Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Quantum limits to optical point-source localization

Open Access Open Access

Abstract

Motivated by the importance of optical microscopes to science and engineering, scientists have pondered for centuries how to improve their resolution and the existence of fundamental resolution limits. In recent years, a new class of microscopes that overcome a long-held belief about the resolution have revolutionized biological imaging. Termed “super-resolution” microscopy, these techniques work by accurately locating optical point sources from far field. To investigate the fundamental localization limits, here I derive quantum lower bounds on the error of locating point sources in free space, taking full account of the quantum, nonparaxial, and vectoral nature of photons. These bounds are valid for any measurement technique, as long as it obeys quantum mechanics, and serve as general no-go theorems for the resolution of microscopes. To arrive at analytic results, I focus mainly on the cases of one and two classical monochromatic sources with an initial vacuum optical state. For one source, a lower bound on the root-mean-square position estimation error is of the order of λ0/N, where λ0 is the free-space wavelength and N is the average number of radiated photons. For two sources, owing to the statistical effect of nuisance parameters, the error bound diverges when their radiated fields overlap significantly. The use of squeezed light to further enhance the accuracy of locating one classical point source and the localization limits for partially coherent sources and single-photon sources are also discussed. The theory presented establishes a rigorous quantum statistical inference framework for the study of super-resolution microscopy and points to the possibility of using quantum techniques for true resolution enhancement.

© 2015 Optical Society of America

1. INTRODUCTION

The resolution limit of optical microscopes is one of the most important problems in science and engineering. The Abbe–Rayleigh criterion with respect to the free-space wavelength λ0 has been a widely used resolution limit [1], but it is now well known that the criterion is heuristic in the context of microscopy and super-resolution microscopy is possible. An important class of super-resolution microscopy, including stimulated-emission-depletion microscopy [2] and photoactivated-localization microscopy [3], relies on the accurate localization of point sources from far field [4]. The localization accuracy, which represents an important measure of the microscope resolution, is then limited by the statistics of the optical measurement [58]. Prior analyses of point-source localization accuracy assumed classical, scalar, and paraxial optics with statistics specific to the measurement methods [58]. On a more fundamental level, however, optics is governed by the quantum theory of electromagnetic fields [9], and the existence of more accurate measurement methods [10] and more fundamental quantum limits remains an open question. For example, the superoscillation phenomenon [11] suggests that super-resolution diffraction patterns can be obtained at the expense of signal power; can it be exploited to improve the resolution of microscopes [1214]?

Using a quantum Cramér–Rao bound (QCRB) [15,16] and the full quantum theory of electromagnetic fields [9], here I derive quantum limits to the accuracy of locating point sources. These quantum resolution limits are more general and fundamental than prior classical analyses in the sense that they apply to any measurement method and take full account of the quantum, nonparaxial, and vectoral nature of photons. To arrive at analytic results, I focus mainly on the cases of one and two monochromatic classical sources and an initial vacuum optical state. The possibility of using squeezed light to further enhance the accuracy of locating one point source will also be discussed. To study partially coherent sources, I model incoherence using the concept of nuisance parameters, which are unknown parameters of no primary interest in the context of statistical inference. Quantum bounds for partially coherent sources are then derived by introducing a new generalized QCRB that accounts for nuisance parameters in a special way.

In quantum optics, there has been a substantial literature on quantum imaging (see, for example, Refs. [15,1738]), but most of them assume certain quantum optical states without considering how they may be generated by objects relevant to microscopy or consider simply the estimation of mirror displacement. Helstrom’s derivation of the QCRB for one point source [15,17] is the most relevant prior work, although he used the paraxial approximation, did not consider the use of squeezed light, and studied two sources only in the context of binary hypothesis testing [15]. There have also been intriguing claims of super-resolution using the nonclassical photon statistics from single-photon sources [3437], but their protocols have not been analyzed using statistical inference, so even though their images appear sharper, the accuracies of their methods in estimating object parameters remain unclear. To investigate their claims, I also derive a quantum bound for locating a single-photon source.

2. QUANTUM PARAMETER ESTIMATION

Let the initial quantum state of a system be |ψ. After unitary evolution U(X,T) with respect to Hamiltonian H(X,t) as a function of parameters X=(X1,X2,), the quantum system is measured with outcome Y. The probability distribution of Y according to Born’s rule can be expressed as [15,16,39]

P(Y|X)=tr[E(Y)U(X,T)|ψψ|U(X,T)],
where E(Y) is the positive operator-valued measure (POVM) that characterizes the quantum measurement and tr denotes the operator trace. Denote the estimator of X using Y as X˜(Y). The estimation error matrix is defined as
Σμν(X)dYP(Y|X)[X˜μ(Y)Xμ][X˜ν(Y)Xν].
For unbiased estimators, the classical Cramér–Rao bound states that [40]
Σ(X)j1(X),
which means that Σj1 is positive semidefinite. j(X) is the Fisher information matrix given by
jμν(X)dYP(Y|X)[θμlnP(Y|X)][θνlnP(Y|X)].
The bound has been used, for example, in Refs. [7,8], to evaluate the point-source localization accuracy for a microscope.

It turns out that for any POVM and thus any measurement in quantum mechanics, another lower bound exists in the form of the QCRB [15,41]:

Σ(X)j1(X)J1(X),
which means that j1J1 and ΣJ1 are positive semidefinite. J is the quantum Fisher information (QFI) matrix; it can be obtained by expressing the fidelity |ψ|U(X,T)U(X+δX,T)|ψ|2 in the interaction picture [42] and expanding it to the second order of δX [43,44]. The result is
Jμν(X)=4ReΔgμ(X)Δgν(X),
where Re denotes the real part, Aψ|A|ψ, ΔAAA, and
gμ(X)10TdtU(X,t)H(X,t)XμU(X,t)
is the generator of the parameter shift in the Heisenberg picture. For M trials, the QFI is simply multiplied by M and at least one component of the QCRB can be attained in an asymptotic M sense [45]. If one wishes to consider a new set of parameters θ related to the original set X and X can be expressed as a function of θ, then the new QFI matrix is simply given by
Jab(θ)=μ,vXμθaJμ,v(X)Xvθb|X=X(θ).
Various generalizations of the QCRB and alternatives are available [15,4649], but the version presented suffices to illustrate the pertinent physics. In Section 6, the QCRB will be generalized to a Bayesian version that treats nuisance parameters separately and is used to study partially coherent sources.

