13
This content has been downloaded from IOPscience. Please scroll down to see the full text. Download details: IP Address: 208.87.72.84 This content was downloaded on 01/10/2014 at 02:43 Please note that terms and conditions apply. Testing the equivalence principle with atomic interferometry View the table of contents for this issue, or go to the journal homepage for more 2012 Class. Quantum Grav. 29 184003 (http://iopscience.iop.org/0264-9381/29/18/184003) Home Search Collections Journals About Contact us My IOPscience

Testing the equivalence principle with atomic interferometry

  • Upload
    claus

  • View
    215

  • Download
    1

Embed Size (px)

Citation preview

Page 1: Testing the equivalence principle with atomic interferometry

This content has been downloaded from IOPscience. Please scroll down to see the full text.

Download details:

IP Address: 208.87.72.84

This content was downloaded on 01/10/2014 at 02:43

Please note that terms and conditions apply.

Testing the equivalence principle with atomic interferometry

View the table of contents for this issue, or go to the journal homepage for more

2012 Class. Quantum Grav. 29 184003

(http://iopscience.iop.org/0264-9381/29/18/184003)

Home Search Collections Journals About Contact us My IOPscience

Page 2: Testing the equivalence principle with atomic interferometry

IOP PUBLISHING CLASSICAL AND QUANTUM GRAVITY

Class. Quantum Grav. 29 (2012) 184003 (12pp) doi:10.1088/0264-9381/29/18/184003

Testing the equivalence principle with atomicinterferometry

Sven Herrmann1, Hansjorg Dittus2, Claus Lammerzahl1,3,(for the QUANTUS and PRIMUS teams)

1 ZARM, Universitat Bremen, Am Fallturm, 28359 Bremen, Germany2 DLR, Koln, Germany3 Institut fur Physik, Universitat Oldenburg, 26111 Oldenburg, Germany

E-mail: [email protected]

Received 9 March 2012, in final form 26 June 2012Published 15 August 2012Online at stacks.iop.org/CQG/29/184003

AbstractThe weak equivalence principle (WEP), that is, the universality of free fall,states that all point-like neutral particles in a gravitational field fall in the sameway. This is the basis of the geometrization of the gravitational interaction.Together with further requirements on the behavior of point particles, lightpropagation and clocks one can show that gravity is modeled by a Riemanniangeometry. Since in the quantum domain all objects are extended, it is not clearwhether the notion of a WEP in the quantum domain makes sense at all. Weshow that for matter wave interferometry the notion of WEP still can be given ameaning. We give a short overview over schemes which allows a violation of theWEP and emphasize that there are also schemes which show that there might beviolations of the WEP in the quantum regime which are not present classically.This makes a test of the WEP with quantum matter necessary. We also givea brief outline of the efforts made for testing the WEP with interferometrywith cold atoms in the Bremen drop tower carried out by the QUANTUS andPRIMUS collaboration.

PACS numbers: 04.80.Cc, 04.60.−m, 03.75.Dg, 37.25.+k

(Some figures may appear in colour only in the online journal)

1. Meaning of the EP

The weak equivalence principle (WEP), also called the universality of free fall (UFF), isone of the cornerstones of general relativity (GR). Together with the universality of thegravitational redshift (UGR) and the local validity of Lorentz invariance (LLI), it forms theEinstein equivalence principle [2]. Here we are dealing mainly with the UFF in the classicaland quantum domain.

Any universality principle, such as the UFF and the UGR, for the dynamics of a systemstates that a class of phenomena is independent from the parameters characterizing the physical

0264-9381/12/184003+12$33.00 © 2012 IOP Publishing Ltd Printed in the UK & the USA 1

Page 3: Testing the equivalence principle with atomic interferometry

Class. Quantum Grav. 29 (2012) 184003 S Herrmann et al

system. As a consequence, these phenomena can be interpreted as an effect of an underlyinggeometry.

1.1. General meaning: geometrization

We are now restricting ourselves to the dynamics of point-like objects, that is, objects whichdo not have any intrinsic structure like, e.g., spin or mass multipoles. This will have to bemodified in the context of quantum mechanics since there we do no longer have point-likeobjects: quantum objects are always extended and, thus, intrinsically violate the UFF. It willbe one task to investigate whether there is still some notion of universality in the quantumregime.

1.2. The notion of UFF

1.2.1. Newtonian framework. The Newtonian framework is given by the Newtonian axioms.The main issue there is the Newton second axiom mix = F, where x(t) is the trajectory of thepoint-like body, mi is the inertial mass and F is the force acting on the body.

