the QCN, CCN, CS-Hom, and QI-OPA radars all
joint probability distribution. Thus we have proven that use phase information, whereas the CS-Het radar does
the QCN radar cannot outperform all classical radars of not, because its receiver performs incoherent (envelope)
the same transmitted energy and identical detector ca- detecton at the intermediate frequency. For the QCN and
pabilities. CCN radars with known θ, phase information is used to
perform coherent detection at their receivers’ interme-
diate frequency. For uniformly-distributed θ, the QCN
ROC COMPARISONS and CCN radars could use incoherent detection at their
heterodyne receivers’ intermediate frequency, and suffer
As a capstone for all that has been presented, this ROC degradation similar to the performance lost in going
section compares the ROCs for five radars in the best from the CS-Hom ROC to the CS-Het ROC. On the other
possible operating scenario: quantum-limited detection, hand, the CS-Hom and QI-OPA radars require phase in-
θ = 0, and κ deterministic and known. All of these radars formation: the CS-Hom radar needs that information to
are trying to detect the presence of a κ = 0.01 roundtrip- lock its local oscillator’s phase to that of the expected
transmissivity target that is embedded in an NB = 20 target return, while the QI-OPA radar needs it to lock
background by irradiating the region of interest with its retained idler’s phase to that of the expected target re-
Ntot = 2 × 104 photons on average. The first two are the turn. One final comment about QCN versus CCN radars
QCN and CCN radars from the preceding section, with is in order: the QCN radar requires a dilution refrigera-
NS = 0.01, M = 2 × 106 , NI = 103 , and NF = 1. The tor to house its Josephson junction parametric amplifier
11

0 ACKNOWLEDGMENTS
CS-Het
-1 Preparation of this article was supported by the Air
log10 (PM ) Force Office of Scientific Research (AFOSR) MURI pro-
gram under Grant No. FA9550-14-1-0052 and by the
-2
MITRE Corporation’s Quantum Moonshot program.
CCN Portions of this paper were previously presented at
-3 CLEO 2017 [63] and COMCAS 2019 [64]. The author
QCN acknowledges the contributions made to his understand-
-4 CS-Hom ing of quantum illumination by his collaborators (in al-
phabetical order): S. Barzanjeh, B. I. Erkmen, V. Gio-
QI-OPA
vannetti, S. Guha, S. Lloyd, L. Maccone, S. Mouradian,
-5
-5 -4 -3 -2 -1 0 S. Pirandola, S.-H. Tan, D. Vitali, C. Weedbrook, F. N.
log10 (PF ) C. Wong, Z. Zhang, and Q. Zhuang. He also acknowl-
edges valuable discussions with and encouragement from
G. Gilbert.
FIG. 5. ROCs, from top to bottom for the CS-Het, CCN,
QCN, CS-Hom, and QI-OPA radars, plotted as log10 (PM )
versus log10 (PF ) for PF , PM ≤ 0.5. All are trying to detect Appendix A: Entangled State for Lloyd’s QI System
the presence of a κ = 0.01 roundtrip-transmissivity target
that is embedded in an NB = 20 background by irradiating
the region of interest with Ntot = 2 × 104 photons on average. For the high time-bandwidth product (M = T W  1)
The QCN and CCN radars have NS = 0.01, M = 2 × 106 , scenario assumed in Lloyd’s QI system, the nth signal-
NI = 103 , and NF = 1. The QI-OPA radar √ has NS = 0.01, idler pulse pair’s temporal modes can be taken to be
M = 2 × 106 , OPA gain G = 1 + NS / NB , and an ideal
photon counter. The CS-Het and CS-Hom radars each trans- e−j(ωSo +2πm/T )(t−nTr )
√ , (A1)
mit a single, coherent-state pulse of average photon number T
Ntot = 2 × 104 , and perform quantum-limited heterodyne
detection and quantum-limited homodyne detection, respec- for the signal, and
tively.
e−j(ωIo −2πm/T )(t−nTr )
√ , (A2)
T
source, but the CCN radar might use a much less expen- for the idler, where |t − nTr | ≤ T /2, |m| ≤ (M − 1)/2,
sive room-temperature microwave noise generator for its and 0 ≤ n ≤ N − 1. Here: ωSo and ωIo are the signal
source. and idler light’s center frequencies; M is an odd integer;
the pulse-repetition period, Tr , obeys Tr > T ; and the
sign difference in the m-dependent part of the exponents
CONCLUDING REMARKS is convenient for treating Tan et al.’s QI system.
Lloyd’s SP and QI analyses restrict the quantum states
of the preceding temporal modes to the state space
In conclusion, despite QI target detection’s not being spanned by the vacuum and single-photon states. Within
an immediate boon to the radar community, it is im- this state space, his QI transmitter’s nth signal-idler
portant to remember that it is the first example of a pulse pair is in the quantum state
bosonic system that offers a performance improvement
on an entanglement-breaking channel when its transmit- (M −1)/2
(n) 1 X
ter is subject to an energy constraint [56]. In other |ψ iSI =√ |1(n) (n)
m iS |1m iI , (A3)
words, we should not dismiss the value of entanglement M m=−(M −1)/2
for the lossy, noisy situations that are the norm in mi- (n) (n)
crowave radar. Indeed, this meta-lesson may lead to where |1m iS (|1m iI ) denotes the nth signal’s (idler’s)
other quantum systems that offer real utility. Two such state—represented as a ket vector |·iS (|·iI )—containing
examples have already arisen in secure communication: a single photon in its mth mode and vacuum in its other
(1) floodlight quantum key distribution (FL-QKD) [59], modes. For comparison, Lloyd’s SP transmitter’s nth
[60], which directly descends from a communication ver- signal pulse is in the quantum state
sion [61] of QI target detection, is capable of Gbps secret- (M −1)/2
key rates over metropolitan-area fiber connections; and (n) 1 X
|ψ iS = √ |1(n)
m iS , (A4)
(2) quantum low probability of intercept [62]—a repur- M m=−(M −1)/2
posed version of FL-QKD—provides security for Gbps ci-
phertext transmissions over metropolitan-area fiber links i.e., a superposition state containing a single photon that
despite the presence of an eavesdropper who holds the is equally likely to be found in any of the M temporal
decryption key. modes.
12

