revtex4-2Repair the float
Enhancing Gaussian quantum metrology with position-momentum correlations
Abstract
Quantum metrology offers significant improvements in several quantum technologies. In this work, we propose a Gaussian quantum metrology protocol assisted by initial position-momentum correlations (PM). We employ a correlated Gaussian wave packet as a probe to examine the dynamics of Quantum Fisher Information (QFI) and purity based on PM correlations to demonstrate how to estimate the PM correlations and, more importantly, to unlock its potential applications such as a resource to enhance quantum thermometry. In the low-temperature regime, we find an improvement in the thermometry of the surrounding environment when the original system exhibits a non-null initial correlation (correlated Gaussian state). In addition, we explore the connection between the loss of purity and the gain in QFI during the process of estimating the effective environment coupling and its effective temperature.
I Introduction
The search for optimal measurement of properties during classical or quantum noisy processes i.e. determining the ultimate precision in which a parameter can be estimated, after eliminating all technical noise, using ideally accurate instruments, and repeatedly preparing the system in the same state, is crucial for the development of quantum technologies [1]. Quantum metrology offers significant applications such as superresolution imaging [2], high-precision clocks [3], estimation of proper times and accelerations in quantum field theory [4], navigation devices [5], magnetometry [6], optical and gravitational-wave interferometry [7, 8], and thermometry [9].
Currently, the capacity to measure low temperatures accurately in quantum systems is important in a wide range of proof of principles experiments and has attracted significant interest recently due to its critical role in optimizing the performance of quantum technologies [10, 11]. For instance, effective two-level atoms can be employed to minimize the undesired disturbance on the sample, and an optimum quantum probe (a small controllable quantum system) can act as a thermometer with maximal thermal sensitivity [12]. The scaling of the temperature estimation precision with the number of quantum probes was investigated [13], and the impact of initial system-environment correlations was examined to enhance the estimation of environment parameters within spin-spin model at low temperatures [14]. Still in the low-temperature limit, as a fundamental aspect of the third law of thermodynamics, the unattainability principle of reaching absolute zero also implies the impossibility of precisely measuring temperatures near absolute zero [15, 16, 17]. A general approach to low temperature quantum thermometry considers restrictions from both the sample and the measurement process [15]. In continuous-variable systems, a non-Markovian quantum thermometer was proposed to measure the temperature of a quantum reservoir, effectively avoiding error divergence in the low-temperature [18]. Additionally, optimizing the interaction time in a bipartite Gaussian state allows for precise estimation of the local temperature of trapped ions [19].
Employing quantum resources such as coherence and entanglement, it is possible to improve, until the Heisenberg limit, the measurement precision beyond the classical shot-noise limit [20, 21, 22]. Recently, it was presented a method for low-temperature measurement that improves thermal range and sensitivity by generating quantum coherence in a thermometer probe [23]. Optimal quantum metrology strategies usually employ entanglement in initialization and readout stages to significantly enhance measurement precision [24, 25]. However, in practice, producing entangled states or measurements with high precision, in high-dimensional or continuous-variable systems is not a trivial task. Relatively small errors are known to ruin the success of such tasks when compared to the optimal classical one, and the utilization of a hybrid quantum-classical approach to automatically optimize the controls has been taken into consideration [26].
In this work, we employ a position-momentum (PM) correlated Gaussian state in quantum metrology. In order to work with it, a real parameter can be controlled such that, when it is null, the state recovers the standard uncorrelated Gaussian wave packet form. PM correlations were originally investigated by considering the quantized operators and , with [27]. These correlations generalize Glauber’s coherent states [28, 29] and minimize the Robertson-Schrödinger uncertainty relation. Practically, this parameter arises from atomic beam propagation along a transverse harmonic potential, acting as a thin lens and causing a quadratic phase shift in the initial state [30]. Gaussian correlated packets have applications in various fields, including quantum optics and double-slit matter-wave interferometers [31, 32, 33, 34, 35]. However, practical applications face challenges due to errors in tuning the parameter, caused by non-ideal incoherent sources of matter waves and the focalization method used to produce it [36, 37, 38].
Here, we propose a scheme to estimate PM correlations. Our protocol consists of three stages, initialization, interaction, and estimation. Moreover, we further apply our protocol to estimate the effective environment coupling and its temperature (thermometry), and we show that the correlated Gaussian system can outperform the standard (uncorrelated) one. More specifically, we estimated the effective temperature of a Markovian bath via the environmental coupling with our quantum thermometer based on a single-mode correlated Gaussian system. Such correlations impact the Quantum Fisher Information (QFI) after the evolution of the initial state through a Markovian bath and establish the conditions to improve our quantum thermometer model assisted by PM correlations.
The dynamics of the QFI and purity will be analyzed to better estimate this initial position-momentum correlation using the Classical Fisher Information (CFI). Purity, like other measures such as entanglement [39], quantum discord [40, 41, 42], and quantum coherence [43], is also a resource, where it is quantified by deviations from the maximally mixed state [44, 39, 45, 46, 43, 22, 47, 48]. Purity can be operationally interpreted as the maximum coherence achievable by unitary operations [48]. It bounds the maximum amount of entanglement and quantum discord that unitary operations can generate, acting then as a fundamental resource for quantum information processing [48]. Unlike other measures, purity is easily accessible in experiments, providing experimental bounds to other quantum quantifiers [49, 50, 51]. In the following, we present our main results. First, we describe our general estimation protocol in Sec. II.1. Then, we investigate the noise environmental effect on PM correlations estimation in Sec. II.2. Finally, in Sec. II.3 we analyze the role of PM correlation in improving the thermal sensitivity in the low-temperature regime, and in Sec. III we discuss our results.