3. ONE CLASSICAL POINT SOURCE

Consider first a classical point source, as depicted in Fig. 1. The Hamiltonian is [9]

H(r,t)=HF+HI(r,t),
HF=sd3kω(k)a(k,s)a(k,s),
HI(r,t)=p(t)·E(r),
E(r)=sd3kω2(2π)3ϵ0[iε(k,s)a(k,s)eik·r+H.c.],
where k=kxx^+kyy^+kyz^ is a wave vector, (x^,y^,z^) denote unit vectors in the Cartesian coordinate system, d3kdkxdkydkz, s is an index for the two polarizations, ε(k,s) is a unit polarization vector, ω(k)=c|k| (c is the speed of light), p(t) is the c-number dipole moment of the source, r=xx^+yy^+zz^ is its position, ϵ0 is the free-space permittivity, a(k,s) is an annihilation operator obeying the commutation relation [a(k,s),a(k,s)]=δssδ3(kk), and H.c. denotes the Hermitian conjugate. Since p(t) is a c-number, HI implements a field displacement operation [9]. The Heisenberg picture of a(k,s) is
a(k,s,t)U(X,t)a(k,s)U(X,t),
=eiωt[a(k,s)+α(k,s,r,t)],
where α is the radiated field. Assuming p(t)=p0eiω0t+c.c., where c.c. denotes the complex conjugate, α becomes
α(k,s,r,T)=ω02(2π)3ϵ0eik·rε*(k,s)·[p0ei(ωω0)T/2sin(ωω0)T/2(ωω0)/2+p0*ei(ω+ω0)T/2sin(ω+ω0)T/2(ω+ω0)/2],
which indicates that only optical modes with ω(k)=ω0 grow in time, corresponding to the far field, whereas all the other near-field modes with ω(k)ω0 oscillate. The ω(k)=ω0 relation specifies the spatial frequencies available to the far optical fields [50,51]. Assuming T2π/ω0, such that sin2[(ω±ω0)T/2]/[(ω±ω0)/2]22πTδ(ω±ω0), using the identity sεμ(k,s)εν*(k,s)=δμνkμkν/|k|2 [9], and switching to the spherical coordinate system for k, it can be shown that the average number of radiated photons for an initial vacuum state is
Nsd3k|α(k,s,r,T)|2|p0|2ω03T3πϵ0c3.

 figure: Fig. 1.

Fig. 1. Classical point source with dipole moment p(t) radiating in free space. Its position r is estimated by measuring the quantum optical field, with a(k,s) denoting its annihilation operator.

Download Full Size | PDF

The far-field limit (ω0T) will be assumed hereafter.

I now focus on two representative cases: a linearly polarized dipole with p0=p0z^ and a circularly polarized dipole with p0=p0(x^+iy^)/2. Taking the unknown parameters to be r, the generators for X=(x,y,z) can be expressed, after some algebra, as

Δgμ(r)=2WμΔPμ(r),μ{x,y,z},
ΔPμ(r)12i[Δbμ(r)Δbμ(r)],
where Δbμ is a normalized annihilation operator defined as
Δbμ(r)Wμsd3k[ikμα*(k,s,r,T)]Δa(k,s),
such that [Δbμ(r),Δbν(r)]=δμν, and the normalization constants Wμ are
Wμ[sd3kkμ2|αμ(k,s,r,T)|2]1/2.
The d3k integrals can again be computed with the help of the far-field assumption and spherical coordinates. The results depend on p0; for p0=p0z^,
Wx=Wy=52λ02πN,Wz=5λ02πN,
and for p0=p0(x^+iy^)/2,
Wx=Wy=103λ02πN,Wz=52λ02πN,
but the important point here is that they are all of the order of λ0/N, where λ02πc/ω0 is the free-space wavelength. The QFI becomes
Jμν(r)=8Wμ2ΔPμ(r)ΔPν(r).
For an initial vacuum state (or any coherent state), ΔPμ(r)ΔPν(r)=δμν/2 and the QCRB is hence
Jμν(r)=4Wμ2δμν,Σμμ(r)Wμ24,
meaning that the quantum resolution limit in terms of the root-mean-square error Σμμ is of the order of λ0/N. I call this limit the quantum shot-noise limit. Generalization to lossless media is straightforward and results simply in λ0 being replaced by the wavelength in the medium. Section 9 shows that a single-photon source also obeys this limit with repeated trials.

Assuming uncorrelated photons, Refs. [57,25] derived a similar limit in the form of σ/N, where σ is the width of the imaging point-spread function. While their limit also scales as 1/N, all those analyses assume the paraxial approximation and measurement by a photon-counting camera, whereas the quantum shot-noise limit here is valid for any numerical aperture and any measurement, including common methods such as photon counting, homodyne/heterodyne detection, and digital holography. The limit in Refs. [57,25] also implies that the quantum shot-noise limit is reasonably tight, as the camera measurement with suitable postprocessing can at least follow the quantum-optimal shot-noise scaling. Section 4 shows that homodyne measurement with a special local-oscillator field can also approach the quantum limit if the radiation is coherent.

For a concrete numerical example, consider the semiclassical paraxial analysis of conventional single-molecule microscopy by Ober et al. [7] who used the classical Cramér–Rao bound and a shot-noise assumption to derive a limit of 2.301 nm on the root-mean-square localization error for a free-space wavelength of 520 nm, a numerical aperture of 1.4, a photon collection efficiency of 0.033, a photon flux of 2×106s1, and an acquisition time of 0.01 s. If the efficiency were 1, their limit would become 0.418 nm. In comparison, if I take the refractive index of the immersion oil to be 1.52, λ0=520nm/1.52=342nm to be the wavelength in the medium, and the photon number to be N=2×104, then the quantum shot-noise limit according to Eqs. (21), (22), and (24) is λ0/(2πN)=0.385nm times a constant factor close to 1.

It remains to be seen whether superoscillation techniques are similarly efficient, but the key point here is that since the quantum bound is valid for any measurement and conventional methods can already get close to it, no other measurement technique is able to offer any significant advantage in resolution enhancement over the conventional methods.