Appendix B: Entangled State for Tan et al.’s QI Photon counting can be used to measure a phase-
System insensitive cross correlation in a conventional interfer-
ometer. In particular, combining the aRm and aIm
For the high time-bandwidth product (M = T W  1) modes on a 50–50 beam √ splitter we can obtain outputs
scenario assumed in Tan et al.’s QI system, the signal- a±m = (aRm ±aIm )/ 2. Given hypothesis Hj , ideal pho-
idler pulse pair’s temporal mode structure can be be ton counting on these outputs yield outputs whose mean
taken to be values, for j = 0, 1, are

h|aRm |2 iHj ± 2Re[ha∗Rm aIm iHj ] + h|aIm |2 i


e−j(ωSo +2πm/T )t h|a±m |2 iHj = .
√ , (B1) 2
T (C1)
For Tan et al.’s QI system we have ha∗Rm aIm iHj = 0, for
for the signal, and j = 0, 1, so no signature of target absence or presence is
provided by the preceding interferometric measurement.
e−j(ωIo −2πm/T )t What is needed for Tan et al.’s QI system is a way to
√ , (B2) measure the phase-sensitive cross-correlation signature,
T
haRm aIm iHj , for j = 0, 1.
for the idler, where |t| ≤ T /2, and |m| ≤ (M −1)/2. Here,
as was the case for Lloyd’s QI system in Appendix A, ωSo Appendix D: Near-Equivalence of the
and ωIo are the signal and idler light’s center frequencies, Quantum-Limited CS-Hom and CCN Radars
and M is an odd integer. Unlike the case for Lloyd’s QI
analysis, Tan et al. do not restrict their system’s temporal
The CCN radar’s source produces M indepen-
modes to the span of their vacuum and single-photon
dent, identically-distributed, signal-idler mode pairs,
states. In particular, the mth temporal modes of their
{(âSm , âIm )}, that are in classically-random mixtures of
signal and idler have a joint state given by
coherent states whose eigenvalues, {(aSm , aIm )}, have the
s joint probability density,

