• Non ci sono risultati.

Colored collapse models from the non-interferometric perspective

N/A
N/A
Protected

Academic year: 2021

Condividi "Colored collapse models from the non-interferometric perspective"

Copied!
7
0
0

Testo completo

(1)

P

HYSICAL

J

OURNAL

D

Regular Article

Colored collapse models from the non-interferometric

perspective

?

Matteo Carlesso1,2,a,Luca Ferialdi3,b,and Angelo Bassi1,2

1Department of Physics, University of Trieste, Strada Costiera 11, 34151 Trieste, Italy 2

Istituto Nazionale di Fisica Nucleare, Trieste Section, Via Valerio 2, 34127 Trieste, Italy

3

Department of Physics and Astronomy, University of Southampton, SO17 1BJ, UK Received 25 May 2018

Published online 18 September 2018 c

The Author(s) 2018. This article is published with open access atSpringerlink.com

Abstract. Models of spontaneous wave function collapse describe the quantum-to-classical transition by assuming a progressive breakdown of the superposition principle when the mass of the system increases, providing a well-defined phenomenology in terms of a non-linearly and stochastically modified Schr¨odinger equation, which can be tested experimentally. The most popular of such models is the continuous sponta-neous localization (CSL) model: in its original version, the collapse is driven by a white noise, and more recently, generalizations in terms of colored noises, which are more realistic, have been formulated. We will analyze how current non-interferometric tests bound the model, depending on the spectrum of the noise. We will find that low frequency purely mechanical experiments provide the most stable and strongest bounds.

1 Introduction

Since the birth of quantum mechanics with its strik-ing differences compared to our classical intuition, the quantum-to-classical transition has puzzled the scientific community. The scientific debate, first confined to con-ceptual arguments, now has an experimental counterpart, thanks to technological developments, which allow to put at test the questions about the boundaries between the classical and quantum realms [1–4].

The quantum-to-classical transition is consistently described in terms of collapse models [5,6]. These are phenomenological models, which modify the standard Schr¨odinger dynamics by introducing new suitable non-linear and stochastic terms, which properly describe the collapse of the wave-function.

The continuous spontaneous localization (CSL) model [7] is the most studied among collapse models, and nowa-days it represents a commonly accepted alternative to standard quantum mechanics. The stochastic terms char-acterizing this model depend on two parameters: the collapse rate λ and the noise correlation length rC. Based

on alternative theoretical considerations, different numer-ical values have been proposed: λ = 10−16s−1 and rC =

10−7m by Ghirardi et al. [8]; λ = 10−8±2s−1 for rC =

10−7m, and λ = 10−6±2s−1for r

C= 10−6m by Adler [9].

?

Contribution to the Topical Issue “Quantum Correlations”, edited by Marco Genovese, Vahid Karimipour, Sergei Kulik, and Olivier Pfister.

a

e-mail:matteo.carlesso@ts.infn.it b

e-mail:l.ferialdi@soton.ac.uk

However, since the CSL model is phenomenological, at present the values of the parameters can only be bounded by experiments.

Although only recently the scientific community has started developing dedicated experiments [10,11], one can infer bounds on the CSL parameters by investigating the predictions of the model for suitable experimental scenarios. One can divide them in two classes: interfero-metric experiments and non-interferointerfero-metric ones. The first class includes those experiments where ideally a spatial superposition is created in an interferometer, and the cor-responding interference pattern is measured. This is the case of cold-atom [12] and molecular [13–16] interferome-try, and entanglement experiments with diamonds [17,18]. Conversely, the experiments falling in the second class are those where no superposition is generated, the collapse being detected indirectly through the random motion which is always associated to it. These experiments involve cold atoms [19], optomechanical systems [10,20–25], X-ray measurements [26], phonon excitations in crystals [27,28]. Note that in non-interferometric experiments one can also consider systems which are (truly) macroscopic. In such a case, due to the amplification mechanism, the collapse can be more significant and easier to detect. This has been shown for example in [10,29] for a micrometer cantilever, and in [30,31] for human-scale gravitational wave detec-tors: these experiments establish the strongest bounds on λ for rC> 10−6m, while X-ray measurements [26], which

also employ human-scale objects, set the strongest bounds for rC< 10−6m.