II Results
II.1 Estimation protocol
Our quantum estimation protocol schematized in Fig. 1 is represented in three parts: initialization, interaction, and estimation. The initialization step is characterized by the generation of fullerene Gaussian wave packets with momentum and an initial coherent length . In the sequence, the parameterization process of unknown parameters is separated into two parts. The first one, here denoted as focalization, describes the unitary parameterization associated with the encoding of the PM correlation parameter employing a real-noisy source (details about this process are discussed in Appendix B). The second one, denoted as interaction, represents a non-unitary parameterization for which the probe interacts with a Markovian bath producing an encoding of the effective environment coupling parameter and in turn its temperature in the correlated Gaussian state. Finally, the estimation procedure of the unknown parameters is carried out during the estimation stage. This stage includes (I) PM correlations and (II) thermometry estimations.
We consider as the initial state the following correlated Gaussian state of transverse width
(1) |
that represents a PM-correlated Gaussian state [52]. The real parameter ensures that the initial state is correlated [52, 53]. Considering the the initial state, Eq. (1), the variance in position becomes , whereas the variance in momentum is , and the initial correlations PM is . Notice that, for we have a simple uncorrelated Gaussian wave packet, i.e., . Exploring the Pearson correlation coefficient between and , i.e., , the parameter turns out to be , illustrating the physical meaning of the as a parameter that encoded the initial correlations between position and momentum for the initial state.
In the standard quantum phase estimation protocols, it is usual to estimate the physical parameter that implements a relative phase shift [54, 55]. Here, our main goal will be to explore whether initial PM correlations can work as a resource to improve quantum metrology under noise. To estimate the initial PM correlations, we consider the production of beam particles from a non-ideal or partially incoherent source (see Fig. 1), where each particle has momentum in the -direction within a certain range of wavenumber and can be modeled by the following initial density matrix [56, 37, 57]
(2) |
with . In this model, we consider wave effects only in -direction since the energy associated with the momentum of the particles in the -direction is high enough such that the momentum component is sharply defined, i.e., . Then, the behavior in the -direction can be considered classical [37]. The geometry of the collimation device is such that it selects only particles with a transverse wave number inside a specific range defined by the width , being described by a classical Maxwell–Boltzmann distribution, . After some algebraic modifications, the initial density matrix of the coherent correlated Gaussian wave packet becomes
(3) |
where , , , , and is related with the coherence level of the initial state. It is important to note that, in the limit (ideal collimation), one has (), meaning that the initial state (3) is completely coherent or pure. On the other hand, if , then one has (), indicating an completely incoherent or mixed state.
According to the protocol depicted in Fig. 1, after the focalization process, the system interacts with an environment that is considered to be a Markovian bath [58] and its evolution is given by
(4) |
where
(5) |
is the propagator which includes the environmental effect [37, 59]. We consider the propagation of fullerene molecules and the decoherence effect produced by air molecules scattering [37]. In this case, the effective scattering constant is given by , where is the total number density of the air, is the mass of the air molecule, is the size of the molecule of the quantum system, is the Boltzmann constant, and is the bath temperature [59]. In the propagator described by Eq. (5), it is assumed that the center-of-mass state of the system remains undisturbed by the scattering events, a valid condition for fullerene molecules () [37]. The environmental scattering decoherence is a ubiquitous effect constantly monitoring the position of quantum systems in the cosmos, that can be caused by air molecules, light (optical photons), solar neutrinos, cosmic muons, background radioactivity, and even the universe’s 3 K cosmic background radiation [59]. It is also essential to emphasize that in the derivation of this model, it was also assumed the following conditions [59, 60]: the system and the surroundings are not initially correlated; when the composite object–environment system is translated, the scattering interaction remains unchanged; the rate of scattering is much faster than the characteristic rate of change in the state of the system; and the distribution of momenta of the particles in the environment is isotropic and obeys a Maxwell–Boltzmann distribution. In physical terms, the parameter quantifies the rate at which spatial coherence over a given distance is suppressed, motivating the introduction of a decoherence timescale given by [59]
(6) |
This equation will be useful for interpreting how PM correlation is related to loss of purity and gain of Fisher information. After integration and manipulating Eq. (4), we get
(7) |
where , , and are parameters that include the interaction with the Markovian bath (see Appendix C).
Employing Eq. (7), we can cast the purity of the state in the concise form
(8) |
where , and . This result produces when (), i.e., for a completely coherent source and no environment effect . Outside this regime, we always have , typical of a noisy scenario represented by mixed states (details in Appendix C).
In the following subsections, we present the dynamics of the QFI and the purity in each part of the parameterization process illustrated in our estimating protocol in Fig. 1. We begin estimating PM correlations in a scenery where the quantum system unavoidably interacts with its surrounding environments and then we explore how these correlations can enhance noisy quantum metrology.
II.2 PM correlations estimation
From the mixed Gaussian state described in Eq. (7), it is easy to check that the first moments and are null and the dimensionless second moments are given by
(9) |
(10) |
Here, is the time at which the distance of the order of the wave packet extension is traversed with a speed corresponding to the dispersion in velocity [61]. Then, the covariance matrix and its inverse can be written as
(11) |
where , is the adjugate matrix of and we have used that, for Gaussian states, [62]. Therefore, we can write the Quantum Fisher Information (QFI) for single-mode Gaussian states (see Eq.(24) in Appendix A) as following
(12) |
The above expression will be useful for interpreting our results, where we can explicitly understand the role of purity and its derivative as a resource for estimating unknown parameters in the presence of noise. Furthermore, we also investigate the dynamics of Classical Fisher Information (CFI) obtained from the density distribution in position space as
(13) |
where . In what follows, the estimated parameters will be the initial position-momentum correlation and the effective environmental coupling constant .