X NSn
|ψm iSI = |niSm |niIm , (B3) e−|αS | /NS
2
(NS + 1)n+1
p
n=0 paSm ,aIm (αS , αI ) = δ(αI − NI /NS αS ),
πNS
(D1)
where |niSm (|niIm ) denotes the mth signal mode’s (mth where δ(·) is the unit-impulse function. In the limit
idler mode’s) state containing n photons. For compari- NI → ∞, quantum-limited heterodyne detection of the
son, the mth temporal mode of Tan et al.’s coherent-state {âIm } modes yields a noise-free measurement of the
transmitter is in the coherent state p
{aIm = NI /NS aSm }. Optimum post-heterodyne pro-

r cessing is then a maximal-ratio combiner [54], which
X NSn e−NS yields the conditional ROC, given aIm = αIm ,
|ψm iS = |niSm . (B4)
n=0
n!
PD =
 s 
P(M −1)/2 |αSm |2
QQ−1 (PF ) − 2κ m=−(M −1)/2 N +
 , (D2)
Appendix C: Phase-Insensitive and Phase-Sensitive B 1
Cross Correlations
p
where αSm = NS /NI αIm . In the limit M → ∞ we
P(M −1)/2
Because the returned light and the retained idler light have that m=−(M −1)/2 |αSm |2 → M NS , by the law of
are in a classical state—loss and noise having destroyed large numbers, hence the CCN radar’s NI  1, M  1
the initial entanglement of the signal and idler—we can unconditional ROC must satisfy
use semiclassical photodetection theory in our analy-
sis. We refer to the ha∗Rm aIm i cross correlation as being
p
PD ≈ Q(Q−1 (PF ) − 2M κNS /(NB + 1)), (D3)
phase-insensitive, because its value is invariant to apply-
ing a phase shift θ to both aRm and aIm . Conversely, we which matches the CS-Hom radar’s ROC when NB  1.
refer to the haRm aIm i cross correlation as being phase- Note that the preceding near-equivalence also applies to
sensitive, because its phase is shifted by 2θ when a phase CCN and CS-Hom radars with NF > 1 noise figures that
shift θ is applied to both aRm and aIm . satisfy NF  NB .

[1] E. Schrödinger, “Die gegenwärtige Situation in der quan- 828, 844–849 (1935).
tenmechanik,” Naturwissenschaften 23, 807–812, 823–
13

