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

Real-time displacement measurement immune from atmospheric parameters using optical frequency combs

Open Access Open Access

Abstract

We propose a direct and real-time displacement measurement using an optical frequency comb, able to compensate optically for index of refraction variations due to atmospheric parameters. This scheme could be useful for applications requiring stringent precision over a long distance in air, a situation where dispersion becomes the main limitation. The key ingredient is the use of a mode-locked laser as a precise source for multi-wavelength interferometry in a homodyne detection scheme. By shaping temporally the local oscillator, one can directly access the desired parameter (distance variation) while being insensitive to fluctuations induced by parameters of the environment such as pressure, temperature, humidity and CO2 content.

© 2012 Optical Society of America

The sensitivity of precise length measurements is commonly limited by dispersive effects. For instance, the dispersion of air is a crucial issue for geodetic surveying [1] or for the optical link between a ground station and a satellite [2]. The lack of knowledge about atmospheric parameters can then act as the main limitation to optical length measurements.

Numerous groups around the world tackle this issue of long range distance measurement. For instance, in atmospheric links such as satellite ranging or Lunar Laser Range, the accuracy provided by time-of-flight measurements reaches the millimeter level or below [2, 3]. More recently, a technique using an optical cross-correlation and femtosecond laser has brought the precision of time-of-flight measurement down to the nanometer regime [4]. It is well known that interferometric distance measurements lead to potentially increased accuracies. However, for simple interferometric experiments, the periodicity of the signal results is ambiguous up to an absolute distance of a half-wavelength. This ambiguity distance can be extended by combining signals from multiple wavelengths [5]. For instance, the Very Large Telescope Interferometer uses dual-field interferometry to reach nanometric accuracy in a 100 m air-filled delay line [6].

It is also possible to combine both time-of-flight and phase measurement to obtain a better sensitivity and an absolute distance measurement. The ideal tool for this is the phase-stabilized mode-locked femtosecond laser which delivers a frequency comb that can be seen as a perfect source for multi-wavelength interferometry [7, 8, 9, 10, 11, 12, 13, 14]. The problem we address here is the way to use this tool in a complex environment, in order to perform a dispersion free measurement of absolute distance variation (displacement). We will more precisely treat the example of displacement in air independent of the variation of physical environmental parameters such as pressure, temperature, CO2 content or humidity.

In a multi-parameter environment, where many physical factors can affect the accuracy of distance measurements, these extra parameters need to be measured and their effects compensated. This is the general strategy of the well-known multicolor schemes [15, 16, 17, 18], which we introduce in the first part of this paper.

In the second part, we derive fundamental limits for optimal measurement schemes (i.e. reaching the Cramér-Rao bound with coherent states [19, 20, 21]) for distance measurements through a dispersive medium, using mode-locked lasers. The technique is based on temporal mode-dependent interferometry. We show that, in contrast to multicolor schemes for instance, only one measurement is necessary whatever the number of parasitic parameters we want to cancel. This is done at the cost of a precise spectral mode shaping of the frequency comb that is used.

We finally propose a general all-optical experimental setup using pulse shaping and homo-dyne detections to reach the previous limit. Our displacement measurement can be made insensitive of the environment parameters such as temperature or pressure. No post-processing is needed to achieve a measurement limited by the laser noise. Another advantage over existing schemes is that mode-locked lasers are remarkable tools for optical measurement because of their intrinsic high stability.

Distance measurement in air: multicolor schemes

Let us start by describing the general technique of multicolor schemes. If one needs to measure a distance in air, the fluctuations of parameters such as pressure or temperature limit the achievable accuracy. The general idea to solve this problem is to perform several measurements at different wavelengths to gain informations about these parameters and compensate the measured length for their variations. For example, in the two-wavelength interferometry (2WI), a given distance L is measured using two wavelengths λ1 and λ2. The two observables Lϕ1 and Lϕ2 deduced from the measurement are such that [15, 16, 17]

Lϕ1=nϕ(λ1)LandLϕ2=nϕ(λ2)L.
For dry air (pressure of water vapor Pw = 0), one finds
L=Lϕ1+α(Lϕ1Lϕ2)withα=K(λ1)K(λ2)K(λ1),
where K(λ) is given in Appendix A, and is calculated from the Edlén model of air. The parameter α is independent from pressure, temperature and CO2 content of air. The standard quantum limit for a distance measurement based on a phase measurement ϕi is given by (δLϕi)shot=c2Nωi. Using the previous Eq. (2), one obtains the standard quantum limit (or equivalently the sensitivity) δL on the distance, which strongly depends on the value of α. Typically, for λ1 = 1064 μm and λ2 = 532 μm, α is of the order of 60. Hence, for a mean photon number N = 4 × 1016 per wavelength (10 mW power and an integration time of 1 s), one gets
(δL)2WIshot3×1014m.
This is the shot noise limit in distance variation measurement (displacement), but one should note that for an absolute distance characterization (absolute ranging), the precise knowledge of α is the main limitation factor (δα/α ∼ 1%).