The gravitational force is assumed to be derived from a potential which for a mass densitydistribution ρ(x, t) is given by a set of moments

Ui1i2···in (x, t) = G∫ ∏n

j=0(x − x′)i jρ(x′, t)

|x − x′|n+1d3x′, n even, (1)

which for n = 0 reduces to the ordinary Newtonian potential U . We also have δi jUi j = U . Forthis general setup the general equation of motion is

xi = ai jk1···kn∂ jU

k1···kn (x, t), (2)

where mi is absorbed into the parameters ai jk1···kn which in general may depend on the particle.

In the case ai jk1···kn = 0 for n > 2 and ai j = δi j + αi j and ai j

k1k2 = δi jαk1k2 we recover thecase described in [3] and for further specialization αi j = αiδ

i j and αk1k2 = αgδk1k2 , we obtain

xi = (1 + αi + αg)δi j∂ jU. (3)

Comparison with mix = mg∇U gives αi+αg = δmg

m − δmim for mg = m+δmg and mi = m+δmi.

Therefore, to the lowest order αi +αg is the Eotvos coefficient. This result [3] can be obtainedfrom the Hamiltonian

H = 1

2m

(δi j + α

i ji

)pi p j + m(δi j + αgi j)U

i j. (4)

1.2.2. General relativistic framework. Within a four-dimensional setting a worldline of apoint-like body is given by a curve xμ(λ), where μ = 0, . . . , 3 and λ is some parameter ofthe curve. In general, this curve might depend on the whole history, that is, in general wehave dx(τ0) = F[x, τ < τ0] dτ , where F[x] is a functional of the whole past curve. It is avery special feature of physics that a worldline is completely determined by its initial positionand initial velocity. Then a worldline is given by a second-order ordinary differential equationxμ = H(x, x; p), where H is some function of xμ, xμ, and of the parameters characterizing theparticle such as charge, mass, etc.

It is now the characteristic of the gravitational interaction that the worldline does notdepend on any of the parameters p characterizing the particle: xμ = H(x, x). The function Hthen can be regarded as a property of a spacetime geometry which in this general setting isgiven by a path structure. This geometrization is the main consequence of the UFF.

2

Page 4: Testing the equivalence principle with atomic interferometry

Class. Quantum Grav. 29 (2012) 184003 S Herrmann et al

Only if one requires that gravity can be transformed away, that is, if there is at eachspacetime point a coordinate system so that the equation of motion for all particles has theform xμ = 0, then the equation of motion in general coordinates has the form vνDνv

μ = αvμ,where α is a function which depends on the parametrization of the worldline and Dμ is acovariant derivative based on some affine connection [4]. With the additional requirement thata particle should not be faster than light and that the ticking rate of clocks should not dependon their history, we end up with a Riemannian geometry as a mathematical description for thegravitational interaction [4].

It is obvious that gravity cannot be transformed in theories with a velocity-dependentconnection like those provided by Finsler geometries. A metrical relativistic framework whichprovides a description of a violation of the UFF is provided by the PPN formalism [2].

This defines the importance of the UFF in the classical domain where one uses idealizedobjects like point particles. The question now is whether it is possible to define a similar notionin the quantum domain where wave packets in principle are spread out and, thus, are nonlocal.

1.3. EP in the quantum domain

There have been many publications where it has been stated that the UFF is violated in thequantum domain. These statements are mainly based on the fact that for nonlocal wave packetsthe inhomogeneity of the gravitational field yields a trajectory of a mean position which differsfrom a geodesics. While this is not surprising, the important question is whether despite thisfeature there might be an operational way to characterize the geometric coupling of gravity(minimal coupling) in the quantum domain.

In fact, it can be shown that interference is an appropriate means to operationallycharacterize that gravity is universally coupled [5]. The phase shift for an interferenceexperiment can be shown to be independent of the mass, if the momentum transfer whichcauses the splitting of the wavefunction is small. In the limit of small enclosed areas thismeans that the phase shift does not depend on the mass for a minimally coupled underlyingfield equation describing the dynamics of the quantum matter. For vanishing curvature weobtain δφ = k ·gT 2 which also holds for large areas. (This is equivalent to the famous Colella–Overhauser–Werner phase shift [5].) One can also show the other way around. If the phaseshift does not depend on the mass, then the field equation is minimally coupled to gravity [6].

A general phenomenological approach to describe violations of the UFF as well as that oflocal Lorentz invariance and local position invariance has been presented in [7]. This is basedon a generalized Dirac equation iγ μ∂μψ −Mψ = 0 where the matrices γ μ are not assumed tofulfil a Clifford algebra and M is some matrix. Such a general equation leads to birefringenceand anisotropic propagation. In the non-relativistic limit, this gives a Hamiltonian of the sameform as in equation (4).

2. Predictions of a violation of the EP

All quantum gravity scenarios predict a violation of the UFF. Most prominent are string-theory-based scenarios for which we refer to the paper by Damour [8].

