Brought to you by:
Paper

Photon-number correlation for quantum enhanced imaging and sensing

, , , , , , and

Published 11 August 2017 © 2017 IOP Publishing Ltd
, , Citation A Meda et al 2017 J. Opt. 19 094002 DOI 10.1088/2040-8986/aa7b27

2040-8986/19/9/094002

Abstract

In this review we present the potentialities and the achievements of the use of non-classical photon-number correlations in twin-beam states for many applications, ranging from imaging to metrology. Photon-number correlations in the quantum regime are easily produced and are rather robust against unavoidable experimental losses, and noise in some cases, if compared to the entanglement, where losing one photon can completely compromise the state and its exploitable advantages. Here, we will focus on quantum enhanced protocols in which only phase-insensitive intensity measurements (photon-number counting) are performed, which allow probing the transmission/absorption properties of a system, leading, for example, to innovative target detection schemes in a strong background. In this framework, one of the advantages is that the sources experimentally available emit a wide number of pair-wise correlated modes, which can be intercepted and exploited separately, for example by many pixels of a camera, providing a parallelism, essential in several applications, such as wide-field sub-shot-noise imaging and quantum enhanced ghost imaging. Finally, non-classical correlation enables new possibilities in quantum radiometry, e.g. the possibility of absolute calibration of a spatial resolving detector from the on-off single-photon regime to the linear regime in the same setup.

Export citation and abstract BibTeX RIS

1. Introduction

Quantum correlations are a subject of deep interest since their exploitation could open unprecedented opportunities in several fields, ranging from the very foundations of quantum mechanics [1] to cosmology [2, 3], and represent a basic resource for the development of quantum technologies such as fundamental metrology [4, 5], quantum communication [6, 7], quantum biology [8, 9], and quantum imaging and sensing [1018].

Quantum enhanced measurement protocols aim at reducing the uncertainty in the estimation of some physical quantities of a system, measuring some modification of an optical probe state, below classical shot-noise limit (or standard quantum limit) scaling as ${n}^{-1/2}$, where n is the number of particles of the probe. Most theoretical investigations have been addressed at using entangled states to change this scaling to a stronger one, up to the ultimate limit imposed by quantum mechanics, ${n}^{-1}$, known as the Heisenberg limit. Many measurement schemes have been proposed [19, 20], and some experimental proofs of principle have been realized [2128] in this direction, typically using an entangled state of the form ${2}^{-1/2}(| n0\rangle +| 0n\rangle )$ (NOON state), where the n photons are distributed over the two paths of an interferometer. While two-photon entangled states are quite routinely produced by low-gain spontaneous parametric down conversion (SPDC) and the Hong–Ou–Mandel effect, in practice generating and detecting a $n\gt 2$ NOON state is more challenging. Even worse, entanglement itself is extremely fragile to losses, for example losing a single photon from a NOON state projects it in a classical mixture. Other quantum states, entangled and squeezed, have been considered, which are more resilient to experimental imperfections [13], but nevertheless reaching the Heisenberg limit for a large number of photons is probably a chimera. In fact, it has been recently shown that in the presence of decoherence, the Heisenberg limit (and, in general, any change of the scaling with the photon number in the uncertainty differing by the shot-noise limit) is out of reach [29, 30]. Rather, an enhancement with respect to the standard quantum limit is at most by a constant factor, for example it takes the form $\sqrt{(1-\eta )/\eta }$ in the presence of a loss factor $(1-\eta )$ [31, 32].

On the other hand, the same advantage can be obtained more easily by exploiting non-classical Gaussian states [33], which are relatively easy to produce experimentally, such as a squeezed vacuum generated by SPDC and optical parametric oscillators (OPO). Single-mode squeezing [3437] in one of the quadratures (generated by OPO) was the first quantum property considered for quantum metrology, in particular for enhanced interferometry [38], and the more successful from the practical point of view, leading to a real sensitivity improvement of modern gravitational wave detectors [35, 39] and also to promising application to photonic force microscopy for biological particle tracking [40, 41] and beam displacement measurements [42].

A fundamental property of the two-mode squeezed vacuum, also referred to as a twin-beam (TWB) state, is that it is entangled in the photon number basically assuming the form ${\sum }_{n=0}^{\infty }{c}_{n}| n\rangle | n\rangle $, meaning that two ideal detectors intercepting each mode respectively always measure the same number of photons. This correlation is strongly non-classical and does not involve any measurement of the phase. Essentially, all the optical measurements, which aim at the estimation of absorption, transmission and reflection can be enhanced by using photon-number correlation. The idea is that using one beam of the pair as a probe and the other as a reference, the strong correlation helps to detect slight modifications in the signals when the two beams are compared. This scheme allows an enhancement in sensitivity that can be exploited in different fields, such as interferometry [43, 44] or imaging [14, 15]. While the direct measurement of the photon number only, i.e. of the intensities, introduces some limitations in the field of applicability, it is experimentally more feasible in many situations, even in a realistic scenario including noise and losses. It is emblematic in this framework that the possibility of detecting partially reflecting objects with significant enhanced sensitivity exactly when the background at the receiver is much more intense than the returning probe, just measuring non-classical photon-number correlations [45]. We mention that bright TWB states with strong non-classical correlation have recently been obtained by four-wave mixing in hot atomic vapor, and some groups have demonstrated the great potentiality of this source in imaging and sensing applications [4651]. These considerations can be extended to the multimode spatial case. Indeed, when TWBs are produced through traveling-wave parametric down conversion, or by four-wave mixing [52], the emission is approximatively a product of a large number of two-mode (spatial) squeezed states, which can be intercepted and detected independently at the same time. Modern high sensitivity multi-pixel detectors, such as charge-coupled device (CCD) cameras can exploit this parallelism for improving the sensitivity of wide-field imaging (WFI) applications. One of the goals, which has recently been demonstrated, is the realization of a wide-field microscope operating below the shot-noise limit [53]. One of the first applications of SPDC entangled photons has been ghost imaging (GI) [54, 55], whose goal is the reconstruction of the spatial transmission/reflection profile of an object by using a single-pixel detector. Even though GI has been also demonstrated exploiting classical correlations and computational methods (exploiting random light patterns generated by a computer and a spatial light modulator), the use of non-classical correlations instead of classical ones can provide sensitivity advantages in very-low illumination regimes [56, 57].

Finally, non-classical correlations have disclosed new possibilities in quantum radiometry [58], e.g. the possibility of absolute calibration of detectors, erasing the need for comparison with calibrated standards. The first proposal for calibrating a single-photon detector was formulated by Zel-Dovich and Klyshko [59] just after the discover of the SPDC process, and nowadays it is an established technique [60] used in metrological institutes. Generalizing the method to the domain of analog detectors and spatially resolving detectors has recently led to the first absolute calibration of electron-multiplying-CCD (EMCCD) and intensified CCD (ICCD) cameras [6163]. An EMCCD camera has been calibrated from the single-photon regime to the linear regime, in the same setup, by tuning the intensity of the SPDC pump laser [61].

One of the main goals of this review is to give the reader all the elements for understanding with a certain level of detail the origin of the quantum advantage in the applications mentioned previously, in particular linking clearly the sensitivity improvement with the degree of non-classicality measured by appropriate parameters. Since losses are unavoidable in optical measurements and they usually affect the performance of quantum strategies quite a lot, we always take them into consideration in the derivation of the results. The review is structured in the following way. In section 2 we introduce some basic elements of the quantum photodetection model. In section 3 we discuss non-classical photon statistics and phododetection statistics, how they can be quantified and the boundary between the classical and quantum worlds. Section 4 presents in some detail the generation of photon-number entangled states in a spatially multimode regime by SPDC, and the issues relating to the efficient detection of non-classical correlation in the far field of the emission. The following sections 5, 6, 7 and 8 are devoted to the presentation of noticeable applications of quantum photon-number correlations, in particular sub-shot-noise (SSN) imaging, target detection against a preponderance of noise (quantum illumination), quantum enhanced GI and the absolute calibration of detectors, respectively.

2. Quantum theory of photodetection and quantum efficiency

The complete and consistent quantum theory of photodetection, relating photo-counts distribution to the intrinsic statistical nature of light and its coherence properties, was introduced by Glauber in 1963 [64, 65]. For a long time only strongly incoherent sources were available and the description of light was confined to the Planck distribution. One of the first and groundbreaking experiments investigating the statistics and coherence of light was realized in 1956 by Hanbury Brown and Twiss, who showed the tendency of starlight to generate a photo-current correlation between two detectors if sufficiently close to each other, demonstrating the photon-bunching effect in thermal light [66, 67]. On the other hand the invention of the laser (and maser before) has led to the measurement of different kinds of statistics, from poissonian (which is a peculiarity of the laser itself) to strongly super-poissonian, for example if a laser is scattered by a rotating ground-glass disc, as reported by Arecchi in [68]. These new observations and phenomenology led to a theoretical effort to deeply understand the relationship between the intrinsic statistical nature of different kinds of sources and the process of photodetection. In a photodetector, the absorption of a photon generates a signal (usually an electric pulse), which represents a photon count. The statistics of these counting events is a faithful representation of the photon statistics only if the detector has ideal characteristics, namely infinite spectral bandwidth (the electric pulse is close to a delta function in time), linear response with the number of photons, and perfect quantum efficiency (each photon impinging on the detector generates a count). Nevertheless, non-idealities of detectors can alter the desired one-to-one relation between the impinging photon and the generated counts.

For instance, in a linear photodetector, the effect of a non-unit quantum efficiency η, can be modeled as the random evolution of the field after passing through a beam splitter (BS) with transmission equal to η [69]. Introducing the bosonic photon annihilation operator $\hat{a}$ of the incoming field, such that $[\hat{a},{\hat{a}}^{\dagger }]=1$, the unitary input–output relations of the BS provide the expression of the transmitted and reflected fields $\hat{{b}_{1}}$ and $\hat{{b}_{2}}$, respectively:

Equation (1)

Equation (2)

where $\hat{v}$ is the mode operator corresponding to the second input port of the BS, which is considered here in the vacuum state $| 0\rangle $.

The photon statistics of the transmitted beam $\hat{{b}_{1}}$ correspond to that of the input beam after the random selection process has taken place (see figure 1). The evolution of the statistics of the number of photons of the incoming field, $\hat{n}={\hat{a}}^{\dagger }\hat{a}$, can be easily calculated from equation (1) using the bosonic commutators [70]:

Equation (3)

where $\langle \hat{N}\rangle $ is the mean value of the measured photon-number operator and $\langle {{\rm{\Delta }}}^{2}\hat{N}\rangle $ is its variance. In equations (3), the definition of the quantum efficiency as the ratio between the detected and the incoming mean number of photons is recovered and the modification of the statistics when $\eta \lt 1$ is clearly expressed. Note that any kind of photon loss, not necessarily at the detection stage, can be treated in the same way [71], so that the formulas in equation (3) can be applied in a broad range of situations where η can assume a different role, which will be specified time by time. Coherent states with $\langle {{\rm{\Delta }}}^{2}\hat{n}\rangle =\langle \hat{n}\rangle $, and thermal states with $\langle {{\rm{\Delta }}}^{2}\hat{n}\rangle \,=\langle \hat{n}\rangle (1+\langle \hat{n}\rangle )$, maintain the same statistical properties, just with a rescaled mean value. On the other hand, losses in the detection process are responsible for the degradation of sub-Poissonian statistics ($\langle {{\rm{\Delta }}}^{2}\hat{n}\rangle \lt \langle \hat{n}\rangle $), which is a signature of the quantum features of light, as we will see in section 3. As a matter of fact, in the expression of the variance, the second term is the Poissonian noise arising from the quantum fluctuation of the vacuum, which occurs even if the incoming field is free from photon-number fluctuation, (${{\rm{\Delta }}}^{2}\hat{n}=0$), such as a Fock state $| n\rangle $ eigenstate of the photon-number operator. In the presence of high losses ($\eta \ll 1$) the photocount statistics tends to the Poissonian one, vanishing the peculiar inner statistical properties of the field.

Figure 1.

Figure 1. Model of a linear photodetector with quantum efficiency η corresponding to the transmissivity of the BS.

Standard image High-resolution image

Correlations are also affected by the process of detection. Considering, for example, the covariance of two modes $\langle {\rm{\Delta }}{\hat{n}}_{1}{\rm{\Delta }}{\hat{n}}_{2}\rangle $, undergoing two independent detection processes with quantum efficiencies ${\eta }_{1}$ and ${\eta }_{2}$, respectively, it evolves as

3. Non-classical photon statistics

Ideal photodetection, without losses, provides precise information on the intrinsic statistical nature of the impinging light. Therefore, an analysis of the fluctuation in the photon counts can be used to trace a discrimination between the quantum and classical nature of the light, where the boundary is represented by the coherent states [65, 72, 73].