For moist air, one can no longer use Eq. (2). If the pressure of water vapor is unknown or changing over time, it leads to a systematic error (see [17] for a discussion). A solution is then to consider a third wavelength λ3 [22] and a third measurement Lϕ3 so that

L=Lϕ1+β(Lϕ2Lϕ1)+γ(Lϕ3Lϕ1).
Expressions of β and γ can be found in [22] and do not depend on pressure, temperature, CO2 content and humidity. Here again these factors can be large, reducing the sensitivity of displacement measurement. For a same total number of photons and typical wavelengths, the shot noise limit is now around 10−12 m. In addition to this degradation, the three-wavelength scheme is experimentally more involved.

From this example one sees that both the required sensitivity and physical characteristics of the medium conditions on the number of extra parameters one has to take into account. For each of these parameters an additional measurement at a different wavelength is necessary.

1. Efficient measurement through dispersive media

We will now give a more general and systematic approach to this problem. First, we derive general equations for the propagation of light through a dispersive medium whose characteristics depend on external parameters. We then elaborate on the ideas developed in [11, 21] and deal with a very general approach on how to efficiently measure parameters affecting the propagation of a light pulse. We derive fundamental limits imposed by the quantum nature of light. One should note nevertheless that we limit ourselves to the study of coherent states, non-classical quantum states being beyond the scope of this article.

1.1. Propagation in a dispersive medium

We consider the propagation of an electromagnetic field along the z direction in a weakly dispersive medium. Its propagation from a source to a detector can be affected by a given set of parameters p⃗ = (p1,..., pi,...) that modify the propagation distance L and/or the characteristics of the dispersive medium: these may be environmental parameters such as air pressure, temperature, etc., or a physical displacement of the source (or the detector). Neglecting polarization effects and using the paraxial approximation, we assume the field to be in a single transverse mode (such as a TEM00 mode), and thus will not write the transverse dependence of the field. We further assume Fourier-limited pulses (assuming perfect temporal coherence) and write the dispersed field, as seen by the detector, as a scalar field:

(t,p)0u(t,p),
where u(t, p⃗) is the normalized mean field mode (integrated over the measurement time of the detector) and ℰ0 is a normalization constant which depends on the number of photons N.

In the following, it will be more convenient to work in the Fourier space:

(ω,p)(t,p)eiωtdt,u(ω,p)u(t,p)eiωtdt.
We define ω0 and Δω as the mean value and variance of the field:
ω0=ω|u(ω,p)|2dω,Δω2=(ωω0)2|u(ω,p)|2dω.
Let us consider an input field ℰi(ω) whose frequency profile is known. For the sake of simplicity, the field is considered Gaussian (the same final result can be reached with a non Gaus-sian field but it involves more complex calculations). This field propagates on a distance L(p⃗) through a dispersive medium with a dispersion relation k(ω, p⃗) which depends on p⃗ through the refractive index nϕ(ω, p⃗). In the frequency space, this propagation is characterized by a spectral phase k(ω, p⃗)L(p⃗):
(ω,p)=i(ω)eik(ω,p)L(p),k(ω,p)=nϕ(ω,p)ωc.
In this paper, we neglect any absorption of the medium, i.e. we consider a real refractive index. The previous Eq. (8) is therefore also valid by replacing fields ℰ by normalized modes u.

For a weakly dispersive medium, the dispersion relation k(ω, p⃗) can be expanded to the second order around the mean frequency ω0:

(ω,p)i(ω)exp[i(ω0tϕ(p)+(ωω0)tg(p)+(ωω0)2ω0tGVD(p))],
where
tϕ(p)=nϕ(ω0,p)L(p)c,
tg(p)=ng(ω0,p)L(p)c=(nϕ(ω0,p)+ω0nϕ(ω0,p))L(p)c,
tGVD(p)=ω0(nϕ(ω0,p)+ω02nϕ(ω0,p))L(p)c.
ng is the group index, and 2(nϕ(ω0,p)+ω02nϕ(ω0,p)) corresponds to the Group Velocity Dispersion (GVD).

In the temporal domain, the field ℰ(t, p) = e0tℰ̃(t, p), the envelope ℰ̃ of the field travels at the group velocity cng while the carrier ω0 moves at the phase velocity cnϕ; a non zero group velocity dispersion leads to a broadening of the envelope.

Any change of a parameter in p⃗ that affects the distance L(p⃗) will contaminate all quantities tϕ, tg and tGVD, as well as any variation of the refractive index of the medium. In Section 2, we show how to uncouple, in air, variation of L from variation due to four different environmental parameters: pressure, temperature, humidity and CO2 content. Note that a generalization to other environmental parameters can be obtained by expanding Eq. (9) to higher orders of the spectral phase and applying the methods developed later in the paper. Nevertheless, these 4 parameters are typically the most relevant for air.

1.2. Detection scheme and Cramér-Rao bound

The general problem of estimating a parameter pip⃗ encoded in a light beam ℰ(p⃗) has been treated in [19, 23]. The ultimate limit of sensitivity in the measurement of pi is given by the so-called quantum Cramér-Rao bound, which can be computed once we specify the quantum state of the light beam (coherent state, squeezed state, entangled state etc...) [21]. For Gaussian states, this Cramér-Rao bound can be reached experimentally with a balanced homodyne detection scheme, as represented in Fig. 1. The general idea is that the homodyne detection signal is proportional to the projection of the input field into the Local Oscillator (LO) mode.

 figure: Fig. 1

Fig. 1 General detection scheme for measuring pi at the Cramér-Rao bound. A specific example of a full experimental setup is given in Fig.3

Download Full Size | PDF

For a small variation of the set of parameters p⃗, the field reads:

(ω,p)(ω,0)+pp(ω,p=0)=0(u(ω,0)+ipiKiwi(ω)),
where wi are normalized modes such as wi(ω)=1Kiupi(ω,p=0), and Ki are dimensional normalization constants. Introducing the standard L2 inner product 〈f, g〉 = ∫ f*(ω)g(ω)dω, one simply has Kiupi,upi.

