You are on page 1of 8

Atom interferometry : principles and applications to fundamental physics

R. Delhuille, C. Champenois, M. Biichner, R. Mathevet, C. Rizzo, C. Robilliard and J. Vigue


Laboratoire Collisions Agregats Reactivite - IRSAMC Universite Paul Sabatier and CNRS UMR 5589 118, Route de Narbonne 31062 Toulouse Cedex, France e-mail: j acques . vigueOirsamc. ups-tlse. f r

Abstract. The first part of this contribution presents an overview of atom interferometry. In a second part, we discuss in greater details the sensitivity of such experiments to weak perturbations and their application to direct test of the electric neutrality of atoms with extreme precision.

GENERAL IDEAS ON ATOM INTERFEROMETRY Matter waves diffraction


Since de Broglie, we know that atoms are described by waves with a wavelength \dB given by the equation

AdB =

mv

(1)

where m is the atom mass and v its velocity. As an example, we consider the helium atom at a typical thermal velocity v 1000 m/s for which the de Broglie wavelength is XdB 10~10 m. As the thermal velocity v is proportional to v oc v/T/m, where T is the temperature of the gas, the scaling of the de Broglie wavelength with T and m is slow, XdB oc (raT)"1/2. Therefore, it is only at extremely low temperatures, available now thanks to laser cooling, that the atom wavelength X^B can reach considerably larger values, up to the micrometer range. The very low values of the de Broglie wavelength at thermal velocity are comparable to the atomic sizes and crystal lattice periods, and it appeared very early that a cleaved crystal surface was adapted to be a grating for atoms. In 1929, Stern and coworkers [1] observed the diffraction of an helium effusive beam by the surface of a LiF crystal (a review appears in [2]). Elastic scattering as well as inelastic

CP564, Quantum Electrodynamics and Physics of the Vacuum, edited by G. Cantatore 2001 American Institute of Physics 0-7354-0000-8/01/$18.00

192

scattering can be observed and inelastic scattering of helium beams has become an important tool for measuring surface properties and surface phonons [3]. However, the step from observing diffraction to building an interferometer was not accomplished for more than 50 years because efficient mirrors and beam splitters for atoms were not available. During the same period, thermal neutron interferometers were built (for a review, see for instance the book of V. F. Sears [4]) : the de Broglie wavelength is comparable for neutron and atoms, but neutrons and atoms interact with matter in a very different way and most of the techniques used to build neutron interferometers cannot be used with atoms.

Main types of atom interferometers


Many types of atom interferometers have appeared since the eighties including polarization interferometers (for reviews see ref. [5,6]). Our goal here is not to give a review of all these works. We will concentrate here on interferometers in which the different atomic paths are spatially separated. From a practical point of view, two types of apparatus can be distinguished, depending if they use hot or cold atoms. The use of cold atoms (in the microkelvin range) provide clearly very long interaction times limited only because the atoms are falling in the Earth's gravity field : these experiments are applied to metrology (for an example, see ref. [7]). Several interferometers working with hot atoms have been operated. Although the interaction times are shorter than with ultra-cold atoms, these interferometers remain very interesting : as far as we know, up to now, it is only with hot atom interferometers that the various atomic paths are sufficiently separated so that a perturbation can be applied to only one path. This feature is fundamental to take full advantage of interferometry.

Three grating Mach-Zehnder interferometers


A very large fraction of the atom interferometry experiments are based on the Mach-Zehnder design, in which the beam splitters and the mirrors are replaced by diffraction gratings. This design, inspired by the perfect crystal interferometer for neutrons [8], was proposed for atoms by Clauser [9]. It is illustrated in figure 1 and presents several advantages :

its high symmetry makes the interferometer almost insensitive to weak defects in its geometry or alignment.
if the first grating G\ and third grating G% are identical, the two beams which interfere have the same amplitudes and the theoretical fringe contrast C (Imax Imin)/(Imax + Imin) is equal to its maximum possible value, C = I

the interference effect is not sensitive to the exact value of the atom wavelength so that the atomic beam may have a rather large velocity spread.

193

G3

FIGURE 1. Scheme of a Mach-Zehnder type interferometer.