We employ the QFI and CFI for the estimation of the initial correlations. They were calculated from Eqs. (12) and (13) by making . After some simplifications, we obtain
(14) |
and
(15) |
where the parameter is explicitly defined in Appendix C and is a monotonic increasing function of the parameters , and (see Appendix D). Since this function is multiplied by the fourth power of , the purity behavior determines the first term in QFI. This can be illustrated using the following parameters: fullerene mass kg, molecular size , width of the initial wave packet nm, mass of the air molecule kg and density of the air molecules molecules/m3 [38, 56, 37]. It is assumed that the collimator apparatus selects wave numbers in the direction, with a transverse wavenumber dispersion of . This corresponds to an initial coherence length of . Parameters of this order of magnitude were previously used in experiments with fullerene molecules by Zeilinger [63, 64]. With the other parameters fixed, the variation in corresponds to the variation in the environmental temperature such that, . For instance, of the order of m-2s-1, corresponds to an effective temperature around 205 K. The value m-2s-1 for the scattering by air molecules estimated in [37] for the experiment with fullerene molecules corresponds to a temperature of 300 K and a density of the air molecules of molecules/m3, smaller than the densities we are considering. In other words, the air molecule density and temperature range we are exploring here are experimentally feasible with the current technology.
To analyze the role of each term in the QFI (12) and the influence of the environmental effect, we present in Fig. 2 the behavior of the QFI (), CFI (), purity , and its absolute relative derivative as a function of the initial correlation for an characteristic interaction time s, and different environmental effect coupling constants. Panels (a) and (b) demonstrate that the QFI behaves similarly to the purity when the environment has a relatively small effect (low-temperature regime). In contrast, panels (c) and (d) demonstrate that the QFI behaves according to the absolute purity variation when the environment effect is stronger. Also, we see that only in the regime of strong environmental effect (quantum to classical transition in the high-temperature regime) does the CFI become comparable with the QFI.
One of the most intriguing features here is that besides the suppression of the QFI when the noise increases, we can find PM correlations that enhance the corresponding QFI compared to the usual Gaussian state. It is also important to note that Fig. 2(d) shows that even a maximal purity value (null PM correlations) does not provide the maximum value for the QFI. This happens because, in the regime of strong environmental effect, the change in purity (second term in Eq. (14)) is dominant over the purity value itself that is close to zero and will be elevated to the fourth power (first term in Eq. (14)). To comprehend in more detail how the QFI behavior is governed by either the purity or by the purity variation, and how this transition occurs, we present in Appendix D these quantities where we fixed and plotted them as a function of the PM correlations and the propagation time.
II.3 Quantum thermometry assisted by PM correlations
Here, we analyze how the PM correlations affect the estimation of the effective environmental coupling constant, showing a specific regime in which the correlated Gaussian state provides a quantum advantage over the standard one in estimating the effective parameter related to an environmental interaction. Using Eqs. (12) and (13) and setting , the QFI and CFI reads
(16) |
and
(17) |
where is another monotonic increasing function of the parameters , and (see Appendix E). Similarly, this function does not determine the profile of QFI in the first term.
In Fig. 3, we exhibit the curves for QFI (), CFI (), purity , and its absolute relative variation as a function of initial correlation for s and considering (a) weak environmental effect ( m-2s-1 or mK) and (b) strong environmental effect ( m-2s-1 or K). Again, we observe two characteristic regimes, the gain in QFI is related either to purity or relative variation in purity when the environmental effect is strong (high-temperature regime) or weak (low-temperature regime), respectively. Such regimes are associated with the first and second term in Eq. (16), respectively. One of the more interesting results of this work, rarely explored in the literature, is investigating the relation between loss of purity and gain in Fisher information. Also, we can observe that in the low-temperature regime, the PM can improve the QFI, and this behavior paves the way for exploring the enhancement of noisy quantum metrology by position-momentum correlations that will be discussed in the following. On the other hand, these PM correlations quantum advantage ceases to be valid in the limit of strong environmental effect ( m-2s-1 or K), with the standard uncorrelated Gaussian state providing the best sensitivity for thermometry in this regime [see Fig. 3 (b)].
Also, it is important to mention that the effect of increasing the propagation time or changing the environmental effect is almost equivalent from a theoretical point of view, they both decrease the purity and coherence of the quantum state. However, from an experimental point of view, the variation of these two parameters is very different. The first is varied just by moving the position of the detector to regions further away from the source. In contrast, the second can be varied by changing, for example, the environmental temperature or the air pressure.
In Fig. 4 we investigate the time dependence of (a) QFI () and (b) purity for different values of the initial correlation and considering m-2s-1 ( mK). We observe a temporal quantum enhancement due to PM correlation . Therefore, employing an initial correlated (position-momentum) Gaussian state can reduce the time required to obtain the maximum amount of information until it is saturated. Moreover, note the relationship [inset of panel (b)] between the maximum temporal rate of purity variations and the time of information saturation.