One should note that in general the modes wi do not form an orthogonal basis. This implies that independent measurements of each parameter become a complex problem that we now discuss in detail extensively. Let us first consider the case where only one parameter pip⃗ is influencing the length measurement. It is shown in [24] that in a homodyne detection scheme, if the LO mode is proportional to wi (and if there is no phase difference between the LO and the signal ℰ), the detected signal S[wi] is given by

S[wi]=1KiRe[u(p),wi]=pji=0pi.
For a coherent state illumination with a mean photon number N, the noise in the measurement is Δpi=2NKi. Therefore, the smallest pi that can be measured (i.e. for a signal to noise ratio equal to one, Sp = 1) is given by
(pi)min=12NKi.
This value coincides with the Cramér-Rao bound with coherent states [21]. This shows that a homodyne detection with a LO shaped in mode wi defines what is called an efficient measurement of pi. Moreover, one sees from the previous expression that the sensitivity of the measurement depends both on the number of photons N and on Ki, the latter reflecting the characteristic variation of the mode with the parameter pi. From now on, the mode wi will be called the detection mode of the parameter pi.

An experimental demonstration of the efficiency of such a scheme has been realized for parameters corresponding to transverse spatial displacement of a beam [25, 24], and a theoretical proposition for a time delay through a dispersion-less medium has been made in [11]. One should stress that, generally speaking, these kind of experiments are sensitive to variation of parameters within the detection bandwidth, limited here by light time travel. On the other hand, this system is immune to any fluctuations, whatever their frequencies, of parameter corresponding to orthogonal modes, given we stay in the linear regime.

Let us now consider the general case where there exists at least another parameter pji such as wi and wj are not orthogonal. In that case, a homodyne detection with LO mode wi will also be sensitive to pj. We show here that this issue can be resolved by ’purifying’ the detection mode wi into a new mode wip that is now orthogonal to wj. This new shape of the LO allows to measure pi independently of pj. Nevertheless, because it differs from the detection mode wi, it leads to a reduced sensitivity in the measurement of pi. In the general situation, the purified mode for a given parameter pi is orthogonal to the hyperplane formed by {wji}, and the normalization factors are given by Kip=Kiwip,wi<Ki. The sensitivity in the measurement of pi is therefore decreased and given by (pi)minp=12NKip.

The choice for the mode of the LO is an experimental trade-off between accuracy (no perturbation coming from pj) and precision. We further elaborate on this point in the following, taking as a simple example the measurement of tϕ, tg and tGVD introduced previously.

To this aim, we introduce a controlled perturbation pϕtϕ of the phase delay tϕtϕ + pϕ, a perturbation pgtg of the group delay tgtg + pg and a perturbation pGVDtGVD of the group velocity dispersion delay tGVDtGVD + pGVD. The corresponding detection modes are given by:

wϕ(ω)=iu(ω)=v0(ω)
wg(ω)=iωω0Δωu(ω)=v1(ω)
wGVD(ω)=i13(ωω0)2Δω2u(ω)=13v0(ω)+23v2(ω)
where we have introduced {vi(ω)} the orthonormal basis of spectral Hermite-Gaussian modes whose expressions are given in Appendix B. The normalization factors are given by Kϕ = ω0, Kg = Δω and KGVD=3Δω2ω0.

It is clear that wϕ and wGVD are not orthogonal, which implies that a measurement using a LO wϕ will not be accurate because of its sensitivity to pGVD, and vice versa. More precisely, measurements over the various detection modes will yield the following signals:

S[wϕ]=pϕ+Δω2ω02pGVD
S[wg]=pg
S[wGVD]=13ω02Δω2pϕ+pGVD.
In order to measure only pϕ, one has to consider the purified mode wϕp introduced previously, which is orthogonal to the detection modes of the other parameters. Formally, wϕp(ω) is orthogonal to the hyperplane generated by {wg, wGVD} in the vector space {wϕ, wg, wGVD} and is given in the present case (up to a scalar factor) by:
wϕp(ω)=23v0(ω)13v2(ω).
A measurement with LO wϕp yields:
S[wϕp]=pϕ.
Defining Kϕp=Kϕwϕp,wϕ=23ω0, the sensitivity of this measurement is:
(pϕ)minp=12N23ω0.

The two previous equations show that it is possible to retrieve a phase delay information insensitive to any group velocity dispersion fluctuations with only one homodyne measurement. This improvement in accuracy is made at the cost of a decreased precision, determined by the overlap between the purified and non purified modes (here the degradation is given by 3/2).

The same process can be applied to obtain the purified mode for measuring pGVD:

wGVDp(ω)=v2(ω)withKGVDp=2Δω2ω0.
The link between the detection modes and the purified modes is shown in Fig. 2 and in Table 1.

 figure: Fig. 2

Fig. 2 Relation between the different LO modes in the vector space {v0, v1, v2}. The modes wϕp, wϕ, wGVD and wGVDp lie in the same plane.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Direct distance measurement with an appropriately shaped LO.

