1 Introduction

Increasing the level of precision achievable in the estimation of physical properties of systems, such as temperatures, optical lengths and magnitude of external fields among others, is one of the multiple applications of quantum technologies that have been extensively studied in the recent years. In particular, the goal of quantum metrology—the field of science laying between quantum mechanics and estimation theory—is to propose and analyse estimation protocols that surpass the precision achievable by classical strategies by employing quantum probes and quantum measurement schemes. In fact it is well known that the classical limit on the precision achievable in the estimation of an unknown parameter when employing N probes, known as shot-noise limit, for which the error is of order \(1/\sqrt{N}\), can be surpassed by quantum strategies achieving the ultimate Heisenberg limit, where the estimation scales as 1/N [1,2,3,4,5,6,7,8,9].

The first proposed protocols reaching Heisenberg-scaling sensitivity heavily employed entanglement as a metrological resource [1,2,3], and several entanglement-based strategies have been recently studied with interesting results, especially in the cases of simultaneous estimation of multiple parameters with non-commuting generators [5,6,7,8,9,10]. Nonetheless, the entanglement fragility and the complicated procedures needed to generate entangled metrological probes, such as NOON or GHZ states [11, 12], are two of the challenges that stimulated the search for more feasible estimation schemes making use of protocols implementing metrological resources that are easier to generate and to manipulate. Squeezed light [13, 14] manifests useful properties (e.g. robustness to decoherence, relatively easy implementation, reduced noise below the vacuum shot-noise) which make it a perfect candidate as a feasible metrological resource [15,16,17,18,19]. Motivated by these favourable properties, many works have recently focused on the analysis and proposal of Gaussian metrological schemes, namely involving squeezed states as probes and homodyne detection as measurement, and the ultimate precision that these can achieve in the estimation of a single localised parameter [15,16,17,18, 20, 21], a function of parameters [22,23,24], or a single distributed parameter, such as the temperature or the electromagnetic field, affecting several component of the network [8, 25,26,27,28,29,30]. These schemes typically perform an analysis based on the study of the Cramér–Rao bound (CRB) [31], or its quantum counterpart [32], to assess the ultimate precision achievable in the estimation, and show whether the Heisenberg scaling can be achieved. Although the CRB analysis does not generally assure that the ultimate bound on the precision found can be achieved globally, namely with an estimation strategy which does not depend on the true value of the parameter to be estimated [33, 34], local estimations are relevant in the typical interferometric framework, in which small deviations of the parameters need to be measured. Interestingly, it has been recently found that it is always possible to reach Heiseneberg-scaling sensitivity regardless of the structure of the network encoding the parameter, only employing a single squeezed vacuum state, a single-homodyne measurement, and an auxiliary network suitably engineered, whose preparation only requires a knowledge on the unknown parameter that can be obtained by a classical measurement [26, 29, 30]. The need for an auxiliary stage in such protocols arises from the fact that in general the probe is scattered by the network in all its output ports, while only a single port is eventually measured through homodyne detection, so that an auxiliary network is required in order to refocus the probe on the only channel observed. A question that naturally arises is whether incrementing the number of observed channels would ease, if not completely lift, the requirement of an auxiliary stage and ultimately the requirement of a prior classical knowledge on the unknown parameter. Moreover, different Gaussian protocols rely on encoding the information about the unknown parameter on the displacement of the probe, requiring that a portion of the resources in the probe are employed in a non-vanishing displacement [28, 35]. Despite concentrating all the photons in the squeezing is known to be the optimal allocation of the resources in the probe [26], encoding the parameter into a non-vanishing displacement can reduce the estimation process into the relatively simple task of inferring the parameter from the expectation value of a Gaussian probability density function [28].