Another scenario is based on spacetime fluctuations in terms of stochastic fluctuationsof the metric around a given background. One starts with a minimally coupled relativisticwave equation characterized by a single mass parameter. Then the metrical fluctuations leadto an apparent violation of the UFF in an effective wave equation. The reason for this isthat the different spatial extensions of particles of different masses lead to an averagingof the fluctuations over different spatial volumes of the wavefunction, ending up in a

3

Page 5: Testing the equivalence principle with atomic interferometry

Class. Quantum Grav. 29 (2012) 184003 S Herrmann et al

particle-dependent effective inertial mass in the corresponding effective Schrodinger equation.For, e.g., a holographic noise scenario, the UFF will be violated on the order of 10−15 if onetakes the geometric extension of the quantum objects as the characteristic length [9]. Suchfluctuations also lead to decoherence [10] and to an additional spreading of the wave packet[11], with both being also accessible through atomic interferometry.

3. Basic concepts of matter wave interferometry

Matter wave interferometry today is an established tool for precision measurements such asmeasurements of the local gravitational acceleration [12], gravity gradients [13], or Newton’sgravitational constant [14, 15]. As such, it may also be a powerful technique to perform a testof the WEP, in particular the UFF of atoms. In addition, ultra-cold atomic ensembles constituteespecially well-defined test masses in terms of spin properties, bosonic or fermionic nature,proton to neutron ratio, etc. If condensed to a Bose–Einstein condensate (BEC), they furtherprovide a test of the WEP with a quantum object, which allows for probing the WEP in apossibly interesting domain, as discussed in the previous section.

Among the many different realizations of atom interferometers [16], the present discussionfocuses on light pulse atom interferometers with neutral atoms, where beam splitters andmirrors are realized by short laser light pulses.

Here, a source of coherent matter waves is provided by the preparation of an ultra-coldensemble of neutral atoms, typically alkali atoms such as Rb or Cs. Initially, the atoms arecooled to μK temperature by standard methods of laser cooling. In many applications itis sufficient to apply subsequent velocity selection of the atoms to provide the necessarycoherence for interferometry. However, the thermal spread of an ensemble of e.g. Rb atomsat μK temperature is on the order of cm s−1. It is thus often desirable to further cool theatoms to ultra-low temperatures and even quantum degeneracy (BEC) at a few tens of nK bysubsequent evaporative cooling in an optical or magnetic trap. For interferometry at extendedfree evolution times of the order of seconds, the use of such ultra-cold atoms may even beconsidered mandatory.

A powerful implementation of a matter wave beam splitter is achieved by means of atwo-photon Raman transition between two hyperfine levels of the ground state [17, 18]. Thisprocess combines the absorption of a photon from one beam with the stimulated emissionof a photon into a second beam. With two beams detuned by the hyperfine splitting andcounter propagating in opposite directions, the transferred photon recoils add up along thesame direction. The corresponding velocity change of the atoms is of the order of severalmm s−1. If each of the two laser beams is sufficiently detuned with respect to the excited state,spontaneous emission from that state can be neglected. The interaction can then effectivelybe described as a two-level atom that undergoes Rabi oscillations between the two hyperfinestates. The effective Rabi frequency is a function of pulse length and power. Typically, squarepulses with durations of the order of a few tens of μs are applied, which provides sufficientspectral Fourier width to address all atoms in the atomic ensemble. By adjusting the pulselength by a factor of 2, either a matter wave beam splitter (π/2-pulse) or a mirror (π -pulse)can be realized. As an alternative to Raman beam splitters, Bragg diffraction of the atoms ona light wave can be used [19] as well. Similar to a stimulated Raman transition, this processcan be described as a two-photon transition, but between different momentum states withinthe same internal state of the atom. The two laser beams in this case are detuned only bythe recoil frequency splitting of the order of several kHz. The benefit of Bragg pulses isthat atoms in both arms remain in the same internal state, which for interferometry providesthe rejection of several noise sources e.g. due to ac Stark shifts or temporal fluctuations of

4

Page 6: Testing the equivalence principle with atomic interferometry

Class. Quantum Grav. 29 (2012) 184003 S Herrmann et al

(a) (b)

Figure 1. (a) Schematic of a Mach–Zehnder light pulse atom interferometer. ToF = time of flight.T = pulse separation time. The relative population of ports n1 and n2 after interference at the instantof the final beam splitter depends on the relative phase of the matter waves in both interferometerarms. (b) Schematic of a two-photon Raman transition between hyperfine states 1 and 2 of theground state.

the magnetic field. Additionally, the low rf frequencies required allow for a rather simpleimplementation e.g. by using frequency modulators. On the other hand, detection of the atomsis not straightforward since the two interferometer ports cannot be separated state selectively,but need to be separated spatially before detection. Also, this process requires narrow velocitydistribution of the ensemble and low Fourier width of the applied pulses, in order to fullyseparate diffraction orders.