Download Full Size | PDF

Tables Icon

Table 1. Summary of the different LO modes and the sensitivities.

2. Application to the measurement of a displacement in air

Let us now consider the specific case of measuring a displacement in air, independently of the fluctuation of environmental parameters such as pressure and temperature. Indeed, induced index refraction fluctuations are the main limitations to precise distance measurement. To access the absolute length, one needs to know precisely the air index variation with these parameters. One solution is to measure precisely these parameters and use an air model to calculate the refractive index, for instance the Edlén [26] or the Ciddor [27] equations. But these techniques are not immune to local variation of the parameters, and some parameters such as the partial pressure of water vapor are very difficult to precisely access. One can also measure directly the local refractive index using a refractometer [28].

Another solution is to make several measurements in order to compensate for these effects. This is the principle of multicolor interferometry [15], whose state of the art is based on second harmonic generation [29, 16], as we have presented in the introduction. Here we develop our new experimental scheme based on mode-locked lasers interferometry as introduced in the previous section. This technique is another kind of multicolor interferometry and allows for direct measurement of displacement in air independently of parameters from the environment.

We apply the technique developed in Section 1.2 to the measurement of displacement, that is to say a variation of the absolute distance L in air. The parameter to be measured with high sensitivity is pL = L. Fluctuations of the environment do perturb this measurement. They can be separated into two groups of parameters. Firstly temperature T, pressure P and CO2 content x affect air index though the same function, as can be seen in the air model developed in Appendix A. They will be described by only one parameter pX = X(T, P, x). Secondly, pressure of water vapor Pw has an independent influence, for which we define the parameter pPw = Pw.

One can calculate the corresponding detection modes using the Edlén model of refractive index recalled in the appendix and the second order development of the electric field introduced in the first section. The distance detection mode is given by:

wL(ω)=1cKL(ω0v0(ω)+Δωv1(ω))
and the two other detection modes are given in the appendix.

These modes are not linearly independent. Thus if we consider a possible experimental scheme with homodyne detection in the detection mode for L (see Fig. 3) the measured signal will be :

S[wL]=pL+KXKLwL,wXpX+KPwKLwL,wPwpPw.
Therefore, the signal will be contaminated by variations of the different parameters pX and pPw. From now, we will characterize the sensitivity only by the shot noise limit of the measurement. This is justified for two reasons. First, we are not interested in measurement of the absolute distance, but only in any variation of this absolute distance. Therefore, the experiment is not limited by the knowledge of the parameters entering the Edlén model of air. Secondly, state-of-the-art mode-locked lasers are indeed shot-noise limited above 1 MHz [30], the frequencies of interest for the measurement of a distance up to a few hundreds of meters.

Let us first compute the shot noise limit in the case where pX and pPw are zero (or sufficiently small). To compare with multicolor schemes, in the remainder of this section the measurement is performed using N = 8 × 1016 photons and assuming a laser bandwidth of Δω=ω06 (corresponding to 3 fs FWHM Fourier-limited pulses). The shot noise limited sensitivity to displacement is about 2 × 10−16 m, comparable to usual interferometric measurement schemes.

One can evaluate the contamination from the other parameters calculating the pre-factor of pX and pPw in Eq. (27). One finds 1LKXKLwL,wX=27×105Pa1 and 1LKPwKLwL,wPw=3.7×1010Pa1. These factors are big compared to shot noise limited pure distance measurements. It means it is necessary in this case to take into account the parasitic parameters in order to preserve the accuracy of the measurement. To solve this issue, we can apply the detection mode purification procedure introduced previously :

wLp(ω)wL(ω)wL,wPwwX,wPwwL,wX1wX,wPw2wPw(ω)
wL,wXwX,wPwwL,wPw1wX,wPw2wX(ω).
The spectral profiles of the purified and non-purified modes are plotted in Fig. 4. The normalization constant as well as the derivation of this mode can be found in Appendix C. It is found that :
KLp=KL1wL,wPw2+wL,wX22wL,wPwwL,wXwX,wPw1wX,wPw2.
In that case, the shot noise limit has the following value :
(δL)1HDshot=c2NKLp=2×1011m.
As a matter of comparison, a purification for pX only (which would be equivalent to the two-color scheme) leads to a shot noise limit of (δL)1HDshot=c2NKLp=3×1013m. Even if the precision is 2 orders of magnitude better, the accuracy of the measurement is not controlled because of the unknown value of humidity introducing a systematic error.

 figure: Fig. 4

Fig. 4 Spectral profile of modes wL and wLp for 3 fs FWHM Gaussian pulses.

Download Full Size | PDF

Using that scheme, one can perform a real-time measurement of displacement immune from environmental parameters fluctuations. Sensitivity is of the same order of magnitude than in multicolor scheme (slightly lower in our examples, but this is simply due to smaller spectral width), and depends on how many parameters one wants to get immune from. Furthermore, the scheme described here could be extended to more parameters. Of course, the mode shaping can be more difficult to produce and in particular the precision required in shaping becomes more stringent when more parameters are considered. In a realistic implementation, one can set up active pulse-shaping with search algorithms to determine the proper shaping. It is significant to notice that all the work is to be done at the detection stage and not on the light sent through the medium, which makes it much easier to handle. Moreover, once the parameter to be measured has been chosen (for example displacement), the corresponding shape of the local oscillator does not have to be changed in real-time to give a measurement of that parameter immune from fluctuations of other parameters (for example pressure, temperature etc...), as long as the experiment stays in the linear regime (fluctuations of the parameter being not too large). Therefore, the pulse shaping has to be configured at the beginning of the experiment, and then remains constant while one performs real-time measurements.