(2)

the collapse noise heats up the system of interest, making its average energy increase linearly. Although such a heat-ing has rather long time-scales (makheat-ing it negligible in most situations), this is something one eventually would like to remove. This has been resolved by the dissipative extension of the model [5,15,16,32–34].

The second weak point concerns the spectrum of the CSL noise, which is flat, being the noise white. If one thinks that the noise providing the collapse has a phys-ical origin, it cannot be white but colored, with a cut off. This extension of the CSL model has already been formulated [15,35–40], and we will refer to it as “col-ored CSL” (cCSL): this is the subject of the present article. In particular, we will investigate the bounds that non-interferometric experiments place on the col-lapse parameters λ and rC, when a colored noise with

exponentially decaying correlation function is considered. The paper is organized as follows: in Section2we intro-duce the cCSL model and we compute its contribution to the density noise spectrum (DNS) of optomechanical systems. In Section3, we quickly present the cCSL predic-tions for other relevant non-interferometric experiments: X-ray measurements [41] and bulk heating [27,28]. In Section4, we use these theoretical formulas to derive the bounds on the cCSL parameters from available experi-mental data, and in Section5 we discuss the results and draw our conclusions.

2 CSL model and optomechanical systems

Before digging into the details of the cCSL model, we start by reviewing the basic features of CSL model, which will be useful for the following analysis. As already briefly described before, the CSL model modifies the standard Schr¨odinger dynamics for the wave-function by adding to it non-linear and stochastic terms [5]. Non-linearity is required in order to suppress quantum superpositions, while stochasticity has to be implemented in order to avoid superluminal signaling and to recover the Born rule in measurement situations [5]. These modifications are devised in such a way that the motion of micro-scopic objects is not significantly affected by them (hence recovering standard quantum mechanics), while a built-in amplification mechanics guarantees that macroscopic bodies behave classically.

A non-linear stochastic collapse equation is rather diffi-cult to solve. However, when one comes to the expectation values of physical quantities, the CSL collapse can be mimicked by the random potential [30]

ˆ VcCSL(t) = − ~ √ λ m0 Z dz ˆM (z)w(z, t), (1)

with w(z, t) a classical Gaussian noise characterized by E[w(z, t)] = 0, E[w(z, t)w(x, s)] = δ(3)(z − x)f (t − s),

(2) where E[ · ] denotes the stochastic average. In the gen-eral case, w(z, t) has a correlation function f (t) with a

non-trivial (colored) spectrum. In particular, the stan-dard CSL model is recovered with f (t) = δ(t), implying a flat spectrum (white noise). In equation (1) we introduced

ˆ

M (z), which is a locally averaged mass density operator:

ˆ M (z) = m0 π3/4r3/2 C X n e− (z−ˆqn)2 2r2 C , (3)

where ˆqn is the position operator of the n-th nucleon of

the system and m0is a reference mass chosen equal to the

one of a nucleon. When the spread in position of the cen-ter of mass is much smaller than rC, under the rigid body

assumption, we can approximate the above expression with [42] ˆ M (z) ' M0(z) + Z dx µ(x) π3/4r7/2 C e− (z−x)2 2r2 C (z − x) · ˆq, (4)

where M0(z) is a complex function, µ(x) is the mass

density of the system and ˆq is the center of mass operator. We now have all the key ingredients to evaluate the effects of the CSL model on optomechanical systems [42–44]. Their dynamics is conveniently described in terms of a Langevin equation, which we here write down in its one dimensional version (along the x direction), in the limit of vanishing optical coupling [45]:

dˆx dt = ˆ p m, dˆp dt = −mω 2 mx − γˆ mp + ˆˆ ξ(t) + FcCSL(t). (5)

Here, m is the mass of the system, ωm is the frequency

of the harmonic trap and γm is the damping constant.