In this work we investigate the ultimate precision achievable in the estimation of a parameter encoded in a generic linear network, when employing a single-mode squeezed coherent Gaussian state and performing homodyne detection on all the output channels (see Fig. 1). We show that, without making any assumption on the structure of the linear network nor on the nature of the parameter, it is always possible to reach Heisenberg-scaling sensitivity with such set-up, without the use of any auxiliary network. This allows for estimation protocols not requiring a preparatory stage nor a prior coarse estimation of the parameter, as opposed to the single-channel homodyne protocols in Refs. [29, 30]. We also show that two independent contributions on the precision arise from our analysis: one originated from the presence of displaced photons in addition to squeezed photons, and the other from the squeezing of the probe. We find that both contributions can reach Heisenberg-scaling sensitivity independently, and this can be achieved expectedly when the local oscillators phases are chosen such that the noise in the outcome is reduced, namely when the squeezed quadratures are observed in each output channel. Thus, differently from the schemes in Refs. [29, 30], it is then possible to employ this set-up to retrieve information on the parameter through measurements of the average value (i.e. the displacement) of the signal observed with homodyne detection, as well as through the modulation of the noise. Although it is not required to reach the Heisenberg-scaling sensitivity, the presence of an auxiliary network in general affects the precision of the estimation through a pre-factor multiplying the scaling. This comes in useful in those cases where priority is given to increasing the precision, at the expenses of engineering an auxiliary network to be added before the estimation protocol is started.

Fig. 1
figure 1

Optical set-up for the estimation at the Heisenberg-scaling precision of a single unknown parameter \(\varphi \) encoded arbitrarily in a generic passive M-channel linear network. The parameter can be either localised in a single element of the network, or represent a global property affecting several components, such as a temperature or a magnetic field. A single source of coherent squeezed states with \(N = N_\mathrm {D}+ N_\mathrm {S}\) average photons, where \(N_\mathrm {D}=d^2\) and \(N_\mathrm {S}=\sinh ^2 r\) are the number of displaced and squeezed photons, respectively, is employed in a single input channel (the first in figure), while homodyne measurements are taken in every output channel

2 Set-up

Let us consider a generic \(M\times M\) passive linear network whose action on any injected photon probe is given by the unitary operator \({\hat{U}}_{\varphi }\), in which the unknown parameter \(\varphi \) to be estimated is encoded in an arbitrary manner. The linearity and passivity of the network allow us to describe it with an \(M\times M\) unitary matrix \(U_{\varphi }\) related to the evolution operator \({\hat{U}}_{\varphi }\) by

$$\begin{aligned} {\hat{U}}_{\varphi }^\dag {\hat{a}}_i {\hat{U}}_{\varphi }= \sum _{j=1}^M (U_{\varphi })_{ij}{\hat{a}}_j. \end{aligned}$$
(1)

The input probe is prepared in a single-mode squeezed coherent state \(\vert \varPsi _{\mathrm {in}}\rangle ={\hat{D}}_1(d){\hat{S}}_1(r)\vert \mathrm {vac}\rangle \) with \(\mathrm{N}=\sinh ^2 \mathrm{r} + \mathrm{d}^2/2 \equiv N_\mathrm {S}+ N_\mathrm {D}\) average number photons, where \({\hat{S}}_1(r) = \exp \bigl (r({\hat{a}}_1^{\dag 2}-{\hat{a}}_1^{2})/2\bigr )\) is the squeezing operator with real squeezing parameter r, and \({\hat{D}}_1(d) = \exp \bigl (d({\hat{a}}^\dag _1-{\hat{a}}_1)/\sqrt{2}\bigr )\) is the displacement operator with real displacement d, and it is injected in one input channel, say the first, of the linear network. In this case, only the first row of \(U_{\varphi }\) is relevant in this protocol

$$\begin{aligned} (U_{\varphi })_{1j} = \sqrt{P_j}\mathrm {e}^{\mathrm {i}{\bar{\gamma }}_j}, \end{aligned}$$
(2)

where we have made explicit the probability \(P_j\) that each photon exits from the jth output port of the network, and the phase \({\bar{\gamma }}_j\) acquired in the process, with j from 1 to M. The unitarity of \(U_{\varphi }\) assures that \(\sum _j P_j = 1\).