3. Conclusion

We have exhibited a novel optimal and all-optical scheme to measure in real time a distance compensated for refractive index changes. It relies on a homodyne detection whose local oscillator projects the measurement on an appropriate mode, hence no post-processing is necessary. We believe this is a simplification compared to existing schemes such as spectral interferometry for example, where derivatives of the spectral phase have to be done after the measurement.

Let us finally mention that this scheme can further be improved in order to go below the standard quantum limit. It is well known that the sensitivity of a measurement can outreach the shot noise limit by using quantum resources such as squeezed or entangled light [31]. In the scheme presented in this paper, this can be achieved by using a multimode signal light beam with squeezing in the detection mode associated to the measurement, as demonstrated in Ref. [21].

A. Index of refraction of air

The common equations used to derive the wavelength dependence of the refractive index of air are given by the Edlén equations [26], modified by different authors since that time [32]. The accuracy of those equations are roughly of the order of a few 10−9 for dry air and 10−8 for moist air and needs to take into account a large number of parameters, usually temperature, pressure, CO2 content, pressure of water vapor. Moreover, different studies do not necessarily agree and do not cover the whole spectrum. A recent precise measurement of the refractive index of air has been done in [33]. Experimentally, measuring all those parameters can be difficult for certain situations, and in addition those parameters have to be measured in real time, in order to compensate for fluctuations of the refractive index.

In this paper we consider the updated Edlén formula of Bönsch and Potulski [28] :

nϕ(σ,T,P,x,Pw)1=K(σ)X(T,P,x)g(σ)Pw
where σ = 1/λ is the wavenumber and
K(σ)=108(A+B130σ2+C38.9σ2),
X(T,P,x)=PD1+108(EFT)P1+GT[1+H(x0.04%)],
g(σ)=1010(IJσ2),
with σ in μm−1, P and Pw in Pascal (Pa), T in degree Celsius (°C) and x, the CO2 content, in percentage. The different coefficients further read
A=8091.37,B=2333983,C=15518,
D=932164.60,E=0.5953,F=0.009876,G=0.0036610,
H=0.5327,
I=3.802,J=0.0384.

The group index is therefore given by

ng(σ,T,P,x,Pw)1=(K(σ)+σK(σ))X(T,P,x)(g(σ)+σg(σ))Pw,
so that ng - nϕ = σ(K′(σ)X(T, P, x) -g′(σ)Pw).

B. Hermite-Gauss set of spectral modes

For a Gaussian mean field mode

u(ω)=1Δω1(2π)1/4e(ωω0)24Δω2,
we introduce the Hermite-Gauss modes
vn(ω)=i12nn!Hn(ωω02Δω)u(ω).
These modes {vn(ω)} form an orthonormal basis of modes.

In order to describe the spectral phase up to the kth order, one needs to use the basis {vn(ω)} up to the mode k. In this paper, we develop the spectral phase to the second order; therefore, we use the following modes:

v0(ω)=iu(ω)
v1(ω)=iωω0Δωu(ω)
v2(ω)=i12((ωω0)2Δω21)u(ω).
The mode spectral profiles are plotted in Fig. 5.

 figure: Fig. 5

Fig. 5 Spectral profile of Hermite-Gaussian modes v0, v1 and v2.

Download Full Size | PDF

C. Detection modes of environmental parameters

Detection modes read:

wL(ω)=1cKL(ω0v0(ω)+Δωv1(ω))wX(ω)=LK(ω0)cKX[(ω0+Δω2ω0(δ1+δ2))v0(ω)+Δω(1+δ1)v1(ω)+2Δω2ω0(δ1+δ2)v2(ω)]wPw(ω)=Lg(ω0)cKPw[(ω0+Δω2ω0(η1+η2))v0(ω)+Δω(1+η1)v1(ω)+2Δω2ω0(η1+η2)v2(ω)]
where we have defined characteristic quantities that do not depend on the environmental parameters T, P, x and Pw:
δ1=ω0K(ω0)K(ω0),δ2=ω022K(ω0)K(ω0),
η1=ω0g(ω0)g(ω0),η2=ω022g(ω0)g(ω0).
Defining:
KL=1cω02+Δω2
KX=K(ω0)Lc(ω0+Δω2ω0(δ1+δ2))2+Δω2(1+δ1)2+2Δω4ω02(δ1+δ2)2
KPw=g(ω0)Lc(ω0+Δω2ω0(η1+η2))2+Δω2(1+η1)2+2Δω4ω02(η1+η2)2
Measurements with the detection modes give:
M[wL]=pL+KXKLwL,wXpX+KPwKLwL,wPwpPw
M[wX]=KLKXwL,wXpL+pX+KPwKXwL,wPwpPw
M[wPw]=KLKPwwL,wPwpL+KXKPwwX,wPwpX+pPw

Calculation of the purified mode wLp: we first calculate a orthonormal basis in the plane {wX, wPw}, which leads for example to a basis {wX, wPwi} where

wPwi(ω)=11wX,wPw2(wPw(ω)wX,wPwwX)
Then we do a Gram-Schmidt orthonormalization of {wX, wPwi, wL} which gives:
wLp(ω)wL(ω)wL,wXwX(ω)wL,wPwiwPwi
The normalization constant is:
1wL,wPw2+wL,wX22wL,wPwwL,wXwX,wPw1wX,wPw2

Acknowledgments

The research is supported by ANR project Qualitime, ERC starting grant Frecquam and by the Australian Research Council Centre of Excellence for Quantum Computation and Communication Technology, project number CE110001027 (OP).