Coherent states, experimentally generated by an ideal laser, can be represented by a displaced vacuum state:

Equation (4)

and are eigenstates of the annihilation operator, $\hat{a}| \alpha \rangle =\alpha | \alpha \rangle $, where α is a complex number. In the photon-number basis, the single-mode state $| \alpha \rangle $ can be expressed as:

Equation (5)

from which it is possible to calculate the photon-number distribution p(n), which turns out to be Poissonian:

Equation (6)

with photon-number variance equal to the mean photon number, $\langle {{\rm{\Delta }}}^{2}\hat{n}\rangle =\langle \hat{n}\rangle $. The relative uncertainty in the mean photon number ${\rm{\Delta }}\hat{n}/\langle \hat{n}\rangle =1/\sqrt{\langle \hat{n}\rangle }$ is usually referred to as the 'shot noise' scaling and establishes a lower bound to the uncertainty of any measurement using classical probes. This point can be clarified through the Glauber–Sudarshan representation [65, 7375]. The phase and modulus of coherent states completely span the phase-space (actually they form an over-complete base), and any arbitrary state with density matrix ρ can be represented as a weighted combination of coherent states

Equation (7)

where $P(\alpha )$ is a quasi-probability distribution. Because ρ is Hermitian and has unit trace, $P(\alpha )$ is real and normalized to unity. However, it does not always behave as a well-defined probability density, for example it can assume negative values or can be more singular than a delta function. Classical states are identified by $P(\alpha )\geqslant 0$ [69], which behave as a true probability density function. This definition is motivated by the fact that for such states the photon statistics predicted by the quantum photodetection theory coincide with the ones derived in the framework of the semiclassical theory of photodetection [76], where the incoming field is considered as classical wave and the shot noise is the result of a random process due to the discreteness of the electron charge [77, 78] generated inside the detector (for all the three paradigms of direct, homodyne and heterodyne detection).

From equation (7) it follows that quantum expectation values of normally ordered operators are expressed through the integral of the corresponding classical quantities weighted with the quasi-probability distributions, as

Equation (8)

In particular, the expression of the photon-number variance within the Glauber–Sudarshan representation is [75]:

Equation (9)

and shows a first term due to the discreteness nature of the light, the shot noise, and a second normally ordered term, usually called second-order Glauber correlation function ${G}^{(2)}$, which can be interpreted as a quasi-classical variance. For classical states, with $P(\alpha )\geqslant 0$, the integral is positive or null, and the fluctuations are Poissonian or super-Poissonian. For non-classical states, in which the quasi-probability assumes negative value (Fock states, squeezed states, or entangled states) it is possible to have a negative integral, allowing SSN fluctuations.

According to the discussion above it is useful to introduce a specific parameter to measure the non-classicality of a state. We consider the Fano factor $F=\langle {{\rm{\Delta }}}^{2}\hat{n}\rangle /\langle \hat{n}\rangle $ [79] or the Mandel Q parameter [80]:

Equation (10)

The value of the Fano factor F = 1 (Q = 0) establishes a bound between classical and non-classical photon statistics; F is lower bounded by the unity for classical states, while specific non-classical states can have $0\leqslant F\lt 1$ ($-1\leqslant Q\lt 0$).

As pointed out in section 2, the statistics of a state are deteriorated by the losses in the photodetection process (including both losses in the optical path and the detector quantum efficiency). The detected Fano factor in the presence of optical losses η becomes ${F}_{{\rm{\det }}}=\eta F+1-\eta $, as it descends from equation (3). Thus, in the presence of losses, the lower bound for a non-classical state is ${F}_{{\rm{\det }}}=1-\eta $.

3.1. Two-mode non-classical statistics

By analogy with equation (7), a classical two-mode (bipartite) state is represented by a Glauber–Sudarshan probability density function $P({\alpha }_{1},{\alpha }_{2})\geqslant 0$

Equation (11)

We can quantify the degree of correlation between the modes and their non-classical features defining the noise reduction factor (NRF) σ as the ratio between the variance of the difference in the number of photons and the noise of two coherent states of equivalent energies [8191]:

Equation (12)

The NRF represents the equivalent of the Fano factor for a bipartite state; in this case, the shot-noise level is given by the sum of the shot noise of the two modes $\langle {\hat{n}}_{1}+{\hat{n}}_{2}\rangle $. For classical bipartite states, σ is larger than or equal to 1. For non-classical beams, quantum correlations can lead to $0\leqslant \sigma \lt 1$. As for the Fano factor, σ is affected by the optical losses experienced by the two fields. From equation (3), considering two modes subject to the same transmission-detection efficiency ${\eta }_{1}={\eta }_{2}=\eta $,

Equation (13)

The lowest bound in the presence of losses is therefore ${\sigma }_{{\rm{\det }}}=1-\eta $.

A demonstration of the classical limit of the correlation is given here in a specific case. Let us consider a two-mode state generated by splitting a single mode $\hat{a}$ with a BS of transmittance τ. In the case of ideal photodetection, the statistics of the output modes can be computed using the input–output relations of the BS in equation (1) (with $\tau =\eta $) as:

Equation (14)

Equation (15)

Equation (16)

The last expression reveals that in order to have a non-null covariance the statistics of the incoming light must be super-poissonian, which also means that a split coherent state does not generate any correlation, while a thermal beam does. Using the relations (14)–(16) in equation (12) one can express the NRF as:

Equation (17)

where $F=\langle {{\rm{\Delta }}}^{2}\hat{n}\rangle /\langle \hat{n}\rangle $. For a balanced 50:50 ($\tau =1/2$) BS this leads to the classical limit $\sigma =1$, irrespective of the statistical properties of the incoming beam, either sub-poissonian or super-poissonian. This means that the fluctuations of the two modes are suppressed by the subtraction, except the shot noise. On the other hand, for unbalanced BS, i.e. $\tau \ne 1/2$, an incoming field with sub-poissonian statistics generates non-classical correlations ($\sigma \lt 1$) at the output ports.

Another parameter that can be used as an indicator of non-classicality for two-mode states is the Cauchy–Schwarz parameter [92]:

Equation (18)

where $\langle ::\rangle $ is the normally ordered quantum expectation value. While σ is deteriorated by the losses, ε is remarkably immune to them, and for this reason it allows experimental access to the non-classical features, even for inefficient detection processes. However, noise added to the detection degrades its value (see section 6). For classical states of light, with a positive Glauber–Sudarshan P function, the Cauchy–Schwarz parameter is $\epsilon \leqslant 1$, while for states with a negative (or singular) P function this limit can be violated. In the case of correlated thermal beams, obtained by a 50:50 BS, the most used classically correlated states (for example, in the classical ghost-imaging protocols, see section 7), $\langle :{{\rm{\Delta }}}^{2}{\hat{n}}_{1}:{\rangle }_{{\rm{TH}}}\,=\langle :{{\rm{\Delta }}}^{2}{\hat{n}}_{2}:{\rangle }_{{\rm{TH}}}=\langle {\rm{\Delta }}{\hat{n}}_{1}{\rm{\Delta }}{\hat{n}}_{2}{\rangle }_{{\rm{TH}}}=\langle \hat{n}{\rangle }^{2}$, as can be simply derived by equations (14) and (16), by introducing the thermal variance $\langle {{\rm{\Delta }}}^{2}\hat{n}\rangle =\langle \hat{n}\rangle (1+\langle \hat{n}\rangle )$. Therefore, the Cauchy–Schwarz parameter for a split thermal beam is ${\varepsilon }_{{\rm{TH}}}=1,$ saturating the classical bounds. This demonstrates that thermal split beams show the best possible correlation allowed for classical states. They represent the classical benchmark for comparing the quantum enhanced performance in some emblematic imaging and sensing protocols, see sections 6 and 7.

4. Spatially multimode photon-number correlation: generation and detection

In fact, the most efficient ways to produce quantum correlations between optical fields are based on SPDC [9397]. This physical phenomenon was discovered at the end of sixties [59, 98] and in recent years, thanks to the development of new kinds of laser systems and photon detectors, it is exploited in the most advanced quantum technologies such as quantum key distribution [99103], quantum computing [104107], tailoring of quantum states [108112], quantum imaging [15, 113] and quantum sensing [114]. Moreover, SPDC is exploited in several experiments concerning the foundation of quantum mechanics [115119]. SPDC is due to the interaction between an intense optical field, usually called 'pump beam', and a non-linear optical medium. Basically, the phenomenon consists in the decay of one photon of the pump beam into two photons, preserving energy and momentum:

Equation (19)

where ${\omega }_{{\rm{p}}}$ is the frequency of the pump photon and ${\omega }_{1}$, ${\omega }_{2}$ are the frequencies of the photons emitted by SPDC, and where ${{\bf{k}}}_{j}$ (with j = p, 1, 2) are the corresponding wave vectors (see figure 2).

Figure 2.

Figure 2. Schematic representation of the spontaneous parametric down conversion.

Standard image High-resolution image

In this section we will describe the physics of the SPDC process, not limited to the low-gain regime (as in [9396]), and consider a multimodal emission both in frequency and momentum [120]. For an alternative treatment see, for example, [97, 121, 122].

4.1. Spontaneous parametric down conversion

In non-linear optics, the dielectric polarization P is expanded as [69, 123]:

Equation (20)

For higher strengths of the electric field (E), the higher-order non-linear terms become important. In addition to ${\chi }^{1}$ being the linear susceptibility coefficient, ${\chi }^{2}$, ${\chi }^{3}$ (...${\chi }^{n}$) are called the non-linear susceptibility coefficients of the medium. Taking into accounts non-linear effects up to second order, the field Hamiltonian in a non-magnetic medium, $H={\int }_{V}\tfrac{1}{2}\vec{E}\cdot ({\epsilon }_{0}\vec{E}+\vec{P})$ can be written as:

Equation (21)

with

Equation (22)

Equation (23)

where the summation on repeated indexes is understood and where here the interaction extends over the volume V of the non-linear medium. The last expression represents the non-linear interaction involving three electric fields and is responsible for fundamental optical non-linear processes like sum- and difference-frequencies generation, second hardmonic generation and also SPDC. The corresponding interaction Hamiltonian is:

Equation (24)

In SPDC, the non-linear effect is small and the probability that a pump photon is down converted into two emitted photons is very low. The pump is usually very intense and not significantly depleted by the interaction, thus can be treated classically, whereas the quantum description of the down-converted fields is essential. They can be written as:

Equation (25)

where j = 1, 2. The electric field of a classical monochromatic pump propagating along the Z-axis direction is:

Equation (26)

where ${\boldsymbol{\rho }}$ is the coordinate vector in the transverse X–Y plane. Considering each wave vector divided into the longitudinal component (pump direction), kjz, and transverse component, ${{\boldsymbol{q}}}_{j}$, the interaction Hamiltonian becomes:

Equation (27)

Initially the down-converted fields are in the vacuum state, and upon the interaction, the evolved state in the Schrödinger picture follows:

Equation (28)

Considering L the length of the crystal, the integral along the z direction results in:

Equation (29)

where ${\rm{\Delta }}k={k}_{{\rm{p}}}-{k}_{1z}-{k}_{2z}$ is the longitudinal phase mismatch. In the limit $L\to \infty $, the sinc function becomes a delta function and the integral term is different from zero for the perfect phase matching condition, i.e ${\rm{\Delta }}k=0$. In the realistic situation, the finite thickness of the crystal allows a certain phase mismatch, whose measure is given by the width of the sinc central peak, inversely proportional to the crystal length.

Similarly, the surface integral in the transverse direction ${\boldsymbol{\rho }}$ leads to the Fourier transform of the pump profile $A({\boldsymbol{\rho }})$. In the approximation of the plane wave, $A({\boldsymbol{\rho }})={A}_{0}$, we have

Equation (30)

In this approximation, the down-converted modes are perfectly correlated in the transverse direction, i.e. the mode with transverse momentum ${\boldsymbol{q}}$ is correlated to a mode with momentum ($-{\boldsymbol{q}}$).