[2] A. Einstein, B. Podolsky, and N. Rosen, “Can quantum- [21] C. Weedbrook, S. Pirandola, R. Garcı́a-Patrón, N. J.
mechanical description of physical reality be considered Cerf, T. C. Ralph, J. H. Shapiro, and S. Lloyd, “Gaus-
complete?,” Phys. Rev. 47, 777–780 (1935). sian quantum information,” Rev. Mod. Phys. 84, 621–
[3] I. M. Georgescu, S. Ashhab, and F. Nori, “Quantum sim- 669 (2012).
ulation,” Rev. Mod. Phys. 86, 153–185 (2014). [22] These inequalities follow from h(zâSm + âIm )† (zâSm +
[4] S. Boixo, T. F. Rennow, S. V. Isakov, Z. Wang, D. âIm )i ≥ 0 and h(zâSm + â†Im )† (zâSm + â†Im )i ≥ 0, for
Wecker, D. A. Lidar, J. M. Martinis, and M. Troyer, all complex-valued z, plus classical states being coherent
“Evidence for quantum annealing with more than one states or their classically-random mixtures.
hundred qubits,” Nat. Phys. 10, 218–224 (2014). [23] These inequalities follow from h(zâSm + âIm )† (zâSm +
[5] J. Biamonte, P. Wittek, N. Pancodtti, P. Rebentrost, âIm )i ≥ 0 and h(zâSm + â†Im )† (zâSm + â†Im )i ≥ 0,
N. Wiebe, and S. Lloyd, “Quantum machine learning,” for all complex-valued z, plus the commutator brackets,
Nature 549, 195–202 (2017). [âKm , â†Km ] = 1 for K = S, I.
[6] P. W. Shor, “Algorithms for quantum computation: Dis- [24] Interestingly, the SPDC source’s output state is also
crete logarithms and factoring,” in Proc. 35th Annu. strongly nonclassical when NS  1, even though that
Symp. on the Foundations of Computer Science pp. 124– regime is of no value to QI target detection. That strong
134 (IEEE Computer Society Press, Los Alamitos, Cali- nonclassicality becomes observable by combining the
fornia, 1994). source’s signal and idler modes on a 50–50 beam √ splitter.
[7] C. H. Bennett and G. Brassard, “Quantum cryptography, The resulting output modes, (âSm ± âIm )/ 2, are then
public key distribution, and coin tossing,” in IEEE Inter- in single-mode squeezed states whose low-noise quadra-
national Conference on Computers, Systems, and Signal tures have variances far below the classical limit of 1/4
Processing (IEEE, 1984), pp. 175–179. when NS  1.
[8] J. Aasi et al. (LIGO Scientific Collaboration), “Enhanced [25] R. Nair, “Discriminating quantum-optical beam-splitter
sensitivity of the LIGO gravitational wave detector by channels with number-diagonal signal states: Applica-
using squeezed states of light,” Nat. Photonics 7, 613– tions to quantum reading and target detection,” Phys.
619 (2013). Rev. A 84, 032312 (2011).
[9] S. Lloyd, “Enhanced sensitivity of photodetection via [26] S. Guha and B. I. Erkmen, “Gaussian-state quantum-
quantum illumination,” Science 321, 1463–1465 (2008). illumination receivers for target detection,” Phys. Rev.
[10] M. F. Sacchi, “Optimal discrimination of quantum oper- A 80, 052310 (2009).
ations,” Phys. Rev. A 71, 062340 (2005). [27] T. Zhong, F. N. C. Wong, A. Restelli, and J. C.
[11] M. F. Sacchi, “Entanglement can enhance the distin- Bienfang, “Efficient single-spatial-mode periodically-
guishability of entanglement-breaking channels,” Phys. poled KTiOPO4 waveguide source for high-dimensional