References and links

1. A. Weiss, M. Hennes, and M. Rotach, “Derivation of refractive index and temperature gradients from optical scintillometry to correct atmospherically induced errors for highly precise geodetic measurements,” Surv. Geophys. 22, 589–596 (2001). [CrossRef]  

2. P. Exertier, P. Bonnefond, F. Deleflie, F. Barlier, M. Kasser, R. Biancale, and Y. Ménard, “Contribution of laser ranging to earth’s sciences,” C. R. Geosci. 338, 958–967 (2006). [CrossRef]  

3. K. Djerroud, O. Acef, A. Clairon, P. Lemonde, C. N. Man, E. Samain, and P. Wolf, “Coherent optical link through the turbulent atmosphere,” Opt. Lett. 35, 1479–1481 (2010). [CrossRef]   [PubMed]  

4. J. Lee, Y. -J. Kim, K. Lee, S. Lee, and S. -W. Kim, “Time-of-flight measurement with femtosecond light pulses,” Nat. Photonics 4, 716–720 (2010). [CrossRef]  

5. R. Dändliker, R. Thalmann, and D. Prongué, “Two-wavelength laser interferometry using superheterodyne detection,” Opt. Lett. 13, 339–341 (1988). [CrossRef]   [PubMed]  

6. F. Delplancke, “The prima facility phase-referenced imaging and micro-arcsecond astrometry,” New Astron. Rev. 52, 199–207 (2008). [CrossRef]  

7. K. Minoshima and H. Matsumoto, “High-accuracy measurement of 240-m distance in an optical tunnel by use of a compact femtosecond laser,” Appl. Opt. 39, 5512–5517 (2000). [CrossRef]  

8. J. Ye, “Absolute measurement of a long, arbitrary distance to less than an optical fringe,” Opt. Lett. 29, 1153–1155 (2004). [CrossRef]   [PubMed]  

9. K. -N. Joo and S. -W. Kim, “Absolute distance measurement by dispersive interferometry using a femtosecond pulse laser,” Opt. Express 14, 5954–5960 (2006). [CrossRef]   [PubMed]  

10. M. Cui, R. N. Schouten, N. Bhattacharya, and S. A. Berg, “Experimental demonstration of distance measurement with a femtosecond frequency comb laser,” J. Eur. Opt. Soc. Rapid Publ. 3, 08003 (2008). [CrossRef]  

11. B. Lamine, C. Fabre, and N. Treps, “Quantum improvement of time transfer between remote clocks,” Phys. Rev. Lett. 101, 123601 (2008). [CrossRef]   [PubMed]  

12. Y. Salvadé, N. Schuhler, S. Lévêque, and S. Le Floch, “High-accuracy absolute distance measurement using frequency comb referenced multiwavelength source,” Appl. Opt. 47, 2715–2720 (2008). [CrossRef]   [PubMed]  

13. I. Coddington, W. C. Swann, L. Nenadovic, and N. R. Newbury, “Rapid and precise absolute distance measurements at long range,” Nat. Photonics 3, 351–356 (2009). [CrossRef]  

14. P. Balling, P. Kren, P. Masika, and S. A. van den Berg, “Femtosecond frequency comb based distance measurement in air,” Opt. Express 17, 9300–9313 (2009). [CrossRef]   [PubMed]  

15. P. L. Bender and J. C. Owen, “Correction of optical distance measurement for the fluctuating atmospheric index of refraction,” J. Geophys. Res. 70, 2461–2462 (1965). [CrossRef]  

16. H. Matsumoto and T. Honda, “High-accuracy length-measuring interferometer using the two-colour method of compensating for the refractive index of air,” Meas. Sci. Technol. 3, 1084–1086 (1992). [CrossRef]  

17. K. Meiners-Hagen and A. Abou-Zeid, “Refractive index determination in length measurement by two-colour interferometry,” Meas. Sci. Technol. 19, 084004 (2008). [CrossRef]  

18. S. Azouigui, T. Badr, J. -P. Wallerand, M. Himbert, J. Salgado, and P. Juncar, “Transportable distance measurement system based on superheterodyne interferometry using two phase-locked frequency-doubled Nd:YAG lasers,” Rev. Sci. Instrum. 81, 053112 (2010). [CrossRef]   [PubMed]  

19. C. Helstrom, “The minimum variance of estimates in quantum signal detection,” IEEE Trans. Inf. Theory 14, 234–242 (1968). [CrossRef]  

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

21. O. Pinel, J. Fade, D. Braun, P. Jian, N. Treps, and C. Fabre, “Ultimate sensitivity of precision measurements with intense gaussian quantum light: a multimodal approach,” Phys. Rev. A 85, 1–4 (2012). [CrossRef]  

22. A. N. Golubev and A. M. Chekhovsky, “Three-color optical range finding,” Appl. Opt. 33, 7511–7517 (1994). [CrossRef]   [PubMed]  

