<<

Perturbative operator approach to high-precision light-pulse atom interferometry (Published as Physical Review A 101, 053615 [2020])

Christian Ufrecht∗ and Enno Giese Institut für Quantenphysik and Center for Integrated Quantum Science and Technology (IQST), Universität Ulm, Albert-Einstein-Allee 11, D-89069 Ulm, Germany

Light-pulse atom interferometers are powerful quantum sensors, however, their accuracy for example in tests of the weak equivalence principle is limited by various spurious influences like stray magnetic fields or blackbody radiation. Improving the accuracy therefore requires a detailed assessment of the size of such deleterious effects. Here, we present a systematic operator expansion to obtain phase shift and contrast analytically in powers of a perturbation potential. The result can either be employed for robust straightforward order-of-magnitude estimates or for rigorous calculations. Together with general conditions for the validity of the approach, we provide a particularly useful formula for the phase including wave-packet effects.

I. INTRODUCTION article we extend the work of Ref. [28] with particular em- phasis on general conditions for the validity of the approach, Since their first implementation [1] in 1991 the accuracy of characterizing the magnitude of the perturbations. light-pulse atom interferometers has been improved consider- In Sec.II we outline our main results and put them into ably, which led to high-precision applications in gravimetry context. Subsequently, in Sec. III we introduce our path- [2, 3], gradiometry [4–6], tests of fundamental such as dependent model and derive the perturbative expansion. The the weak equivalence principle [7–10], measurements of the conditions under which our method is valid will be discussed fine-structure constant [11, 12] and proposals for gravitational- in Sec.IV. Finally in Sec.V, as a simple example, we apply wave detection [13]. However, as the accuracy is pushed further, the formalism to the cubic potential appearing in the Taylor an increasing number of formerly negligible influences such expansion of the gravitational potential of Earth. as magnetic field gradients, blackbody radiation inducing a spatially-dependent a.c. Stark shift [14, 15], or gravitational fields of the laboratory environment have to be included into the II. BRANCH-DEPENDENT DESCRIPTION error budget. In this article we present a systematic approach to account for such spurious effects. Light-pulse atom interferometers consist of a sequence of Although phase shifts caused by the corresponding in gener- light pulses which coherently split the initial wave packet and alanharmonic potential shifts can be small, they might neverthe- subsequently direct the atoms along the two branches of the less be non-negligible which calls for a systematic perturbative interferometer. After recombination by a final laser pulse, the approach. Over the years, several powerful analytic methods number of atoms at each exit port displays an interference for the calculation of phase and contrast of light-pulse atom pattern from which the relative phase accumulated between interferometers have been developed based on the Feynman the branches of the interferometer can be inferred. In this path integral [16, 17], descriptions in phase space [18, 19] work we assume that the Hamiltonian describing the motion as well as in the form of representation-free descriptions on through the interferometer can be decomposed into a dominant the operator level for linear gravity [20], path-independent part (linear gravity, laser pulses) and a weak perturbation quadratic Hamiltonians [21, 22] or within a local-harmonic (e.g. gravity gradients, blackbody radiation, etc.). As illustrated [23, 24] approximation. Within the latter approach, it was by the Mach-Zehnder (MZ) gravimeter shown in Fig.1, the also possible to obtain wave-packet effects to lowest order two branches of the interferometer are mainly caused by the [25]. However, these methods are either applicable to at most dominant part of the Hamiltonian (thick solid lines in the quadratic potentials or lack a comprehensive discussion of figure). These trajectories are only slightly disturbed by the consistency in the case of more general applications. perturbation potentials (leading to the thin dashed lines in the In this article we derive a systematic perturbative description figure), which can in general be different for each branch and for phase and contrast including effects due to wave-packet time dependent. The hypothetical interferometer sequence with dynamics based on two formal , the Magnus [26–29] vanishing perturbation will be referred to in the following as and the cumulant [30] expansion. They have already been unperturbed interferometer. arXiv:2003.02042v2 [quant-ph] 10 May 2020 applied in the context of light-pulse atom interferometry [28] Neglecting wave-packet effects, it has been shown [16, 23, to determine the quality of magnetic shielding [31] and to cal- 25, 28] that the phase φ of such an interferometer can be culate relativistic effects for interferometric redshift tests [32]. obtained from the classical trajectories through Recently, the Magnus expansion was furthermore used to take into account the effect of finite pulse duration [33]. In this φ = ∆S/~ + φs , (1) where ∆S = S(u) − S(l) is the classical action difference be- tween the upper and lower branch (superscripts u and l) in ∗ [email protected] combination with a separation phase φs in case the classical 2