Besides this, to quantify such an enhancement in the time required to saturate the QFI due to PM correlation , we introduce the Temporal Gain of Information (TGI), defining it on the decibel scale as follows
(18) |
where the time corresponds to the maximization point of the relative purity variation, i.e., the maximum of [see inset of Fig. 4(b)]. In this definition, the reduction in the time to saturate the QFI is compared with the correspondent time needed for the standard uncorrelated Gaussian state (), such that the uncorrelated state is always associated with a null temporal gain. After some approximations (see more details in Appendix F) the TGI can be explicitly given by
(19) |
In Fig. 5, we examine the enhancement in TGI as a function of PM correlation . The solid orange line corresponds to the approximate gain calculated from Eq. (19). The full circles correspond to particular TGI values determined by a precise number that optimizes the function without any approximation in the purity expression. These points are obtained directly by probe inspection in the inset of Fig 4. (b) and are summarized in Tab. 1. Moreover, this procedure was applied in order to obtain an approximate and simple expression for the TGI and its dependence on the PM correlations . Note that for PM correlation on the order of 150, the gain achievable is almost 15 dB over the standard uncorrelated Gaussian state. In experimental terms, this parameter can be controlled, for example, by changing the laser frequency or intensity in the initialization step (see Appendix B).
III Discussion
Concepts such as purity distillation and purity cost, analogous to entanglement distillation and entanglement cost, were introduced to quantify this resource [65, 66]. Transformations from pure to mixed states are usually associated with information loss and irreversibility [65]. However, this work explores the relationship between Quantum Fisher Information (QFI) and purity rate, focusing on a regime where purity loss primarily drives QFI behavior. It examines how position-momentum (PM)-correlated probes can enhance noisy quantum metrology of the environmental thermal bath and demonstrates that PM-correlated Gaussian states can outperform standard uncorrelated Gaussian states in estimating the effective environmental coupling constant.
We started presenting the probe initialization procedure and how the PM correlation can be controlled and produced at this stage. We then calculated the dynamics of QFI and purity for the probe as it evolves through a decoherence channel, representing the scattering model, one of the most common sources of decoherence in physical systems [59, 60]. This decoherence model describes how the entanglement of the system with the environmental particles (fermionic or bosonic bath), caused by countless scattering events, delocalizes local phase relations between spatially separated wave-function components, leading to decoherence in position space (i.e., to localization) [59]. Our findings demonstrate the role of purity and purity “velocity” as a resource for estimating unknown parameters in the presence of noise [see Eq. (12)].
We observed that, in estimating the PM correlation, when the decoherence factor is relatively small (i.e., low-temperature regime), the QFI and purity curves have similar behaviors. However, for high temperatures, the QFI is dependent on the rate of change of purity. In this configuration, maximum purity does not necessarily imply maximum values for the QFI due to the dynamics of the purity rate of change term for regimes with an effective strong coupling with the Markovian bath. As expected, QFI tends to match CFI under conditions of quantum-classical transition (high-temperature regime). Furthermore, we found scenarios where the QFI can be asymptotically estimated by the value of CFI by a suitable preparation of the correlated initial state [see, for example, Fig. 2(a)]. From a practical standpoint, this result is both surprising and useful, since obtaining the QFI is challenging because it requires maximizing all possible POVMs to ensure that estimations are independent of the chosen POVM and to achieve optimal measurement precision.
Regarding the estimation of the effective environmental coupling constant, we observed an improvement in the thermometry of the surrounding environment when the probe was prepared in a non-null initial correlated Gaussian state. Specifically, in this part, we explored the acceleration in the loss of purity, due to correlations, as a resource for increasing the QFI [second term in Eq. (16)]. Once the parameter quantifies the decrease rate in spatial coherence, and since PM correlation values imply that high momentum values are correlated with high spatial values, as a result, we have an extended wavepacket for the probe state. This consequently causes an increase in the decoherence rate [see Eq. (6)], reducing the purity of the state. In our model, we considered the long-wavelength limit [59], where the wavelengths of scattering air molecules are much longer than the wave packet extension of the probe. Consequently, a large number of scattering events are needed to resolve the fullerene position by encoding substantial which-path information in the environment, ensuring weak coupling and resulting in an effective Markovian bath. Once we consider that the momentum of environment particles obeys a Maxwell–Boltzmann distribution, the temperature regime investigated here is still more than one million times greater than the ultra-low temperatures obtained, for example, in a Bose-Einstein condensate on the order of 170 nK [67]. Similarly, for example, the amplitude-damping channel [68], which models physical processes such as spontaneous emission, occurs at K, and even so, Markovianity is maintained since the coupling of the system with the infinite vacuum modes is weak.
The findings discussed here can be extended to more general scenarios. For instance, recently it was shown that incorporating additional degrees of freedom in molecules, such as rotations and vibrations, can significantly enhance sensitivity when using molecular sensors as probes [69, 70], creating new opportunities for their application in quantum protocols. Building on this, one could integrate these degrees of freedom into our estimation protocol to fully explore the advantages of molecular probes. Additionally, our findings could also be expanded to include more complex dynamics through a generalized channel that accounts for non-Markovian memory processes. This would offer deeper insights into non-Markovian decoherence features in structured reservoirs, such as decoherence-free states [71]. By pursuing this approach, one could also advance in the exploration at even lower temperatures by relaxing the Born-Markov approximation [18, 72].
Acknowledgements.
J. C. P. Porto acknowledges Fundação de Amparo à Pesquisa do Estado do Piauí (FAPEPI) for financial support. L.S.M. acknowledges the Federal University of Piauí for providing the workspace. P.R.D acknowledges support from the NCN Poland, ChistEra-2023/05/Y/ST2/00005 under the project Modern Device Independent Cryptography (MoDIC). I.G.P. acknowledges Grant No. 306528/2023-1 from CNPq. C.H.S.V acknowledges the São Paulo Research Foundation (FAPESP), Grant. No. 2023/13362-0, for financial support and the Federal University of ABC (UFABC) to provide the workspace.Appendix A Fisher information
Here, we review the definitions and differences between classical and quantum Fisher information, as well as the definition of purity and its relation to resource theory. Our focus is on describing valid relations specifically for Gaussian states.