The observation of interference fringes in such an interferometer can be done either by inducing a phaseshift on only one of the two interfering atomic paths (see below) or simply by displacing one of the three gratings in its plane : the displacement needed to sweep one fringe is just the grating period, not the atomic wavelength. We can explain this effect [10] with an argument developed for amplitude gratings by D. Pritchard and coworkers [11]. Two coherent atomic beams (propagating along the lower and upper paths shown in figure 1) recombine on the third grating where they form a stationary atomic wave which has exactly the grating period. The observation of fringes is therefore explained by a Moire filtering effect of the stationary atomic wave by the third grating and the displacement of this grating modulates the intensity of the output beams : these intensities are minimum (maximum) when the grating wires are at the antinodes (respectively the nodes) of the atomic stationary wave.

Beam splitters for atomic waves


The general type of beam splitter [12] for an atomic wave is obtained by a potential V(r,t), taken as an harmonic function of space and time : l/(r, z) VQ(Z) cos [uot
(2)

where VQ(Z) is the spatial envelope. An incident atom in the internal state a > (internal energy Ea) and momentum Kk{ absorbs a quasi-particle of energy hu and momentum /ik^ : this corresponds to an inelastic diffraction process where the atom is excited in the state |6>, of internal energy Eb = Ea + fiuo and final momentum Kkf Fiki + Kk.g. This formalism is very general and it can be used to describe almost any diffraction process. It must be slightly modified if the gratings are pulsed in time as for instance in the experiments of Kasevich and Chu [7]. We are now going to consider the simplest case of beam-splitters, where the initial final internal states a > and |6> are identical, i.e. u = 0. In this case, two types of beam-splitters can be used :
grids of very thin wires : the potential is not harmonic but it remains periodic. A simple approximation is to assume that the potential is infinite inside the material of the grid, zero outside. This approximation is not sufficient and

194

the van der Waals interaction of the atom with the material can play a role as shown by theory [10,13] and experiment [14]. This type of diffraction gratings has been used by D. Pritchard and coworkers [11] for sodium, by J. P. Toennies and coworkers with helium atoms, dimers and small clusters [14,15] and by A. Zeilinger and coworkers for C6o and other fullerenes [16].
laser standing waves close to atomic resonance. The potential due to the laser is the dynamical Stark effect. Provided that the laser is sufficiently detuned from resonance, the probability of spontaneous photon emission while crossing the laser standing wave can be very small : such a real excitation process must be avoided as it induces incoherent scattering of the atomic wave. Experiments on this type of diffraction process have been made first by D. Pritchard and coworkers [17,18] and more recently by several groups. The Bragg scattering regime first observed by the group of D. Pritchard [19] has been studied in detail by Siu Au Lee and coworkers [20] and is now commonly used to manipulate Bose Einstein condensates (see for example ref. [21]).

Using these two types of beam splitters, three atom interferometers have been operated since 1991 :
an interferometer using nanofabricated grids as gratings and a supersonic beam of sodium was built by D. Pritchard and coworkers [22,11]. The maximum count rate was 50000 atoms per second and the best fringe contrast was 50 %. an interferometer using Raman-Nath diffraction on laser standing waves and an atomic beam of metastable argon was built by A. Zeilinger and coworkers [23]. The typical signal count rate was 15000 atoms per second when the fringe contrast was close to 10 %. an interferometer using Bragg diffraction on laser standing waves and an atomic beam of metastable neon was built by Siu Au Lee and coworkers [20]. The typical signal count rate was 1500 atoms per second and the best fringe contrast was 62 %.

APPLICATION TO THE MEASUREMENTS OF WEAK PERTURBATIONS


Principle
We can apply a weak constant perturbation V to one of the atomic path inside the interferometer and not to the other paths, as illustrated in figure 1. The induced phase shift is simply given by :
Vr

(3)

195

where the interaction time r is r D/v, D being the length over which the perturbation is applied and v is the atomic velocity. With the values corresponding to the experiment of D. Pritchard and coworkers, D ~ 0.1 m and v ~ 1000 ra/s, the interaction time r is equal to r 10~4 s. From the fringe contrast and the count rate, the minimum detectable phaseshift appears to be A^ m ^ n = 10 milliradians for a 1 minute measurement time [25]. The resulting minimum detectable perturbation Vmin is quite small :

Vmin 10~32 J w 7 x 10~14 eV

(4)

This technique was applied by D. Pritchard and coworkers to measure with high accuracy the electric polarizability of atomic sodium in its ground electronic state [25]. This quantity which measures the absolute energy shift of the ground state cannot be measured by laser spectroscopy, which is only sensitive to the energy difference between two levels.