calculating their value to desired accuracy. Denoting the perturbation potential on the upper and lower branch by V(u) and V(l), respectively, we will find for a closed unperturbed interferometer sequence 1 ¼ φ = φ0 − dt V(t) ~ 1 ¼ − dt Vij (t) hrˆi(t)rˆ j (t)i (2) 2~ for the phase including the leading-order phase shifts from the perturbation, where repeated indices are summed over. In Eq. (2) we defined φ0 as the phase of the unperturbed interfer- Figure 1. Mach-Zehnder gravimeter. Initially released from an atomic ometer which can be calculated for example with the general formula provided in Ref. [34]. The perturbation potential trap, the free fall of the atoms is interrupted by a π/2 pulse at t = ti = 0 ( ) (α)( (α)( )) to split the wave packet into two components transferring a momentum V t = V r0 t might be different on the upper (α = u) of ~k to one of them, which gains the recoil velocity vr = ~k/m. After and lower branch (α = l) and is evaluated at the unperturbed (α) a time T both components are redirected by a π pulse at t1 = T and ( ) trajectories r0 t . The integrals run from the initial time ti, finally recombined by a second π/2 pulse at t2 = 2T. At the detection where the atoms are released from the trap, up to the detection t time d the number of particles is measured at one of the exit ports, time t on the upper branch and return along the lower branch which forms an interference pattern dependent on the relative phase d t accumulated between the two branches. The unperturbed branches back to i. We stress that the perturbation potential and its (thick solid lines) are determined by the analytic expressions shown second derivative Vij = ∂i ∂jV are evaluated at the two unper- next to them, where g is the local gravitation acceleration. The turbed trajectories, which obviates the solution of a possibly perturbation V(α) slightly disturbs the atoms, leading to the deviating involved differential equation. The initial conditions for the branches of the actual interferometer (thin dashed lines). trajectories are r0(ti) = hrˆi and rÛ0(ti) = hpˆi/m, where the ex- pectation value is taken with respect to the initial wave packet. In Eq. (2) we furthermore defined the operator describing the free evolution of the wave packet trajectories do not coincide upon detection. In this case we refer to the interferometer as open. The calculation of the pˆ − hpˆi rˆ(t) = rˆ − hrˆi + t . (3) phase via Eq. (1) therefore consists of (i) solving the differential m equation for the classical trajectories including all perturbations and (ii) evaluating the action difference. In the case of weak Consequently, hrˆi(t)rˆ j (t)i, where the indices label the com- perturbing potentials, however, this approach is unfavorable for ponent of the vectors, provides a measure for the width of the following reasons: In general, analytic expressions for the the wave packet for vanishing perturbation as detailed further trajectories including the perturbation do not exist. Therefore, below. Solving the classical trajectories and the integral in the trajectories can be solved iteratively to desired order in Eq. (1) to first order in the perturbation, the first line in Eq. (2) the perturbation [23] and are then substituted into Eq. (1), can also be derived [35] directly from Eq. (1). The second possibly resulting in cumbersome expressions, whereas a di- line, however, describes wave-packet effects which are not rect perturbative expansion in powers of the weak perturbing taken into account by Eq. (1). In neutron interferometry small potential would be much more convenient. Furthermore, if the perturbations can also be taken into account to first order in a perturbation is only available numerically, in an integration of WKB-like treatment [36]. Eq. (1) one has to account for both the dominant contribution as well as the perturbation, which can be difficult numerically since they likely differ in size by multiple orders of magni- III. PERTURBATIVE TREATMENT tude. Finally, the validity of Eq. (1) is premised on negligible wave-packet effects. However, in the presence of anharmonic We now derive Eq. (2) from a full quantum-mechanical perturbation potentials the two components of the wave packet description. In order to calculate the phase measured by a will experience different local expansion dynamics along the light-pulse atom interferometer, one has to start from the multi- branches, leading to a slight mismatch and therefore resulting in level Hamiltonian including the virtual states necessary for the additional phase contributions upon detection. Consequently, it diffraction process. However, after adiabatic elimination of the is a priori not obvious if these phases are negligible compared auxiliary states [37, 38], neglecting atom-atom interactions, to those of Eq. (1) originating from the perturbation. The and assuming infinitely short laser pulses, the evolution can be approach presented in this article is based on a systematic reduced to a branch-dependent description [20, 28] in which the operator expansion derived from a full quantum-mechanical phase φ and contrast C of the interferometer after projection on description of the interferometer to overcome these problems. one exit port is defined by the expectation value of the overlap It allows formulating conditions for its validity determining operator exactly when wave-packet effects are negligible or, in turn, hUˆ (l)†Uˆ (u)i = Ceiφ (4) 3 with respect to the initial wave function, where Uˆ (α) generates the time evolution along branch α. In this article we assume that the Hamiltonian for each branch allows the decomposition

ˆ (α) ˆ (α) (α)( ) H = H0 + V rˆ, t . (5)