A widely applied configuration of a light pulse matter wave interferometer is an analogonto the optical Mach–Zehnder interferometer (MZI, see figure 1). In this configuration, theatomic ensemble is split by a beam splitter laser pulse, redirected by a mirror pulse andrecombined by a final beam splitter pulse, all equally spaced in time by the pulse separationtime T. With pulse separation times being of the order of hundreds of ms, as obtained in atomicfountains, and recoil velocities of the order of mm s−1; the separation of the two interferometerarms is typically of a few mm only.4 The phase difference of the interferometer is determinedby detecting the relative proportion of atoms emerging from the two ports of the interferometerafter the final beam splitter is applied. With Raman beam splitters, the distinction between portscan be done by state selective detection, since the atoms in either port are in different hyperfinestates. When scanning the phase difference of the interferometer paths over subsequent runs,an oscillation of the emerging atom numbers in either port can be observed. These oscillationsare typically referred to as the interferometer fringes.

The phase shift between the paths of a Mach–Zehnder light pulse atom interferometerhas been derived e.g. in [20]. Neglecting the effect of gravity gradients, it can be dividedinto contributions from (i) the propagation of the matter wave (ii) and the interaction ofthe light with the matter wave. In a semiclassical treatment it can be shown that the firstcontribution vanishes for symmetry reasons (again no gravity gradient considered), while thesecond contribution adds up to a term

= keff aT 2, (5)

wherein keff = |�k1 −�k2| is the effective wave vector of the laser pulses, T is the pulse separationtime and a is the acceleration the atoms undergo. For counter propagating beams the effectivewave vector is approximated as keff = 2k, where k = k1 ≈ k2. An intuitive description of

4 Note that to reduce the thermal spreading of the atomic ensemble below the splitting separation ultra-cold atoms atsub-μK are required.

5

Page 7: Testing the equivalence principle with atomic interferometry

Class. Quantum Grav. 29 (2012) 184003 S Herrmann et al

equation (1) is to state that the acceleration a of the atoms is effectively measured against thebeam splitter laser wavefronts. Aside other sources of phase noise, the shot noise limit of thedetection of n1 and n2 then determines the resolution in acceleration. For the detection of N(= n1 + n2) atoms and a pulse separation time T we can estimate the resolution as

δa = 1√NkT 2

. (6)

From the above equations we can estimate the basic sensitivity of a matter wave interferometertoward accelerations. For Rb atoms we use laser light at 780 nm, that is keff ≈ 4π/λ = 1.6×107

m−1. With a pulse separation time of T = 100 ms and freely falling atoms at a = g this resultsin a phase shift of = 1.6 × 106 rad. If the signal-to-noise ratio of the interferometer fringesallows for a phase resolution of δ = 10 mrad per shot, which has been achieved e.g. by [12],this allows for a single shot resolution in g of

δg

g= δ

= 10−8 @T = 100 ms (7)

or since the phase scales quadratically with free fall time T ,δg

g= δ

= 10−10 @T = 1 s. (8)

4. Atom interferometry tests of the WEP

4.1. Past tests and tests currently in preparation

A test of the UFF may be obtained by comparing the local gravitational accelerations gi andg j of two atomic species i and j as determined from an atom interferometer measurement. TheEotvos parameter then can be expressed as

ηi j = 2(gi − g j)

gi + g j. (9)

Along the above lines, Peters et al [12] have measured the local gravitational accelerationof Cesium atoms in an atomic fountain of 0.5 m in height. This allowed them to use a pulseseparation time of T = 160 ms and to achieve a resolution of 3 × 10−9 g in 2 min, that is,averaging over 40 data points. They also used this to perform a comparison with a FG5 fallingcorner cube gravimeter located nearby the atomic fountain. Thus, they were able to performa test of the UFF of Cesium atoms as compared to a macroscopic body, which they found toagree to within (7 ± 7) × 10−9 g, limited by the measurement of the local gravity gradient.The latter needed to be taken into account as a correction to the measurement result, since bothexperiments were located at a separation of the order of 2 m (with 0.5 m difference in height).Indeed, this is a critical issue of this comparison since any gravity gradient not accounted forcould potentially mask a violation of the WEP.

An approach that can alleviate this issue has been followed by Fray et al [21]. They haveused an atom interferometer in subsequent runs with 87Rb and 85Rb to compare the free fall ofthese isotopes at the same location with an accuracy of 10−7 g. They prepared atoms by lasercooling to a temperature of 6 μK and used diffraction of the atoms on a light grating, pulseseparation times up to T = 40 ms. This limit was set by the loss of interferometer visibility atlonger interaction times presumably due to vibrations.