The integral over the interaction time of equation (28) leads to:

Equation (31)

This allows us to express the frequencies as ${\omega }_{1}=\tfrac{{\omega }_{p}}{2}+{\rm{\Omega }}$ and ${\omega }_{2}=\tfrac{{\omega }_{p}}{2}-{\rm{\Omega }}$, where $\tfrac{{\omega }_{{\rm{p}}}}{2}$ is the degenerate frequency. With this simplification, the evolution operator becomes:

Equation (32)

where the phase matching function $f({\boldsymbol{q}},{\rm{\Omega }})$, contains information about the strength of interaction (proportional to the length of the non-linear medium and the pump amplitude) and the spatio-temporal bandwidth of the down-converted fields:

The quantum state of SPDC modes at the start of the process is a vacuum state and due to the time evolution becomes:

Equation (33)

Considering discrete values of ${\bf{q}}$, Ω, the integral can be replaced by the summation:

Equation (34)

Since the operators appearing in equation (34) corresponding to different pairs of modes $({\boldsymbol{q}},{\rm{\Omega }})\ne ({\boldsymbol{q}}^{\prime} ,{\rm{\Omega }}^{\prime} )$ commute with each other, following the Baker–Campbell–Hausdorff formula, i.e ${{\rm{e}}}^{x(\hat{A}+\hat{B})}={{\rm{e}}}^{x\hat{A}}\cdot {{\rm{e}}}^{x\hat{B}}$ for $[\hat{A},\hat{B}]=0$, the above state can be written in the direct product form as follows:

Equation (35)

In the plane-wave pump approximation, SPDC can be seen as a collection of independent states, each one involving two-modes with correlated transverse momenta and frequencies. Expanding the exponential it is possible to rewrite the state as a product of two-mode entangled states in the photon number (multimode TWB) [120]:

Equation (36)

where the probability amplitude ${c}_{{\boldsymbol{q}},{\rm{\Omega }}}(n)\propto \sqrt{{\mu }^{n}/{(\mu +1)}^{n+1}}$ is a coefficient that can be considered constant and is related to the mean number of photons in the mode $({\boldsymbol{q}},{\rm{\Omega }})$, $\mu \,={\sinh }^{2}| f({\boldsymbol{q}},{\rm{\Omega }})| $.

4.2. SPDC photon statistics

We are now interested in the statistical distribution of photons for a couple of conjugated modes, indicated by ${\hat{a}}_{({\boldsymbol{q}},{\rm{\Omega }})}\to {\hat{a}}_{1}$ and ${\hat{a}}_{(-{\boldsymbol{q}},-{\rm{\Omega }})}\to {\hat{a}}_{2}$. To calculate this it is convenient to consider one of the evolution operators in equation (35) (the so-called two-mode squeezing operator):

Equation (37)

acting only on conjugated modes. For simplicity, we rewrite the complex amplitude as $f({\boldsymbol{q}},{\rm{\Omega }})={{r}{\rm{e}}}^{{\rm{i}}\theta }$ where $r({\boldsymbol{q}},{\rm{\Omega }})\,={A}_{0}L\ \mathrm{sinc}\left(\tfrac{{\rm{\Delta }}k({\boldsymbol{q}},{\rm{\Omega }})L}{2}\right)$ and $\theta ({\boldsymbol{q}},{\rm{\Omega }})={\rm{\Delta }}k({\boldsymbol{q}},{\rm{\Omega }})z/2$. The real quantity r is usually referred to as the squeezing parameter. The input–output relations for mode 1 and mode 2 follows as [120]:

Equation (38)

Equation (39)

where:

Equation (40)

Equation (41)

Now we are able to calculate the mean photon number for the mode j (j = 1, 2):

Equation (42)

where we have used the unitary condition ${\hat{S}}^{\dagger }\hat{S}=1$.

It is possible to derive the statistical momenta of superior orders by following the same steps as in the previous calculation. In particular we are interested in the second order moments (normally ordered):

Equation (43)

Equation (44)

and in the variance of single modes and their covariance:

Equation (45)

Equation (46)

From equation (45) it can be seen that the single mode of the SPDC radiation has thermal statistics with a super-poissonian component equal to ${\mu }^{2}$ (excess noise).

4.3. Detected photon statistics

In the previous paragraph we derived the statistical behavior of SPDC photons emitted in two conjugated modes. Here we are interested in the statistical behavior of the detected photons $\langle {\hat{N}}_{j}\rangle $.

According to the simple detection model of section 2, equations (3), and the results of section 4.2, it easily follows that:

Equation (47)

Equation (48)

and the measured covariance is:

Equation (49)

Now it is possible to calculate the parameters that quantify the degree of correlation between two modes and, in particular, the NRF σ defined in equation (12). As described in section 3.1, the NRF allows discrimination between classical states of light and quantum states of light. If $\sigma \geqslant 1,$ we have classical light such as thermal light or coherent light if $\sigma =1$, and if $\sigma \lt 1$ we have quantum correlated light.

Substituting the statistics of two conjugated modes of SPDC, equations (45) and (46), in the definition of equation (12) we can obtain perfect correlations in the ideal lossless case i.e.:

Equation (50)

while when losses are considered, according to equation (13), we have:

Equation (51)

where we assume ${\eta }_{1}={\eta }_{2}=\eta $. For unbalanced losses, the NRF becomes:

Equation (52)

where μ is the mean number of photons per mode and $\bar{\eta }$ is the mean quantum efficiency. Equation (51) shows how the measured NRF between two conjugated modes is always smaller than 1 in the case of identical quantum efficiency. Otherwise, if we have ${\eta }_{1}\ne {\eta }_{2}$ there is an additional positive term, proportional to the mean value of photons per mode, which arise from a non-perfect cancellation of the excess noise of the thermal fluctuation. This can lead to the measure ${\sigma }_{{\rm{\det }}}\gt 1$ losing the non-classical signature, even in case of perfectly correlated quantum light.

These results are valid in the plane-wave pump approximation. In the experiments, the momentum distribution of the pump which cannot be a delta function, generates an uncertainty in the relative momentum (direction of propagation) of correlated photons. Therefore, a full study of the mode collection inside finite detection areas is needed for describing the experimental results and some issues relating to the detection of non-classical features of a multimode squeezed vacuum.

4.4. Mode collection in the far field

In the far field region, obtained at the focal plane of a thin lens in an ff configuration, any transverse mode ${\boldsymbol{q}}$ is associated with a single position ${\boldsymbol{x}}$ according to the geometric transformation $(2{cf}/\omega ){\boldsymbol{q}}\to {\boldsymbol{x}}$, where c is the speed of light. The exact condition ${{\boldsymbol{q}}}_{1}+{{\boldsymbol{q}}}_{2}=0$ for correlated photons, which comes from the integral in equation (30) in the plane wave pump approximation, becomes in the far field a strict condition on their positions, ${{\boldsymbol{x}}}_{1}+{{\boldsymbol{x}}}_{2}=0$. For degenerate frequencies, ${\omega }_{1}={\omega }_{2}={\omega }_{p}/2$, correlated photons reach symmetric positions with respect to the pump intersection point (${\boldsymbol{x}}=0$). A more realistic Gaussian distributed pump with angular spread ${\rm{\Delta }}{\boldsymbol{q}}$ leads to uncertainty on the position of the correlated photon, ${{\boldsymbol{x}}}_{1}+{{\boldsymbol{x}}}_{2}=0\pm {\rm{\Delta }}{\boldsymbol{x}}$, where ${\rm{\Delta }}{\boldsymbol{x}}\,=(2{cf}/{\omega }_{p}){\rm{\Delta }}{\boldsymbol{q}}$ represents the size, in the far field, of the coherence area ${{ \mathcal A }}_{{\rm{coh}}}$ in which it is possible to collect photons from correlated modes. It is possible to visualize coherence areas in the high gain regime ($r\gt 1$) where they appear as correlated spots (speckles) around symmetrical positions ${\boldsymbol{x}}$ and $-{\boldsymbol{x}}$, where the center of symmetry (CS) is basically the pump-detection plane interception. These correlations in photon numbers can be seen in figure 3.

Figure 3.

Figure 3. Far-field emission of TYPE II SPDC in the non-linear high gain regime, in which super-poissonian fluctuation is responsible for the speckled structure. The rings correspond to a spectral selection of 10 nm around the degeneracy.

Standard image High-resolution image

It is possible to measure the size of this coherence area by performing the spatial cross correlation between the two beams:

where $\xi =(x,y)$ is the shift.

In order to collect most of the correlated photons, two symmetrically placed detectors must have sensitive areas larger than the coherence area Acoh. Referring to figure 4, we consider two equal and symmetric areas ${{ \mathcal A }}_{{\rm{\det }},j}$ $(j=1,2)$ containing a large number of transverse spatial modes ${{ \mathcal M }}_{c}={A}_{{\rm{\det }},j}/{A}_{{\rm{coh}}}$, represented by the light blue circles in the figure. However, there are modes ${{ \mathcal M }}_{{\rm{b}}}$ on the detector border that are only partially detected, namely with efficiency β, which can be assumed to be equal to 1/2 on average. Moreover, experimental misalignment, δ, can lead to the collection of some uncorrelated modes ${{ \mathcal M }}_{{\rm{u}}}$. Even if it is possible to optimize the experiment in order to reduce the contributions of ${{ \mathcal M }}_{{\rm{b}}}$ and ${{ \mathcal M }}_{{\rm{u}}}$, it is necessary to take them into account for a complete description of the physical scenario.

Figure 4.

Figure 4. A scheme of the correlated modes ${{ \mathcal M }}_{{\rm{c}}}$, the uncorrelated modes ${{ \mathcal M }}_{{\rm{u}}}$ and the partially correlated modes ${{ \mathcal M }}_{{\rm{b}}}$, when we assume that there is a misalignment with respect to the CS indicated by the blue dot.

Standard image High-resolution image

Since each SPDC couple of modes is independent of the others, the variance and covariance of a state with ${ \mathcal M }$ pairs are ${ \mathcal M }$ times the values of a single pair. Therefore, taking into account the contribution of the different kinds of involved modes and the single/two-mode statistics in equations (47), (48), and (49) one has:

Equation (53)

Equation (54)

Equation (55)

where μ is the mean photon number per mode and ${\eta }_{1}$ and ${\eta }_{2}$ are the detection efficiencies of the two channels.

Substituting the previous expressions into the definition of the NRF in equation (12) we have (${\eta }_{1}={\eta }_{2}=\eta $):

Equation (56)

The quantity $0\lt A\lt 1$ can be interpreted as the collection efficiency of correlated photons pairs (or modes) and it assumes the form:

Equation (57)

It is possible to evaluate this collection efficiency using just some basic geometrical considerations: in figure 4 we call δ the misalignment, r the coherence radius at the detection plane (the radius of the light blue circles) and L the linear size of a detection region. Under the conditions $L\gt 2r$ and $\delta \ll L$, the counts of the different types of modes are related to the geometrical measurable parameters as:

Equation (58)

Equation (59)

Equation (60)

By introducing the dimensionless parameters $X=L/2r$ and $D=\delta /2r$, the collection efficiency becomes:

Equation (61)

Thus, in the limit $\mu \to 0$, the measured NRF in equation (56) does not depend on the mean number of photons. In the asymptotic limit $X\gg 1$, i.e. when the detection size is much larger than the correlation area, A approaches unity and the NRF reaches the value in equation (51), the one of two correlated modes in the monochromatic plane-wave pump approximation.

5. Sub shot noise absorption imaging

Absorption measurements are used in many fields of science, ranging from spectroscopy, for estimating the chemical concentration of compounds of gases and solutions using the Beer–Lambert law, astronomy, atomic and molecular physics to biological microscopy. Wide-field absorption microscopy, is the simplest, fastest, least expensive and oldest imaging modality used, for example, for live-cell imaging. It has the advantage of requiring the lowest photon dose, especially for absorption light imaging. It is recognized by biologists that the lowest photon dose should be used to probe and investigate biological processes [124], since the bright illumination can affect the regular biochemistry pathway or induce phototoxicity and damage [8]. As a drawback at low level of illumination, where few hundreds (or thousands) of photons per pixel (or frame) are collected, the photon shot noise starts to be an issue for the image quality and limits the information retrieved on the sample.

SSN absorption measurement has been demonstrated some time ago [125] using an SPDC source, and recently re-proposed with the help of modern and more efficient devices and exploiting heralded single-photon sources [126, 127]. However, these works focus on the estimation of a single value of the absorption, because only two correlated modes are exploited, see section 5.1. SSN wide field imaging (SSNWFI), where the whole spatial structure of the absorption profile is reconstructed in a single frame, requires the exploitation of many, namely thousands, of pair-wise correlated spatial modes, which must be efficiently detected by a matrix of pixels. Multimode quantum correlations generated by SPDC described in section 4, represent a valid tool for reaching SSN sensitivity in each pixel of the image [128, 129]. The next two sub-sections will provide a detailed description of SSN absorption measurements and SSNWFI, presenting the latest achievements in the field [53, 130].