4. QUANTUM ENHANCEMENT

Even though the source is classical, quantum enhancement is possible if the initial state |ψ is nonclassical, as I now show. If Δgμ were independent of the parameter, the accuracy could be enhanced by squeezing and measuring the conjugate quadrature [52]. Although Δgμ(r) depends on the unknown r here, the radiated field can be approximated as α(k,s,r,T)α(k,s,r0,T), resulting in Δgμ(r)Δgμ(r0), provided

|rr0|λ0,
with respect to a known reference position r0. The acquisition of such prior information will require a fixed amount of overhead resource; but once it is done, one can squeeze the quadrature,
ΔQμ(r0)12[Δbμ(r0)+Δbμ(r0)],
in the initial state and perform a homodyne measurement of ΔQμ(r0) to estimate r much more accurately. Since [ΔQμ(r0),ΔQν(r0)]=0, all three quadratures can be squeezed and measured simultaneously in principle. The estimation error becomes
Σμμ(r)Wμ22ΔQμ2(r0)
and the error reduction below the shot-noise limit is determined by the squeezing factor, which is limited by the average photon number N0 in the initial state (not to be confused with N). Using ΔQμ2(r0)+ΔPμ2(r0)2N0+1 and the uncertainty relation ΔQμ2ΔPμ21/4, it can be shown that
ΔQμ2(r0)f(N0)2,ΔPμ2(r0)12f(N0),
f(N0)(2N0+1)[11(2N0+1)2],
where
f(0)=1,f(N0)14N0forN01.
With a zero-mean minimum-uncertainty state and all initial photons in the Δbμ(r0) mode, the estimation error becomes
Σμμ(r)Wμ22ΔQμ2(r0)=Wμ24f(N0).
The enhancement factor f(N0) is optimal as the QCRB can be further bounded by
Σμμ(r)Jμμ1(r)=Wμ28ΔPμ2(r)Wμ24f(N0).
This means that squeezed light with average photon number N0 can beat the quantum shot-noise limit to the mean-square error by roughly a factor of N0.

The optical mode to be squeezed has a profile ikμα(k,s,r0,T). This means that, in real space, the electric field profile of the mode should be the spatial derivative of the radiated field. This kind of squeezing and measurement has actually been demonstrated experimentally, albeit in the paraxial regime, by Taylor et al. in the context of particle tracking [38] where the weak scatterer under a strong pump can be modeled as a classical source, similar to the implementation of field displacement by a beam splitter [53], and the spatial mode profile of the squeezed light and the local oscillator is a spatial derivative of the scattered field. To realize an enhancement in practice, accurate phase locking of the squeezed light and the local oscillator to the radiated field and a high measurement efficiency are crucial. Phase locking cannot be achieved with incoherent point sources such as fluorescent markers, but can be done with dielectric particles via Rayleigh scattering [38] or second-harmonic nanoparticles [54,55]; the latter are especially promising for biological imaging applications [56].

5. TWO CLASSICAL POINT SOURCES

Next, consider two classical point sources at r and r, as shown in Fig. 2. The Hamiltonian is now

H(r,r,t)=HF+HI(r,r,t),
HI(r,r,t)=p(t)·E(r)p(t)·E(r).

 figure: Fig. 2.

Fig. 2. Two classical point sources with dipole moments p(t) and p(t) at r and r, respectively, with quantum optical radiation.

Download Full Size | PDF

The Heisenberg picture of a(k,s) becomes a(k,s,t)=eiωt[a(k,s)+α(k,s,r,t)+α(k,s,r,t)], where α and α are the radiated fields from the two sources, α is the same as before, and α has the same expression as α, except that p is replaced by p and r by r. One can then follow the preceding procedure to obtain the QCRB for estimating r and r. To highlight the important physics, however, consider here the estimation of just two parameters X=(x,x). The generators Δgx and Δgx may not commute and the QFI matrix for an initial vacuum or any coherent state now has off-diagonal components,

Jxx(X)=Jxx(X)=4Resd3kkx2α*(k,s,r,T)α(k,s,r,T),
whereas Jxx remains the same and Jxx has a similar expression to Jxx. Jxx and Jxx still obey a shot-noise scaling with the average photon number, but the nonzero off-diagonal components mean that the parameters act as nuisance parameters to each other and the QCRB with respect to, say, x is always raised:
Σxx(X)1Jxx[1κ(X)],
where the resolution degradation factor, defined as
κ(X)Jxx2(X)JxxJxx=(Resd3kkx2α*α)2sd3kkx2|α|2sd3kkx2|α|2,
is within the range 0κ1 and is determined by the overlap between the two differential mode profiles. The nuisance-parameter effect generalizes the Rayleigh criterion and other classical results [8] by revealing a fundamental measurement-independent degradation of resolution for two point sources with overlapping radiation. For example, Fig. 3 plots κ against |xx|/λ0, assuming p=p=p0eiω0t+c.c., T2π/ω0, p0=p0x^, y=y, and z=z. κ0 for |xx|λ0, as expected, but it approaches 1 and leads to a diverging QCRB when |xx|λ0. Section 8 shows that the degradation effect should still exist for two partially coherent sources.

 figure: Fig. 3.

Fig. 3. Plot of the resolution degradation factor κ versus the true separation |xx|/λ0 between two in-phase point sources, assuming p=p=p0eiω0t+c.c., T2π/ω0, p0=p0x^, y=y, and z=z. At |xx|=0, the Fisher information matrix is singular [57,58]. κ remains the same for two out-of-phase sources with otherwise the same assumptions.

Download Full Size | PDF

The degradation effect can be avoided by minimizing the overlap before each source is located independently. The overlap can be reduced by making the radiated fields separate in space, time, frequency, quadrature, or polarization; the time multiplexing of point sources has especially been the key driver in current super-resolution microscopy [24].

Note that κ also depends on the relative phase between α and α. For example, under the assumptions in the caption of Fig. 3, it can be shown that the QFI matrix transformed with respect to the average position (x+x)/2 and the separation xx is diagonal. The QFI component with respect to the average position still obeys a shot-noise scaling, while the increase in κ can be traced to the increased error in estimating the separation. Similarly, when α and α are 180° out of phase but otherwise obey the same assumptions, κ remains the same but its increase is now due to the increased error for the average position. An interesting scenario occurs when α and α are 90° out of phase: the two fields are in orthogonal quadratures, κ becomes 0, and they can be measured separately using homodyne or heterodyne detection. This phenomenon suggests that structured illumination [51,5961] can be used to excite the sources, such that their relative phase and amplitude can be controlled to some degree and the overlap can be reduced for certain ranges of parameters.

When the overlap is unavoidable or when the generators do not commute, heterodyne measurements can still be used to measure both quadratures of a(k,s,T) and should have a classical Fisher information within a factor of 1/2 of the QFI. Quantum enhancement may also be possible using entangled squeezed states [62]; the specific experimental design will be left for future studies.