Further improvements to these precision tests of the WEP are expected along the followinglines: first of all, comparisons of different atomic species might be pursued that are better suitedto reveal a violation of the WEP in a particular scenario. Besides atoms differing substantiallyin mass (e.g.133Cs and 6Li) comparisons of other isotopes that differ largely with respect to

6

Page 8: Testing the equivalence principle with atomic interferometry

Class. Quantum Grav. 29 (2012) 184003 S Herrmann et al

their composition of neutrons and protons might be pursued. In particular, a comparison ofbosonic species to fermionic species might be of special interest, e.g. 87Rb and 40K.

Furthermore, there has been a lot of progress toward the application of large momentumtransfer (LMT) beam splitters in recent years. Such beam splitters transfer multiple photonmomenta to the atoms and allow us to significantly increase the enclosed area and thus thesensitivity of the interferometer. In [22] the coherent beam splitting by 102 �k has alreadybeen demonstrated and in [23] an interferometer with much increased enclosed spacetime area(T = 250 ms and 10 �k momentum beam splitters) has been reported.

Finally, from equation (5) it is evident that phase sensitivity scales quadratically with thepulse separation time. To significantly increase the latter, one needs to increase the availabletime of free fall. Several research groups have already taken on this challenge.

The group of Kasevich in Stanford is currently working on an apparatus that will allow for10 m of free fall, enabling them to expand the free fall time up to 1.4 s [24]. They are aimingto compare the free fall of 87Rb and 85Rb evaporatively cooled to sub-μK temperature. Thechoice of these two similar isotopes allows them to use a single laser for the implementationof the beam splitters which will possibly apply multiple photon transfer as in [22]. Using thesame laser in the two simultaneously applied interferometers will allow for the suppression oflaser phase noise and systematic effects.

The French ICE project operates an apparatus onboard the zero-g Airbus during parabolicflights, where they are aiming to perform measurements on a two-species atom interferometer[25]. The environment onboard the airplane provides them with a free fall time of up to20 s at residual accelerations of a < 10−2 g. The two species chosen for ICE are 87Rb and40K atoms, which require different wavelengths of 780 and 767 nm, respectively. These areprovided by frequency doubling of fiber lasers at 1560 and 1534 nm, respectively. Ultimately,they aim to perform a test of the UFF on a mixture of 87Rb and 40K atoms. As discussed in[26] the implementation of the interferometer with two separate lasers of different wavelengthonly allows for the limited suppression of common vibrational phase noise in the differentialmeasurement.

The QUANTUS collaboration5 finally is also headed toward a test of the UFF based onthe comparison of 87Rb and 40K. More details on this activity at the Bremen drop tower arepresented in the following.

4.2. Cold-atom experiments at the Bremen drop tower

The benefit of the drop tower over other microgravity platforms is the excellent weightlessnessenvironment it provides at relatively low cost, with residual accelerations and vibrations at the10−6 g level. This makes it a pristine platform to carry out pioneering atom interferometryexperiments in zero-gravity, ultimately aiming to test the UFF. The drop tower is a totalof 141 m in height and allows for 4.7 s of free fall, or 9.4 s if operated in catapult mode(see figure 2). The experimental payload is integrated into a drop capsule and kept underatmospheric pressure therein. Prior to each drop, the drop tube is evacuated to a residualpressure of 10 Pa by pumping for 90 min. After the release of the capsule and subsequent freefall from a height of 110 m, it is captured on the ground level in a deceleration container filledwith styrofoam particles. The deceleration therein occurs over 250 ms reaching a peak valueof up to 50 g.

5 The QUANTUS project is a collaboration between the Institute of Quantum Optics at the Leibniz University ofHanover, ZARM at the University of Bremen, the Institute of Physics at the Humboldt-University of Berlin, theInstitute of Laser-Physics at the University of Hamburg, the University of Birmingham, the Max-Planck-Institute ofQuantum Optics, Munich, the Institute of Quantum Physics at the University of Ulm, and the Institute of AppliedPhysics at the University of Darmstadt.

7

Page 9: Testing the equivalence principle with atomic interferometry

Class. Quantum Grav. 29 (2012) 184003 S Herrmann et al

(a)

(b)

Figure 2. (a) The drop tower at ZARM Bremen [27]. The total height is 141 m. The inner vacuumtube is 120 m in height and fully decoupled from the outer structure. The capsule is released fromthe top at 110 m and capture in a deceleration container on ground. (b) Capsule and decelerationcontainer (left) inside the deceleration chamber at the ground level.

(a) (b)

Figure 3. (a) Schematic of the QUANTUS-I setup in the drop capsule [27]. The total height of thecapsule is 215 cm, with an inner diameter of 60 cm. (b) Schematic of the QUANTUS-II vacuumsetup [38]. The total height of the capsule is reduced to 140 cm. The total payload weight isapproximately 150 kg.