Rev. A 72, 014305 (2005). entanglement-based quantum key distribution,” Opt. Ex-
[12] The quantum Chernoff bound [13], [14], for N in- press 20, 26868–26877 (2012).
dependent, identically-distributed observations is an [28] F. N. C. Wong, J. H. Shapiro, and T. Kim, “Efficient gen-
exponentially-tight upper bound on the error probability, eration of polarization-entangled photons in a nonlinear
Pr(e) ≤ e−N E /2, of the minimum error-probability quan- crystal,” Laser Phys. 16, 1517–1524 (2006).
tum receiver, i.e., it satisfies limN →∞ [ln(Pr(e))/N ] = [29] Z. Zhang, S. Mouradian, F. N. C. Wong, and J. H.
−E. Shapiro, “Entanglement-enhanced sensing in a lossy
[13] K. M. R. Audenaert, J. Calsamiglia, R. Muñoz-Tapia, and noisy environment,” Phys. Rev. Lett. 114, 110506
E. Bagan, Ll. Masanes, A. Acin, and F. Verstraete, (2015).
“Discriminating states: The quantum Chernoff bound,” [30] E. D. Lopaeva, I. Ruo Berchera, I. P Degiovanni, S. Oli-
Phys. Rev. Lett. 98, 160501 (2007). vares, G. Bride, and M. Genovese, “Experimental real-
[14] J. Calsamiglia, R. Muñoz-Tapia, Ll. Masanes, A. Acin, ization of quantum illumination,” Phys. Rev. Lett. 110,
and E. Bagan, “Quantum Chernoff bound as a measure of 153603 (2013).
distinguishability between density matrices: Application [31] D. G. England, B. Balaji, and B. J. Sussman, “Quantum-
to qubit and Gaussian states,” Phys. Rev. A 77, 032311 enhanced standoff detection using correlated photon
(2008). pairs,” Phys. Rev. A 99, 023828 (2019).
[15] J. H. Shapiro, “The quantum theory of optical communi- [32] J. Calsamiglia, J. I. de Vicente, R. Muñoz-Tapia, and
cations,” IEEE J. Sel. Top. Quantum Electron. 15, 1547– E. Bagan, “Local discrimination of mixed states,” Phys.
1569 (2009). Rev. Lett. 105, 080504 (2010).
[16] J. H. Shapiro and R. W. Boyd, “The physics of ghost [33] S. Bandyopadhyay, “More nonlocality with less purity,”
imaging,” Quantum Inf. Process. 11, 949–993 (2012). Phys. Rev. Lett. 106, 210402 (2011).
[17] R. J. Glauber, “Coherent and incoherent states of the [34] D. Bacon, I. L. Chuang, and A. W. Harrow, “Efficient
radiation field,” Phys. Rev. 131, 2766–2788 (1963). quantum circuits for Schur and Clebsch-Gordan trans-
[18] J. H. Shapiro and S. Lloyd, “Quantum illumination ver- forms,” Phys. Rev. Lett. 97, 170502 (2006).
sus coherent-state target detection,” New J. Phys 11, [35] Q. Zhuang, Z. Zhang, and J. H. Shapiro, “Opti-
063045 (2009). mum mixed-state discrimination for noisy entanglement-
[19] S.-H. Tan, B. I. Erkmen, V. Giovannetti, S. Guha, S. enhanced sensing,” Phys. Rev. Lett. 118, 040801 (2017).
Lloyd, L. Maccone, S. Pirandola, and J. H. Shapiro, [36] S. J. Dolinar, “An optimum receiver for the binary
“Quantum illumination with Gaussian States,” Phys. coherent-state quantum channel,” Research Laboratory
Rev. Lett. 101, 253601 (2008). of Electronics, MIT, Quarterly Progress Report No. 111,
[20] S. Pirandola and S. Lloyd, “Computable bounds for the pp. 115-120, 1973.
discrimination of Gaussian states ,” Phys. Rev. A 78,
012331 (2008).
14