The perturbation potential V(α)(rˆ, t) only slightly disturbs the dominant part of the evolution which is caused by the unper- turbed Hamiltonian Figure 2. Time contour. To merge the two time-evolution operators corresponding to the two branches of the interferometer, we introduce 2 (α) pˆ (α) the time contour depicted in the figure over which all integrals extend. Hˆ = + mgzˆ + V (rˆ, t) (6) t t 0 2m em Starting at i the integrals run to the detection time d over the pertur- bation potential corresponding to the upper branch and subsequently describing the motion of an atom with mass m in the linear return along the lower branch back to the initial time ti. gravitational field, where g is the local gravitational acceleration. The interaction with the laser pulses is modeled by the potentials

( ) Õ ( ) ( ) V α (rˆ, t) = − [k α rˆ + ϕ α ]δ(t − t ) Recalling Eq. (7), we identified the phase of the closed unper- em ~ ` ` ` ( ) ` turbed interferometer and moved the exponential exp iφ0 to the left as it is only a c-number. Writing the two interaction (α) picture time-evolution operators explicitly which transfer the momentum ~k` on branch α at time t = t` (α)  td   td  and imprint the laser phase ϕ` evaluated at the time of the i ¹ i ¹ ˆ (l)† ˆ (u) T ˆ (l) T − ˆ (u) pulse on the wave packet. In this article we assume that UI UI = exp dt VI exp dt VI , ~ t ~ t the unperturbed interferometer is closed, translating into the i i condition we note that the time-ordering operator T orders times (reading (l)† (u) from right to the left) from the initial time ti to the final time Uˆ Uˆ = eiφ0 , (7) 0 0 td while the anti-time-ordering operator T orders from td back to ti. Thus, we merge the two time-evolution operators to Uˆ (α) where 0 is the time-evolution operator with respect to Eq. (6). one path-ordered exponential by introducing the path-ordering The phase of the unperturbed interferometer φ , is merely a 0 operator Tp so that we obtain c-number, implying perfect wave-packet overlap at the end of the unperturbed interferometer sequence. However, the  ¼  (l)† (u) iφ0 i interferometer including the perturbation is in general not Uˆ Uˆ = e Tp exp − dt VˆI(t) . (9) ~ closed. The time-evolution operator with respect to Hamiltonian (5) Resorting to the concept of path ordering, initially introduced can be decomposed into by Schwinger and Keldysh [39, 40] in the context of thermal field theory, Tp orders time along the contour illustrated in ( ) (α) (α) Uˆ α Uˆ Uˆ (u) = 0 I (8) ˆ ( ) ˆ ( ) Fig.2. On this contour we define VI t = VI t for t on the (l) (α) upper path and Vˆ (t) = Vˆ (t) for t on the lower path. With by transforming into the interaction picture with respect to Hˆ I I 0 these definitions in mind, we disregard the explicit labeling (α) so that the operator of the branch if not necessary. t ( )  i ¹ d  In order to associate the operator rˆ t with the unperturbed ˆ (α) T − ˆ (α)( ) UI = exp dt VI t trajectory of the atoms, the initial mean momentum and position ~ t i can be included by introducing r0(t) = r˜0(t) + hrˆi + hpˆit/m so that ˆ (α)( ) (α)( (α)( ) ) only includes the potential VI t = V rˆ t , t which is a (α)( ) ˆ (α)† ˆ (α) r(t) = r (t) + rˆ(t) function of rˆ t = U0 rˆU0 , the solution of the Heisenberg ˆ 0 ( ) equations of motion generated by Hˆ α . An explicit expression 0 ˆ( ) for the solution is straightforwardly obtained for our form of for each branch, where r t was defined in Eq. (3). Conse- (α) (α) (α) quently, with Eq. (9) the overlap operator can be written as a Hˆ , resulting in rˆ(α)(t) = rˆ + pˆt/m + r˜ (t). Here, r˜ (t) are 0 0 0 path-ordered exponential in which the perturbation potential the classical trajectories caused by the unperturbed Hamiltonian is evaluated at the classical unperturbed trajectory plus an r(α)(t ) = rÛ (α)(t ) = (6) with the initial conditions ˜0 i 0 and ˜0 i 0 since operator part with vanishing expectation value. The expecta- the Schrödinger and Heisenberg picture coincide at t = ti. 2 2 2 2 2 tion value hrˆ j (t) i = ∆r + ∆p t /m , where ∆rj , ∆pj are the Inserting the decomposition shown in Eq. (8) into the overlap j j initial position and momentum widths of the wave packet in operator in Eq. (4) then yields the jth direction, is therefore a measure for the width of the (l)† (l)† (u) (u) (l)† (u) expanding wave packet as long as any distortion effects from the Uˆ (l)†Uˆ (u) = Uˆ Uˆ Uˆ Uˆ = eiφ0 Uˆ Uˆ . I 0 0 I I I perturbation potential are negligible. Note that we disregarded 4 for the moment possible initial correlations between rˆ and pˆ. Thus, if the change of the potential over the size of the wave packet on each interferometer branch is sufficiently small, the Taylor expansion around the classical trajectory r0(t) 1 Vˆ (t) = V + Virˆi + Vijrˆirˆ j + ..., (10) I 2 where indices of the potential again denote derivatives, accu- rately approximates the potential by taking into account only a few terms. Note that we omitted the time dependence on Figure 3. Comparison between different scales. (a) Perturbation the right-hand side for the sake of readability and again use potential evaluated at the two unperturbed trajectories spanning the summation convention. interferometer as a function of time (blue and orange curve). Whereas ∆V The overlap operator from Eq. (9) still contains the formal corresponds to the maximal potential difference probed by the atoms, δV is the maximal potential difference between the branches path-ordering operator T , prohibiting any further manipulation p of the interferometer, which is generally smaller. (b) The space-time of the contour-ordered exponential. To remove the former, diagram