In a joint effort, the QUANTUS collaboration for the first time realized a cold-atomexperiment in this challenging environment, demonstrating the preparation of a degeneratequantum gas, a BEC, as a suitable source for further matter wave interferometer experimentsin extended free fall. The QUANTUS-I apparatus is fitted into a drop capsule, with a payloadcompartment of 2.15 m in height and 0.7 m in diameter (figure 3(a)). It comprises an ultra-high

8

Page 10: Testing the equivalence principle with atomic interferometry

Class. Quantum Grav. 29 (2012) 184003 S Herrmann et al

vacuum chamber containing an atom chip with optical access from several view ports. Laserlight for cooling, preparation and detection of the atoms is provided via optical fibers from aminiaturized laser system based on robust, frequency stabilized DFB laser diodes at 780 nm.Laser control electronics as well as computer control and data acquisition are all onboard thedrop capsule. The complete setup can be remotely operated on a 24 V, 40 Ah battery powersupply.

As the source of atoms, Rubidium dispensers are fired prior to each drop and approximately108 atoms are loaded into a magneto-optical trap (MOT) over 15 s from the backgroundpressure. The MOT is realized as a mirror MOT by using the chip surface for reflection oftwo MOT beams incident on 45◦. The release of the capsule then triggers the further coolingsequence. During the subsequent 4.7 s of free fall the atoms are cooled in an optical molassesand transferred to the magnetic chip trap (n ≈ 4 × 106). There, forced evaporation cooling isused to further increase phase space density and cool the atoms down to phase transition to aBEC. Typically, after 2.9 s of free fall we are left with 104 atoms in a BEC, which is carefullyreleased from a shallow magnetic trap of trap frequencies in the low Hertz regime. This leavesapproximately 1 s for the observation of the freely expanding BEC.

Details of the measurements performed with this setup have been reported in [27]. Asdescribed there, atoms were initially prepared in a magnetically sensitive mF = 2 state and thefree expansion was distorted by residual magnetic field gradients. Meanwhile, atoms can betransferred to a non-magnetic state mF = 0 via an adiabatic sweep of a radio frequency signalthat couples the different mF levels and the effect of residual magnetic field gradients is thuseliminated. With that the QUANTUS-I apparatus now provides a well-established chip-basedsource for coherent matter waves and can now be used to carry out the first atom interferometermeasurements in a microgravity environment. The first such measurements have already beenperformed and are currently prepared for publication [28].

In 2012, the QUANTUS-I apparatus will be superseded by an even more compactapparatus QUANTUS-II, which is currently being completed (3(b)). This apparatus will beoperated in catapult mode extending the time of free fall to 9.4 s. It is equipped with a newgeneration atom chip, greatly extending the possibilities and efficiency of atom preparation. Itwill also use a 2D-MOT for faster loading of more atoms into the 3D mirror MOT, which willbe realized solely with magnetic fields generated from mesoscopic chip structures, withoutlarge external magnetic coils. The layout of this new apparatus is specifically adapted for theoperation of an interferometer, including Raman lasers for beam splitting of the matter waves.Ultimately, QUANTUS-II shall then be used for the preparation of an ultra-cold mixture of87Rb and 40K atoms, and also for dual species atom interferometry.

4.3. Toward a test of the WEP at the drop tower

QUANTUS-II shall then provide the basis for an atom interferometer test of the WEP in thedrop tower. Such a test will be pursued in a dedicated effort, PRIMUS, within the QUANTUScollaboration. The experiment will apply two simultaneous interferometers with 87Rb and40K. This choice of atoms offers several advantages. First of all, the concepts for cooling theseatoms to degeneracy have already been demonstrated e.g. in [29–32]. Also, the interactionproperties of a rubidium–potassium mixture have been studied extensively e.g. in [33, 34].In addition 87Rb/40K features a larger difference in the mass and proton–neutron ratio ascompared to rubidium isotopes, and there also exist different isotopes of potassium to choosefrom, fermionic 40K or bosonic 39K and 41K. Furthermore, the similar wavelengths of Rband K D2 lines at 780 and 767 nm permit us to use the same optics (fibers, mirror coatings,etc) throughout large parts of the laser system. Last but not least, a small difference in the

9

Page 11: Testing the equivalence principle with atomic interferometry

Class. Quantum Grav. 29 (2012) 184003 S Herrmann et al

wavelength of 13 nm also favors the enhanced suppression of common laser phase noise inthe differential interferometer measurement.