We introduced two stochastic terms of different origin: ˆ

ξ(t) and FcCSL(t). The former is quantum, and describes

the thermal action of the surrounding environment, which is supposed to be in equilibrium at temperature T . Its average and correlation are h ˆξ(t)i = 0 and

h ˆξ(t) ˆξ(s)i = ~mγm Z 2π e −iω(t−s)ωh1 + coth ~ω 2kBT i . (6) The second stochastic contribution is classical and describes the CSL action on the system, according to FcCSL(t) =

i

~[ ˆVcCSL(t), ˆp]. Given the form of ˆVcCSL(t) and

ˆ

M (z) respectively in equations (1) and (4), we find that [30] FcCSL(t) = ~ √ λ π3/4m 0 Z dzdxµ(x) rC7/2 e− (z−x)2 2r2 C (z − x)xw(z, t). (7) We note that, with reference to equation (3), without approximations the stochastic force would be an operator. However, since in the Taylor expansion of equation (4) we considered only the terms up to the first order in the cen-ter of mass position operator, FcCSL(t) becomes a function

(3)

Equation (5) allows us to derive the density noise spectrum, which quantifies the overall noise on the system. 2.1 Density noise spectrum

The DNS S(ω) characterizes the motion of an optome-chanical system in its steady state [46]. It is defined as the Fourier transform of the fluctuation δ ˆx(t) = ˆ

x(t) − ˆxSS of the position operator around its steady

state ˆxSS: S(ω) = 1 2 R+∞ −∞ dτ e −iωτ E [h{δ ˆx(t), δ ˆx(t + τ )}i] , where h · i denotes the quantum expectation value. The DNS can also be written in terms of δ ˜x(ω), which is the Fourier transform of δ ˆx(t): S(ω) = 1 4π Z +∞ −∞ dΩ E [h{δ˜x(ω), δ ˜x(Ω)}i] . (8) Following the standard prescription [46,47], one can derive from equation (5) the equations of motion of δ ˆx(t) and of the similarly defined momentum fluctuations δ ˆp(t) = ˆ

p(t) − ˆpSS. By solving these equations in the Fourier space,

we get: δ ˜x(ω) = 1 m ˜ ξ(ω) + ˜FcCSL(ω) (ω2 m− ω 2) − iγ mω , (9)

where as general notation rule, we denote with tilde the Fourier transform of a quantity. The correlation functions of the noises, in Fourier space, read:

h ˜ξ(ω) ˜ξ(Ω)i = 2π~mγmω h 1 + coth ~ω 2kBT i δ(ω + Ω), E h { ˜FcCSL(ω), ˜FcCSL(Ω)} i = 2π~2η ˜f (ω)δ(ω + Ω), (10) where ˜f (ω) is the Fourier transform of f (t) and

η = λr 3 C π3/2m2 0 Z dk|˜µ(k)|2kx2e−k2r2C, (11)

where ˜µ(k) is the Fourier transform of the mass density. By exploiting equations (8)–(11) to evaluate the DNS, in the high temperature approximation we obtain the following identity: S(ω) = 2mγmkBT + ScCSL(ω) m2[(ω2 m− ω 2)2+ γ2 mω 2], (12) where ScCSL(ω) = 1 4πR dΩ E[{ ˜FcCSL(ω), ˜FcCSL(Ω)}] is the CSL contribution to the DNS.

In the standard CSL model, whose noise is white, one gets: SCSL= ~ 2 λr 3 C π3/2m2 0 Z dk|˜µ(k)|2kx2e−k2r2C, (13)

which is independent of ω, being the noise spectrum flat. Conversely, in its colored extension, one has:

ScCSL(ω) = SCSL× ˜f (ω). (14)

The frequency dependent contribution due to the noise correlation function potentially can lead to relevant mod-ifications of the bounds on the collapse parameters λ and rC.

3 Other non-interferometric tests

We briefly report the theoretical analysis of other two sig-nificant non-interferometric tests of collapse models: X-ray measurements [41] and low temperature measurements of phonon vibrations [27,28], which we will use to infer the bounds on the cCSL parameters. Also cold atom experi-ments [19] are significant for testing CSL; as they have been fully analyzed in [40], we refer to that paper for their derivation, and we will simply report the result in the following section.