5.1. Absorption measurements

In wide field absorption imaging the sample is illuminated by a probe state and the transmitted pattern is detected by the pixels of a 2D matrix, e.g. a CCD camera. The absorption coefficient α in a point of the sample, is estimated by the measurement of the photon number $\langle \hat{N}\rangle $ detected by a corresponding pixel [131]. The uncertainty is:

Equation (62)

In the following we will consider the uncertainty ${\rm{\Delta }}\alpha $ in two different measurement schemes, referred as to direct (DR) and differential (DIFF) imaging, respectively.

The DR imaging scheme is represented in figure 5(a). A single probe beam is addressed to the object and the transmitted part is collected by the detector.

Figure 5.

Figure 5. Simple sketch of the different imaging schemes. (a) DR imaging. (b) DIFF classical imaging, where classical correlated beams are generated by a balanced BS. (c) DIFF imaging with quantum correlated beams generated by SPDC.

Standard image High-resolution image

The losses due to the sample can be modeled by a BS with transmission coefficient $1-\alpha $. Referring to equation (3) of section 2 and substituting η with $1-\alpha $, the mean detected photon number is $\langle \hat{N}\rangle =(1-\alpha )\langle \hat{n}\rangle $, where $\langle \hat{n}\rangle $ is the mean number of the detected photons, as it would be without the object. The variance becomes:

Equation (63)

Substituting equation (63) in equation (62), the uncertainty in the absorption estimation for the DR imaging scheme is

Equation (64)

where F is the Fano factor defined in section 3, as it would be measured in absence of the object. For a classical probe state (lower bounded by F = 1), the sensitivity scales as ${\rm{\Delta }}{\alpha }_{{\rm{DR}}}=\sqrt{(1-\alpha )/\langle \hat{n}\rangle }$ which represents the shot-noise limit. Furthermore, from equation (64), it is evident that a probe state with non-classical statistics, i.e. a value of F smaller than the unit, allows going beyond the shot-noise limit. It can be demonstrated that a Fock state $| n\rangle $, with F = 0, allows one to reach the ultimate quantum limit in the precision of absorption estimation. A Fock state with n = 1 can be approximated experimentally by a heralding single-photon source and it has been used recently in an absorption spectroscopy experiment [127]. However, as discussed in section 3, the Fano factor is deteriorated by the detection loss η. Thus, the non-classical behavior in terms of noise reduction is lower bounded by ${F}_{{\rm{\det }}}=1-\eta $. It is important to note that splitting a single-mode beam into ${ \mathcal N }$ pixels leads to a detection probability of the order of $\eta \leqslant 1/{ \mathcal N }$ for each of them, ruling out the possibility of using a single mode for reaching SSN sensitivity in WFI. Even if sub-Poissonian light ($F\lt 1$) in single mode or few modes have been obtained, experimental complications in their generation and simultaneous detection limit their use for imaging, where higher number of non-classical spatial modes are needed, each mode addressing a single pixel. On the other hand, as we showed in section 4, the SPDC process naturally produces a pair of beams, which are (individually) spatially incoherent (containing thousands of independent spatial modes) but are locally correlated in the photon number. Even if fluctuations of a single spatial mode in one beam are super-poissonian, due to photon-number entanglement these fluctuations are perfectly replicated in the correlated mode of the second beam. This property can be applied in a DIFF imaging scheme as first proposed in [129] and described in the following.

Differential imaging exploits the correlation properties of two beams instead of one. These can be, for example, TWBs generated by SPDC as represented in figure 5(c), or a thermal beam split by a 50:50 BS, depicted in figure 5(b). The scheme is the following: one of the two beams impinges on an absorbing object, with transmittance ($1-\alpha $), before being detected. The other beam is directly detected, playing the role of reference for the noise. Assuming for simplicity the same detection losses along the two optical paths, the difference in mean photon numbers is proportional to the absorption coefficient:

Equation (65)

The variance in the photon-number difference can be expressed in terms of the Fano factor and the NRF, σ, defined in equation (12), in the absence of the object, as:

Equation (66)

Therefore, the sensitivity in a DIFF scheme can be evaluated according to equation (62), where $N\to {N}_{-}$, as:

Equation (67)

The expression of the differential classical (DC) scheme can be obtained from equation (67) by substituting $\sigma =1$. For a weakly absorbing object, $\alpha \to 0$, the term ${\alpha }^{2}(F-1)$ is very small even for super-Poissonian sources and can be neglected. Thus, the uncertainty in the DC scheme becomes ${\rm{\Delta }}{\alpha }_{{\rm{DC}}}=\sqrt{(2-\alpha )/\langle \hat{n}\rangle }$, which is a factor of $\sqrt{2}$ larger than the DR imaging for small α. The uncertainty achieved by the quantum correlations with $\sigma \lt 1$, namely ${\rm{\Delta }}{\alpha }_{{\rm{SSN}}}\,={\rm{\Delta }}{\alpha }_{{\rm{DIFF}}}(\sigma \lt 1)$ can be compared to both the direct and the classical DIFF imaging by using equation (64) and equation (67) in the relevant limits discussed before:

Equation (68)

Here, we have introduced the signal-to-noise ratio (SNR), ${SNR}=\alpha /{\rm{\Delta }}\alpha $, as an equivalent figure of merit of the measured sensitivity. From equation (68), it is clear that the advantage of quantum correlation can be quantified by the value of the non-classical parameter σ, which for the TWB is lower bounded only by the loss factor $\sigma =1-\eta $ (see section 4.3). In particular, the SSN condition, $\sigma \lt 1$, guarantees an advantage with respect to the DC scheme, while a more restrictive condition, $\sigma \lt 1/2$, is needed for the SSN scheme to beat the direct (shot-noise limited) one. This condition corresponds to the requirement of an overall loss in the detection of correlated photons smaller than 50%.

In fact one of the difficulties of the technique, when addressed to SSNWFI, is to achieve a good collection efficiency of the correlated modes in the far field without sacrificing the spatial resolution. This is due to the trade-off between the collection efficiency and the pixel size, as discussed in section 4.4.

5.2. SSNWFI: experimental results

The first experimental demonstration of the SSNWFI involving many spatial modes was given in 2010 [130]. Figure 6 presents the advantages of quantum DIFF imaging over DR and DC schemes. The weak absorbing 'π'-shaped object is hidden in the noise for both the classical imaging techniques, whereas its shape can be clearly identified in the image obtained using the SSN schemes of figure 5(c).

Figure 6.

Figure 6. Two sets of typical images taken from the experiment in [130] are shown: SSN image (left) obtained by subtracting the quantum correlated noise; DC image (middle); direct classical image (right). The pixel size is $L=480\,\mu {{\rm{m}}}^{2}$, obtained by hardware binning of the physical pixels of the CCD to fulfill the condition $L\gt 2r$. For both sets of images the mean number of photons per pixel is $\langle \hat{N}\rangle \approx 7000$.

Standard image High-resolution image

It is important to mention that, in the experiment discussed in [130], the image was obtained without any imaging lenses, basically revealing the shadow of the object placed closed to the detection plane. Thus, the resolution was not high enough for any potential application in the real world, especially in microscopy, where the technique would be naturally addressed. Moreover, the average NRF achieved was just slightly below 0.5, enough to surpass DC imaging but not sufficient to provide a real exploitable advantage with respect to DR imaging in realistic conditions.

Very recently, an important step forward was made with the realization of SSNWFI in a real microscopic configuration [53]. A noise reduction such as $\sigma =0.8$ was obtained for each pixel in a matrix of approximately 8000 pixels, and a spatial resolution of 5 μ m at the sample. This noise reduction is enough to beat DC imaging, and the resolution is sufficient for the imaging of complex structures, such as cells. Reducing the resolution by one third, allows the performance of the DR imaging scheme to be easily overcome. The trade-off between the noise reduction, represented by the value of the paramater $\sigma $, and the resolution, according to the model of the collection efficiency developed in section 4.4, is reported in figure 7. The corresponding improvement of the SNR with respect to both differential and direct shot-noise limited classical imaging schemes. The main difficulties in comparison to classical microscopy are that the imaging systems should be able to reduce the aberration without introducing any losses. Figure 8 shows the experimental image profile of the sample (a 'ϕ'-shaped few-nanometer-thick deposition with absorption coefficient $\alpha =1 \% $) at different resolution scale L ($L=d\cdot 5\,\mu {\rm{m}}$). From $L=15\mu $ (d = 3) the object starts to appear in the SSN image, while remains almost undefined in the classical images.

Figure 7.

Figure 7. Experimental NRF and SNR as a function of the resolution in the focal (object) plane L. The black dots represent the NRF. The red dots are the SNR of the SSN images normalized to the one of the DR images. For $L\geqslant 15\,\mu {\rm{m}}$ there is the advantage of the quantum protocol. Analogously, the blue series shows that the advantage of SSN imaging with respect to DC imaging is present at any spatial resolution and reaches values of more than 80%. Solid lines correspond to the quantum enhancement predicted by equation (68), when the estimated values of the NRF are considered.

Standard image High-resolution image
Figure 8.

Figure 8. The DR, DC, and SSN images are compared in each panel for the same value of spatial resolution d. The upper-right panel is the image of the object after the average over 300 shots.

Standard image High-resolution image

Finally, we mention that, different from previous proofs of principle of quantum enhanced phase-contrast microscopy exploiting NOON states (with N = 2) [2326], the SSN wide-field microscope can offer the possibility of dynamic imaging without scanning the whole sample.

6. Target detection in preponderant noise

The mainstream of quantum enhanced measurement protocols focuses on the reduction of the uncertainty below shot-noise limit (or standard quantum limit), which derives from the intrinsic quantum fluctuation of the probe beam and scales as ${(n)}^{-1/2}$, with n the mean photon number. In this context it is recognized that quantum strategies, which in the ideal case outperform classical counterparts, are highly penalized in the real world by the unavoidable decoherence processes such as noise and losses. In particular, it has been shown that in the presence of decoherence the Heisenberg limit $\propto {(n)}^{-1}$ and, in general, any chance of a more favorable scaling of the uncertainty with the photon number, cannot be achieved. Rather, the enhancement is of the form $k{(n)}^{-1/2}$ where k is a constant factor, for example $k=\sqrt{(1-\eta )/\eta }$ in the presence of a loss factor of $(1-\eta )$. From this viewpoint, it seems that there is not much to do but technologically reducing the losses and noise in the experiments as much as possible.

A completely different paradigm is the one proposed by Lloyd in 2008 [132], named quantum illumination (QI), where the goal is to provide a quantum improvement in target detection (radar-like configuration) in the presence of a strong, dominant thermal background. The goal is to discriminate between the presence (H1 hypothesis) and the absence (H0 hypothesis) of a partially reflecting target (${\eta }_{{\rm{P}}}$ being the reflection coefficient). In this case, the preponderant source of noise is not the one affecting the probe, but is given by the background. Indeed the works in [133, 134] have shown that a scheme, as represented in figure 9, where one beam from SPDC is used as a probe and a joint measurement is performed on the returned probe and the second entangled beam, delivers a 6 dB (a factor 4) improvement in the error probability exponent with respect to the best classical strategy. Further improvements can, in principle, be obtained by using photon-subtracted two-mode-squeezed states [135], although their production is experimentally extremely challenging. Two outstanding features of quantum illumination are that its advantage does not depend either on losses or on the noise the probe experiences during the propagation and interaction with the target. It is important to note that both these processes cause decoherence and therefore the initial entanglement is completely lost at the detection stage. This property is very valuable, since it represents the first example of a quantum protocol robust to noise and losses.

Figure 9.

Figure 9. Scheme of the QI protocol proposed in [133], whose aim is to establish the presence of a target. One beam of the TWB is used as a reference, while the other, dubbed a probe, interacts with the target, if present. The reflected part of the probe mixes with a strong thermal background and goes to a detector, where a joint measurement is performed with the reference beam. ${\eta }_{{\rm{R}}}$ models losses on the reference path.