For a test of the WEP, an ultra-cold mixture of these atoms will be prepared and thetwo interferometers will be applied simultaneously during free fall in the drop tower. Thesingle shot resolution at which the interferometer phase can be resolved depends on the freefall time T and the signal-to-noise level of the detection of the interferometer fringes. Thisagain is fundamentally limited by the particle number N. We aim to operate with T = 1 s androughly N = 105 particles, which sets a fundamental limit of approximately 10 mrad per shot.According to equation (8) this corresponds to 10−10 g single shot resolution in acceleration.Since each interferometer (atoms and lasers) is in free fall, ideally there should be no phaseshift in both interferometers. Still, residual accelerations of the order of 10−6 g due to residualair drag, vibrations or residual capsule rotation are present. Without further measures tosuppress these effects, this results in a phase shift of the order of 100 rad varying from dropto drop. Such phase noise will prevent the straightforward extraction of valuable accelerationdata from a single interferometer. However, the operation of a dual species interferometerallows for a differential measurement which is expected to alleviate this problem substantially.Since both interferometers are subject to the same vibrations, the resulting phase noise inthe interferometer is common mode and should affect the differential signal much less. Thecommon mode suppression is however not perfect because both interferometers apply laserbeam splitters of different wavelengths. An estimate of the suppression factor has been given in[26]. In the limit of low-frequency vibrations, the suppression is (k1−k2)/k. With k1 = k = 780nm and k2 = 767 nm this results in approximately 60-fold suppression of common phasenoise. An estimate of the differential phase can then be obtained by ellipse fitting or by using aBayesian estimate as described in [35] and [26]. The latter method generally provides accuratephase estimates already from a limited number of measurements. Since the drop tower allowsfor only three drops per day, this is an especially useful feature.

Ultimately, a drop tower test of the WEP may thus be capable of demonstrating adifferential single shot resolution of the order of δa = 10−10 g. Based on such a singleshot performance we extrapolate that from about 30 data points a sensitivity estimate on theEotvos parameter of the order of η < 5 × 10−11 may ultimately be obtained. While this is notcompetitive with current torsion pendulum experiments, and a careful study of systematicswould certainly require substantially more than 30 drops, it could already exceed the sensitivityof the current best Galilean free fall test by Kuroda et al [36].

5. Outlook: toward a space-based test of the WEP with atoms

The full potential of atom interferometry to test the WEP will finally only be exploited in aspace-based experiment. This will combine the benefit of extended free evolution times of theorder of seconds, i.e. largely increased single shot resolution, with integration over a largenumber of measurements. This potential has been recognized by the European Space Agencyas well. Following previous activities such as the Space Atom Interferometer project [37],ESA has now selected the STE-QUEST mission proposal (Space time Explorer–QuantumEquivalence Principle Test) within the Cosmic Vision program call for 2015–2025 as apossible M-class mission [38]. Besides clock comparison experiments to test special andgeneral relativity, this mission also comprises operation of a dual species atom interferometeronboard a satellite in the Earth’s orbit to compare the free fall of Rubidium isotopes. Such asatellite-based interferometer may operate with interferometer pulse separation times of T = 5 sand a cycle time of 20 s. The mission goal is to obtain a single shot resolution in the differential

10

Page 12: Testing the equivalence principle with atomic interferometry

Class. Quantum Grav. 29 (2012) 184003 S Herrmann et al

acceleration of 10−12 m s−2, and an overall sensitivity of η < 10−15 after integrating over thefull mission lifetime.

6. Conclusion

Matter wave interferometry is a very active and rapidly advancing field of research thatprovides an interesting avenue to improved precision tests of the WEP. Currently, the mostaccurate free fall tests of the WEP involving atom interferometry obtain a sensitivity ofη ∼ 10−9 limited by vibrations and the available time of free fall. Within the near future,dedicated WEP precision tests are expected to substantially improve on this sensitivity byusing differential atom interferometers, the application of multi-photon beam splitting andparticularly by making use of extended free fall times. An excellent platform to demonstratethe feasibility of the latter approach is provided by the Bremen drop tower, where experimentswith cold atoms are already being performed on a regular basis. Such drop tower experimentsmay achieve sensitivities of η < 10−11 and thereby provide a valuable stepping stone towarda satellite-based atom interferometer that may even aim for η ∼ 10−15 sensitivity [38]. Sincethe latter sensitivity estimate up to now requires extrapolation over many orders of magnitudein sensitivity as well as the suppression of systematic effects, such stepping stone experimentsas pursued at the drop tower are urgently required.

Acknowledgments

The QUANTUS and PRIMUS projects are supported by the German Space Agency DLRwith funds provided by the Federal Ministry of Economics and Technology (BMWi) due toan enactment of the German Bundestag under grant numbers DLR 50WM1131-1137 andDLR 50WM0842, respectively. We also thank the German Research Foundation for fundsprovided by the Center of Excellence QUEST, Center for Quantum Engineering and Space-Time Research.