3.1 X-ray emission

In testing CSL with X-ray measurements [39,41,48–51], the basic idea is that electrons and protons in the sam-ple material are accelerated by the collapse noise and emit radiation, which can be detected. Under suitable approximations the associated photon emission rate reads [34] dΓ (ω) dω = e2~η 2π2 0c3m2e 1 ω ˜ f (ω), (15)

where e is the unitary charge, 0the dielectric constant of

vacuum and c the speed of light, meis the electron mass

and η = λm2

e/(2m

2

0rC2). The standard CSL expression is

obtained by setting ˜f (ω) = 1. 3.2 Phonon excitation

Recently a novel way to test CSL was proposed, setting strong bounds on λ for rC< 10−6m [27,28]. The idea is

that the CSL noise modifies the phonon spectrum in a material, while heating it. This modification is quantified by the energy gain rate per mass unit:

dE dtdM = 3 4 ~2 r2 Cm 2 0 λeff, (16) where λeff= 2λr5 C 3π3/2 Z dq e−q2rC2q2f (ω˜ L(q)) (17)

with ωL(q) denoting the longitudinal phonon frequency,

which explicitly depends on the momentum q. In the white noise case λeff= λ, recovering the standard CSL result for

the energy heating [27].

4 Experimental bounds

(4)

experimental data. As an explicit example, we consider an exponentially decaying noise correlation function with correlation time Ω−1c :

f (t − s) = Ωc 2 e

−Ωc|t−s|. (18)

Such a choice, besides being physically reasonable, allows to easily recover the white noise limit for Ωc→ ∞. This

form of the correlation function was already considered in [37,40] in the context of colored modifications of collapse models.

For the optomechanical setting discussed in Section 2, our choice of noise corresponds to ScCSL(ω) in the Drude–

Lorentz form ScCSL(ω) = SCSL Ω2 c Ω2 c+ ω 2, (19)

where we see that Ωcplays the role of frequency cutoff on

the noise spectrum.

In a similar way, we can straightforwardly derive the photon emission rate for this specific time correlation function of the cCSL noise:

dΓ(ω) dω = e2 ~η 2π2 0c3m2e 1 ω Ω2 c Ω2 c + ω 2. (20)

Again, in the limit of Ωc→ +∞ we recover the expression

for the standard CSL.

As for CSL induced phonon excitation, having a col-ored noise leads to more involved modifications of the energy gain rate, as the frequency ωL(q) appearing in

equation (17) in general depends in a non-trivial way on the momentum. For a monoatomic crystal, the dispersion relation is [52]

ωL2(q) = 4C mA

sin2 12|q|a , (21) where mA is the atomic mass, C is the force constant

between the nearest-neighbor crystal planes, whose dis-tance a is of the order of 10−10m. In the limit of q  1/a, which is valid for rC 10−9m1, one can approximate the

sine with its argument. Thus, we obtain

ωL(q) ' vS|q|, (22)

where vS is the speed of sound in the crystal (Eq. (22) is

valid also for more general crystals). Using this expression, equation (17) becomes λeff= 4λr2 CΩ 2 c 3v2 S " 1 2− r2 CΩ 2 c v2 S +√πr 3 CΩ 3 c v3 S e r2CΩ2c v2 S erfc  rCΩc vS  # , (23) where erfc(x) = 1 − erf(x).

Equations (19), (20) and (23) give the theoretical pre-dictions for the cCSL model, which can be tested against 1The relation in equation (17), shows that the wave number

density is peaked near to ∼ rC.

experimental data. The resulting bounds are reported in Figure 1for different choices of Ωc.

4.1 Details of the experimental setups

Before discussing the results, we briefly describe the experimental setups, which we used in computing the bounds.

AURIGA is an aluminum cylindrical bar of length 3 m, radius 30 cm and mass 2300 cooled down at 4.2 K, whose resonant elongation is magnetically monitored at a fre-quency of ωm/2π ∼ 900 Hz [21]. The contribution to the

noise that can be atribuited to cCSL is ScCSL(ω) ' 1.4 ×

10−22N2/Hz at ω/2π = 931 Hz, and the corresponding

bound is reported in Figure1 in red.