of a given interferometer sequence determines the domain we apply the Magnus expansion, detailed in AppendixA, of the potential probed by the atoms (colored in bright red). The facilitating an exponential representation of a time-ordered suppression factor of consecutive terms in the Magnus expansion as exponential in terms of a formal series, namely well as wave-packet effects are crucially dependent on the value of ξ which is the length scale on which the potential changes. If we assign ( ∞ ) a wave number to the oscillatory behavior of the potential shown in (l)† (u) iφ +iφˆ Õ Uˆ Uˆ = e 0 = exp i φˆn , (11) the figure, ξ is given by its inverse (colored in darker red). In contrast, n=0 in the case of a simple polynomial form of the potential, the value of ξ would scale with the extent of the interferometer itself. A third length ˆ Í∞ ˆ where φ = n=1 φn. For a more convenient notation, the scale is determined by the characteristic size of the wave packet d. operator φˆ0 is in fact defined as the phase of the unperturbed interferometer φ0, i.e. a c-number. The other contributions φˆn for n > 0 are determined by the Magnus expansion. In the appendix we provide the terms explicitly to third order; and in For the moment let us assume that these conditions are general φˆn contains n nested integrals along the time contour satisfied, allowing a truncation of the Magnus and cumulant over nested of order n − 1 between the potential expansion at first order. Consequently, with the help of Eq. (11) evaluated at different times. and Eq. (B2) the phase is φ = φ0 + hφˆ1i. Inserting the Taylor The Magnus expansion therefore leads to an operator ex- polynomial of the potential up to quadratic order into Eq. (A1) pansion in powers of the perturbation potential V. If small as and recalling that hrˆi = 0, we arrive at Eq. (2), the main result defined in the following section, the Magnus expansion can be of our article. truncated at desired order and we have already succeeded in finding an approximate exponential representation of the over- lap operator. However, according to Eq. (4), it is still necessary IV. CONDITIONS FOR VALIDITY to evaluate the expectation value of this operator. In case of a path-independent harmonic potential it can be shown [22] We now derive conditions for the validity of the perturbative that the exponent of the overlap operator depends only linearly treatment presented in the section above. First, we introduce a on rˆ and pˆ, making the calculation of the expectation value characteristic length scale ξ with straightforward. However, in general such a simple representa- tion does not exist and the overlap operator will contain various ∆V ∼ ξn∆V(n) , (12) powers of rˆ and pˆ. In this case, a suitable approach lies in the cumulant expansion, a formal series, casting the expectation where ∆V is the difference between the extremal values of value of an exponential operator into the form the potential probed by the atoms over the course of the inter- ferometer. Furthermore, ∆V(n) is the typical value of the nth ( ∞ ) derivative of the potential. For example in case of a power-law φˆ Õ κn hei i = exp , ξ n! dependence of the potential, corresponds to the size of the n=1 atomic fountain in which the experiment is performed while where the cumulants κn, defined in AppendixB, are functions for oscillating potentials the value of ξ can be much smaller. of the first n moments of φˆ. Consequently, the phase of an The parameters ∆V and ξ are visualized in Fig.3 together with interferometer can at least formally be expressed in powers of further quantities defined below. First, a perturbative treatment the perturbation potential. is only valid, if the deviation of the unperturbed trajectories Truncating the series at some desired order, however, requires caused by the perturbation potential is small compared to the a detailed assessment of conditions characterizing the magni- characteristic length scale ξ on which the potential varies. tude of the potential and the size of wave-packet effects. These Therefore, identifying an acceleration a ∼ −∆V/(mξ) of the conditions for the validity of our approach will be discussed in atom due to the perturbation potential, where ∆V/ξ ∼ ∆V(1), the subsequent section. 5 we require that the distance aT 2 is much smaller than the ξ  d/ξ ηd/ξ characteristic length ξ, that is Table I. Explicit values for , , , and for different sources of the perturbation. In the table we show that all these examples satisfy ∆V ∆VT 2 the conditions for the validity of our approach. The parameters ,  =  1, (13) δV and ξ used for the calculation of  and η are estimated from the ξ2m potentials stated in the references. The width of the wave packet is assumed to be d = 50 µm. where T denotes the characteristic interferometer time which can be chosen to be equal to the interrogation time of the gravity magnetic field blackbody mass defect interferometer. As shown more rigorously in AppendixD, the gradients [41] gradients [31] radiation [15] in quantum parameter  constitutes the factor by which subsequent orders clocks [32] of the Magnus expansion are suppressed. ξ [m] 0.1 0.1 0.01 10 We now consider the leading-order correction to the unper- turbed phase, that is the second term on the right-hand side of  10−8 10−11 10−5 10−12 Eq. (2). This term can be estimated to be of the size d/ξ 5 · 10−4 5 · 10−4 5 · 10−3 5 · 10−6 δVT η = , (14) ηd/ξ 10−4 10−9 10−3 10−9 ~ where δV is the maximal potential difference between the branches at one instance of time, which can be much smaller V. EXAMPLE: GRAVITATIONAL POTENTIAL than ∆V, the difference