6. BAYESIAN QCRB WITH NUISANCE PARAMETERS

Incoherent sources are characterized by the statistical fluctuations of the fields [9]. For point-source radiation, incoherence originates from the randomness of the amplitude, phase, and direction of the point dipole. In the context of statistical inference, these random variables are most suitably modeled as nuisance parameters. There are many ways to generalize the Cramér–Rao bounds when nuisance parameters are present [63]. The previous sections show one way, which includes the nuisance parameters as part of the wanted parameters X. To derive tighter bounds for other types of nuisance parameters, here I start with a Bayesian QCRB and generalize a classical approach by Miller and Chang [63,64]. Let Z be a set of nuisance parameters and suppose first that Z is given. The Bayesian estimation error matrix is

Σ¯μν(Z)dXdYP(Y|X,Z)PX|Z(X|Z)[X˜μ(Y,Z)Xμ]×[X˜ν(Y,Z)Xν],
where PX|Z(X|Z) is the prior distribution of X conditioned on Z. Note that this Bayesian definition of error regards X as a random parameter by averaging over its prior and is different from the frequentist definition in Eq. (2) [40]. A Bayesian QCRB valid for any estimator is given by [46,47]
Σ¯(Z)J¯1(Z),
J¯(Z)=EX|Z[J(X|Z)]+j(Z),
where J(X|Z) is the same QFI as before, except that it is now conditioned on Z, EX|Z denotes expectation over PX|Z(X|Z), and j(Z) is a prior Fisher information defined as
jμν(Z)dXPX|Z(X|Z)[XμlnPX|Z(X|Z)]×[XνlnPX|Z(X|Z)].
If Z is a random parameter with prior distribution given by PZ(Z), the estimation error is
ΠμνEZ[Σ¯(Z)],
Σ¯(Z)dXdYP(Y|X,Z)PX|Z(X|Z)[X˜μ(Y)Xμ]×[X˜ν(Y)Xν],
where EZ denotes expectation over PZ and the estimator X˜μ(Y) can no longer depend on Z. The lower bound in Eq. (39) still holds for Σ¯(Z), so one can obtain a lower bound on Π given by
ΠEZ[J¯1(Z)].
The important mathematical feature of this bound is that the expectation with respect to the nuisance parameter Z is taken after the inverse of the conditional QFI matrix. This can sometimes lead to a tighter bound than a QCRB that includes Z as part of X. Note also that this Bayesian bound is valid for any estimator, not just the unbiased ones, unlike the claim in Ref. [64]. The tightness of the bound should depend on whether the nuisance parameters can be accurately estimated from the measurements.

7. ONE PARTIALLY COHERENT SOURCE

I now use the new bound to study partially coherent sources. First, consider the example of one point source in Section 3, but suppose that the complex dipole amplitude p0 is unknown. Assuming Z=p0, the quantum state before measurement is

ρ=d2p0PZ(p0)U(X,p0,T)ρ0U(X,p0,T).
If ρ0 is a vacuum state, Uρ0U is a coherent state and ρ is a classical mixed state of light with PZ(p0) determining the Sudarshan–Glauber P function [9]. The random p0, therefore, gives rise to a classical partially coherent source model. For an initial vacuum, J¯(p0) is given by
J¯μν(p0)=4Wμ2(p0)δμν+jμν=N(p0)Cμλ02δμν+jμν,
where Wμ and N now depend on the unknown dipole moment and Cμ is a constant of the order of 1 that can be determined from Eqs. (21) or (22). Assuming that j is diagonal and independent of p0 and taking the inverse and then the expectation according to Eq. (44), I obtain
ΠμμEp0[14/Wμ2(p0)+jμμ]=Ep0[1N(p0)/(Cμλ02)+jμμ].
For example, if PZ(p0) corresponds to the P function of a thermal source, then the bound can be written in terms of the average radiated photon number N¯ as
Ep0[1N(p0)/(Cμλ02)+jμμ]=Cμλ02N¯0dNexp(NN¯)1N+Cμλ02jμμ,
Cμλ02N¯lnN¯Cμλ02jμμ,N¯Cμλ02jμμ.
An alternative Bayesian QCRB can be obtained by including p0 as part of X. In that case, the off-diagonal QFI elements between p0 and r turn out to be zero, and the expectation with respect to p0 is taken before the inverse, leading to a bound given by Cμλ02/N¯. The separate treatment of p0 as a nuisance parameter here involves taking the expectation after the inverse, giving rise to an additional factor lnN¯ and, thus, a tighter bound for a large N¯. In general, Jensen’s inequality can be used to show that the shot-noise scaling with respect to the average photon number Ep0[N(p0)] cannot be beat for any non-negative P function.

8. TWO PARTIALLY COHERENT SOURCES

Next, consider the example of two point sources in Section 5, and Z=(p0,p0) is now assumed to be unknown to model partially coherent sources. Assuming again that j is diagonal and X=(x,x) is independent of Z,

J¯xx(p0)=4Wx2(p0)+jxx,
J¯xx(p0)=4Wx2(p0)+jxx,
J¯xx(p0,p0)=EX[Jxx(p0,p0)].
J¯xx is now the expectation of Jxx over X=(x,x), conditioned on the dipole moments. If the two point sources are a priori known to be close relative to λ0, then J¯xx(p0,p0) can still have a significant magnitude for certain (p0,p0). The bound given by Eq. (44) becomes
ΠxxE(p0,p0){1J¯xx(p0)[1κ¯(p0,p0)]},
where a new resolution degradation factor is defined as
κ¯(p0,p0)J¯xx2(p0,p0)J¯xx(p0)J¯xx(p0).
The important point here is that κ¯(p0,p0) can still be close to 1 for certain values of (p0,p0) if the two sources are a priori known to be close to each other relative to λ0, 1/[1κ¯(p0,p0)]1 is possible, and the expectation over (p0,p0) will then be dominated by those large values. In other words, the resolution degradation effect derived for coherent sources must still exist for partially coherent sources if their radiated fields may have significant overlap.

An alternative Bayesian QCRB that includes (p0,p0) as part of X can again be computed, but at some stage it involves taking the expectation of J¯xx with respect to (p0,p0) before the inverse is taken. For incoherent sources, this can reduce the off-diagonal components significantly; the resulting bound, while still valid, would be less tight and no longer demonstrate the resolution degradation effect.