LIGO is a Michelson interferometer, whose two 4 km arms are configured as a Fabry-Perot cavity [22]. At each extreme of the two arms, a cylindrical silica mir-ror (density 2200 kg/m3, radius 17 cm and length 20 cm)

are suspended and oscillate at a frequency below 1 Hz, while its noise is monitored in the 10 − 103Hz band. The

cCSL compatible contribution to the DNS is ScCSL(ω) '

9 × 10−27N2/Hz at ω/2π = 30 − 35 Hz, which constrains

the CSL parameters as reported in blue in Figure1. LISA Pathfinder is a space-based experiment which monitors the relative distance between two identical cubic masses (length 4.6 cm, average distance 37.6 cm, mass 1.928 kg) at low frequencies [24,25]. We can attribuite to cCSL a noise contribution of ScCSL(ω) = 3.15 ×

10−30N2/Hz just above the mHz regime. The

correspond-ing bound is highlighted in green in Figure1.

Two cantilever experiments were performed with masses of 3.5 × 10−13kg [20,29] and 1.2 × 10−10kg [10]. Here, we focus on the second experiment, which consists in a silica cantilever of dimension 450 × 57 × 2.5 µm3and

stiffness 0.4 N/m, to which is attached a ferromagnetical sphere of radius 15.5 µm and density 7.43 × 103kg/m3.

The harmonic motion of the latter, which is character-ized by a frequency ωm/2π = 8174.01 Hz, is monitored

with a SQUID. A CSL-like non-thermal contribution to the DNS was measured, taking the value ScCSL(ω) =

1.87 ± 0.16 aN2/Hz. Assuming that the measured extra noise is due to cCSL, the values of the cCSL parameters lie on the upper purple line in Figure1. Conversely, if such a noise can be attributed to standard sources, the exper-iment sets an upper bound corresponding with the lower purple line in figure.

The case of X-ray measurements [39,41,48–51] is slightly different from previous ones because it relies on a dif-ferent physical mechanism, the spontaneous emission of radiation rather than the Brownian motion. The expres-sion for the photon emisexpres-sion rate in equation (20) is compared with the experimental measure, which gives 4π2 0c3m20ω d Γ (ω)dω/e 2 ~ . 803 s−1m−2 [41]. The corre-sponding upper bound is reported in light blue in Figure1. In low temperature experiments [53] there is a residual heating of about 10−11 W/kg. This value should be com-pared with the energy rate in equation (16), where λeff is

estimated via equation (23) with vS= 3000 m/s (the speed

(5)

Fig. 1. Upper and lower bounds on the cCSL parameters λ and rC for three values of the frequency cutoff: Ωc= +∞

(top left panel) corresponding to standard CSL, Ωc= 1015Hz (top right), Ωc= 104Hz (bottom left), and Ωc= 101Hz (bottom

right). Red, blue and green lines (and respective shaded regions): upper bounds (and exclusion regions) from AURIGA, LIGO and LISA Pathfinder, respectively. Purple region: upper bound from cantilever experiment. Light blue region: upper bound from X-ray measurements. Orange and grey regions: upper bound from cold atom experiment [19,40] and from bulk heating experiments [27]. The black dashed line shows the lower bound based on theoretical arguments [15].

bound corresponding to these experiments is reported in grey in Figure1.

5 Discussion and conclusion

The bounds reported in Figure 1 refer to four values of the cutoff frequency Ωc: 1, 10

4, 1015s−1 and ∞, the latter

case corresponding to the standard CSL model.

For Ωc = 1015s−1, we notice the first change in the

parameter space due to the colored extension of the model. The bound from X-ray measurements becomes weaker: the reason is that frequency of the X-rays (ω ∼ 1019 s−1), at which the collapse noise is sampled, exceeds the cutoff frequency. The next to vanish is the bound on phonon excitations, which samples the noise at frequen-cies ∼1011s−1. Similarly, the bound from the cantilever

(6)

last panel of Figure1. The same happens for the bounds from AURIGA, LIGO and LISA Pathfinder when the cut-off frequency takes values Ωc . 103, 102 and 10−2s−1,

respectively. Eventually, for a cutoff at Ωc= 10

1s−1, the

only relevant bounds are those coming from cold atom experiments and LISA Pathfinder.