between the two extremal values of the potential probed by the atoms during the course of the interferometer. In this section we illustrate the formalism derived in the In Eq. (10) we Taylor expanded the potential about the unper- previous sections by the example of an MZ-interferometry turbed trajectory over the size of the wave packet represented by experiment conducted in the Newtonian gravitational potential of Earth the operator rˆ. Heuristically replacing the position operators by the wave-packet width d, the Taylor expansion can be truncated mM ( ) V(r) = −G , after a few terms if ∆V n dn  ∆V which translates with the r help of Eq. (12) into where G is Newton’s constant and M is the mass of Earth. d/ξ  1. (15) Choosing the z axis of a new coordinate system in direction of Earth’s radius R, we Taylor expand the potential around a point − Finally, if this condition holds, the leading-order operator- on Earth’s surface in powers of R 1. This calculation yields valued term in φˆ is due to the first-order term in the Taylor ∮ 1 ( ) 1 ( ) tV rˆ / ( ) Γ 1 Γ 2 O( −3) expansion of the potential, taking the form d i i ~. To V r = mgz + m ij rirj + m ijl rirjrl + R guarantee the validity of the cumulant expansion, which is a 2 6 function of the moments of φˆ, this term should be much smaller after omitting the irrelevant constant. The only non-vanishing than unity. This requirement can be expressed by components of the fully symmetric first and second gravity- gradient tensors Γ(1) and Γ(2) then are given by ηd/ξ  1 (16) (1) (1) g (1) g Γxx = Γyy = , Γzz = −2 , after again replacing the operator rˆ by d, the integration in time R R by the characteristic interferometer time T, and by making use and (including all possible permutations) of Eq. (12). A more rigorous derivation of these conditions can be found in AppendixD. ( ) ( ) g ( ) g Γ 2 = Γ 2 = −3 , Γ 2 = 6 , In tableI we give approximate values for ξ, , and ηd/ξ for xxz yyz R2 zzz R2 some experiments or recent proposals. As shown in the table, the accuracy achieved by truncation of the Magnus expansion where we identified g = GM/R2. Phase corrections caused by at first order often is already sufficient and our formalism is the first gradients (described by Γ(1)) have been calculated to all well suited to be applied to these situations. orders in Γ(1) [22, 42]. However, note that phase contributions Obviously, a scaling such as in Eq. (12) cannot be guaranteed to second order in Γ(1) and to first order in Γ(2) both scale in general. For this case we apply the Magnus expansion to with R−2 and may therefore be of the same size, making the the Taylor polynomial of the perturbation potential explicitly calculation inconsistent when disregarding Γ(2). in AppendixC from which the scaling of phase shifts can In this example we focus on the second gravity gradients be inferred. However, for many applications a parameter ξ (described by Γ(2)). Phase shifts from this contribution have does exist, satisfying Eq. (12) at least approximately, which is been calculated before [23] but here we will additionally sufficient for an order-of-magnitude estimation of the size of include the contribution of wave-packet effects. Furthermore, phase shifts beyond Eq. (2). we stress how straightforward the calculation becomes with our formalism and put particular emphasis on the application 6 of the conditions for the validity of a perturbative description.

Check of conditions for validity First we check condition (13) for the validity of the Magnus expansion. Because of the polynomial form of the potential, ξ is given by the total extent of the interferometer. Assuming the recoil velocity vr = ~k/m, where ~k is the effective momentum transfer of the lasers as well as the initial velocity of the atoms vi to be much smaller than gT, the size of the interferometer scales with ξ = gT 2/2 and we choose ∆V = mΓ(2)ξ3, where T is half of the interferometer time (see Fig.1). Consequently, with the help of Eq. (13)

 = Γ(2)gT 4 ≈ 10−12 for T ≈ 1 s. Consecutive terms in the Magnus expansion corresponding to the same power of the operators rˆ and pˆ are therefore suppressed by this factor and a consideration to first Figure 4. Plot of phase shifts caused by the gravitational field of order is sufficient. Next we calculate the potential difference Earth. As shown in the figure, the dominant contribution kgT2 to between the interferometer branches by inserting the analytic the phase (blue line) is due to the unperturbed interferometer. The expressions for the unperturbed trajectories from Fig.1 into phase shift caused by the first gravity gradients [41] scales with R−1 the potential so that (purple line). Phase contributions from the second gravity gradients scale with R−2 and can be divided into the first term inside the bracket (2) 2 3 2 3 (2) 2 5 in Eq. (19) (orange line) and the second term which originates from δV ∼ mΓ [(vrT − gT /2) − (−gT /2) ] ∼ mΓ g vrT , wave-packet effects (yellow line). For this figure the parameters are: ' · 6 where we replaced t ∼ T. Consequently, the leading-order The radius of Earth is R 6 10 m, the effective wave vector is given by k = 4π/(780 nm), an interferometer time of T = 1 s is assumed, the phase shift from the second gravity gradients is of the order of initial trapping frequency is ω = 2π · 60 Hz, and a mass corresponding to rubidium 87 is chosen. mg2v T 6 η = Γ(2) r = Γ(2)kg2T 6 ≈ 10−4 , ~ where, as the momentum transfer stems from a two-photon where we recalled Eq. (3) and defined the central expectation process, the value k = 4π/(780 nm) of the effective wave value h·ic with respect