A.1 Classical Fisher Information
The Fisher information is a measure of information that an observable random variable carries about an unknown parameter [73]. It expresses the level of uncertainty of a measured physical quantity. Let be the conditioned probability of measuring data given a specific value of . The Classical Fisher Information (CFI) for estimation of the parameter is defined by
(20) |
If exhibits a peak as a response to variations in , then, the data provides the information needed to estimate the parameter . On the other hand, when is flat, numerous samples of would be required to estimate the value of , which would be determined only by using the complete sample population. In this context, the Cramér-Rao inequality [74] establishes the lower bound on standard deviation of the estimated parameter
(21) |
where is the number of times that the experiment is repeated. Therefore, the Fisher information determines the reachable accuracy of the estimated quantity and represents the figure of merit in parameter estimation problems. Import to mention, that this inequality is valid only for unbiased estimators, i.e., estimators for which .
In quantum theory, we use a set of positive operator-values measure (POVM) , parameterized by , to describe the measurement procedure. These operators are positive and satisfy to ensure normalization. The probability can be expressed as . Using as the probability distribution, the CFI is defined as
(22) |
where is the density operator describing the system. Naturally, optimal POVMs are those characterized by a statistical distribution of measurement results that is maximally sensitive to changes in the parameter [75].
A.2 Quantum Fisher Information
The Quantum Fisher information (QFI), here denoted by , is defined by maximizing Eq. (22) over all possible POVMs as follows [76, 77, 78, 79]
(23) |
The quantum estimation is then related to the optimal possible measurement precision. Note that this maximization procedure turns the QFI an upper bound for the classical one, i.e., [78]. Quantum metrology is not the only application of quantum Fisher information, alternatives include quantum cloning [80], entanglement detection [81, 82], and quantum phase transition [83]. Feng and Wei [84] have highlighted the relationship between quantum coherence and Quantum Fisher Information (QFI), demonstrating that QFI is useful for quantifying quantum coherence. This relationship is established because QFI satisfies monotonicity under typical incoherent operations and convexity when quantum states are mixed. In Ref. [85], it was demonstrated how QFI can be used to identify quantum correlations such as steering, showing that such correlations can be useful for quantum-enhanced measurement protocols.
For a single-mode Gaussian state, with covariance matrice and first moments , the QFI is given by [62, 86]
(24) |
where is the purity of the quantum state and denotes differentiation with respect to the parameter . The first term is associated with the dynamical dependence of the covariance matrix with the parameter. The second one is the dynamic of purity under variation, and the third is the contribution of the moments dynamics of the Gaussian state for the estimated parameter. For example, this result has been utilized in metrological protocols that take advantage of the superradiant phase transition in the Rabi model [87]. For our system under investigation, we will apply the result (24) throughout the work to quantify the QFI.
Appendix B How generate PM-correlated Gaussian states?
We review the procedure that can generate correlated Gaussian states for pedagogical reasons and to clarify the work from an experimental point of view. We will explicitly show the parameters determining the initial position-momentum correlation in Eq. (1). The traditional method known as atom lens [88, 89] can be applied to produce these initial position-momentum correlations. This method uses the atom-light interactions to create spatially dependent AC-Stark shift of the electronic ground state of the atoms which is induced in the vicinity of an intensity maximum of a sub-resonant standing wave laser field [30]. The standing wave is generated using a cavity perpendicular to the atom direction (see Fig. 1). If the atoms are modeled as two-level systems, the atom-laser interaction’s Rabi frequency is given by
(25) |
to describe a standing wave with period (light wavelength) in the -direction and a Gaussian beam profile in the -direction. Here, is the atomic center-of-mass velocity, and is the effective interaction time, assumed smaller than the radiative lifetime of the excited state of the atoms so that spontaneous emission can be neglected. This approach also applies to the focusing of molecules, since we can consider their vibrational states practically degenerate as long as the energy difference between them is much smaller than the excitation energy of the light beam inside the optical cavity. Due to the Stark shift, the optical potential for incoming atoms in the ground state is given by
(26) |
where , with is the laser frequency and is the resonance transition [30]. In the vicinity of a maximum intensity region, the atoms roughly feel the harmonic potential , with photon momentum . For a short interaction time, the harmonic potential changes the initial state to
(27) |
where is the curvature radius of the wavefronts associated with the beam propagation. Then, it acts as a thin lens with focal length [30]
(28) |
where is the de Broglie wavelength of the atoms. We see therefore from this analysis which experimental parameters determine the initial position correlation characterized by the parameter encoded in the quadratic phase of the initial state (1). From an experimental point of view, the focal length can be varied, for example, by modulating the laser power. Also, note that positive (negative) values of are associated with a diverging (converging) beam with curvature radius ().
Appendix C Parameters of density matrix for a correlated Gaussian state and purity
Here, we describe the parameters obtained through the action of the time evolution propagator over the density matrix shown in Eq. (7)
(29) |
(30) |
(31) |
and
(32) |
Appendix D QFI parameters for correlation estimation
In this appendix, we treat the parameters associated with the function on Eq.(14), representing the QFI for correlation estimation.
(34) |
where , are expressed as follows
(35) |
and
(36) |
We show in Fig. 6 (a) QFI (), (b) purity and (c) its absolute relative derivative as a function of initial correlation and the propagation time , considering m-2s-1 ( K). Observe that the QFI behavior is directly influenced by the purity (absolute variation in purity) in the regime of small (larger) propagation times , allowing us to describe how the transition occurs between a regime determined by purity or its rate of variation in the behavior of the QFI.