Standard image High-resolution image

The optimal classical illumination is known to be a product of ${ \mathcal K }$ identical coherent states $| \alpha \rangle $ and a homodyne-detection receiver [133, 136]. Homodyne detection measures the quadratures of the incoming field, in particular $\hat{x}=(\hat{a}+{\hat{a}}^{\dagger })/2$. In this case $\langle \hat{x}\rangle =\langle {\hat{x}}_{{\rm{P}}}\rangle +\langle {\hat{x}}_{{\rm{B}}}\rangle $, where $\langle {\hat{x}}_{{\rm{P}}}\rangle =\sqrt{{\eta }_{{\rm{P}}}n}$ ($n=| \alpha {| }^{2}$) is the quadrature of the coherent probe after the object interaction and $\langle {\hat{x}}_{{\rm{B}}}\rangle =0$ is the quadrature of the thermal noise, which has zero-mean value. In the limit ${n}_{{\rm{B}}}\gg n$ the noise of the measurement is dominated by the noise on $\langle {\hat{x}}_{{\rm{B}}}\rangle $, equal to $\langle {\hat{x}}_{{\rm{B}}}^{2}\rangle =(2{n}_{{\rm{B}}}+1)/4$ . The SNR in the discrimination of the object presence is therefore:

As mentioned, entanglement, in particular the multimode SPDC state (see for example equation (36)), provides an advantage of a factor of 4 in the exponent of the error probability, which is proportional to the SNR [137]. The structure of the optimal 6 dB enhancement receiver is not known, however, sub-optimal receivers with 3dB advantage has already been proposed and realized. They can be based on a non-linear interferometer, in particular a phase-sensitive low-gain ($G-1\gg 1$) optical parametric amplifier (OPA). The idea is that the OPA output depends on the phase relation between the returning probe and the reference beam, while a completely dephased thermal beam does not. This has enabled the experimental demonstration of the advantage of QI both in the detection of a low reflection phase object [137] (a shift of the probe phase of 0 (π) corresponding to the H1(H0) hypotheses, respectively), and for defeating a passive eavesdropping attack in quantum communication [138, 139]. In particular, the difference between the output signal in the two cases is proportional to the so-called phase-sensitive cross correlation $\langle {\hat{a}}_{1}{\hat{a}}_{2}\rangle $ between the signal and idler fields, which for the two-mode squeezed state is $\sqrt{n(n+1)}$, largely exceeding the classical limit of correlation for a source with the same mean photon number n, in the limit $n\ll 1$. Recently a microwave/optical QI proposal was also reported [140]. Two electro-optomechanical converters were used to entangle a microwave signal, which was sent to the target region and an optical field was retained at the source. The microwave radiation reflected by the target was then phase conjugated and upconverted into a second optical field that was jointly detected with the retained one.

Both the quantum sub-optimal and classical optimal receiver described previously are phase-sensitive measurements, requiring the probe to arrive at the receiver with a precise, unperturbed phase relation with a local oscillator or/and the reference beam. This may not be practicable in many contexts, also because it requires a precise mode matching at the receiver. Moreover, a quantum memory is needed, for example realized by an adjustable optical delay line (difficult to make if the distance of the object is not known a priori), to store the reference beam while the probe is propagating forth and back from the target object.

On the other hand, in [45, 141] a version has been proposed of QI considering a restricted scenario in which only intensity measurements (phase-insensitive) are exploited. The scheme is the one of figure 10. Here a photon-number measurement was performed independently in the reference arm and in the probe arm, then the covariance of the two quantities was evaluated. Another difference with respect to the scheme of [133] is that the background field is not necessarily mixed to a BS with the probe but, more realistically, independently reaches the detector. It is important to highlight that in this specific framework, even the classical benchmark is different with respect to the optimal one obtained in the more general context using homodyne detection. Similarly the quantum strategy cannot aim at achieving the optimal bounds of [133]. However, also in the contest where only intensity measurements are allowed, the quantum protocol maintains most of the appealing features of the original idea, such as a huge quantum enhancement under similar conditions, $n\ll 1$ and ${n}_{{\rm{B}}}\gg 1$, and a robustness against noise and losses. Moreover, even in this case, the advantage surprisingly survives when the quantumness at the detection state is broken. As we will show in detail in the next section, the SNR improvement provided by exploitation of quantum correlation in the SPDC state with respect to the classical benchmark of a direct measurement of the mean photon number of the returned probe is ${\eta }_{{\rm{R}}}/\sqrt{n}$, where, in this context, ${\eta }_{{\rm{R}}}$ represents the losses on the reference channel. Moreover, introducing a further limitation, which is that a measurement of the background alone is not possible, i.e. the background and the reflected probe always come together at the receiver, the best classical strategy cannot be the direct measurement while it is arguably the use of classical correlations. In this case the quantum advantage scales as $M/n=1/\mu $, where M is the number of identical and undistinguished modes collected in the single measurement and μ is the mean number of photons in each mode. Interestingly, this corresponds to the ratio of the total mutual information of classical and quantum correlated states [142].

Figure 10.

Figure 10. Scheme of the QI protocol proposed in [45], where only intensity measurements are performed. As in figure 9 one beam of the TWB is used as reference, while the other is the probe and they interact with the target if it is present. The beams are collected by two detectors: n1 is the number of the n photons of the reference beam that arrive at the first detector, n2 is the sum of the photons eventually reflected from the target and the photons from the thermal background, nB.

Standard image High-resolution image

In the spirit of this review, which explores the non-classical photon-number correlation and their application in quantum enhanced measurement, in the next section we will describe in detail the realization of the QI protocol based on photon-number/intensity correlation measurement.

6.1. Intensity-correlation based QI

Let us consider first the direct scheme in which a probe with mean photon number n is addressed to the object and its reflected (transmitted) part reaches a detector together with a much stronger background with mean photon number nB. Note that in this case only one beam is used. In hypothesis H0 only the background reaches the detector and the measured photon number is nB, while for H1 one has ${\eta }_{{\rm{P}}}n+{n}_{{\rm{B}}}$, their difference being the signal. The variance of the measurement is assumed to be dominated in both cases by the background fluctuations $\langle {\delta }^{2}{\hat{n}}_{{\rm{B}}}\rangle $. Thus, the SNR obtained by ${ \mathcal K }$ measurements is:

Equation (69)

Even if the strategy described above seems the simplest and the most natural approach, it assumes implicitly that it is possible to have a separate estimation of the mean photon number of the background, for example by a measurement made in the absence of the object. If the background cannot be measured separately, for example because the object is constantly present (of course this information is not available a priori), the previous method, simply based on the discrimination of two average intensity levels cannot be applied. A second-order measurement of the intensity is required instead, therefore we consider $\langle \delta {\hat{n}}_{1}\delta {\hat{n}}_{2}\rangle $, where $\delta {\hat{n}}_{1}$ and $\delta {\hat{n}}_{2}$ are the fluctuations on the reference and 'probe+noise' beams, respectively. In the absence of the target, the background and reference are uncorrelated, thus the covariance $\langle \delta {\hat{n}}_{1}\delta {\hat{n}}_{2}{\rangle }_{{H}_{0}}$ is null and it establishes the natural zero-offset for the measurement. In the presence of the target $\langle \delta {\hat{n}}_{1}\delta {\hat{n}}_{2}{\rangle }_{{H}_{1}}$ is, in general, different from zero and depends on the exploited state, i.e. on the correlations in the photon-number fluctuations between the reference and the reflected probe beam. In order to calculate the SNR it is necessary to also consider the uncertainty of the covariance. The fluctuation of this quantity is, by definition, for i = 0, 1:

Equation (70)

As before, if we consider the fluctuation in n2 dominated by the background, both in the presence and in the absence of the target, i.e. $\delta {n}_{2}{| }_{{H}_{1}}\approx \delta {n}_{2}{| }_{{H}_{0}}=\delta {n}_{B}$, it immediately follows that $\langle {\delta }^{2}(\delta {\hat{n}}_{1}\delta {\hat{n}}_{2})\rangle =\langle {\delta }^{2}{\hat{n}}_{1}\rangle \langle {\delta }^{2}{\hat{n}}_{{\rm{B}}}\rangle $. Exploiting the quantum correlation in photon-number fluctuations of the TWB state, according to section 4.3, $\langle \delta {\hat{n}}_{1}\delta {\hat{n}}_{2}{\rangle }_{{H}_{1}}\,=n{\eta }_{p}{\eta }_{R}(1+n/M)$, where M is the number of modes. Since each beam of the TWBs is multithermal (with M modes): $\langle {\delta }^{2}{\hat{n}}_{1}\rangle \,=n{\eta }_{{\rm{R}}}(1+{\eta }_{{\rm{R}}}n/M)$, the SNR can be evaluated as:

Equation (71)

Equation (72)

where the last approximation holds for $n/M\ll 1$, i.e. when the mean photon number per mode $n/M=\mu $ is small. We can now compare this result with the SNR obtained with the direct measurement of the probe mean photon number (when this is possible). This results in an improvement for $n\lt 1$ as large as ${{SNR}}_{{\rm{SPDC}}}/{{SNR}}_{{\rm{Dr}}}={({\eta }_{{\rm{R}}}/n)}^{1/2}$.

It is also interesting to evaluate the advantage of the quantum correlation with respect to the possible use of classically correlated states. The first line of equation (71) shows that the classical and quantum schemes with the same local statistics only differ for the strength of the correlation, quantified by the covariance $\langle \delta {\hat{n}}_{1}\delta {\hat{n}}_{2}{\rangle }_{{H}_{1}}$. According to the generalized Cauchy–Schwarz inequality presented in section 3, the covariance for classical beams is bounded by $\varepsilon =\langle \delta {\hat{n}}_{1}\delta {\hat{n}}_{2}\rangle \,/{(\langle :{\delta }^{2}{\hat{n}}_{1}:\rangle \langle :{\delta }^{2}{\hat{n}}_{2}:\rangle )}^{1/2}\leqslant 1$. Split thermal beams saturate the inequality, ${\varepsilon }_{{\rm{TH}}}=1$, with $\langle \delta {\hat{n}}_{1}\delta {\hat{n}}_{2}{\rangle }_{{\rm{TH}}}\,={\eta }_{{\rm{R}}}{\eta }_{{\rm{P}}}{n}^{2}/M$, thus representing the best classical strategy. On the other hand the SPDC quantum correlation provides ${\varepsilon }_{{\rm{SPDC}}}=M/n+1$ with $\langle \delta {\hat{n}}_{1}\delta {\hat{n}}_{2}{\rangle }_{{\rm{SPDC}}}={\eta }_{{\rm{P}}}{\eta }_{{\rm{R}}}n(1+n/M)$. Therefore, the comparison of the SNR with classical and quantum correlation immediately gives:

Equation (73)

A dramatic quantum enhancement for a photon number per mode $\mu =n/M\ll 1$ is evident. It is important to notice that, as anticipated, the enhancement does not depend on the background intensity and it is also immune to the losses.

Finally we would like to trace a connection between the QI using the OPA receiver of [137] and the intensity-measurement-based scenario described above, showing that they have the same non-classicality/entanglement breaking condition. Indeed, for a zero-mean Gaussian distributed bipartite state, the moment-factoring theorem allows writing the photon-number covariance in terms of the modulus of the phase-sensitive cross correlation (which is the quantity measured by the OPA receiver in [137]): $\langle \delta {\hat{n}}_{1}\delta {\hat{n}}_{2}\rangle =| \langle {\hat{a}}_{1}{\hat{a}}_{2}\rangle {| }^{2}$. On the other hand the normal ordered variance for a gaussian mode can be written in terms of the mean photon number: $\langle :{\delta }^{2}{\hat{n}}_{j}:\rangle =\langle {\hat{a}}_{j}^{\dagger }{\hat{a}}_{j}{\rangle }^{2}$. Therefore, the non-classicality breaking condition, represented in general by the violation of the Cauchy–Schwarz inequality, in the framework of Gaussian states, coincides with the entanglement breaking condition $| \langle {\hat{a}}_{1}{\hat{a}}_{2}\rangle {| }^{2}\leqslant \langle {\hat{n}}_{1}\rangle \langle {\hat{n}}_{2}\rangle $ reported in [137], valid for two conjugated modes. Substituting in the Cauchy–Schwarz inequality the explicit expression of the photon statistics at the detectors, in the general multimode case, the condition becomes:

Equation (74)

In the limit of $\mu =n/M\ll 1$ the condition simplifies as ${n}_{{\rm{B}}}\geqslant {\eta }_{{\rm{P}}}{({{MM}}_{{\rm{B}}})}^{1/2}$. For example when single modes are detected $M={M}_{{\rm{B}}}=1$, a mean number of background photons ${n}_{{\rm{B}}}\gt 1$ is enough to destroy Gaussian entanglement and more in general non-classical photon statistics; nevertheless the enhancement in the SNR remains.

6.2. Experimental implementation of QI

The experimental setup used in [45] for the realization of the intensity-correlation-based QI protocol is represented in figure 11(a). Type II SPDC generates pairs of correlated 5 ns pulses with an average number of photons per spatio-temporal mode of $\mu \sim 0.1$, which are then addressed to a high quantum efficiency CCD camera. In the QI protocol (figure 11(a)) one beam (the reference) is directly detected, while a target object (a 50:50 BS) is placed in the path of the other one (the probe), where it is superimposed with a pseudothermal background produced by a laser beam scattered by an Arecchi rotating ground glass. When the object is removed, only the background reaches the detector. The CCD camera detects, in different regions, both the optical paths. In the classical illumination (CI) protocol (figure 11(b)), the TWBs are substituted with classical correlated beams, obtained by splitting a single arm of SPDC, that is a multithermal beam, and by adjusting the pump intensity to ensure the same local statistics and spatial coherence properties for the quantum and the classical source.