Test of the neutrality of lithium atom


To test the neutrality of an atom, one can place a parallel plate capacitor C of length D between two gratings of a Mach-Zehnder interferometer so that the two atomic paths propagate symmetrically inside the field region. The relevant perturbation term V here is given the multipolar expansion applied to the atom :

V - V0 + Vi + V2 + ... = qV - d- E - -aE2 + ...

(5)

where V and E stand for respectively the electric potential and field, d is a possible permanent electric dipole moment induced by the existence of a residual charge and a is the ground state static polarisability. If the atomic charge q is not vanishing, the permanent electric dipole moment can be cancelled by an appropriate choice of the origin and we will forget the corresponding energy term. The phase shift, A0, induced between the two waves can be written as :

hv

fa\ (6)

We have assumed an exact cancellation of the polarisability term as a consequence of the symmetry of the two paths. The mean separation between the two trajectories Jis proportional to the diffraction angle 9^ 2^dB/^L for first order diffraction by a laser standing wave operating of wavelength A/,.

d = Odzmean

(7)

Zmean = (zout + ^m)/2 where zin and zout represent the distances from the first grating to the two ends of the capacitor. Then

196

A JL EDzmeanOd ^ = qX Hv

.
(8)

The interferometer must be designed such as to give the largest possible phase shift. Besides an optimization of the geometrical dimensions D and zmean, the largest gain will be obtained by increasing the ratio Od/v which is proportional to v~2. For our numerical estimates, we have considered the case of lithium, a beam velocity v 10 ra/s, first order diffraction on a laser standing wave (the laser wavelength is 671 nm and corresponds to the resonance line) giving a quite large diffraction angle 6d = 17 mrad. The separation between consecutive gratings is taken equal to L 0.6 m with a capacitor of length D = 0.5 m located at middistance between the first and second gratings so that zmean L/2 = 0.3 m. The applied electric field E is taken equal to E 106 V/m. The resulting phase shift
A0 = qLi x 2.4 x 1036 rad. (9)

To reach a limit on the value of |qLi of the order of 10~21 x \qe (i.e. theexistinglim.it on \qn limit [26]), we must measure a phase shift A0min as small as 4 x 10~4 rad. This seems feasible with a reasonable acquisition time [27].

The experimental limit on the neutron charge viewed as an interferometry experiment


The experimental limit on the neutron charge was established by a direct measurement [26], which is a neutron optics experiment. In this experiment, a neutron absorbing grating G\ is imaged on a similar grating G^ by an achromatic grazing incidence cylindrical mirror with unit magnification (the distance from GI to G<2 is noted L). This experiment is schematically represented in figure 2, the mirror being replaced by a lens for simplicity of drawing. Two capacitors of total length D are used to apply an electric field E perpendicular to the optical axis and the measurement consists in detecting the resulting displacement of the image of grating GI on grating G^- This displacement, which should result in a modified neutron transmission and count rate, is given by :
A,= f^z

^mnv

(10)

where qn, mn and v are the charge, mass and velocity of the neutron. For more details, we refer the reader to the paper by Baumann et al. [26]. We are going to describe this experiment as an interferometry experiment. The first grating GI diffracts the neutron beam in various diffraction orders p and these beams recombine in a coherent fashion to produce a standing wave on G<2 which has exactly the period a of this grating. In the experiment, the grating period is a = 60 /xm, the neutron wavelength XdB is about 1.74 nm (this is a weighted

197

FIGURE 2. Schematical view of the experiment testing the neutron charge.

average value corresponding to the sensitivity of the experiment proportional to \2dB) resulting in a first order diffraction angle 64 \dB/a 29 /jirad. Fringes are detected by displacing this grating in its plane and the fringe contrast (as shown in figure 3 of ref. [26]) is C ~ 30 %, with a neutron count rate of 30000 s~l. For the diffraction order p, the phase shift induced by the electric field is given by an equation very similar to equation 8 :

with

qnEDL9d

It is easy to calculate the interference pattern of the series of diffracted beams and to see that a phase shift proportional to the diffraction order simply shifts the image in its plane by a quantity
(12)

2?r

This quantum point of view is complementary to the classical trajectory point of view developed in ref. [26] and it gives the same result. The achieved sensitivity can be expressed in terms of a phase sensitivity A(/>(1). A neutron charge equal to 10~21 x \qe\ induces an image shift AT/ = 4.8 x 10~10 m for the mean neutron wavelength X^B 1-74 nm. The corresponding phase shift A(/>(1) is equal to A0(l) = 5 x 10~5 rad. This extremely high sensitivity was obtained after several hundred hours of data acquisition and it is limited by some systematics [26].