If one assumes that cCSL has a cosmological origin, which is reasonable considering the universality of the noise, a hint on the value of Ωc could come from the

behaviour of cosmological fields [6]. If one takes the cosmic microwave background (CMB) radiation, or the relic neu-trino background, one has Ωc∼ 1012s−1 [54]. For such a

value, the cCSL bound from X-rays is completely washed away. Thus, the bound from bulk heating effects [27,28] becomes the strongest. For Ωc = 10

11s−1 also the

lat-ter vanishes and the cold atom bound [19,40] prevails for rC < 10

−7m2, while for r

C > 10

−7m the strongest

bounds are provided by cantilever experiments and LISA Pathfinder.

We can conclude that low frequency, purely mechani-cal experiments provide the most robust bounds on the CSL parameters. These are resistant against changes in the spectrum of the noise, unless of course for some reason the low frequency part of the noise spectrum is suppressed, which cannot be identified at the present stage. However, this possibility would compromise the reduction process, which requires non-vanishing low frequency components [35,36].

The authors wish to thank SL Adler for useful comments on a preliminary draft of the paper. MC and AB acknowl-edge financial support from the University of Trieste (FRA 2016), INFN, the COST Action QTSpace (CA15220) and the H2020 FET project TEQ (grant n. 766900). AB acknowl-edges hospitality from the IAS Princeton, where part of this work was carried out and partial financial support from FQXi. LF acknowledges funding from the Royal Society under the Newton International Fellowship No. NF170345.

Author contribution statement

MC and AB conceived the presented idea. MC and LF developed the theory and derived the experimental bounds. AB coordinated the research project. All authors contributed to the writing of the manuscript.

Open Access This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

2We believe that the CSL contribution to the

posi-tion diffusion of the cold atom cloud in the colored case, described by equation (55) of [40], should be substituted with

3λA2~2 2m2r2 C h t3 2 − t2τ 2 + τ 2 τ − (t + τ )e−t/τi

, where A is the atomic number, m is the mass of the sigle atom and τ = Ω−1c . We used such

an expression to draw the corresponding bounds in Figure1.

References

1. E. Joos, H.D. Zeh, Z. Phys. B 59, 223 (1985) 2. W.H. Zurek, Phys. Today 44, 36 (1991)

3. E. Joos, H.D. Zeh, C. Kiefer, D.J. W. Giulini, J. Kupsch, I.-O. Stamatescu, Decoherence and the Appearance of a Classical World in Quantum Theory, 2nd edn. (Springer-Verlag, Berlin, Heidelberg, 2003)

4. M.A. Schlosshauer, Decoherence and the Quantum-To-Classical Transition, 1st edn. (Springer-Verlag, Berlin, Heidelberg, 2007)

5. A. Bassi, G.C. Ghirardi, Phys. Rep. 379, 257 (2003) 6. A. Bassi, D.-A. Deckert, L. Ferialdi, Europhys. Lett. 92,

50006 (2010)

7. G.C. Ghirardi, P. Pearle, A. Rimini, Phys. Rev. A 42, 78 (1990)

8. G.C. Ghirardi, A. Rimini, T. Weber, Phys. Rev. D 34, 470 (1986)

9. S.L. Adler, J. Phys. A 40, 2935 (2007)

10. A. Vinante et al., Phys. Rev. Lett. 119, 110401 (2017) 11. A. Bassi, C. Curceanu, Nucl. Phys. News 28, 37 (2018) 12. T. Kovachy et al., Nature 528, 530 (2015)

13. S. Eibenberger et al., Phys. Chem. Chem. Phys. 15, 14696 (2013)

14. K. Hornberger, J.E. Sipe, M. Arndt, Phys. Rev. A 70, 053608 (2004)

15. M. Toroˇs, G. Gasbarri, A. Bassi, Phys. Lett. A 381, 3921 (2017)

16. M. Toroˇs, A. Bassi, J. Phys. A 51, 115302 (2018) 17. K.C. Lee et al., Science 334, 1253 (2011)

18. S. Belli et al., Phys. Rev. A 94, 012108 (2016)

19. T. Kovachy et al., Phys. Rev. Lett. 114, 143004 (2015) 20. O. Usenko, A. Vinante, G. Wijts, T.H. Oosterkamp, Appl.

Phys. Lett. 98, 133105 (2011)

21. A. Vinante et al. (AURIGA Collaboration), Class. Quan-tum Grav. 23, S103 (2006)

22. B.P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collaboration), Phys. Rev. Lett. 116, 131103 (2016)

23. B.P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collaboration), Phys. Rev. Lett. 116, 061102 (2016)

24. M. Armano et al., Phys. Rev. Lett. 116, 231101 (2016) 25. M. Armano et al., Phys. Rev. Lett. 120, 061101 (2018) 26. C.E. Aalseth et al. (The IGEX Collaboration), Phys. Rev.