9. SINGLE-PHOTON SOURCE

Consider now an initially excited two-level atom in free space. A detailed analysis of atom–photon interaction is formidable [9,65], but when the initial optical state is vacuum, spontaneous emission can be treated more easily, as the atom must decay to ground state in the long-time limit and the final optical state must contain exactly one photon. Using the continuous Fock space [9], the final optical state in the Schrödinger picture can be written with respect to the vacuum |0 as

|Ψ=c|0,csd3kϕ(k,s)a(k,s),
where
ϕ(k,s)=k,s|Ψ=0|a(k,s)|Ψ
is the one-photon configuration-space amplitude. Following Chap. III.C of Ref. [65], it can be expressed as
ϕ(k,s)=k,s|g|HI|e|0[(ωω˜0)+i/(2T1)]eiωT,
HI=iω0(μ12σμ12*σ)·A(r),
A(r)=sd3k2(2π)3ωϵ0[a(k,s)ε(k,s)eik·r+h.c.],
T1=3πϵ0c3|μ12|2ω03,
where |e and |g are the excited and ground atomic states, respectively, ω0 is the atomic resonance frequency, ω˜0 is the Lamb-shifted atomic frequency, T1 is the decay time, μ12 is the off-diagonal atomic dipole moment, and σ|ge| is the atomic lowering operator. The result is
ϕ(k,s)=1i(ωω˜0)+1/(2T1)ω022(2π)3ωϵ0μ12·ε*(k,s)×eik·riωT.
Consider the fidelity
F=|0|c(X)c(X+δX)|0|2,
=|[c(X),c(X+δX)]|2,
1+μ,νδXμδXν{Re[c,2cXμXν]+Im[c,cXμ]Im[c,cXν]},
where the fact
Re0|ccXμ|0=Re[c,cXμ]=0,
due to 0|c(X+δX)c(X+δX)|0=0|c(X)c(X)|0=1, has been used. It can further be shown that
[c,cXμ]=sd3kϕ*(k,s)ϕ(k,s)Xμ=0,
leading to a QFI given by
Jμν=4Re[c,2cXμXν],
=4Resd3kϕ*(k,s)2ϕ(k,s)XμXν,
=4Resd3kkμkν(ωω˜0)2+1/(4T12)ω022(2π)3ωϵ0×|μ12·ε*(k,s)|2.
Assuming that the decay time is much longer than the optical period (T12π/ω˜0) and the Lamb shift is much smaller than the optical ω0 (ω˜0ω0), the QFI becomes
Jμν4Resd3kkμkν2πT1δ(ωω0)ω022(2π)3ωϵ0×|μ12·ε*(k,s)|2.
This turns out to be identical to the QFI derived in Section 3 for an N=1 classical source:
Jμν=4δμνNWμ2δμνλ02,ΣμμJμμ1=NWμ24λ02,
where N is defined in Eq. (16) and Wμ is defined in Eq. (20) and given by Eqs. (21) or Eqs. (22), such that NWμ2 is of the order of λ02. This result shows that a single-photon source offers no fundamental advantage over a classical source that emits one photon on average. Super-resolution beyond the classical Abbe–Rayleigh limit can still be obtained, however, if the experiment can be repeated. The QFI is then multiplied by the number of trials M, which is also the total number of emitted photons, and the resulting QCRB is identical to that for a classical source with M replacing N. The experiments reported in Refs. [3537] certainly involved a large number of measurements of many photons in total, which can explain the apparent super-resolution, but it remains to be seen whether their methods are accurate or efficient in estimating object parameters.

For two atoms with a large separation (|rr|λ0), the one-atom analysis is expected to be applicable to each atom independently. The analysis of two close atoms is much more challenging because of atomic cooperative effects such as Dicke super-radiance [9] and Förster resonance energy transfer [4]. Beyond the current assumption of spontaneous emission, it will also be interesting, although highly nontrivial, to analyze the interaction between two-level atoms and other states of light, such as coherent states or squeezed states, and investigate their quantum localization limits and the possibility of quantum enhancement.

10. CONCLUSION

I have derived quantum limits to point-source localization using quantum estimation theory and the quantum theory of electromagnetic fields. These results not only provide general no-go theorems concerning microscope resolution, but should also motivate further progress in microscopy through classical or quantum techniques beyond the current assumptions. For example, the theory presented may be applied to other more exotic quantum states of light interacting with quantum sources, such as multilevel atoms [9,66], quantum dots [35,67], diamond defects [36,37,68], and second-harmonic nanoparticles [5456]. It is also possible to generalize the current formalism for open quantum systems [6972] to account for mixed states, decoherence, optical losses, background noises, and imperfect measurement efficiency. The phenomenon of fluorescence intermittency or blinking represents another interesting challenge to the statistical analysis, as a fluorescent source can turn itself on and off randomly and the localization of two blinking sources offers another route to super-resolution [73]. If done with care, the application of quantum information science to microscopy is destined to yield sound insights and opportunities in both fields.

Funding

National Research Foundation-Prime Minister’s office, Republic of Singapore (NRF) (NRF-NRFF2011-07).

REFERENCES

1. M. Born and E. Wolf, Principles of Optics: Electromagnetic Theory of Propagation, Interference and Diffraction of Light (Cambridge University, 1999).

2. S. W. Hell, “Far-field optical nanoscopy,” Science 316, 1153–1158 (2007). [CrossRef]  

3. E. Betzig, G. H. Patterson, R. Sougrat, O. W. Lindwasser, S. Olenych, J. S. Bonifacino, M. W. Davidson, J. Lippincott-Schwartz, and H. F. Hess, “Imaging intracellular fluorescent proteins at nanometer resolution,” Science 313, 1642–1645 (2006). [CrossRef]  

4. W. E. Moerner, “New directions in single-molecule imaging and analysis,” Proc. Natl. Acad. Sci. USA 104, 12596–12602 (2007).

5. N. Bobroff, “Position measurement with a resolution and noise-limited instrument,” Rev. Sci. Instrum. 57, 1152–1157 (1986). [CrossRef]  

6. R. E. Thompson, D. R. Larson, and W. W. Webb, “Precise nanometer localization analysis for individual fluorescent probes,” Biophys. J. 82, 2775–2783 (2002). [CrossRef]  

7. R. J. Ober, S. Ram, and E. S. Ward, “Localization accuracy in single-molecule microscopy,” Biophys. J. 86, 1185–1200 (2004). [CrossRef]  

8. S. Ram, E. S. Ward, and R. J. Ober, “Beyond Rayleigh’s criterion: a resolution measure with application to single-molecule microscopy,” Proc. Natl. Acad. Sci. USA 103, 4457–4462 (2006).