References

[1] Will C M 1993 Theory and Experiment in Gravitational Physics (Cambridge: Cambridge University Press)(revised edition)

[2] Haugan M P 1973 Energy conservation and the principle of equivalence Ann. Phys. 118 156[3] Ehlers J, Pirani F A E and Schild A 1972 The geometry of free fall and light propagation General Relativity

(Oxford: Clarendon) (papers in honour of J L Synge)[4] Lammerzahl C 1996 Gen. Rel. Grav. 28 1043[5] Lammerzahl C 1998 Acta Phys. Pol. 29 1057[6] Lammerzahl C 1998 Class. Quantum Grav. 14 13[7] Damour T 2012 Theoretical aspects of the equivalence principle arXiv:1202.6311 [gr-qc][8] Goklu E and Lammerzahl C 2008 Class. Quantum Grav. 25 105012[9] Breuer H-P, Goklu E and Lammerzahl C 2009 Class. Quantum Grav. 26 105012

[10] Goklu E, Lammerzahl C, Camacho A and Macias A 2009 Class. Quantum Grav. 26 225010[11] Peters A, Chung K and Chu S 1999 Measurement of gravitational acceleration by dropping atoms Nature 400 849[12] McGuirk J M et al 2002 Sensitive absolute-gravity gradiometry using atom interferometry Phys. Rev.

A 65 033608[13] Fixler J B et al 2007 Atom interferometer measurement of the Newtonian constant of gravity Science 315 74[14] Lamporesi G et al 2008 Determination of the Newtonian gravitational constant using atom interferometry Phys.

Rev. Lett. 100 050801[15] Cronin A D et al 2009 Optics and interferometry with atoms and molecules Rev. Mod. Phys. 81 1051–129[16] Kasevich M and Chu S 1991 Atomic interferometry using stimulated Raman transitions Phys. Rev. Lett. 67 181–4[17] Moler K et al 1992 Theoretical analysis of velocity selective Raman transitions Phys. Rev. A 45 342–8

11

Page 13: Testing the equivalence principle with atomic interferometry

Class. Quantum Grav. 29 (2012) 184003 S Herrmann et al

[18] Muller H et al 2008 Atom-wave diffraction between the Raman–Nath and the Bragg regime: effective Rabifrequency, losses, and phase shifts Phys Rev. A 77 023609

[19] Storey P and Cohen-Tanoudji C 1994 The Feynman path integral approach to atomic interferometry. A tutorialJ. Phys. II 4 1999–2027

[20] Fray S et al 2004 Atomic interferometer with amplitude gratings of light and its applications to atom based testsof the equivalence principle Phys. Rev. Lett. 93 240404

[21] Chiow S-W et al 2011 102 �k large area atom interferometers Phys. Rev. Lett. 107 130403[22] Lan S-Y et al 2012 Influence of the Coriolis force in atom interferometry Phys. Rev. Lett. 108 090402[23] Dimopoulos S et al 2007 Testing general relativity with atom interferometry Phys. Rev. Lett. 98 111102[24] Stern G et al 2009 Light-pulse atom interferometry in microgravity Eur. Phys. J. D 53 353–7[25] Varoquaux G et al 2009 How to estimate the differential acceleration in a two-species atom interferometer to

test the equivalence principle New J. Phys. 11 113010[26] van Zoest T et al 2010 Bose–Einstein condensation in microgravity Science 328 1540–3[27] Muentinga H et al Atom interferometry in microgravity in preparation[28] Modugno S et al 1999 Sub-Doppler laser cooling of fermionic 40K atoms Phys. Rev. A 67 R3373–6[29] Modugno S et al 2002 Bose–Einstein condensation of Potassium atoms by sympathetic cooling Science 297 2240[30] Goldwin J et al 2002 Two-species magneto-optical trap with 40K and 87Rb Phys. Rev. A 65 021402[31] Roati G et al 2007 39K Bose–Einstein condensate with tunable interactions Phys. Rev. Lett. 99 010403[32] Ospelkaus S et al 2007 Degenerate KRb Fermi–Bose gas mixtures with large particle numbers degenerate KRb

Fermi–Bose gas mixtures with large particle numbers J. Mod. Opt. 54 661–73[33] Ospelkaus C et al 2006 Interaction-driven dynamics of 40K-87Rb fermion–boson gas mixtures in the large-

particle-number limit Phys. Rev. Lett. 96 020401[34] Stockton J K et al 2007 Bayesian estimation of differential interferometer phase Phys. Rev. A 76 033613[35] Kuroda K and Mio N 1989 Test of a composition-dependent force by a free-fall interferometer Phys. Rev.

Lett. 62 1941–4[36] Sorrentino F et al 2010 A compact atom interferometer for future space missions Microgravity Sci.

Technol. 22 551–61[37] http://www.exphy.uni-duesseldorf.de/Publikationen/2010/STE-QUEST-final.pdf[38] Rudolph J et al 2011 Degenerate quantum gases in microgravity Microgravity Sci. Technol. 23 287–92

12