C 59, 2108 (1999)

27. S.L. Adler, A. Vinante, Phys. Rev. A 97, 052119 (2018) 28. M. Bahrami, Phys. Rev. A 97, 052118 (2018)

29. A. Vinante et al., Phys. Rev. Lett. 116, 090402 (2016) 30. M. Carlesso, A. Bassi, P. Falferi, A. Vinante, Phys. Rev.

D 94, 124036 (2016)

31. B. Helou, B.J.J. Slagmolen, D.E. McClelland, Y. Chen, Phys. Rev. D 95, 084054 (2017)

32. A. Smirne, B. Vacchini, A. Bassi, Phys. Rev. A 90, 062135 (2014)

33. A. Smirne, A. Bassi, Sci. Rep. 5, 12518 (2015) 34. J. Nobakht et al.,arXiv:1808.01143(2018)

is still in preparation status.

35. S.L. Adler, A. Bassi, J. Phys. A 40, 15083 (2007) 36. S.L. Adler, A. Bassi, J. Phys. A 41, 395308 (2008) 37. A. Bassi, L. Ferialdi, Phys. Rev. A 80, 012116 (2009) 38. L. Ferialdi, A. Bassi, Phys. Rev. A 86, 022108 (2012) 39. S.L. Adler, A. Bassi, S. Donadi, J. Phys. A 46, 245304

(7)

40. M. Bilardello, S. Donadi, A. Vinante, A. Bassi, Physica A 462, 764 (2016)

41. K. Piscicchia et al., Entropy 19, 319 (2017)

42. S. Nimmrichter, K. Hornberger, K. Hammerer, Phys. Rev. Lett. 113, 020405 (2014)

43. M. Bahrami et al., Phys. Rev. Lett. 112, 210404 (2014)

44. L. Di´osi, Phys. Rev. Lett. 114, 050403 (2015)

45. M. Carlesso et al., (2017) New J. Phys. 20, 083022 (2018)

46. C. Gardiner, P. Zoller, Quantum Noise (Springer-Verlag, Berlin, Heidelberg, 2004)

47. M. Paternostro et al., New J. Phys. 8, 107 (2006)

48. Q. Fu, Phys. Rev. A 56, 1806 (1997)

49. S.L. Adler, F.M. Ramazano˘glu, J. Phys. A 40, 13395 (2007)

50. A. Bassi, S. Donadi, Phys. Lett. A 378, 761 (2014) 51. S. Donadi, D.-A. Deckert, A. Bassi, Ann. Phys. 340, 70

(2014)

52. C. Kittel, Introduction to Solid State Physics, 8th edn. (John Wiley and Sons, Hoboken, NJ, 2005)

53. F. Pobell, Matter and Methods at Low Temperatures, 3rd edn. (Springer, Berlin, 2007)

Riferimenti

Documenti correlati

There- fore, we conclude that the identified targeted genes do contain NF-Y binding CCAAT boxes.. To further check whether the predicted CCAAT boxes were correctly evaluated,

We show that the TCM enables a faster convergence of the tuning process and that it can be used as a wavelength locking strategy to preserve the tuned condition of the filter in

The comparison is done in terms of average recharge cost gap (i.e. the percent increase of the overall energy cost due to the online approaches w.r.t. the optimal day-ahead

The inhibition of the proliferation of human thyroid carci- noma cells lines by myc-specific antisense oligonucleotides in- dicates a role of c-myc overexpression in

Supplementary Materials: The following are available online at http://www.mdpi.com/2072-6694/12/6/1507/s1 , Figure S1: Assessment of the predictivity of sncRNAs from plasma EVs and

As above described we also assessed the influence of the protecting group on the side chain of the adjacent amino acid on the nucleophilic attack initiated by amide backbone