Figure 11.

Figure 11. Experimental setup and examples of acquired frames as presented in [45]. (a) QI, (b) CI, (c) detected TWBs, in the presence of the object, without thermal bath. The region of interest is selected by an interference filter centered around the degeneracy wavelength (710nm) and bandwidth of 10 nm. After selection the filter is removed. (d) Detected field for split thermal beams in the presence of the object, without thermal bath. (e) A typical frame used for the measurement, where the interference filter has been removed and a strong thermal bath has been added to the object branch. The color scales on the right correspond to the number of photons per pixel. Reprinted figure with permission from [45], Copyright 2013 by the American Physical Society.

Standard image High-resolution image

In this scheme, n1 and n2 are the photon numbers detected by pairs of spatially correlated pixels in a single 5 ns shot of the pump laser, as presented in figures 11 (c)–(e). Since K = 80 correlated pixel pairs are present, it is possible to perform spatial statistics, which allows the evaluation of the covariance $\langle \delta {\hat{n}}_{1}\delta {\hat{n}}_{2}\rangle $ in a single shot, thus reducing the measurement time needed for asserting the presence or the absence of the target.

Figure 12 reports the measured ε versus the theoretical prediction. It can be seen that for the TWB ${\varepsilon }_{{\rm{QI}}}$ is in the quantum regime (${\varepsilon }_{{\rm{QI}}}\gt 1$) for small intensities of the thermal background, reaching the value ${\varepsilon }_{{\rm{QI}}}\simeq 10$ when ${n}_{{\rm{B}}}=0$. It rapidly decreases below the classical threshold according to the condition in equation (74) when the background increases. For classical correlation of split thermal beams, ${\varepsilon }_{{\rm{TH}}}$ is always in the classical regime, starting from ${\varepsilon }_{{\rm{TH}}}=1$ for ${n}_{{\rm{B}}}=0$, as expected.

Figure 12.

Figure 12. Generalized Cauchy–Schwarz parameter ε in the case of QI exploiting TWB state from SPDC source, ${\varepsilon }_{{\rm{QI}}}$, and for the correlated thermal beams, ${\varepsilon }_{{\rm{TH}}}$, as a function of the average number of background photons nB (with ${M}_{{\rm{B}}}=1300$). The solid lines represent the theoretical prediction for the estimated value of the mean photon number per mode $\mu =0.075$.

Standard image High-resolution image

In figure 13 an experimental comparison of the SNR for quantum and classical illumination is presented. While the SNR unavoidably decreases when the noise increases for both QI and CI (see equation (71)), the ratio between them is constant regardless of the value of nB, in agreement with the theoretical prediction provided in equation (71), ${{SNR}}_{{\rm{SPDC}}}/{{SNR}}_{{\rm{TH}}}={\varepsilon }_{{\rm{SPDC}}}\simeq 10$. In turn, the measurement time, i.e. the number of repetitions ${ \mathcal K }$ needed for discriminating the presence/absence of the target, is dramatically reduced by 100 times when quantum correlations are exploited.

Figure 13.

Figure 13. SNR versus the number of background photons $\langle {n}_{{\rm{B}}}\rangle $ normalized by the square root of K (number of correlated pixel pairs in a single-shot image). The red and black markers refer to QI and CI, respectively. The curves correspond to the theoretical model. Each experimental point is extracted from a statistic over a set of 6000 shots.

Standard image High-resolution image

7. Ghost imaging (GI)

GI is an imaging technique theoretically proposed in 1994 [54] and experimentally realized in 1995 by Pittman et al [55], using non-classical states of light. Since then, this technique has attracted great interest [143152, 154] for the wide field of its possible applications and many GI schemes have been investigated accordingly [155167].

The aim of this protocol is to retrieve the transmittance profile of an unknown object without a direct spatially resolved measurement. To perform GI, two beams, whose intensity fluctuations are correlated, are used. As shown in figure 14 the first beam (beam 1), without interacting with the object, illuminates a spatial resolving detector, like a camera. The second beam (beam 2), after the interaction with the object, is sent to a bucket detector without spatial resolution (e.g. a single-pixel photodetector). The procedure is repeated and ${ \mathcal K }$ frames of the camera, in correspondence with the ${ \mathcal K }$ signals of the bucket detector, are collected. It is not possible to obtain the image of the object through the signal from detectors 1 or 2 separately, since the first one has not interacted with the object, while the other has no spatial resolution. Anyway, as we will see, correlating the signals from the two detectors makes it possible to retrieve the image.

Figure 14.

Figure 14. GI schematic representation: two beams, beam 1 and beam 2, whose intensity fluctuations are correlated, are sent to two distinct optical paths: one containing a spatial resolving detector ${\text{}}1$, and the other one containing the object to be imaged and a bucket detector ${\text{}}2$. The image of the sample is retrieved by correlating the output of the two detectors.

Standard image High-resolution image

In the first experimental realization of GI, SPDC correlated photon pairs were used, measuring their coincidences by single-photon detectors. Nevertheless, later it was shown, both theoretically and experimentally, that split thermal light can also be used to perform GI, although with a smaller visibility [145150], as well as intense TWBs [151]. In [152] it is shown that even sunlight can be used; an interesting result in view of future practical applications.

These results started an intense debate and a lot of work was addressed to understanding the differences between GI using classical (i.e. split thermal light) or quantum (i.e. TWB state) light and to establish the usefulness of quantum resources, in particular entanglement. To clarify the boundary between classical and quantum GI, different configurations were implemented and various measurements considered in order to find any evidence that clearly distinguishes between the two cases. An exhaustive discussion about the 'quantumness' of GI is presented in [154]. In that review, competing interpretations of this technique are unified in a unique theoretical frame and misunderstandings about the role of entanglement are clarified. In particular, the equivalence of the interpretations in terms of intensity-fluctuation correlations and two-photon interference are shown, both for pseudothermal and PDC light. It follows that experiments cannot distinguish between these two interpretations and therefore any GI experiment can be reproduced both with classical and quantum light. The only difference between these two schemes is in terms of visibility [56], or better SNR, for an equal number of measurements [57]. This is a consequence of the stronger correlations present in TWBs and, in low-illumination conditions, this enhancement becomes important. We note that the origin of this advantage is the same at the base of the quantum enhanced target detection protocol described in section 6.1. In this work we focus on this aspect, presenting a simple theoretical model of GI, evaluating the SNR in the quantum and classical cases and discussing the quantum enhancement in different regimes.

Before going into the details of the GI technique, let us review some of its possible applications and recent developments in order to appreciate how this technique offers important opportunities in a lot of different fields.

Since the image is retrieved from beam 1, which does not interact with the sample, this method can be extremely useful in the presence of phase distortions on beam 2. This means that GI is particular interesting in the presence of an object in a diffusive medium, a condition that appears in several significant cases (such as open-air conditions or biological samples, where tissues represent the diffusive medium). Several works have analyzed the performance of GI in turbid media, among others [155157]. GI performs better than in standard non-correlated DR imaging when the turbulence is between the sample and the bucket detector, while in [158] it is experimentally demonstrated that turbulence affects GI if it is between the source and the object. In the same article a possible solution for diminishing the effect of turbulence slightly changing the GI apparatus is present. A concise but exhaustive theoretical treatment of turbulence and other aspects of non-ideality is also presented in [153].

In addition, GI can be useful in particular experimental conditions, for example if the accessible volume in the proximity of the sample is limited: in this case the light beam interacting with the object can be collected simply with a single-pixel detector as an optical fiber connected with a photodiode. This possibility is applied in [159] to magneto-optical imaging by Faraday microscopy, where magnetic samples are usually embedded in a small cryostat, with intense magnetic fields generated by superconducting magnets. In this case the basic GI setup is opportunely modified by inserting a polarizer in front of the bucket detector.

Using conventional GI, then, it is possible to retrieve the image of an object from reflected photons instead of the transmitted ones. This protocol was experimentally realized in [160] and can offer interesting opportunities; in particular GI in reflection could find application as an alternative to the conventional laser radar for standoff sensing. To this aim, in [161] the vulnerability of reflective GI to atmospheric turbulence is studied.

Another possible GI configuration is so-called 'computational GI'. In conventional GI with (pseudo) thermal light the two beams are usually obtained by sending a laser beam to a time-varying (rotating ground-glass) diffuser and then to a BS. In [162] it was argued that the ground-glass diffuser can be replaced with a programmable spatial light modulator (SLM) and even a single beam and a single-pixel detector is sufficient for GI. Applying deterministic modulation to the SLM and then correlating this precomputed modulation, opportunely processed, to the output of the bucket detector it is possible to retrieve the image of the object. Notice that, in this case, only the bucket detector is used. Computational GI was experimentally implemented by Bromberg et al [163]. The same authors further developed this technique introducing the compressive GI method [164], also used later in [161].

Different works, for example [165167], have explored the possibility of the so-called 'two-wavelength GI', taking advantage of the SPDC peculiarity to generate correlated beams even with very different optical frequencies, providing the energy conservation in equation (19). Performing GI with beams at a significantly different wavelengths can offer advantages: on the one hand high spatial resolving and/or efficient detectors are not available at all wavelengths, on the other hand atmospheric turbulence and scattering effects strongly depend on the wavelength. Therefore one can chose a suitable wavelength for the spatial resolving detector operating in the reference protected channel and the most appropriate one for open-air propagation through turbulence or scattering media. Note that split thermal beams do not offer this possibility.

In conclusion, several applications and extensions have been proposed. The list provided above is far from being exhaustive. For instance, the use of higher-order correlations to form ghost images and the use of homodyne detection instead of direct detection has also been considered [149, 168, 169].

In the following we will focus on the quantum enhancement provided by TWBs.

Quantum correlations are particularly effective at low-illumination level. In [170] the authors obtained a high-quality image of an object using less than 0.5 photons per pixel exploiting by down-converted photons pairs from SPDC To achieve this result a GI-like protocol has been implemented, in which an intensified CCD camera (ICCD) is gated by the bucket detector counts. Hence, a photon is measured by the resolving detector only if its correlated one hits the bucket detector. To improve the quality of the image a post-processing reconstruction technique is applied, in particular exploiting the natural sparsity in the spatial frequency domain of typical images and the Poissonian nature of the noise on the experimental data. This method has been tested on a biological sample (a wasp wing); as a matter of fact, biological imaging could be one of the most important applications of imaging at low-illumination level since in this case samples can be sensitive to high fluxes. Developing new techniques in this direction is therefore of extreme interest.

7.1. Theory of conventional GI

A scheme of a conventional GI technique experimental setup is shown in figure 14. The image of the object is retrieved by measuring a certain function $S({x}_{j})$, where xj is the position of the pixel j of the resolving detector, in arm ${\text{}}1$. In general $S({x}_{j})$ involves the correlation functions of the output of the two detectors:

Equation (75)

where ${{\mathbb{N}}}_{2}$ is the total number of photons collected at the bucket detector and ${N}_{1}({x}_{j})$ is the number of photons collected in the jth pixel of the resolving detector.

Experimentally these quantities are evaluated averaging the number of acquisitions ${ \mathcal K }$: $E[X]=\tfrac{1}{{ \mathcal K }}{\sum }_{k=1}^{{ \mathcal K }}{X}_{k}$.

The ghost image can be retrieved by exploiting different GI protocols, namely, different expressions for $S({x}_{j})$ [171]. We focus on the covariance between the two outputs (note that the covariance between the outputs has also been considered in the QI protocol based on intensity correlations described in section 6.1):

Equation (76)

Here, we consider that the portion of beam 1 detected by the pixel in ${x}_{j}^{(1)}$ is locally correlated only with the corresponding portion of beam 2 at the object plane position ${x}_{j}^{(2)}$. This can be obtained, for example, if the point-to-point far-field correlations of SPDC (described in section 4.4) are imaged in beam 1 at the detection plane, while in beam 2 at the object plane. In any case, pairs of correlated spatial modes in split pseudothermal beams can be likewise used in GI experiments. Hereinafter we will omit the suffixes 1 and 2.

In the following we compare spatially incoherent, locally correlated pseudothermal beams and TWB states. The first state is usually obtained by splitting a single pseudothermal beam through a BS. In this case it holds (derived from equations (47) and (48) considering M independent modes and with the substitution ${\eta }_{j}\to {T}_{j}$):

Equation (77)

Equation (78)

where μ is the mean number of photons per mode, M is the number of modes detected by each pixel and Tj is the transmission coefficient in correspondence with pixel j, which can also include the detection efficiency. For $M\gg \langle \hat{N}({x}_{j})\rangle $ ($\mu \ll 1$) we have ${\langle {\delta }^{2}\hat{N}({x}_{j})\rangle }_{{\rm{TH}}}={\langle \hat{N}({x}_{j})\rangle }_{{\rm{TH}}}$: in this limit thermal light can be described by a Poisson distribution, hence approaching the shot noise.