A homodyne detection is then performed at each of the output channels, and the quadratures \({\hat{x}}_{i,\theta _{i}}\) are measured, where \(\theta _i\) is the ith local oscillator reference phase, from which we want to infer the value of \(\varphi \). Due to the Gaussian nature of the scheme, the joint probability distribution \(p({\varvec{x}}|\varphi )\) associated with the M-mode homodyne measurement is Gaussian

$$\begin{aligned} p({\varvec{x}}|\varphi ) = \frac{1}{\sqrt{(2\pi )^M \left|\varSigma \right|}}\exp \Bigl [-\frac{({\varvec{x}}-\varvec{\mu })^\mathrm {T}\varSigma ^{-1}({\varvec{x}}-\varvec{\mu })}{2}\Bigr ]. \end{aligned}$$
(3)

Here, \(\varSigma \) is the \(M\times M\) covariance matrix with elements (see “Appendix A”)

$$\begin{aligned} \varSigma _{ij} = \frac{\delta _{ij}}{2} + \sqrt{P_i P_j}\bigl (\cos (\gamma _i-\gamma _j)\sinh (r)^2 +\cos (\gamma _i+\gamma _j)\cosh (r)\sinh (r)\bigr ),\quad \end{aligned}$$
(4)

where \(\delta _{ij}\) is the Kronecker delta, \(\gamma _i={\bar{\gamma }}_i-\theta _i\) is the phase delay at the output of the ith channel relative to the correspondent local oscillator, and \(\left|\varSigma \right|\) is the determinant of \(\varSigma \), which reads (see “Appendix A”)

$$\begin{aligned} \left|\varSigma \right| =&\frac{1}{2^M} + \frac{\sinh (r)}{2^{M-1}}\sum \limits _{i=1}^M P_i (\sinh (r) + \cos (2\gamma _i)\cosh (r)) \nonumber \\&-\frac{\sinh ^2(r)}{2^{M-2}}\sum \limits _{i = 1}^M \sum \limits _{j = i+1}^M P_i P_j \sin ^2(\gamma _i-\gamma _j), \end{aligned}$$
(5)

and \(\varvec{\mu }\) is the mean vector

$$\begin{aligned} \mu _i = d\sqrt{P_i}\cos \gamma _i. \end{aligned}$$
(6)

For any given unbiased estimator \({\tilde{\varphi }}\), the statistical error in the estimation of \(\varphi \) after \(\nu \) iterations of the measurement is limited by the Cramér-Rao bound (CRB) [31]

$$\begin{aligned} \mathrm {Var}[{\tilde{\varphi }}] \ge \frac{1}{\nu {\mathcal {F}}(\varphi )}, \end{aligned}$$
(7)

where \({\mathcal {F}}(\varphi )\) is the Fisher information

$$\begin{aligned} {\mathcal {F}}(\varphi ) = \int \mathrm {d}{\varvec{x}}\ p({\varvec{x}}|\varphi )\bigl (\partial _\varphi \log p({\varvec{x}}|\varphi )\bigr )^2, \end{aligned}$$
(8)

associated with the Gaussian distribution (3), and reads (see “Appendix B”)

$$\begin{aligned} {\mathcal {F}}(\varphi ) = \frac{1}{\left|\varSigma \right|}\partial _\varphi \varvec{\mu }^\mathrm {T}C\partial _\varphi \varvec{\mu } + \frac{1}{2}\left( \frac{\partial _\varphi \left|\varSigma \right|}{\left|\varSigma \right|}\right) ^2 - \frac{1}{2\left|\varSigma \right|}\mathrm{Tr}[(\partial _\varphi \varSigma )(\partial _\varphi C)], \end{aligned}$$
(9)

where \(C=\left|\varSigma \right|\varSigma ^{-1}\) is the cofactor matrix of \(\varSigma \) and \(\mathrm{Tr}[\cdot ]\) denotes the trace. In the following we will discuss in detail expression (9) in the asymptotic limit of large N, showing which condition must be met in order for this set-up to reach Heisenberg-scaling precision in the estimation of \(\varphi \), and compare differences and advantages with respect to the Heisenberg-scaling single-homodyne schemes [29, 30].