Appendix E QFI parameters for the effective environmental coupling estimation
In this appendix, let us show the constants associated with the estimation of the environment coupling through the QFI [].
(37) |
where , expressed as follows
(38) |
(39) |
In the Fig. 7, we analyze the dependence of the QFI () and purity as a function of environmental effect . The inset in Fig. 7(b) shows how the absolute relative variation directly affects the way the QFI behaves. Furthermore, note that initial correlated states with have more QFI when compared with the standard uncorrelated Gaussian state. Also, although for the standard Gaussian state (), the purity remains almost unchanged, preserving the value of this quantity does not translate into any gain in QFI. Therefore, the PM-correlations can be employed to enhance the thermal sensitivity of the probe in the regime of small environmental effects (low-temperature regime). Moreover, the range of thermal precision improvement is broad, albeit decreases with temperature growth.
Appendix F Temporal Gain of Information Gain (TGI)
Here we introduce some approximations to obtain a simplified and closed expression for the TGI as a function of PM-correlations . Considering the propagation times in the order of some microsecond and since the cubic term in Eq.(33) becomes dominant, and therefore the purity can be approximately expressed as
(40) |
(s) | TGI (dB) | ||||
---|---|---|---|---|---|
-50.0 | 17.1 | 0.563 | 58 488 | 0.246 | 11.28 |
-25.0 | 27.2 | 0.563 | 36 861 | 0.247 | 9.24 |
-1.0 | 183.2 | 0.563 | 5 472 | 0.247 | 0.97 |
0.0 | 228.4 | 0.563 | 4 377 | 0.247 | 0 |
35.0 | 21.7 | 0.563 | 46 117 | 0.247 | 10.22 |
70.0 | 13.7 | 0.563 | 73 191 | 0.248 | 12.23 |
150.0 | 8.2 | 0.563 | 121 646 | 0.247 | 14.45 |
Then, the time that maximizes the relative purity variation, i.e., the maximization point of is given by
(41) |
Note that this time can be reduced by increasing the initial correlation . In this way, using from Eq. (41), the approximate TGI() can be written explicitly as
(42) |
References
- Wiseman and Milburn [2010] H. Wiseman and G. Milburn, Quantum Measurement and Control (Cambridge University Press, 2010).
- Köse and Braun [2023] E. Köse and D. Braun, Phys. Rev. A 107, 032607 (2023).
- Rosenband et al. [2008] T. Rosenband, D. B. Hume, P. O. Schmidt, C. W. Chou, A. Brusch, L. Lorini, W. H. Oskay, R. E. Drullinger, T. M. Fortier, J. E. Stalnaker, S. A. Diddams, W. C. Swann, N. R. Newbury, W. M. Itano, D. J. Wineland, and J. C. Bergquist, Science 319, 1808 (2008).
- Ahmadi et al. [2014] M. Ahmadi, D. E. Bruschi, C. Sabín, G. Adesso, and I. Fuentes, Scientific Reports 4, 4996 (2014).
- Grace et al. [2020] M. R. Grace, C. N. Gagatsos, Q. Zhuang, and S. Guha, Phys. Rev. Appl. 14, 034065 (2020).
- Budker and Romalis [2007] D. Budker and M. Romalis, Nature Physics 3, 227 (2007).
- Demkowicz-Dobrzański et al. [2015] R. Demkowicz-Dobrzański, M. Jarzyna, and J. Kołodyński, Progress in Optics 60, 345 (2015).
- Tse and et. al. [2019] M. Tse and et. al., Phys. Rev. Lett. 123, 231107 (2019).
- Weng et al. [2014] W. Weng, J. D. Anstie, T. M. Stace, G. Campbell, F. N. Baynes, and A. N. Luiten, Phys. Rev. Lett. 112, 160801 (2014).
- De Pasquale and Stace [2018] A. De Pasquale and T. M. Stace, Quantum thermometry, in Thermodynamics in the Quantum Regime: Fundamental Aspects and New Directions, edited by F. Binder, L. A. Correa, C. Gogolin, J. Anders, and G. Adesso (Springer International Publishing, Cham, 2018) pp. 503–527.
- Vieira et al. [2023] C. H. Vieira, J. L. de Oliveira, J. F. Santos, P. R. Dieguez, and R. M. Serra, Journal of Magnetic Resonance Open 16, 100105 (2023).
- Correa et al. [2015] L. A. Correa, M. Mehboudi, G. Adesso, and A. Sanpera, Phys. Rev. Lett. 114, 220405 (2015).
- Stace [2010] T. M. Stace, Phys. Rev. A 82, 011611 (2010).
- Mirza and Al-Khalili [2024] A. R. Mirza and J. Al-Khalili, The role of initial system-environment correlations in the accuracies of parameters within spin-spin model (2024), arXiv:2407.03584 [quant-ph] .
- Patrick P. Potts and Brunner [2019] J. B. B. Patrick P. Potts and N. Brunner, Quantum 3 (2019).
- Masanes and Oppenheim [2017] L. Masanes and J. Oppenheim, Nature Communications 8, 14538 (2017).
- Wilming and Gallego [2017] H. Wilming and R. Gallego, Phys. Rev. X 7, 041033 (2017).
- Zhang et al. [2022] N. Zhang, C. Chen, S.-Y. Bai, W. Wu, and J.-H. An, Phys. Rev. Appl. 17, 034073 (2022).
- de Sá Neto et al. [2022] O. P. de Sá Neto, H. A. S. Costa, G. A. Prataviera, and M. C. de Oliveira, Scientific Reports 12, 6697 (2022).