As described in section 4, a TWB state is a quantum state of light that can be produced by the non-linear optical phenomenon of SPDC and presents perfect correlation in photon-number fluctuation. This perfect correlation is intrinsically quantum. The single beam fluctuations follow the same thermal statistics of equation (77) and equation (78).

The difference between split thermal light and TWB states arises when considering the expressions for the covariance between photon-number fluctuations in the two beams (derived from the two-modes equation (49) considering here the contribution of M pairs of independent modes):

Equation (79)

Equation (80)

where the Kronecker delta function ${\delta }_{i,j}$ takes into account that only pairs of positions in the two beams are correlated. The transmission coefficient T1 on channel 1 is considered uniform over the spatial resolving detector. The two statistics become asymptotically identical for $\mu \gg 1$ with a dependence $\sim {\mu }^{2}$, whereas for a small number of photons per mode ($\mu \ll 1$) TWB scales more favorably, proportional to $\sim \mu $. This means that TWB, even the shot-noise component of the fluctuations, proportional to μ, is correlated among the two beams: this is the basis of quantum enhancement.

It is now evident that, by measuring locally the covariance of each pixel with the bucket detector, it is possible to retrieve the object transmittance profile ${T}_{2}({x}_{i})$. Here, we report the demonstration for classical GI but the principle is the same for quantum GI. In fact, writing the bucket detector signal as ${{\mathbb{N}}}_{2}={\sum }_{i}{\hat{N}}_{2}({x}_{i})$ and then using equation (79), one has:

Equation (81)

7.2. Thermal GI and quantum GI performance

In order to quantify the quality of the reconstructed image of the object and to compare quantum and classical GI we consider the SNR. For the sake of simplicity we assume the object to be characterized by two levels of transmission, the lower one $0\leqslant {T}_{2-}\leqslant 1$, and the higher one $0\leqslant {T}_{2+}\leqslant 1$. Correspondingly, Rm is the number of pixels that are locally correlated with object points of transmission T2m ($m=+,-$). The SNR, which represents the possibility of discerning between the two transmission coefficients, is:

Equation (82)

where Sm, $m=+,-$, is the value of the correlation function (for example the covariance in equation (79)–(80)) in correspondence with T2m. Experimentally, both the mean value, $\langle {S}_{m}\rangle $, and its variance ${\delta }^{2}{S}_{m}$ can be estimated by performing spatial averages over the regions + and − of the reconstructed image. The theoretical values of these quantities depend on the state of the light used. For evaluating $\langle {S}_{+}\rangle $ and $\langle {S}_{-}\rangle $, we use equations (79) and (80) for the classical and quantum cases, respectively. Similar to what was done in QI in equation (70) the variance of these quantities are obtained by:

Equation (83)

The first equality in equation (83) is the definition of variance, while the second one holds under the hypothesis of ${R}_{m}\gg 1$; in this case the uncorrelated components dominate. Considering equation (78) and that:

Equation (84)

we have:

Equation (85)

It is important to note that this expression is the same both in the thermal and quantum cases; this is a consequence of the thermal nature of the single beam of a TWB state.

From the definition of SNR (equation (82)):

Equation (86)

Equation (87)

As expected the SNR increases with the number of acquisitions ${ \mathcal K }$ (namely the total number of frames collected by the spatial resolving detector), while the dependence on other parameters is more complex. For a better understanding of these expressions, the relevant physical limits are analyzed.

  • Let us consider the case of $\mu \gg 1$, which is the situation of a high number of photons per mode. In order to further simplify the expressions we also consider the case of ${T}_{2+}=1$ and ${T}_{2-}=0$ (the detector detects all the photons that do not hit the object, modeled as perfectly absorbing mask):
    Equation (88)
    Equation (89)
    In this limit the same expressions are found both in the classical or quantum case. This result is not surprising and comes from the fact that for $\mu \gg 1$ the expressions for the covariances converge asymptotically. Looking at equation (88) it results that in this limit the SNR does not depend on the transmittance on channel ${\text{}}1$, T1.
  • In the opposite case, for $\mu \ll 1$, the expressions become:
    Equation (90)
    Equation (91)
    Equation (92)
    In this limit the difference between the two cases is evident: while in the classical case the SNR decreases proportionally with μ and approaches 0 as soon as $\mu \longrightarrow 0$, in the quantum case the SNR converges to a constant value. It is important to notice that this constant depends on ${ \mathcal K }$, and can therefore be arbitrary increased with a longer acquisition time experiment.

Moreover, in both regimes considering ${T}_{2+}=1$ and ${T}_{2-}=0$, ${SNR}\propto \tfrac{1}{\sqrt{{R}_{+}}}$. For a fixed total area, a greater ${R}_{+}$ implies a smaller ${R}_{-}$, hence a small object; this result therefore explains the reasonable fact that it is more difficult to image a small object. We also recall that all these expressions for the SNR are obtained in the limit ${R}_{+}\gg 1$.

To conclude the comparison between classical and quantum GI we consider the ratio, G, of the two SNRs. From the exact expressions in equations (86) and (87), it follows that:

Equation (93)

  • For $\mu \gg 1$, $G\longrightarrow 1$: in this regime, the quantum and classical case are equivalent in terms of SNR. The quantum enhancement is in this case negligible.
  • For $\mu \ll 1$, $G\longrightarrow \infty $: this is the regime where the quantum enhancement is more important. Only using TWB states it is possible to retrieve the object profile.

As we pointed out in the introduction the analogy between GI and QI is confirmed by equation (93), which is exactly the same as the one in equation (73).

Sometimes, in practical situations it could be helpful to compare the SNR for quantum and classical GI, as a function of the detected photons instead of the photon per mode emitted by the source. For this purpose we consider for example the detected photons per pixel of the spatially resolved detector $\langle {\hat{N}}_{1}\rangle ={T}_{1}M\mu $. This can be the quantity of interest if there is a strong limitation on the total photon that can be used per acquisition time, for example in the case of a detector with a low level of saturation, or to not exceed some damage-level of a photosensitive sample. In terms of $\langle {\hat{N}}_{1}\rangle $ equation (93) becomes:

Equation (94)

A clear advantage of the quantum GI appears when less than one photon per pixel is detected (for ideal efficiency, ${T}_{1}=1$). However, the higher the number of spatio-temporal modes M collected by each pixel, the higher the quantum enhancement effect.

Here, we have presented the derivation of the SNR for the specific case in which the covariance of the bucket and the spatial resolving detector is used for image reconstruction. Different correlation functions can be used and can be more advantageous in a particular case, for example in the case of a slightly absorbing object [172]. The case treated here is sufficient for the purpose of this review, which is to identify the regimes in which quantum light is advantageous and quantifying this effect.

This consideration paves the way for a lot of interesting applications of quantum GI in situations where a low light level is needed, like in the case of imaging of certain biological samples.

Finally, we note that similarities between GI and QI performed using intensity measurements (described in section 6.1) arise. Comparing figures 10 and 14 the analogy is evident. In particular GI can be seen as a specific case of QI intensity protocol, where the background at the bucket detector comes from those spatial modes of the source which are not correlated with the single pixel of the spatial resolving detector. Of course, in GI the spatial resolution of the reference arm allows a full reconstruction of the object transmission profile, while the QI goal is just discriminating its presence. Despite this difference, the quantum enhancement in terms of SNR assumes the same form.

8. Detector absolute calibration

Quantum correlations find a special application in the field of quantum radiometry [173]. Specifically, they allow the estimation of the quantum efficiency of a photon detector in an absolute way, i.e. without comparison with pre-calibrated devices or standards.

The quantum efficiency η represents one of the most important figures of merit for photon detectors and it is defined in the most general case as the overall probability of detecting a single photon impinging on the detector, in other words it is the loss component, which can be exclusively ascribed to the detector (see chapter 2).

Absolute techniques for quantum efficiency become fundamental at low illumination regimes (i.e. single or few photons), where it is difficult to provide a metrological traceability to the standards and units that are usually developed for macroscopic quantities. Indeed, currently, at the single/few photon level there are no absolute detectors (detectors with predictable quantum efficiency) or standard sources (deterministic single-photon sources) that have stability and accuracy suitable for metrological purposes. However, calibrated detectors at the level of single or a few photons are fundamental for the increase in quantum technologies exploiting quantum states of light, such as quantum computation [174], quantum key distribution [175] quantum imaging [15], and in fundamental tests of quantum mechanics, for instance to ensure that Bell's inequalities are violation free from the fair-sampling assumption [117, 118].

Quantum photon-number correlations offer the possibility of absolute calibration methods of single-photon detectors and in this section we review the most significant aspects of this field. Moreover, some alternative techniques aimed at absolute characterization have been recently demonstrated, based on different input states such as squeezed light [176] or coherent states [177]. For completeness, we mention that quantum efficiency does not represent the whole behavior of a detector, which can be described completely by a set of measurement operators known as the positive operator-valued measure (POVM). The POVM reconstruction has been realized in many experiments [178184], even if they can not be considered absolute techniques since they exploit intense calibrated sources and calibrated attenuators to provide well-known and controlled input states.

In order to review the calibration techniques exploiting quantum correlation, it is useful to divide the light detectors into two main categories: analogical detectors and single-photon detectors. Analogical detectors provide a signal proportional to the radiant flux impinging on the sensor; normally they are not able to detect single photons due to the high noise level and are designed to work at medium/high light intensity. By contrast, single-photon detectors have a resolution that allows the discrimination of single photons, but usually they are limited to working at low light intensity due to saturation effects. We can further divide the single-photon detectors into click/no-click detectors and photon number resolving (PNR) detectors. The former can only discriminate between zero photons detected (no-click) and one or more photons detected (click) in a time window depending on the detector characteristics. PNR detectors are able to provide the number of impinging photons as they arrive simultaneously [185].

8.1. Klyshko's method for absolute calibration of single-photon detectors

The first calibration method exploiting quantum correlations was proposed in the seventies by Klyshko [98, 186], but only in the nineties did the technological development allow accurate calibrations to be performed [187]. Nowadays, Klyshko's calibration technique is recognized as a fundamental metrological tool by the international radiometric community [60, 188194].

The Klyshko method is based on the SPDC process described in section 4. Photons emitted by SPDC are always produced in pairs almost simultaneously, the presence of n photons, in a particular set of optical modes, guarantees the presence of exactly n photons in the set of conjugated optical modes. We label the two sets of conjugated optical modes, and the corresponding paths, as 1 and 2.

In the basilar Klyshko technique two click/no-click single-photon detectors are used, as shown in figure 15. The detector in path 1 is the device under test (DUT) while the second, on path 2, is the reference. In a real measurement, the clicks on the reference channel are used to trigger a coincidence circuit. If we assume that the losses in the system are only due to the non-ideal quantum efficiency of the detectors and that dark counts and the background photon level are negligible, then the calibration technique is quite simple: by definition the number of photons detected by the DUT and by the trigger detector are, respectively, ${n}_{1}={\eta }_{1}n$ and ${n}_{2}={\eta }_{2}n$, and the number of times both detectors click (coincidences) is simply $C={\eta }_{1}{\eta }_{2}n$. It is possible to determine the efficiency ${\eta }_{1}$ simply by taking the ratio ${\eta }_{1}=C/{n}_{2}$.

Figure 15.

Figure 15. Scheme of typical apparatus for Klyshko's calibration technique.

Standard image High-resolution image

When real devices are considered, this simple formula has to be modified to account for the presence of noise and losses. The number of measured counts should be corrected by the sum nB of the electronic dark counts and spurious photo-counts that are not correlated among the paths. The true coincidences have to be obtained from the measured ones Cm by subtracting the accidental coincidences CA occurring between two dark counts, a dark count and a photon, or between two uncorrelated photons. Concerning the losses, it is important to note that the methods cannot distinguish between losses due to the optical elements along the path, parametrized here by τ ($0\leqslant \tau \leqslant 1$), and the ones occurring at the DUT, which represent the quantum efficiency ${\eta }_{1}$. In any case, an independent calibration of the transmissivity of all the optical elements allows an independent estimation of the path transmissivity τ to be performed.