9. L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University, 1995).

10. C. J. R. Sheppard, “The optics of microscopy,” J. Opt. A 9, S1 (2007). [CrossRef]  

11. M. V. Berry and S. Popescu, “Evolution of quantum superoscillations and optical superresolution without evanescent waves,” J. Phys. A 39, 6965–6977 (2006). [CrossRef]  

12. N. I. Zheludev, “What diffraction limit?” Nat. Mater. 7, 420–422 (2008). [CrossRef]  

13. F. M. Huang and N. I. Zheludev, “Super-resolution without evanescent waves,” Nano Lett. 9, 1249–1254 (2009). [CrossRef]  

14. H. J. Hyvärinen, S. Rehman, J. Tervo, J. Turunen, and C. J. R. Sheppard, “Limitations of superoscillation filters in microscopy applications,” Opt. Lett. 37, 903–905 (2012). [CrossRef]  

15. C. W. Helstrom, Quantum Detection and Estimation Theory (Academic, 1976).

16. A. S. Holevo, Statistical Structure of Quantum Theory (Springer-Verlag, 2001).

17. C. W. Helstrom, “Estimation of object parameters by a quantum-limited optical system,” J. Opt. Soc. Am. 60, 233–239 (1970). [CrossRef]  

18. M. I. Kolobov, “The spatial behavior of nonclassical light,” Rev. Mod. Phys. 71, 1539–1589 (1999). [CrossRef]  

19. C. Fabre, J. B. Fouet, and A. Maître, “Quantum limits in the measurement of very small displacements in optical images,” Opt. Lett. 25, 76–78 (2000). [CrossRef]  

20. N. Treps, N. Grosse, W. P. Bowen, C. Fabre, H.-A. Bachor, and P. K. Lam, “A quantum laser pointer,” Science 301, 940–943 (2003). [CrossRef]  

21. S. Barnett, C. Fabre, and A. Maıtre, “Ultimate quantum limits for resolution of beam displacements,” Eur. Phys. J. D 22, 513–519 (2003).

22. A. N. Boto, P. Kok, D. S. Abrams, S. L. Braunstein, C. P. Williams, and J. P. Dowling, “Quantum interferometric optical lithography: exploiting entanglement to beat the diffraction limit,” Phys. Rev. Lett. 85, 2733–2736 (2000). [CrossRef]  

23. H. Li, V. A. Sautenkov, M. M. Kash, A. V. Sokolov, G. R. Welch, Y. V. Rostovtsev, M. S. Zubairy, and M. O. Scully, “Optical imaging beyond the diffraction limit via dark states,” Phys. Rev. A 78, 013803 (2008). [CrossRef]  

24. R. W. Boyd and J. P. Dowling, “Quantum lithography: status of the field,” Quantum Inf. Process. 11, 891–901 (2012). [CrossRef]  

25. M. Tsang, “Quantum imaging beyond the diffraction limit by optical centroid measurements,” Phys. Rev. Lett. 102, 253601 (2009). [CrossRef]  

26. H. Shin, K. W. C. Chan, H. J. Chang, and R. W. Boyd, “Quantum spatial superresolution by optical centroid measurements,” Phys. Rev. Lett. 107, 083603 (2011). [CrossRef]  

27. L. A. Rozema, J. D. Bateman, D. H. Mahler, R. Okamoto, A. Feizpour, A. Hayat, and A. M. Steinberg, “Scalable spatial superresolution using entangled photons,” Phys. Rev. Lett. 112, 223602 (2014). [CrossRef]  

28. V. Giovannetti, S. Lloyd, L. Maccone, and J. H. Shapiro, “Sub-Rayleigh-diffraction-bound quantum imaging,” Phys. Rev. A 79, 013827 (2009). [CrossRef]  

29. R. Nair and B. J. Yen, “Optimal quantum states for image sensing in loss,” Phys. Rev. Lett. 107, 193602 (2011). [CrossRef]  

30. S. Pirandola, “Quantum reading of a classical digital memory,” Phys. Rev. Lett. 106, 090504 (2011). [CrossRef]  

31. C. A. Pérez-Delgado, M. E. Pearce, and P. Kok, “Fundamental limits of classical and quantum imaging,” Phys. Rev. Lett. 109, 123601 (2012). [CrossRef]  

32. P. R. Hemmer and T. Zapata, “The universal scaling laws that determine the achievable resolution in different schemes for super-resolution imaging,” J. Opt. 14, 083002 (2012). [CrossRef]  

33. P. C. Humphreys, M. Barbieri, A. Datta, and I. A. Walmsley, “Quantum enhanced multiple phase estimation,” Phys. Rev. Lett. 111, 070403 (2013). [CrossRef]  

34. O. Schwartz and D. Oron, “Improved resolution in fluorescence microscopy using quantum correlations,” Phys. Rev. A 85, 033812 (2012). [CrossRef]  

35. O. Schwartz, J. M. Levitt, R. Tenne, S. Itzhakov, Z. Deutsch, and D. Oron, “Superresolution microscopy with quantum emitters,” Nano Lett. 13, 5832–5836 (2013). [CrossRef]  

36. J.-M. Cui, F.-W. Sun, X.-D. Chen, Z.-J. Gong, and G.-C. Guo, “Quantum statistical imaging of particles without restriction of the diffraction limit,” Phys. Rev. Lett. 110, 153901 (2013). [CrossRef]  

37. D. Gatto Monticone, K. Katamadze, P. Traina, E. Moreva, J. Forneris, I. Ruo-Berchera, P. Olivero, I. P. Degiovanni, G. Brida, and M. Genovese, “Beating the Abbe diffraction limit in confocal microscopy via nonclassical photon statistics,” Phys. Rev. Lett. 113, 143602 (2014). [CrossRef]  

38. M. A. Taylor, J. Janousek, V. Daria, J. Knittel, B. Hage, H.-A. Bachor, and W. P. Bowen, “Biological measurement beyond the quantum limit,” Nat. Photonics 7, 229–233 (2013). [CrossRef]  

39. H. M. Wiseman and G. J. Milburn, Quantum Measurement and Control (Cambridge University, 2010).

40. H. L. Van Trees, Detection, Estimation, and Modulation Theory, Part I (Wiley, 2001).

41. S. L. Braunstein and C. M. Caves, “Statistical distance and the geometry of quantum states,” Phys. Rev. Lett. 72, 3439–3443 (1994). [CrossRef]  

42. M. Tsang and R. Nair, “Fundamental quantum limits to waveform detection,” Phys. Rev. A 86, 042115 (2012). [CrossRef]  

43. M. G. A. Paris, “Quantum estimation for quantum technology,” Int. J. Quantum Inform. 7, 125–137 (2009). [CrossRef]  

44. A. De Pasquale, D. Rossini, P. Facchi, and V. Giovannetti, “Quantum parameter estimation affected by unitary disturbance,” Phys. Rev. A 88, 052117 (2013). [CrossRef]  