[37] Q. Zhuang, Z. Zhang, and J. H. Shapiro, “Entanglement- n̂amp


Km , is in a thermal state with average photon
enhanced Neyman-Pearson target detection using quan- number Namp . Non-quantum-limited heterodyne de-
tum illumination,” J. Opt. Soc. Am. B 34, 1567–1572 tection of âamp
Km results in a classical complex-valued
(2017). random variable, aKm , whose statistics are equivalent
[38] Q. Zhuang, Z. Zhang, and J. H. Shapiro, “Quantum illu- to those of the operator âamp det†
Km + n̂Km for K = R, I,
mination for enhanced detection of Rayleigh-fading tar- where the noise mode, n̂det Km , is in a thermal state
gets,” Phys. Rev. A 96, 020302(R) (2017). with average photon number Ndet . The covariance
[39] N. S. Kopeika and J. Bordogna, “Background noise in matrices for the QCN and CCN radars then folllow
optical communication systems,” Proc. IEEE 58, 1571– directly from Tan et al.’s target-detection scenario
1577 (1970). with GA NF ≡ (GA − 1)(Namp + 1) + Ndet + 1. This
[40] J. H. Shapiro, S. Guha, and B. I. Erkmen, “Ultimate measurement is quantum limited when both n̂amp Km and
channel capacity of free-space optical communications,” n̂det
Km are in their vacuum states so that Namp = Ndet = 0
J. Opt. Netw. 4, 501–515 (2005). lead to NF = 1.
[41] S. Barzanjeh, S. Guha, C. Weedbrook. D. Vitali, J. H. [54] H. L. Van Trees, Detection, Estimation, and Modulation
Shapiro, and S. Pirandola, “Microwave quantum illumi- Theory, Part I. New York, Wiley, 1968.
nation,” Phys. Rev. Lett. 114, 080503 (2015). [55] J. H. Shapiro, “An extended version of Van Trees’s
[42] The high-brightness background condition for microwave receiver operating characteristic approximation,” IEEE
operation is equivalent to h̄ω  kTs , where ω is the radar Trans. Aerospace and Electron. Systems 35, 709–716
frequency, k is Boltzmann’s constant, and Ts is the sky (1999).
temperature at frequency ω. [56] Bennett et al. [57] showed that use of prior shared entan-
[43] J. Hewitt, “Quantum radar can de- glement can increase the classical information capacity
tect what’s invisible to regular radar,” of the lossy, noisy, bosonic channel in comparison with
https://www.extremetech.com/extreme/200161- a classical (coherent-state source, coherent-detection re-
quantum-radar-can-detect-whats-invisible-to-regular- ceiver) system of the same transmitted energy. Because
radar. Bennett et al. did not account for the energy needed
[44] S. Pirandola, B. Roy Bardhan, T. Gehring, C. Weed- for entanglement distribution, their system does not rep-
brook, and S. Lloyd, “Advances in photonic quantum resent a true benefit from using entanglement over an
sensing,” Nat. Photonics 12, 224–233 (2018). entanglement-breaking channel, i.e., the classical infor-
[45] C. Macklin, K. O’Brien, D. Hover, M. E. Schwartz, V. mation capacity of the lossy, noisy, bosonic channel that
Bolkhovsky, X. Zhang, W. D. Oliver, and I. Siddiqi, is subject to an average energy constraint on all its
“A near-quantum-limited Josephson traveling-wave para- transmitted light—including whatever light is transmit-
metric amplifier,” Science 350, 307–310 (2015). ted during entanglement distribution—is achieved with
[46] Thus far in this paper we have ignored—and, except here, coherent-state transmission without entanglement [58].
we will continue to ignore—a target’s effects on its re- [57] C. H. Bennett, P. W. Shor, J. A. Smolin, and A. V.
turn’s polarization and frequency. These effects, however, Thapliyal, “Entanglement-assisted capacity of a quan-
should be included in a more complete QI analysis. For tum channel and the reverse Shannon theorem,” IEEE
example, they place additional burdens—polarization Trans. Inform. Theory 48, 2637–2655 (2002).
tracking and Doppler compensation—on Guha and Erk- [58] V. Giovannetti, R. Garcı́a-Patrón, N. J. Cerf, and A. S.
men’s OPA receiver. Holevo, “Ultimate classical communication rates of quan-
[47] U. Las Heras, R. Di Candia, K. G. Fedorov, F. Deppe, tum optical channels,” Nat. Photon. 8, 796–800 (2014).
M. Sanz, and E. Solano, “Quantum illumination reveals [59] Q. Zhuang, Z. Zhang, J. Dove, F. N. C. Wong, and J. H.
phase-shift induced cloaking,” Sci. Rep. 7, 9333 (2017). Shapiro, “Floodlight quantum key distribution: A prac-
[48] G. De Palma and J. Borregaard, “Minimum error proba- tical route to gigabit-per-second secret-key rates,” Phys.
bility of quantum illumination,” Phys. Rev. A 98, 012101 Rev. A 94, 012322 (2016).
(2018). [60] Z. Zhang, C. Chen, Q. Zhuang, F. N. C. Wong, and J.
[49] D. Luong, B. Balaji, C. W. S. Chang, V. Manjunath, A. H. Shapiro, “Experimental quantum key distribution at
Rao, and C. Wilson, “Microwave quantum radar: An ex- 1.3 gigabit-per-second secret-key rate over a 10 dB loss
perimental validation,” in Proc. 2018 IEEE International channel,” Quantum Sci. Technol. 3, 25007 (2018).
Carnahan Conference on Security Technology (ICCST), [61] J. H. Shapiro, “Defeating passive eavesdropping with
New York, NY, USA (IEEE, 2018), pp. 110–115. quantum illumination,” Phys. Rev. A 80, 022320 (2009).
[50] D. Luong, C. W. S. Chang, A. M. Vadiraj, A. Damini, [62] J. H. Shapiro, D. M. Boroson, P. B. Dixon, M. E. Grein,
C. M. Wilson, and B. Balaji, “Receiver operating char- and S. A. Hamilton, “Quantum low probability of inter-
acteristics for a prototype two-mode squeezing radar,” cept,” J. Opt. Soc. Am. B 36, B41–B50 (2019).
arXiv:1903.00101 [quant-ph]. [63] J. H. Shapiro, “Quantum illumination: From enhanced
[51] S. Barzanjeh, S. Pirandola, D. Vitali, and J. M. target detection to Gbps quantum key distribution,” in
Fink, “Experimental microwave quantum illumination,” Conference on Lasers and Electro-Optics (CLEO), Tech-
arXiv:1908.03058v2 [quant-ph]. nical Digest, Washington, DC, USA (OSA, 2017), paper
[52] Here, and in the next two sections, we forego considera- FTu3F.1.
tion of QI with SFG or FF-SFG reception owing to the [64] J. H. Shapiro, “The quantum illumination story,” in
dim prospects for implementing those receivers with cur- IEEE Conference on Microwaves, Antennas, Communi-
rent technology. √ cations, and Electronic Systems (COMCAS), Technical
[53] Pre-amplification yields âamp Km = GA âKm + Digest, New York, NY, USA (IEEE, 2019).

GA − 1 n̂amp†
Km for K = R, I, where the noise mode,

You might also like