23. P. Réfrégier, Noise Theory and Application to Physics: From Fluctuations to Information (Springer Verlag, 2004).

24. V. Delaubert, N. Treps, C. Harb, P. Lam, and H. Bachor, “Quantum measurements of spatial conjugate variables: displacement and tilt of a gaussian beam,” Opt. Lett. 31, 1537–1539 (2006). [CrossRef]   [PubMed]  

25. M. Hsu, V. Delaubert, P. Lam, and W. Bowen, “Optimal optical measurement of small displacements,” J. Opt. B: Quantum Semiclassical Opt. 6, 495–501 (2004). [CrossRef]  

26. B. Edlén, “The refractive index of air,” Metrologia 2, 71–80 (1966). [CrossRef]  

27. P. E. Ciddor, “Refractive index of air: new equations for the visible and near infrared,” Appl. Opt. 35, 1566–1573 (1996). [CrossRef]   [PubMed]  

28. G. Bönsch and E. Potulski, “Measurement of the refractive index of air and comparison with modified Edlén’s formulae,” Metrologia 35, 133–139 (1998). [CrossRef]  

29. A. Ishida, “Two-wavelength displacement-measuring interferometer using second-harmonic light to eliminate air-turbulence-induced errors,” Jpn. J. Appl. Phys. 28, L473–L475 (1989). [CrossRef]  

30. O. Pinel, Pu. Jian, R. Medeiros de Araújo, J. Feng, B. Chalopin, C. Fabre, and N. Treps, “Generation and characterization of multimode quantum frequency combs,” Phys. Rev. Lett. 108, 083601 (2012). [CrossRef]   [PubMed]  

31. V. Giovannetti, S. Lloyd, and L. Maccone, “Quantum-enhanced measurements: beating the standard quantum limit,” Science 306, 1330–1336 (2004). [CrossRef]   [PubMed]  

32. Formulas for refractive index of air are available on http://emtoolbox.nist.gov/Wavelength/Documentation.asp

33. R. Macovez, M. Mariano, S. D. Finizio, and J. Martorell, “Measurement of the dispersion of air and of refractive index anomalies by wavelength-dependent nonlinear interferometry,” Opt. Express 17, 13881–13888 (2009). [CrossRef]   [PubMed]  

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 (5)

Fig. 1
Fig. 1 General detection scheme for measuring pi at the Cramér-Rao bound. A specific example of a full experimental setup is given in Fig.3
Fig. 2
Fig. 2 Relation between the different LO modes in the vector space {v0, v1, v2}. The modes w ϕ p , wϕ, wGVD and w GVD p lie in the same plane.
Fig. 3
Fig. 3 Direct distance measurement with an appropriately shaped LO.
Fig. 4
Fig. 4 Spectral profile of modes wL and w L p for 3 fs FWHM Gaussian pulses.
Fig. 5
Fig. 5 Spectral profile of Hermite-Gaussian modes v0, v1 and v2.

Tables (1)

Tables Icon

Table 1 Summary of the different LO modes and the sensitivities.

Equations (57)