45. A. Fujiwara, “Strong consistency and asymptotic efficiency for adaptive quantum estimation problems,” J. Phys. A 39, 12489–12504 (2006). [CrossRef]  

46. H. P. Yuen and M. Lax, “Multiple-parameter quantum estimation and measurement of nonselfadjoint observables,” IEEE Trans. Inform. Theory 19, 740–750 (1973). [CrossRef]  

47. M. Tsang, H. M. Wiseman, and C. M. Caves, “Fundamental quantum limit to waveform estimation,” Phys. Rev. Lett. 106, 090401 (2011). [CrossRef]  

48. M. Tsang, “Ziv–Zakai error bounds for quantum parameter estimation,” Phys. Rev. Lett. 108, 230401 (2012). [CrossRef]  

49. D. W. Berry, M. Tsang, M. J. W. Hall, and H. M. Wiseman, “The quantum Bell–Ziv–Zakai bounds and Heisenberg limits for waveform estimation,” arXiv:1409.7877 (2014).

50. J. W. Goodman, Introduction to Fourier Optics (McGraw-Hill, 2004).

51. R. Heintzmann and M. G. L. Gustafsson, “Subdiffraction resolution in continuous samples,” Nat. Photonics 3, 362–364 (2009). [CrossRef]  

52. S. L. Braunstein, C. M. Caves, and G. J. Milburn, “Generalized uncertainty relations: theory, examples, and Lorentz invariance,” Ann. Phys. 247, 135–173 (1996). [CrossRef]  

53. M. G. Paris, “Displacement operator by beam splitter,” Phys. Lett. A 217, 78–80 (1996). [CrossRef]  

54. Y. Pu, M. Centurion, and D. Psaltis, “Harmonic holography: a new holographic principle,” Appl. Opt. 47, A103–A110 (2008). [CrossRef]  

55. C.-L. Hsieh, R. Grange, Y. Pu, and D. Psaltis, “Three-dimensional harmonic holographic microcopy using nanoparticles as probes for cell imaging,” Opt. Express 17, 2880–2891 (2009). [CrossRef]  

56. W. P. Dempsey, S. E. Fraser, and P. Pantazis, “SHG nanoprobes: advancing harmonic imaging in biology,” BioEssays 34, 351–360 (2012). [CrossRef]  

57. P. Stoica and T. L. Marzetta, “Parameter estimation problems with singular information matrices,” IEEE Trans. Signal Process. 49, 87–90 (2001). [CrossRef]  

58. A. Rotnitzky, D. R. Cox, M. Bottai, and J. Robins, “Likelihood-based inference with singular information matrix,” Bernoulli 6, 243–284 (2000). [CrossRef]  

59. M. G. L. Gustafsson, D. A. Agard, and J. W. Sedat, “I5M: 3D widefield light microscopy with better than 100 nm axial resolution,” J. Microsc. 195, 10–16 (1999). [CrossRef]  

60. M. G. L. Gustafsson, “Surpassing the lateral resolution limit by a factor of two using structured illumination microscopy,” J. Microsc. 198, 82–87 (2000). [CrossRef]  

61. R. Heintzmann and C. G. Cremer, “Laterally modulated excitation microscopy: improvement of resolution by using a diffraction grating,” Proc. SPIE 3568, 185–196 (1999). [CrossRef]  

62. M. G. Genoni, M. G. A. Paris, G. Adesso, H. Nha, P. L. Knight, and M. S. Kim, “Optimal estimation of joint parameters in phase space,” Phys. Rev. A 87, 012107 (2013). [CrossRef]  

63. H. L. Van Trees and K. L. Bell, eds., Bayesian Bounds for Parameter Estimation and Nonlinear Filtering/Tracking (Wiley-IEEE, 2007).

64. R. W. Miller and C. B. Chang, “A modified Cramér–Rao bound and its applications (corresp.),” IEEE Trans. Inform. Theory 24, 398–400 (1978). [CrossRef]  

65. C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, Atom–Photon Interactions (Wiley-VCH, 2004).

66. M. O. Scully and M. S. Zubairy, Quantum Optics (Cambridge University, 1997).

67. X. Michalet, F. F. Pinaud, L. A. Bentolila, J. M. Tsay, S. Doose, J. J. Li, G. Sundaresan, A. M. Wu, S. S. Gambhir, and S. Weiss, “Quantum dots for live cells, in vivo imaging, and diagnostics,” Science 307, 538–544 (2005). [CrossRef]  

68. L. T. Hall, G. C. G. Beart, E. A. Thomas, D. A. Simpson, L. P. McGuinness, J. H. Cole, J. H. Manton, R. E. Scholten, F. Jelezko, J. Wrachtrup, S. Petrou, and L. C. L. Hollenberg, “High spatial and temporal resolution wide-field imaging of neuron activity using quantum NV-diamond,” Sci. Rep. 2, 401 (2012). [CrossRef]  

69. B. M. Escher, R. L. de Matos Filho, and L. Davidovich, “General framework for estimating the ultimate precision limit in noisy quantum-enhanced metrology,” Nat. Phys. 7, 406–411 (2011). [CrossRef]  

70. R. Demkowicz-Dobrzański, J. Kołodyński, and M. Guţă, “The elusive Heisenberg limit in quantum-enhanced metrology,” Nat. Commun. 3, 1063 (2012). [CrossRef]  

71. M. Tsang, “Quantum metrology with open dynamical systems,” New J. Phys. 15, 073005 (2013). [CrossRef]  

72. S. I. Knysh, E. H. Chen, and G. A. Durkin, “True limits to precision via unique quantum probe,” arXiv:1402.0495 (2014).

73. K. Lidke, B. Rieger, T. Jovin, and R. Heintzmann, “Superresolution by localization of quantum dots using blinking statistics,” Opt. Express 13, 7052–7062 (2005). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (3)

Fig. 1.
Fig. 1. Classical point source with dipole moment p(t) radiating in free space. Its position r is estimated by measuring the quantum optical field, with a(k,s) denoting its annihilation operator.
Fig. 2.
Fig. 2. Two classical point sources with dipole moments p(t) and p(t) at r and r, respectively, with quantum optical radiation.
Fig. 3.
Fig. 3. Plot of the resolution degradation factor κ versus the true separation |xx|/λ0 between two in-phase point sources, assuming p=p=p0eiω0t+c.c., T2π/ω0, p0=p0x^, y=y, and z=z. At |xx|=0, the Fisher information matrix is singular [57,58]. κ remains the same for two out-of-phase sources with otherwise the same assumptions.