to the initial state but displaced by hrˆi number corresponds to twice the wave length of the D2 line of and hpˆi to the origin of phase space so that e.g. hrˆirˆj ic = h(rˆi − rubidium 87. hrˆii)(rˆj −hrˆj i)i. In Eqs. (17) and (18) we made furthermore use Finally, it is left to examine condition (16) for the validity of the symmetry of Γ(2). The chosen interferometer sequence of the cumulant expansion. Note that in principle one should only enters the expression through the functions f , f , f and consider operator-valued contributions in the overlap operator φ rr r p f defined in AppendixE and also explicitly evaluated for the from the first and second gravity gradients but for the sake pp MZ geometry depicted in Fig.1. In order to keep the expressions of a simple presentation we restrict the discussion to latter. compact, we assume the initial position r (t ) = z δ without Assuming a maximal size d = 200 µm of the wave packet, we 0j i i jz an initial velocity. If necessary, however, a more general find calculation is straightforward. As shown in the appendix, the ηd/ξ ≈ 10−9 phase then depends on the initial position zi of the state. This result is a direct consequence of the operator-valued form of so that Eq. (2) can be confidently applied. the overlap operator. In contrast, if the interferometer is closed, the overlap operator reduces to a c-number and its expectation Calculation of phase value is independent of the wave function and therefore of the We now consider a general interferometer sequence, but for initial conditions. simplicity we assume the laser pulses to be aligned with the To obtain compact explicit expressions for the phase shift due direction of linear gravity so that the classical unperturbed to wave-packet effects, we assume that the first laser pulse acts trajectories take the form r0j (t) = z0(t)δz j . Insertion of the right after releasing the atoms from the trap. Thus, assuming the perturbing potential into Eq. (2) then yields for the phase shift initial state as the ground state of a harmonic trap with frequency due to Γ(2) ωi in the ith direction results in hrˆirˆj ic = δij ~/(2mωi) and hpˆ pˆ i = δ mω /2 as well as vanishing correlations between ¼ i j c ij ~ i 1 m (2) momentum and position operators. Specifying further ω ≡ dt V(t) = Γzzz fφ (17) ~ 6~ ωx = ωy = 2ωz, the final result for the lowest-order correction and ¼ 1 1 (2) h dt Vij (t) hrˆi(t)rˆ j (t)i = Γ mhrˆirˆj ic frr 2~ 2~ zij 1 i + hrˆ pˆ + pˆ rˆ i f + hpˆ pˆ i f , (18) i j j i c r p m i j c pp 7 due to the cubic potential is The application of the Magnus expansion to the overlap operator has resulted in nested contour integrals whoose an-  2   g m vrT 3 7 2 alytical evaluation becomes cumbersome for large orders of φ = φ0 − fφ + − (ωT) . (19) R2 ~ ω 2 8 the expansion. Even though, the integrals can be reordered [28] to streamline analytical calculations, the loop structure Interestingly, if the atoms expanded out of a symmetric har- of the integral is particularly useful in a numerical implemen- monic trap, the wave-packet induced contributions would van- tation. Here, the perturbation potential (and its derivatives) ish. are discretized exactly on the time contour so that a numerical In Fig.4 we illustrate Eq. (19) for reasonable experimental integration algorithm will automatically account for the loop parameter values. In the figure we compare the phase of the properties of the integrals. unperturbed interferometer sequence (blue line) and the phase Our approach is applicable to a variety of situations with ( ) due to Γ 1 (purple line) [41] to the phase shift calculated in different sources of the perturbation, including e.g. blackbody Eq. (19) from the cubic contribution of the Taylor expansion radiation, gravity gradients, inhomogeneities in the gravita- ( ) (orange line). The effect of phase shifts caused by Γ 1 are tional potential of the laboratory environment, magnetic field generally relevant in state-of-the-art precision measurements gradients, relativistic effects, violation parameters of the uni- [41]. As a consequence, such phase contributions either have versality of free fall, and finite laser pulse lengths. to be included into the analysis, or have to be compensated In this work we proposed a new perturbative tool to assess through differential schemes [8] (for example used for test of small phases due to spurious influences for light-pulse atom the weak equivalence principle) and mitigation techniques [43]. interferometers. Making use of a path-dependent description ( ) Phase shifts originating from second gradients Γ 2 , however, formalized by the introduction of the path-ordering operator, are of the order of magnitude to possibly limit future spaceborne we emphasized how the method solves the problems of previous missions if not appropriately accounted for. Finally, the phase results. Based on two formal series, the Magnus and cumulant shift originating from different expansion dynamics along expansion, we derived Eq. (2) for the leading-order phase the branches (yellow line) is beyond any accessible value for shifts originating from the perturbation including wave-packet light-pulse atom interferometric experiments in the mid future. effects and obtained detailed conditions for the validity of our Such phases, however, can be much larger when taking into approach which are stated in Eq. (13), Eq. (15) and Eq. (16). account the inhomogeneous gravitational field of the laboratory Finally, we commented on straightforward generalizations and environment. the numerical implementation in case analytic calculations are not possible.