We conclude this section by remarking that, in the case of a single channel, \(M=1\), the last term in the right-hand side of (9) vanishes, and thus the only relevant terms are the first two, containing the derivative of the mean \(\varvec{\mu }\) and of the determinant of \(\varSigma \), which reduces to the variance of a single-homodyne measurement. Interestingly enough, we will show that also in the multi-homodyne case, only the first two terms are relevant for the Heisenberg scaling in the asymptotic regime.

3 Heisenberg scaling of the Fisher information

In order to investigate the asymptotic behaviour of the Fisher information (9), it is convenient to express the elements of the cofactor matrix C in terms of the squeezing factor r (see “Appendix B”):

$$\begin{aligned} C_{ss}&= \frac{1}{2^{M-1}} + \frac{1}{2^{M-2}}\sum \limits _{\begin{array}{c} i=1 \\ i \ne s \end{array}}^M \Big (\varSigma _{ii} - \frac{1}{2}\Big ) - \frac{1}{2^{M-3}}\sum \limits _{\begin{array}{c} i=1 \\ i \ne s \end{array}}^M\sum \limits _{\begin{array}{c} j=i+1 \\ j \ne s \end{array}}^M S_{iij},\nonumber \\ C_{st}&= -\frac{1}{2^{M-2}}\varSigma _{st} + \frac{1}{2^{M-3}}\sum \limits _{\begin{array}{c} i=1 \\ i\ne s,t \end{array}}^M S_{sti},\quad s\ne t, \end{aligned}$$
(10)

where

$$\begin{aligned} S_{sti} = \sinh ^2{r}\sqrt{P_s P_t} P_i \sin (\gamma _s - \gamma _i)\sin (\gamma _t-\gamma _i). \end{aligned}$$
(11)

Notice that every element of C in the previous expressions, and of \(\varSigma \) in (4), scale at most as quick as N, namely \(C_{st}=O(N_\mathrm {S})\) and \(\varSigma _{st}=O(N_\mathrm {S})\), while the mean vector \(\varvec{\mu }\) is of order \(O(\sqrt{N_\mathrm {D}})\) (see “Appendix C”), and the same asymptotic bounds hold for their derivatives with respect to \(\varphi \), since neither \(P_i\) nor \({\bar{\gamma }}_i\) depend on N. For this reason, in order for the Fisher information in (9) to asymptotically grow with Heisenberg scaling, it is essential to study the asymptotics of the determinant \(\left|\varSigma \right|\) in (5) and find the conditions for which it does not grow with N.

In fact, it is evident from Eq. (5) that in general \(\left|\varSigma \right|=O(N_\mathrm {S})\), and we show in “Appendix C” that the necessary condition for it to scale slower than \(N_\mathrm {S}\) is that the relative phases \(\gamma _i\) tend to \(\pm \pi /2\) for large \(N_\mathrm {S}\): in other words, the larger the number of photons employed in the squeezing of the probe to reach higher precisions, the closer the local oscillator phase needs to be tuned to the minimum-variance quadrature of each mode.

More precisely, as shown in “Appendix C”, the conditions to reach Heisenberg scaling in the Fisher information (9), read

$$\begin{aligned} \gamma _i = \pm \frac{\pi }{2} + O(N_S^{-1}),\qquad i=1,\dots , M. \end{aligned}$$
(12)

When these conditions hold, we can introduce the finite quantities \(k_i =\lim _{N_\mathrm {S}\rightarrow \infty } N_\mathrm {S}(\gamma _i\mp \pi /2)\), and the determinant \(\left|\varSigma \right|\) reduces to

$$\begin{aligned} \left|\varSigma \right| = \frac{1}{2^{M-2}N_\mathrm {S}}\biggl (\Bigl (\sum \limits _{i=1}^M P_i k_i\Bigr )^2 +\frac{1}{16}\biggr ), \end{aligned}$$
(13)

while \(\partial _\varphi \left|\varSigma \right|\), \(\partial _\varphi \varSigma \), \(\partial _\varphi C\) and C tend to constant values, and \(\partial _\varphi \varvec{\mu }\) scales as \(\sqrt{N_\mathrm {D}}\), thus making only the first two terms of the Fisher information dominant for large N.