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

L ϕ 1 = n ϕ ( λ 1 ) L and L ϕ 2 = n ϕ ( λ 2 ) L .
L = L ϕ 1 + α ( L ϕ 1 L ϕ 2 ) with α = K ( λ 1 ) K ( λ 2 ) K ( λ 1 ) ,
( δ L ) 2 W I shot 3 × 10 14 m .
L = L ϕ 1 + β ( L ϕ 2 L ϕ 1 ) + γ ( L ϕ 3 L ϕ 1 ) .
( t , p ) 0 u ( t , p ) ,
( ω , p ) ( t , p ) e i ω t d t , u ( ω , p ) u ( t , p ) e i ω t d t .
ω 0 = ω | u ( ω , p ) | 2 d ω , Δ ω 2 = ( ω ω 0 ) 2 | u ( ω , p ) | 2 d ω .
( ω , p ) = i ( ω ) e i k ( ω , p ) L ( p ) , k ( ω , p ) = n ϕ ( ω , p ) ω c .
( ω , p ) i ( ω ) exp [ i ( ω 0 t ϕ ( p ) + ( ω ω 0 ) t g ( p ) + ( ω ω 0 ) 2 ω 0 t GVD ( p ) ) ] ,
t ϕ ( p ) = n ϕ ( ω 0 , p ) L ( p ) c ,
t g ( p ) = n g ( ω 0 , p ) L ( p ) c = ( n ϕ ( ω 0 , p ) + ω 0 n ϕ ( ω 0 , p ) ) L ( p ) c ,
t GVD ( p ) = ω 0 ( n ϕ ( ω 0 , p ) + ω 0 2 n ϕ ( ω 0 , p ) ) L ( p ) c .
( ω , p ) ( ω , 0 ) + p p ( ω , p = 0 ) = 0 ( u ( ω , 0 ) + i p i K i w i ( ω ) ) ,
S [ w i ] = 1 K i Re [ u ( p ) , w i ] = p j i = 0 p i .
( p i ) min = 1 2 N K i .
w ϕ ( ω ) = i u ( ω ) = v 0 ( ω )
w g ( ω ) = i ω ω 0 Δ ω u ( ω ) = v 1 ( ω )
w GVD ( ω ) = i 1 3 ( ω ω 0 ) 2 Δ ω 2 u ( ω ) = 1 3 v 0 ( ω ) + 2 3 v 2 ( ω )
S [ w ϕ ] = p ϕ + Δ ω 2 ω 0 2 p GVD
S [ w g ] = p g
S [ w GVD ] = 1 3 ω 0 2 Δ ω 2 p ϕ + p GVD .
w ϕ p ( ω ) = 2 3 v 0 ( ω ) 1 3 v 2 ( ω ) .
S [ w ϕ p ] = p ϕ .
( p ϕ ) min p = 1 2 N 2 3 ω 0 .
w GVD p ( ω ) = v 2 ( ω ) with K GVD p = 2 Δ ω 2 ω 0 .
w L ( ω ) = 1 c K L ( ω 0 v 0 ( ω ) + Δ ω v 1 ( ω ) )
S [ w L ] = p L + K X K L w L , w X p X + K P w K L w L , w P w p P w .
w L p ( ω ) w L ( ω ) w L , w P w w X , w P w w L , w X 1 w X , w P w 2 w P w ( ω )
w L , w X w X , w P w w L , w P w 1 w X , w P w 2 w X ( ω ) .
K L p = K L 1 w L , w P w 2 + w L , w X 2 2 w L , w P w w L , w X w X , w P w 1 w X , w P w 2 .
( δ L ) 1 HD shot = c 2 N K L p = 2 × 10 11 m .
n ϕ ( σ , T , P , x , P w ) 1 = K ( σ ) X ( T , P , x ) g ( σ ) P w
K ( σ ) = 10 8 ( A + B 130 σ 2 + C 38.9 σ 2 ) ,
X ( T , P , x ) = P D 1 + 10 8 ( E F T ) P 1 + G T [ 1 + H ( x 0.04 % ) ] ,
g ( σ ) = 10 10 ( I J σ 2 ) ,
A = 8091.37 , B = 2333983 , C = 15518 ,
D = 932164.60 , E = 0.5953 , F = 0.009876 , G = 0.0036610 ,
H = 0.5327 ,
I = 3.802 , J = 0.0384 .
n g ( σ , T , P , x , P w ) 1 = ( K ( σ ) + σ K ( σ ) ) X ( T , P , x ) ( g ( σ ) + σ g ( σ ) ) P w ,
u ( ω ) = 1 Δ ω 1 ( 2 π ) 1 / 4 e ( ω ω 0 ) 2 4 Δ ω 2 ,
v n ( ω ) = i 1 2 n n ! H n ( ω ω 0 2 Δ ω ) u ( ω ) .
v 0 ( ω ) = i u ( ω )
v 1 ( ω ) = i ω ω 0 Δ ω u ( ω )
v 2 ( ω ) = i 1 2 ( ( ω ω 0 ) 2 Δ ω 2 1 ) u ( ω ) .
w L ( ω ) = 1 c K L ( ω 0 v 0 ( ω ) + Δ ω v 1 ( ω ) ) w X ( ω ) = L K ( ω 0 ) c K X [ ( ω 0 + Δ ω 2 ω 0 ( δ 1 + δ 2 ) ) v 0 ( ω ) + Δ ω ( 1 + δ 1 ) v 1 ( ω ) + 2 Δ ω 2 ω 0 ( δ 1 + δ 2 ) v 2 ( ω ) ] w P w ( ω ) = L g ( ω 0 ) c K P w [ ( ω 0 + Δ ω 2 ω 0 ( η 1 + η 2 ) ) v 0 ( ω ) + Δ ω ( 1 + η 1 ) v 1 ( ω ) + 2 Δ ω 2 ω 0 ( η 1 + η 2 ) v 2 ( ω ) ]
δ 1 = ω 0 K ( ω 0 ) K ( ω 0 ) , δ 2 = ω 0 2 2 K ( ω 0 ) K ( ω 0 ) ,
η 1 = ω 0 g ( ω 0 ) g ( ω 0 ) , η 2 = ω 0 2 2 g ( ω 0 ) g ( ω 0 ) .
K L = 1 c ω 0 2 + Δ ω 2
K X = K ( ω 0 ) L c ( ω 0 + Δ ω 2 ω 0 ( δ 1 + δ 2 ) ) 2 + Δ ω 2 ( 1 + δ 1 ) 2 + 2 Δ ω 4 ω 0 2 ( δ 1 + δ 2 ) 2
K P w = g ( ω 0 ) L c ( ω 0 + Δ ω 2 ω 0 ( η 1 + η 2 ) ) 2 + Δ ω 2 ( 1 + η 1 ) 2 + 2 Δ ω 4 ω 0 2 ( η 1 + η 2 ) 2
M [ w L ] = p L + K X K L w L , w X p X + K P w K L w L , w P w p P w
M [ w X ] = K L K X w L , w X p L + p X + K P w K X w L , w P w p P w
M [ w P w ] = K L K P w w L , w P w p L + K X K P w w X , w P w p X + p P w
w P w i ( ω ) = 1 1 w X , w P w 2 ( w P w ( ω ) w X , w P w w X )
w L p ( ω ) w L ( ω ) w L , w X w X ( ω ) w L , w P w i w P w i
1 w L , w P w 2 + w L , w X 2 2 w L , w P w w L , w X w X , w P w 1 w X , w P w 2
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.