- Kitagawa and Ueda [1993] M. Kitagawa and M. Ueda, Phys. Rev. A 47, 5138 (1993).
- Wineland et al. [1994] D. J. Wineland, J. J. Bollinger, W. M. Itano, and D. J. Heinzen, Phys. Rev. A 50, 67 (1994).
- Baumgratz et al. [2014] T. Baumgratz, M. Cramer, and M. B. Plenio, Phys. Rev. Lett. 113, 140401 (2014).
- Ullah et al. [2023] A. Ullah, M. T. Naseem, and O. E. Müstecaplıoğlu, Phys. Rev. Res. 5, 043184 (2023).
- Giovannetti et al. [2006] V. Giovannetti, S. Lloyd, and L. Maccone, Phys. Rev. Lett. 96, 010401 (2006).
- Huang et al. [2024] J. Huang, M. Zhuang, and C. Lee, Entanglement-enhanced quantum metrology: from standard quantum limit to heisenberg limit (2024), arXiv:2402.03572 [quant-ph] .
- Yang et al. [2021] X. Yang, X. Chen, J. Li, X. Peng, and R. Laflamme, Scientific Reports 11, 672 (2021).
- Bohm [1951] D. Bohm, Quantum Theory, Dover Books on Physics Series (Dover Publications, 1951).
- Dodonov et al. [1980] V. Dodonov, E. Kurmyshev, and V. Man’ko, Physics Letters A 79, 150 (1980).
- Glauber [1963] R. J. Glauber, Phys. Rev. 131, 2766 (1963).
- Janicke and Wilkens [1995] U. Janicke and M. Wilkens, Journal of Modern Optics 42, 2183 (1995).
- Campos [1999] R. A. Campos, Journal of Modern Optics 46, 1277 (1999).
- de Araujo et al. [2019] O. R. de Araujo, L. S. Marinho, H. A. S. Costa, M. Sampaio, and I. G. da Paz, Modern Physics Letters A 34, 1950017 (2019).
- Lustosa et al. [2020] F. R. Lustosa, P. R. Dieguez, and I. G. da Paz, Phys. Rev. A 102, 052205 (2020).
- Marinho et al. [2020a] L. S. Marinho, I. G. da Paz, and M. Sampaio, Phys. Rev. A 101, 062109 (2020a).
- Marinho et al. [2024] L. S. Marinho, P. R. Dieguez, C. H. S. Vieira, and I. G. da Paz, Scientific Reports 14, 12223 (2024).
- Nairz et al. [2002] O. Nairz, M. Arndt, and A. Zeilinger, Phys. Rev. A 65, 032109 (2002).
- Viale et al. [2003] A. Viale, M. Vicari, and N. Zanghì, Phys. Rev. A 68, 063610 (2003).
- Marinho et al. [2018] L. S. Marinho, H. A. S. Costa, M. Sampaio, and I. G. da Paz, Europhysics Letters 122, 50007 (2018).
- Horodecki et al. [2009] R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki, Rev. Mod. Phys. 81, 865 (2009).
- Henderson and Vedral [2001] L. Henderson and V. Vedral, J. Phys. A 34, 6899 (2001).
- Ollivier and Zurek [2001a] H. Ollivier and W. H. Zurek, Phys. Rev. Lett. 88, 017901 (2001a).
- Dieguez and Angelo [2018] P. R. Dieguez and R. M. Angelo, Quantum Inf. Process. 17, 1 (2018).
- Streltsov et al. [2017] A. Streltsov, G. Adesso, and M. B. Plenio, Rev. Mod. Phys. 89, 041003 (2017).
- Vedral et al. [1997] V. Vedral, M. B. Plenio, M. A. Rippin, and P. L. Knight, Phys. Rev. Lett. 78, 2275 (1997).
- Ollivier and Zurek [2001b] H. Ollivier and W. H. Zurek, Phys. Rev. Lett. 88, 017901 (2001b).
- Streltsov [2014] A. Streltsov, Quantum Correlations Beyond Entanglement: and Their Role in Quantum Information Theory, SpringerBriefs in Physics (Springer International Publishing, 2014).
- Winter and Yang [2016] A. Winter and D. Yang, Phys. Rev. Lett. 116, 120404 (2016).
- Streltsov et al. [2018] A. Streltsov, H. Kampermann, S. Wölk, M. Gessner, and D. Bruß, New Journal of Physics 20, 053058 (2018).
- Ekert et al. [2002] A. K. Ekert, C. M. Alves, D. K. L. Oi, M. Horodecki, P. Horodecki, and L. C. Kwek, Phys. Rev. Lett. 88, 217901 (2002).
- Pichler et al. [2013] H. Pichler, L. Bonnes, A. J. Daley, A. M. Läuchli, and P. Zoller, New Journal of Physics 15, 063003 (2013).
- Islam et al. [2015] R. Islam, R. Ma, P. M. Preiss, M. Eric Tai, A. Lukin, M. Rispoli, and M. Greiner, Nature 528, 77 (2015).
- Dodonov [2002] V. Dodonov, J. Opt. B: Quantum Semiclass. Opt. 4, R1 (2002).
- Dodonov and Dodonov [2014] V. Dodonov and A. Dodonov, Russ. Laser Res. 35, 39 (2014).
- Hradil and Řeháček [2005] Z. Hradil and J. Řeháček, Physics Letters A 334, 267 (2005).
- Lee et al. [2012] C. Lee, J. Huang, H. Deng, H. Dai, and J. Xu, Frontiers of Physics 7, 109 (2012).