As expected, the determinant of the covariance matrix \(\varSigma \) reaches its minimum value when \(\gamma _i = \pi /2\), or \(k_i=0\) for \(i=1,\dots ,M\), namely when the squeezed quadratures are measured. When conditions (12) are met, we can neglect the trace term in Eq. (9), and we can write

$$\begin{aligned} {\mathcal {F}}(\varphi )&\simeq \frac{1}{\left|\varSigma \right|}\partial _\varphi \varvec{\mu }^\mathrm {T}C\partial _\varphi \varvec{\mu } + \frac{1}{2}\left( \frac{\partial _\varphi \left|\varSigma \right|}{\left|\varSigma \right|}\right) ^2 \nonumber \\&\simeq 8(\partial \gamma )_\mathrm {avg}^2\left( 2 \zeta (\mathrm{k}_\mathrm {avg}) \mathrm{N}_\mathrm{D} \mathrm{N}_\mathrm{S} + \varrho \left( k_\mathrm {avg}\right) N_S^2\right) , \end{aligned}$$
(14)

where \(k_\mathrm {avg}\equiv \sum \nolimits _{i=1}^M P_i k_i\), \((\partial \gamma )_\mathrm {avg} \equiv \sum \nolimits _{i=1}^M P_i \partial _\varphi \gamma _i\), and \(\varrho (x) = (8x)^2/(16x^2 + 1)^2\) and \(\zeta (x)=(16x^2 + 1)^{-1}\) are positive, even function which reach their maxima at \(x=\pm 1/4\) and \(x=0\), respectively, namely \(\varrho (1/4)=1\) and \(\zeta (0)=1\). The Cramér-Rao bound (7) with the Fisher information (14) is saturated for large \(\nu \) by the maximum-likelihood estimator [36,37,38], and thus, Heisenberg-scaling precision can be achieved. The expression of the maximum-likelihood estimator for this estimation scheme can be found in “Appendix D”.

Noticeably, both terms in the asymptotic Fisher information (14) give a Heisenberg-scaling precision in the estimation of the parameter \(\varphi \), provided that both the average number of photons in the displacement \(N_\mathrm {D}\) and in the squeezing \(N_\mathrm {S}\) scale with the total average number of photons \(N=N_\mathrm {S}+N_\mathrm {D}\), namely \(N_\mathrm {S}= \beta N\) and \(N_\mathrm {D}= (1-\beta )N\), for any value \(0<\beta \le 1\) independent of N.

Moreover, it is worth noticing that the first term in Eq. (9), and thus in Eq. (14), depends on the information encoded in the displacement of the probe, and thus it vanishes if \(\varvec{\mu } = 0\), namely if the probe is a squeezed vacuum and \(N_\mathrm {D}=0\). The second term instead depends on the information on \(\varphi \) encoded in the variance of the measurement itself: it arises only from the interaction with the squeezed photons and vanishes if \(\partial _\varphi \left|\varSigma \right| = 0\), namely when \(k_\mathrm {avg} = 0\) in Eq. (14), and in particular when \(\gamma _i = \pm \pi /2\), for \(i=1,\dots ,M\), in Eq. (12), corresponding to quadratures with minimum squeezed variances, and thus locally insensible to the variations of the parameter.

Interestingly, this latter case is similar to the single squeezed vacuum and single-homodyne scenario found in the literature [29, 30]: in fact, the second term in (14) represents a generalisation of the single-homodyne Fisher information \({\mathcal {F}}_1(\varphi ) = 8\varrho (k)(\partial _\varphi \gamma )^2 N^2\), and it can be obtained by substituting k and \(\partial _\varphi \gamma \) with their averages over the probabilities \(P_i\), namely \(k\rightarrow \sum _i P_i k_i\) and \(\partial _\varphi \gamma \rightarrow \sum _i P_i \partial _\varphi \gamma _i\).