Equations (71)

Equations on this page are rendered with MathJax. Learn more.

P(Y|X)=tr[E(Y)U(X,T)|ψψ|U(X,T)],
Σμν(X)dYP(Y|X)[X˜μ(Y)Xμ][X˜ν(Y)Xν].
Σ(X)j1(X),
jμν(X)dYP(Y|X)[θμlnP(Y|X)][θνlnP(Y|X)].
Σ(X)j1(X)J1(X),
Jμν(X)=4ReΔgμ(X)Δgν(X),
gμ(X)10TdtU(X,t)H(X,t)XμU(X,t)
Jab(θ)=μ,vXμθaJμ,v(X)Xvθb|X=X(θ).
H(r,t)=HF+HI(r,t),
HF=sd3kω(k)a(k,s)a(k,s),
HI(r,t)=p(t)·E(r),
E(r)=sd3kω2(2π)3ϵ0[iε(k,s)a(k,s)eik·r+H.c.],
a(k,s,t)U(X,t)a(k,s)U(X,t),
=eiωt[a(k,s)+α(k,s,r,t)],
α(k,s,r,T)=ω02(2π)3ϵ0eik·rε*(k,s)·[p0ei(ωω0)T/2sin(ωω0)T/2(ωω0)/2+p0*ei(ω+ω0)T/2sin(ω+ω0)T/2(ω+ω0)/2],
Nsd3k|α(k,s,r,T)|2|p0|2ω03T3πϵ0c3.
Δgμ(r)=2WμΔPμ(r),μ{x,y,z},
ΔPμ(r)12i[Δbμ(r)Δbμ(r)],
Δbμ(r)Wμsd3k[ikμα*(k,s,r,T)]Δa(k,s),
Wμ[sd3kkμ2|αμ(k,s,r,T)|2]1/2.
Wx=Wy=52λ02πN,Wz=5λ02πN,
Wx=Wy=103λ02πN,Wz=52λ02πN,
Jμν(r)=8Wμ2ΔPμ(r)ΔPν(r).
Jμν(r)=4Wμ2δμν,Σμμ(r)Wμ24,
|rr0|λ0,
ΔQμ(r0)12[Δbμ(r0)+Δbμ(r0)],
Σμμ(r)Wμ22ΔQμ2(r0)
ΔQμ2(r0)f(N0)2,ΔPμ2(r0)12f(N0),
f(N0)(2N0+1)[11(2N0+1)2],
f(0)=1,f(N0)14N0forN01.
Σμμ(r)Wμ22ΔQμ2(r0)=Wμ24f(N0).
Σμμ(r)Jμμ1(r)=Wμ28ΔPμ2(r)Wμ24f(N0).
H(r,r,t)=HF+HI(r,r,t),
HI(r,r,t)=p(t)·E(r)p(t)·E(r).
Jxx(X)=Jxx(X)=4Resd3kkx2α*(k,s,r,T)α(k,s,r,T),
Σxx(X)1Jxx[1κ(X)],
κ(X)Jxx2(X)JxxJxx=(Resd3kkx2α*α)2sd3kkx2|α|2sd3kkx2|α|2,
Σ¯μν(Z)dXdYP(Y|X,Z)PX|Z(X|Z)[X˜μ(Y,Z)Xμ]×[X˜ν(Y,Z)Xν],
Σ¯(Z)J¯1(Z),
J¯(Z)=EX|Z[J(X|Z)]+j(Z),
jμν(Z)dXPX|Z(X|Z)[XμlnPX|Z(X|Z)]×[XνlnPX|Z(X|Z)].
ΠμνEZ[Σ¯(Z)],
Σ¯(Z)dXdYP(Y|X,Z)PX|Z(X|Z)[X˜μ(Y)Xμ]×[X˜ν(Y)Xν],
ΠEZ[J¯1(Z)].
ρ=d2p0PZ(p0)U(X,p0,T)ρ0U(X,p0,T).
J¯μν(p0)=4Wμ2(p0)δμν+jμν=N(p0)Cμλ02δμν+jμν,
ΠμμEp0[14/Wμ2(p0)+jμμ]=Ep0[1N(p0)/(Cμλ02)+jμμ].
Ep0[1N(p0)/(Cμλ02)+jμμ]=Cμλ02N¯0dNexp(NN¯)1N+Cμλ02jμμ,
Cμλ02N¯lnN¯Cμλ02jμμ,N¯Cμλ02jμμ.
J¯xx(p0)=4Wx2(p0)+jxx,
J¯xx(p0)=4Wx2(p0)+jxx,
J¯xx(p0,p0)=EX[Jxx(p0,p0)].
ΠxxE(p0,p0){1J¯xx(p0)[1κ¯(p0,p0)]},
κ¯(p0,p0)J¯xx2(p0,p0)J¯xx(p0)J¯xx(p0).
|Ψ=c|0,csd3kϕ(k,s)a(k,s),
ϕ(k,s)=k,s|Ψ=0|a(k,s)|Ψ
ϕ(k,s)=k,s|g|HI|e|0[(ωω˜0)+i/(2T1)]eiωT,
HI=iω0(μ12σμ12*σ)·A(r),
A(r)=sd3k2(2π)3ωϵ0[a(k,s)ε(k,s)eik·r+h.c.],
T1=3πϵ0c3|μ12|2ω03,
ϕ(k,s)=1i(ωω˜0)+1/(2T1)ω022(2π)3ωϵ0μ12·ε*(k,s)×eik·riωT.
F=|0|c(X)c(X+δX)|0|2,
=|[c(X),c(X+δX)]|2,
1+μ,νδXμδXν{Re[c,2cXμXν]+Im[c,cXμ]Im[c,cXν]},
Re0|ccXμ|0=Re[c,cXμ]=0,
[c,cXμ]=sd3kϕ*(k,s)ϕ(k,s)Xμ=0,
Jμν=4Re[c,2cXμXν],
=4Resd3kϕ*(k,s)2ϕ(k,s)XμXν,
=4Resd3kkμkν(ωω˜0)2+1/(4T12)ω022(2π)3ωϵ0×|μ12·ε*(k,s)|2.
Jμν4Resd3kkμkν2πT1δ(ωω0)ω022(2π)3ωϵ0×|μ12·ε*(k,s)|2.
Jμν=4δμνNWμ2δμνλ02,ΣμμJμμ1=NWμ24λ02,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.