Combining losses and noise effects, the DUT detector efficiency can be estimated as:

Equation (95)

in which all the quantities are directly measurable. In particular, we note that the efficiency of the trigger detector and the number of photon pairs generated do not appear in the final equation, confirming that Klyshko's calibration is an absolute technique in which no pre-calibrated references are needed. Finally, up to now we have not pointed out issues arising from the saturation of click/no-click detectors and from the typical inactivity time (dead time) of these detectors after a counting event. For an accurate calibration of these devices it is necessary to take into account, and correct for, the dead time and operate the source at very low intensity so that the probability of emitting more than one pair of photons in the detection window is negligible.

Recently, many variants and extensions of the Klyshko technique for different intensity regimes and other type of detectors have been studied. The possibility of exploiting the quantum correlations at higher photon fluxes for calibrating analog and linear devices [195198] have been investigated and application to spatial resolving detectors, such as multi-pixel cameras, has been demonstrated with good accuracy. We will discuss these advancements in the next sections.

8.2. Extension of the Klyshko method to PNR detectors

PNR detectors play an important role in many fields of science and technology [199, 200]. PNR capabilities can be built upon multiplexing click/no-click detectors [201205] or using intrinsically PNR detectors such as photo-multipliers [206209], visible light photon counters [210, 211], transition edge sensors (TESs) [212] and inductive superconducting transition edge detectors [213].

The basic version of the Klyshko technique can also be used to calibrate PNR detectors. However, the direct application of such a method does not exploit the full potential of a PNR detector because it does not take into account the possibility of having more photons simultaneously. An extension of the Klyshko method that involves contribution of more than one photon pair at a time has been developed recently [214]. The apparatus, shown in figure 16, is similar to the apparatus of the basic Klyshko technique. The main differences are: the DUT is a TES, i.e. a superconducting PNR detector, and the intensity of SPDC emission is sufficient to produce more then one pair of photons in the detection window. The detector used as a trigger is still a click/no-click detector with unknown quantum efficiency. The typical histogram representing the output of a PNR detector in terms of relative frequency of detection events as a function of the electric pulse amplitude (corresponding to the number of photons) is shown in figure 17.

Figure 16.

Figure 16. Experimental scheme for the PNR Klyshko technique: a pulsed laser beam is used to pump a non-linear crystal in which SPDC takes place. The heralding signal from the trigger detector announces the presence of the conjugated photon, which is coupled in the single-mode optical fiber and sent towards the PNR detector (identified by the dotted line) starting from the fiber end. Both the detectors are gated by the laser trigger.

Standard image High-resolution image
Figure 17.

Figure 17. A typical histogram of TES counts. In the abscissa we have the amplitude of the output signal and in the ordinate axis the number of events.

Standard image High-resolution image

To perform an absolute calibration it is necessary to acquire two separate sets of measures, one in the presence and one in the absence of heralding photons (i.e. photons detected by the trigger detector). It is then possible to estimate the probabilities of observing i counts in the presence and in the absence of the heralded photon, indicated respectively by P(i) and ${ \mathcal P }(i)$.

Now it is useful to define the following quantities: the probability of having a true heralding count ξ (i.e. not due to noise), the overall quantum efficiency γ (including detector efficiency and channel losses), the transmissivity of the optical channel τ from the crystal to the PNR detector, and the quantum efficiency of the detector itself η. From these definitions we can write the relation: $\gamma =\tau \eta $.

The PNR detector has a probability of observing no photons and i photons given, respectively, by:

Equation (96)

Equation (97)

Inverting these two equations it is possible derive the overall quantum efficiency of the PNR detector, which includes the losses on the path. Note that, different to the original Klyshko technique, there are several ways of calculating the quantum efficiency: one for each peak observed in the histogram:

Equation (98)

Each of these derivations of the quantum efficiency exploits a different number of simultaneously detected photons and can be calculated independently. However, all the ${\gamma }_{i}$ represent the same physical quantity and should have the same value for a linear detector. Therefore, this extension of the Klyshko technique allows the consistency of the detection model with data comparing different ${\gamma }_{i}$ to be checked. In addition, in this case, to obtain the quantum efficiency of the detector, $\eta =\gamma /\tau $, it is necessary to estimate independently the losses on the optical path.

8.3. Absolute calibration for analog spatial resolving detectors

The techniques described in sections 8.1 and 8.2 are based on true single-photon and PNR detectors, which allow us to measure temporal coincidences between the time tagged photo-counting events. However, these devices cannot be used for detecting beams with relatively high photon flux, they are expensive and require complex coincidence electronics. On the other hand, the majority of optical detectors operate in the analog regime, providing an output signal that is proportional to the intensity of the light, namely to the number of photons, and have a large dynamic range. In contrast, the high electronic background noise does not allow the individual photons to be discriminated and coincidence measurements to be performed. This limitation does not prevent the possibility of observing true quantum effects such as SSN intensity correlations, when the shot-noise fluctuations proportional to $\sqrt{N}$ emerge from the background noise. On the one hand, this enables the use of analog detectors for quantum enhanced imaging and sensing applications (see for example sections 5, 6, and 7) and on the other hand it allows the absolute calibration of the devices [120, 195198, 215].

The calibration method discussed here is strongly based on the detection of a set of pair-wise correlated spatio-temporal modes of twin beams impinging the detectors areas, according to the model described in section 4.4. There, we demonstrated the relation between the measured noise reduction factor ${\sigma }_{{\rm{\det }}}$ and the detection efficiencies in the case where they are perfectly balanced, see equation (56). Here we report the more general formula in which the detectors have different efficiencies ${\eta }_{1}$ and ${\eta }_{2}$ [90, 120]:

Equation (99)

where $\alpha =\langle {\hat{n}}_{1}\rangle /\langle {\hat{n}}_{2}\rangle ={\eta }_{1}/{\eta }_{2}$ is the measurable ratio between the beams intensities and A is the collection efficiency introduced in equations (57) and (61), a function of geometrical parameters (essentially the size of the detector area and the coherence area) that can be measured independently. Moreover, in equation (99) we introduced a slightly modified NRF parameter ${\sigma }_{\alpha }$, to compensate unbalancing [216]:

Equation (100)

The absolute value of the efficiency ${\eta }_{1}$ obtained by inverting equation (99) is a mean value over the whole area of detection. Note that in the first approximation this equation is valid whatever photon flux impinges the detectors and even for high-gain SPDC, where many photon pairs are generated in a single spatio-temporal mode. A more accurate analysis discloses that there is a certain dependence of the collection efficiency A from the mean occupation number μ of the uncorrelated modes ${{ \mathcal M }}_{{\rm{u}}}$, as reported in equation (57), but this dependence can be minimized as long as a careful matching of the conjugated mode between the detectors is performed.

The technique described above is particularly suitable for calibrating an analog (but not only analog, as we will see in the next section) spatial resolving detector, such as CCD cameras. One of the reasons is the possibility of setting arbitrarily the detection areas and to have fine control in the collection of the conjugated spatial modes. Nevertheless, calibration techniques for spatial resolving detectors are essential for many applications, among which imaging represents the most significant. Due to their importance, in the following we focus our attention on CCD cameras, but, in principle, the technique can be applied to any spatial resolving detector that provides, point by point, an analogical signal proportional to the impinging light flux. Standard CCD cameras, i.e. without any avalanche electro-multiplication, are able to count the number of photoelectrons generated in each pixel for a given exposure time, providing an output proportional to the intensity of the adsorbed light (analogical regime). There are two main sources of noise in CCD cameras: thermal noise and read noise. Thermal noise is due to charges generated by thermal excitations, which is proportional to the exposure time and it is strongly dependent on the temperature. Read noise is generated in the electronic reading process and it is independent of the exposure time and other physical parameters. The read noise contribution is not avoidable and its presence implies that CCDs cannot distinguish single photons from the background. In 2010, the first absolute calibration of a standard CCD camera was realized by exploiting a bright squeezed vacuum [90] and, after several improvements, this technique reached a level of accuracy suitable for metrological application [216], aligned with the state of the art of the absolute calibration of single-photon detectors using Klyshko's method.

8.4. Electro-multiplied CCD (EMCCD) as a link between the single-photon level and the high-intensity level

An EMCCD is a camera able to detect single photons with high quantum efficiency. This capability is achieved exploiting an electron multiplication structure, built into the sensor, that can be activated or not, giving the possibility to switch from the analog to the single-photon regime using the same device.

The statistical distribution of the output counts for this device is well understood [217, 218]. Given n photoelectrons generated inside a pixel, the multiplication stage provides a random number of electron counts x following the distribution:

Equation (101)

Equation (102)

where g is the multiplication gain. The total number of counts per pixel is due to the contribution of photoelectrons and noise. Therefore, the count distribution at the output is the convolution of ${ \mathcal P }(x| n)$ with the noise distribution:

Equation (103)

The statistical distribution of the output counts implies that, despite the possibility of detecting single photons, each pixel of an EMCCD is not a native photon-counting detector. It is possible to use an EMCCD in a photon counting regime by means of proper data processing, after which each pixel can be considered as a click/no-click detector. This behaviour is achieved applying a discriminating threshold T on the electron counts x at each pixel: a photon is detected when $x\gt T$. In this regime, a detection area ${{ \mathcal A }}_{{\rm{\det }}}$ on the sensor can be used as a non-linear photon-number-resolving detector, counting the number of pixels that have $x\gt T$ (spatial multiplexing).

As demonstrated in a recent experimental work [61], exploiting the spatially multimode quantum correlations in squeezed vacuum states, it is possible to calibrate an EMCCD, both in an analog regime and in a photon counting regime, obtaining two different quantum efficiencies, indicated respectively by ${\eta }_{0}$ and $\eta (T)$. The quantum efficiency in a single-photon counting regime is strongly dependent on the threshold T in a predictable way. Moreover, it has been demonstrated that there exists a relation between the quantum efficiencies ${\eta }_{0}$ and $\eta (T)$, by providing a radiometric link between the low illumination range to the mesoscopic and to the macroscopic range [61]. This result represents an important step in the field of quantum radiometry, in particular because it allows the metrological traceability of measurements at the few-photon level, which is essential for most emerging quantum technologies.

The calibration of the analog quantum efficiency ${\eta }_{0}$ is identical for EMCCDs and for standard CCDs, and was described in section 8.3. In addition, the calibration method for EMCCDs in the photon-counting regime is based on the same principle of calibration as in the analog regime, therefore the experimental apparatus is the same as in figure 18. Indeed, EMCCD calibration is also based on the measurement of the corrected NRF reported in equation (100). However, we have to take into account that, in this case, n1 and n2 are the numbers of pixels that have $x\gt T$ in two correlated areas and they depend on the threshold:

Equation (104)

This quantity satisfies the relation with the quantum efficiency as reported in equation (99):

Equation (105)

Therefore, it is possible to use the same absolute calibration technique, both for the photon-counting regime and the analogical regime.

Figure 18.

Figure 18. Schematic representation of an experimental apparatus for absolute calibration of spatial resolving detectors such as CCDs.

Standard image High-resolution image

In principle, for an EMCCD operating in a photon-counting regime, it is also possible to exploit directly the Klyshko method (as in section 8.1) to perform an absolute calibration of the quantum efficiency. However, there are two main practical reasons that prevent its use. First of all, Klyshko's technique needs a few coincidences per frame and unfortunately, in this configuration, the noise in EMCCDs becomes dominant, preventing any possible coincidences being counted. The second reason is that the read time of an EMCCD is much higher with respect to typical single-photon detectors, and as a consequence the Klyshko technique would be much too slow for practical applications.

In this section we focused our attention on the most diffuse spatial resolving devices that allow single-photon counting: EMCCD cameras. However, other types of spatial resolving detectors are able to work in the photon-counting regime. An important commercial device is the ICCD, for which similar absolute calibration techniques have been developed [62, 63]. Most recently, a spatial resolving detector based on arrays of true 'click/no click' single photon detectors has been developed [219]. This type of device, largely used in recent quantum optics experiments [220, 221], can directly exploit the calibration techniques based on squeezed vacuum correlations.

9. Conclusions

Quantum correlations have emerged as a fundamental tool for developing quantum technologies.

In particular, quantum correlations of optical fields are the most exploited resource for these new technologies, whose applications ranges from quantum imaging and sensing to quantum communication and quantum computation.

In this review paper we summarized the main properties of photon statistics and photon-number correlations of technologically relevant optical fields, describing in some detail the use of TWBs in a few quantum enhanced protocols. Our main message is that the relatively easy production of the TWB states and their demonstrated advantages in various protocols make them a fundamental tool for getting past the death valley between proof of principle experiments and commercial systems. They will therefore represent a source of the utmost importance for the approaching second quantum revolution.

Acknowledgments

We acknowledge the support of the MIUR Project Premiale P5 and of the John Templeton Foundation (Grant ID 43467). The opinions expressed in this publication are those of the authors and do not necessarily reflect the views of the John Templeton Foundation.

Please wait… references are loading.