CONCLUSION
In this talk, we have tried to give a brief historical perspective of atom interferometry and to show that the extreme sensitivity of this technique to very weak perturbations. We have illustrated this discussion by a proposed measurement of a possible non zero charge of the lithium atom and we have shown that the direct measurement of the neutrality of the neutron can be viewed as an interferometry experiment.

198

REFERENCES
1. 2. 3. 4. 5. 6. 7. O. Stern, Naturwiss. 17, 391 (1929) N. F. Ramsey, Molecular Beams, Oxford University Press (1956) J. P. Toennies, Springer series in Surface Science 14, 248 (1988), and 21, 111 (1991) V. F. Sears, Neutron optics, edited by Oxford University Press (1989) Atom interferometry, edited by P. R. Berman (Academic Press 1997) J. Baudon, R. Mathevet and J. Robert, J. Phys. B, 32, R173 (1999) B. Young, M. Kasevich and S. Chu, in Atom interferometry edited by P. R. Berman (Academic Press 1997) p 363 H. Rauch, W. Treimer and U. Bonse, Phys. Lett. 47A, 369 (1974) J. F. Clauser, Physica B 151, 262 (1988) C. Champenois, M. Blichner and J. Vigue, Eur. Phys. J. D, 5, 363 (1999) J. Schmiedmayer, M. S. Chapman, C. R. Ekstrom, T. D. Hammond, D. A. Kokorowski, A. Lenef, R.A. Rubinstein, E. T. Smith and D. E. Pritchard, in Atom interferometry edited by P. R. Berman (Academic Press 1997) p 1 C. J. Borde, in Atom interferometry edited by P. R. Berman (Academic Press 1997)

8. 9. 10. 11.

12.

p 257 13. G. C. Hegerfeldt and T. Kohler, Phys. Rev. A 61, 023606 (2000) 14. R. E. Grisenti, W. Schollkopf, J. P. Toennies G. C. Hegerfeldt and T. Kohler, Phys. Rev. Lett. 83, 1755 (1999) 15. W. Schollkopf and J. R Toennies, Science 266, 1345 (1994) 16. M Arndt, O. Nairz, J. Voss-Andreae, C. Keller, G. van der Zouw and A. Zeilinger Nature 401, 680 (1999) 17. P.E. Moskowitz, P. L. Gould, S. R. Atlas and D. E. Pritchard, Phys. Rev. Lett., 51, 370 (1983) 18. P. L. Gould, G. A. Ruff and D. E. Pritchard, Phys. Rev. Lett., 56, 827 (1986) 19. P. J. Martin, B. G. Oldaker, A. H. Miklich and D. E. Pritchard, Phys. Rev. Lett., 60, 515 (1988) 20. D.M. Giltner, R. W. McGowan and Siu Au Lee, Phys. Rev., A 52, 3966 (1995) 21. M. Kozuma, L. Deng, E. W. Hagley, J. Wen, R. Lutwak, K. Helmerson, S. L. Rolston and W. D. Philipps, Phys. Rev. Lett. 82, 871 (1999) 22. D. W. Keith, C. R. Ekstrom, Q. A. Turchette and D. E. Pritchard, Phys. Rev. Lett. 66, 2693 (1991) 23. E. M. Rasel, M. K. Oberthaler, H. Batelaan, J. Schmiedmayer and A. Zeilinger, Phys. Rev. Lett., 75, 2633 (1995) 24. D.M. Giltner, R. W. McGowan and Siu Au Lee, Phys. Rev. Lett., 75, 2638 (1995) 25. C. R. Ekstrom, J. Schmiedmayer, M. S. Chapman, T. R. Hammond and D. E. Pritchard, Phys. Rev. A 51, 3883 (1995) 26. J. Baumann, R. Gahler, J. Kalus and W. Mampe Phys. Rev. D 37, 3107 (1988) 27. C. Champenois, M. Biichner, R. Delhuille, R. Mathevet, C. Robilliard and J. Vigue, Proc. of the workshop Hydrogen atom II, S. G. Karshenboim and F. Pavone editors (Springer 2000) to be published

199

You might also like