VI. DISCUSSION AND CONCLUSION VII. ACKNOWLEDGEMENT Obviously, Hamiltonian (5) does not account for atom-atom interactions. However, state-of-the-art atom interferometers We thank F. Di Pumpo, A. Friedrich, A. Roura, and employ Bose-Einstein condensates as highly-coherent atom W. P. Schleich for helpful discussions. This work is supported sources which are intrinsically interacting many-body systems. by the German Aerospace Center (Deutsches Zentrum für Luft- Nonetheless, numerical propagation of the initial mean-field und Raumfahrt, DLR) with funds provided by the Federal Min- state with the help of the Gross-Pitaevskii equation, including istry for Economic Affairs and Energy (Bundesministerium release from the trap and possibly magnetic lensing [44], shows für Wirtschaft und Energie, BMWi) due to an enactment of that due to the dynamical expansion, the strength of interactions the German Bundestag under Grant Nos. DLR 50WM1556 quickly decreases. Thus, if the initial expansion time before and 50WM1956. We thank the Ministry of Science, Research the first laser pulse is sufficiently large, any further evolution and Art Baden-Württemberg (Ministerium für Wissenschaft, will be accurately described by the Schrödinger equation and Forschung und Kunst Baden-Württemberg) for financially sup- our formalism is valid from this instance of time. Then, the porting the work of IQST. state right before the first laser pulse is used as input [24] and the expectation value of the overlap operator is calculated with respect to this state. Even more if interactions are negligible Appendix A: Magnus expansion during the whole experiment, we include the time between release and first laser pulse into the unperturbed trajectories so The Magnus expansion [26–28] is a formal series for the that the influence of the perturbation on the wave packet during exponential representation of a time-ordered exponential this initial expansion time is automatically accounted for. It ( ∞ ) then suffices to calculate expectations values with respect to  i ¹ td  Õ the ground state of the trap. Uˆ = T exp − dt Hˆ (t) = exp i φˆn , In our article we have restricted the discussion to pertur- ~ ti n=1 bations which only depend on the position operator rˆ.A where Hˆ (t) is a time-dependent Hamiltonian. Applied to the generalization, however, of our results for the application to path-ordered exponential in Eq. (9), we obtain for the first three pˆ-dependent perturbations, present for example in rotating frames, is straightforward. 8 elements of the series in Eq. (11) Appendix C: Phase to second order 1 ¼ φˆ1 = − dt VˆI(t) (A1) In this appendix we apply the Magnus expansion to second ~ order to the overlap operator. To this end, we insert the Taylor ¼ ¼ t i 0 0 series of the potential into Eq. (11), where φˆn is determined by φˆ2 = dt dt [VˆI(t), VˆI(t )] (A2) 2~2 the Magnus expansion. Making use of the ¼ ¼ t ¼ t0 1 0 00  0 00 φˆ = dt dt dt [Vˆ (t), [Vˆ (t ), Vˆ (t )]] 0 i~ 0 3 3 I I I [rˆ (t), rˆ (t )] = (t − t)δ , (C1) 6~ i j m ij 00 0  + [VˆI(t ), [VˆI(t ), VˆI(t)]] . where δij is the Kronecker symbol, we obtain with the help of Eqs. (A1) and (A2) In general, φˆn consists of n nested integrals over (n − 1)th- order commutators between the potential evaluated at different φˆ0 =φ0 times. The perturbation potential is a function of the solution 1 ¼  1 1 of the Heisenberg equations of motion rˆ(t) generated by the φˆ1 = − dt V + Virˆi + Vijrˆirˆ j + Vijkrˆirˆ jrˆk unperturbed Hamiltonian. ~ 2 6 1  + Vijklrˆirˆ jrˆkrˆl + ... 24 Appendix B: Cumulant expansion ¼ ¼ t  1 0 0 0 0 ˆ0 0 ˆ φˆ2 = − dt dt (t − t) ViVi + ViVij r j + Vi Vij r j 2~m The cumulant expansion [30, 45] is a formal series for 1 0 1 0  0 0  an exponential representation of the expectation value of an + V Vijk rˆ j rˆk + VikV rˆ rˆi + rˆi rˆ 2 i 2 k j j j exponential operator which is defined by  1 0 ˆ0 ˆ0 ( ∞ ) + ViVijk r j rk + ... . Õ κn 2 heiφζˆ i = exp ζ n , (B1) n! 0 n=1 Note that all quantities depend on time t or t . If dependent on the latter, this dependence is abbreviated by a prime on the ζ where one introduces a formal expansion parameter , which respective quantity. is set to unity after the calculation, and the coefficients κn are referred to as cumulants. By taking the logarithm on both sides, we find the definition of the cumulants as Appendix D: Derivation of conditions for validity n d iφζˆ κn = lnhe i , dζ n ζ=0 In order to complement the validity discussion of Sec.IV, we proceed in two steps. First, we derive the factor  by which where the nth cumulant is function of the first n moments of φˆ. subsequent terms in the Magnus expansion are suppressed. Here, we state explicitly the first three cumulants Second, we investigate the scaling of different orders in the cumulant expansion. κ1 = ihφˆi For the sake of simplicity, in the following we choose one 2 2 typical direction x and suppress the time dependence of xˆ(t) κ2 = −hφˆ i + hφˆi except when appearing in the commutator. Within this simpli- ˆ3 ˆ2 ˆ ˆ 3 κ3 = −i[hφ i − 3hφ ihφi + 2hφi ] . fication we replace the potential and its