We have then found that, also when employing multiple homodyne detections, one for each output port of the interferometer, the Heisenberg-scaling precision obtained through measurements of the squeezed noise (i.e. \(\partial _\varphi \varvec{\mu }=0\)) is only reached when the quantum fluctuations of the observed quadratures are reduced to their quantum limit, i.e. \(\left|\varSigma \right| = O(N_\mathrm {S}^{-1})\), while the variations of the unknown parameter \(\varphi \) still yield a visible effect on the outcomes of the measurements, i.e. \(\partial _\varphi \left|\varSigma \right|\) is not vanishing.

However, at the expense of introducing a nonzero displacement in the probe, it is possible to relax the condition \(\partial _\varphi \left|\varSigma \right|\ne 0\), thus allowing us to choose \(k_i=0\) in Eq. (13), thus effectively measuring the maximally squeezed quadratures at \(\gamma _i= \pm \pi /2\). Indeed in such a case, even if the contribution to the Fisher information associated with only the squeezed photons in Eq. (14) is vanishing, it is still possible to reach Heisenberg-scaling precision through the information on the parameter encoded in the displacement of the probe.

An important feature of this protocol, which differentiate it from its single-homodyne counterpart, is that it does not require any adaptation of the network to the value of the unknown parameter, namely no auxiliary networks needs to be added at the input nor the output of \({\hat{U}}_{\varphi }\) to reach Heisenberg-scaling precision. The only condition (12) can be thought as a minimum-resolution requirement on the local oscillators phases, which can thus be achieved without adding further auxiliary networks.

However, this does not mean that the form of the network \(U_{\varphi }\) does not affect the precision of our protocol in the estimation of \(\varphi \): the terms \(k_\mathrm {avg}\) and \((\partial \gamma )_\mathrm {avg}\) appearing in the constant factor in the Fisher information in Eq. (14) depend on the transition probabilities \(P_i\) and on the derivatives of the relative phases \(\partial _\varphi \gamma _i\). In particular, an exceptionally poorly conceived network, e.g. one for which \(\gamma _i\) is independent on \(\varphi \) for every i such that \(P_i \ne 0\), can be associated with a null factor \((\partial \gamma )_\mathrm {avg}\) that sets to zero the Fisher information. In this case, adding a \(\varphi \)-independent auxiliary network V, either at the input or at the output of \(U_{\varphi }\), might modify both \(P_i\) and \(\gamma _i\), and thus \((\partial \gamma )_\mathrm {avg}\).

4 Conclusions

We have shown that performing homodyne measurements at each output channel of an arbitrary linear network encoding an unknown distributed parameter \(\varphi \) to be estimated allows us to reach Heisenberg-scaling precision for a single-mode squeezed probe with no prior information on \(\varphi \). The information on \(\varphi \) is encoded both in the displacement and in the squeezing of the probe, leading to two independent contributions which can both provide Heisenberg-scaling sensitivity. We have shown that the determinant of the covariance matrix associated with the measurement outcomes plays an important role in the enhanced sensitivity: in particular, we demonstrated that the conditions to reach Heisenberg scaling in either of the two contributions, which can be met manipulating the phases of the local oscillators, correspond to imposing that the determinant of the covariance matrix is of order \(N^{-1}\) for large N. Differently from protocols involving only homodyne measurements at a single channel, here there is no need for a refocusing auxiliary stage: the procedure is independent of the network and of the value of the parameter. This allows us to safely entrust the measurement operation to an independent party without sharing any information on the structure of the network, possibly opening up a further path towards secure sensing and cryptographic quantum metrology [39,40,41]. On the other hand, we showed that, despite not required to achieve the Heisenberg limit, one can still employ an auxiliary stage to further enhance the estimation precision by a constant factor. As a future step, it would be interesting to extend the results hereby presented beyond the Cramér–Rao bound analysis, also in the framework of a parameter distributed arbitrarily in a network.