- Marinho et al. [2023] L. S. Marinho, O. P. de Sá Neto, M. Sampaio, H. A. S. Costa, and I. G. da Paz, Europhysics Letters 143, 40001 (2023).
- Silva et al. [2023] P. P. d. Silva, C. H. S. Vieira, M. Sampaio, J. F. G. Santos, and I. G. d. Paz, The European Physical Journal Plus 138, 210 (2023).
- da Silva et al. [2024] P. P. da Silva, C. H. S. Vieira, J. F. G. Santos, L. S. Marinho, M. Sampaio, and I. G. da Paz, The role of position momentum correlations in coherence freezing and purity behavior (2024), arXiv:2404.15473 [quant-ph] .
- Schlosshauer [2007] M. A. Schlosshauer, Decoherence: and the Quantum-To-Classical Transition (Springer, Berlin, Heidelberg, 2007).
- Giulini et al. [2003] D. Giulini, E. Joos, C. Kiefer, J. Kupsch, I.-O. Stamatescu, and H. D. Zeh, Decoherence and the Appearance of a Classical World in Quantum Theory, Physics and astronomy online library (Springer, 2003).
- Marinho et al. [2020b] L. S. Marinho, I. G. da Paz, and M. Sampaio, Phys. Rev. A 101, 062109 (2020b).
- Serafini [2017] A. Serafini, Quantum Continuous Variables: A Primer of Theoretical Methods (CRC Press, 2017).
- Brezger et al. [2002] B. Brezger, L. Hackermüller, S. Uttenthaler, J. Petschinka, M. Arndt, and A. Zeilinger, Phys. Rev. Lett. 88, 100404 (2002).
- Hornberger et al. [2003] K. Hornberger, S. Uttenthaler, B. Brezger, L. Hackermüller, M. Arndt, and A. Zeilinger, Phys. Rev. Lett. 90, 160401 (2003).
- Horodecki et al. [2003] M. Horodecki, P. Horodecki, and J. Oppenheim, Phys. Rev. A 67, 062104 (2003).
- Gour et al. [2015] G. Gour, M. P. Müller, V. Narasimhachar, R. W. Spekkens, and N. Yunger Halpern, Physics Reports 583, 1 (2015), the resource theory of informational nonequilibrium in thermodynamics.
- Anderson et al. [1995] M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E. Wieman, and E. A. Cornell, Science 269, 198 (1995), https://www.science.org/doi/pdf/10.1126/science.269.5221.198 .
- Nielsen and Chuang [2011] M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information: 10th Anniversary Edition (Cambridge University Press, 2011).
- Hutzler [2020] N. R. Hutzler, Quantum Science and Technology 5, 044011 (2020).
- DeMille et al. [2024] D. DeMille, N. R. Hutzler, A. M. Rey, and T. Zelevinsky, Nature Physics 20, 741 (2024).
- Xiong et al. [2015] H.-N. Xiong, P.-Y. Lo, W.-M. Zhang, D. H. Feng, and F. Nori, Scientific Reports 5, 13353 (2015).
- Einsiedler et al. [2020] S. Einsiedler, A. Ketterer, and H.-P. Breuer, Phys. Rev. A 102, 022228 (2020).
- Fisher [1925] R. A. Fisher, Mathematical Proceedings of the Cambridge Philosophical Society 22, 700–725 (1925).
- Cramér [1946] H. Cramér, Mathematical Methods of Statistics, Goldstine Printed Materials (Princeton University Press, 1946).
- Pezzè et al. [2018] L. Pezzè, A. Smerzi, M. K. Oberthaler, R. Schmied, and P. Treutlein, Rev. Mod. Phys. 90, 035005 (2018).
- Helstrom [1969] C. W. Helstrom, Quantum detection and estimation theory, Vol. 1 (Journal of Statistical Physics, 1969) pp. 231–252.
- Holevo [1982] A. S. Holevo, Probabilistic and statistical aspects of quantum theory (Edizioni della Normale Pisa, North-Holland, Amsterdam, 1982).
- Braunstein and Caves [1994] S. L. Braunstein and C. M. Caves, Phys. Rev. Lett. 72, 3439 (1994).
- Braunstein et al. [1996] S. L. Braunstein, C. M. Caves, and G. Milburn, Annals of Physics 247, 135 (1996).
- Song et al. [2013] H. Song, S. Luo, N. Li, and L. Chang, Phys. Rev. A 88, 042121 (2013).
- Pezzé and Smerzi [2009] L. Pezzé and A. Smerzi, Phys. Rev. Lett. 102, 100401 (2009).
- Li and Luo [2013] N. Li and S. Luo, Phys. Rev. A 88, 014301 (2013).
- Salvatori et al. [2014] G. Salvatori, A. Mandarino, and M. G. A. Paris, Phys. Rev. A 90, 022111 (2014).
- Feng and Wei [2017] X. N. Feng and L. F. Wei, Scientific Reports 7, 15492 (2017).
- Yadin et al. [2021] B. Yadin, M. Fadel, and M. Gessner, Nature Communications 12, 2410 (2021).
- Monras [2013] A. Monras, Phase space formalism for quantum estimation of gaussian states (2013), arXiv:1303.3682 [quant-ph] .
- Garbe et al. [2020] L. Garbe, M. Bina, A. Keller, M. G. A. Paris, and S. Felicetti, Phys. Rev. Lett. 124, 120504 (2020).
- Sleator et al. [1992] T. Sleator, T. Pfau, V. Balykin, and J. Mlynek, Applied Physics B 54, 375 (1992).
- Adams et al. [1994] C. Adams, M. Sigel, and J. Mlynek, Physics Reports 240, 143 (1994).