derivatives in Eq. (10) k k ∞ by their typical size and use Eq. (12) to find Vˆ ∼ ∆V Í xˆ /ξ . Since φˆ = Í φˆ is calculated from the overlap operator by I k n=1 n With the help of this form of the potential, the commutator Magnus expansion and is therefore Hermitian, we separate becomes Eq. (B1) into phase and amplitude. By comparing to Eq. (4), we find the phase φ of the interferometer k ∆V2 T Õ xˆ [Vˆ (t), Vˆ (t0)] ∼ ~ , 1 I I mξ2 ξk φ = φ + hφˆi − [hφˆ3i − 3hφˆi2hφˆi + 2hφˆi3] + ..., (B2) k 0 6 where we replaced the time difference in Eq. (C1) by the char- where we included the phase of the unperturbed interferometer acteristic interferometer time T and suppressed any numerical φ0 from Eq. (11), and the contrast C is factors. This result is easily generalized to the (n − 1)th order 1 nested commutator (which contains the potential evaluated at lnC = − (hφˆ2i − hφˆi2) + .... n different times) as 2 n−1 ˆ k 0 ∆V~ n−1 Õ x [Vˆ (t), [Vˆ (t ), [...]]]n− ∼  , (D1) I I 1 T n−1 ξk k 9 where  is given in Eq. (13). According to AppendixA the expectation value of any power of xˆ(t) can indeed be expressed nth-order term φˆn of the Magnus expansion contains (n − 1)th- in terms of the width d. order commutators, a factor 1/~n, and n integrals over time However, expectation values of the commutators in the which we replace by T n. Hence, together with Eq. (D1) one Magnus expansion or higher-order terms of the cumulant again obtains an infinite series expansion involve products between powers of xˆ evaluated at different times. The expectation value of such expressions Õ k φˆn ∼ ckn xˆ , (D2) can be calculated for example in Wigner phase space where k one has to take care of the correct operator ordering [47, 48]. Nevertheless, the scaling with the size of the wave packet where remains similar so that we will e.g. assume η n−1 l k 0 l+k ckn ∼ (D3) hxˆ (t)xˆ (t )i ∼ d . ξk and η was defined in Eq. (14). Note that, since the final integral With the help of Eqs. (D2) and (D3) we therefore find in the nested sequence extends along the whole contour, we Õ  d k replaced one factor of ∆V by the maximal potential difference hφˆ i ∼ η n−1 n ξ over the separation of the branches δV. Consequently, consecu- k tive terms in the Magnus expansion corresponding to the same n ˆ for the expectation value of the th order of the Magnus power of x are suppressed by  independently of the power. expansion.  Hence, if  1, the Magnus expansion can be truncated as the We now investigate the behavior of the cumulant expansion prefactors ckn quickly decrease order by order of n. by considering only the dominant operator-valued term in the After performing the Magnus expansion, it remains to calcu- overlap operator which is due to n = 1 and k = 1 in Eq. (D2) late the expectation value of the overlap operator. Because this provided   1 and d/ξ  1. As explained in AppendixB is in general not possible in an exact manner, we resort to the the lth order of the cumulant expansion is a function of the cumulant expansion for which expectation values of powers of first l moments which consequently scales as (ηd/ξ)l and we ˆ r j at different times have to be evaluated. To estimate the size therefore require of such expectation values independently of the explicit form of the initial wave function, we truncate the corresponding d η  1, probability density outside some region with characteristic ξ width d, where the probability to find a particle is vanishing. This approximation will allow us to express any moment in which is condition (16). If satisfied, we also truncate the terms of the finite width d of the wave function. Note that cumulant expansion at first order and obtain (considering only calculating phase and contrast with the help of the Magnus and terms up to harmonic order in the Taylor expansion of the cumulant expansions might lead to divergent series but there potential) Eq. (2), the main result of this article after recalling ˆ will exist a finite number of terms after which truncating the hr(t)i = 0. formal series leads to the best approximation in the spirit of an asymptotic expansion. In order to estimate the expectation value of powers of Appendix E: Gravitational potential rˆ j (t), we define the centered wave function |ψci, the initial state displaced by hrˆi and hpˆi to the origin of phase space. Equally, evolving with the free time-evolution operator U(ˆ t) = In this appendix we give the explicit form of the functions 2 f f f f exp[−ipˆ t/(2m~)], the freely expanding centered wave function φ, rr , r p and pp defined in Sec.V to calculate the phase (2) is denoted by |ψc(t)i. Thus, and wave-packet effects arising from Γ . For an arbitrary pulse sequence encoded in the unperturbed trajectory z0(t), the k k coefficients take the form hxˆ (t)i = hψc | (xˆ + ptˆ /m) |ψci † ¼ ¼ = hψ | Uˆ (t)xk U(ˆ t) |ψ i = hψ (t)| xk |ψ (t)i c ˆ c c ˆ c f = dt z (t)3, f = dt z (t), ¹ φ 0 rr 0 3 2 k k k = d r |ψc(r, t)| x = xd(t) ∼ d . ¼ ¼ 2 fr p = dt z0(t)t, and fpp = dt z0(t)t . First, we removed the expectation values hrˆi and hpˆi appearing in the definition of rˆ(t), see Eq. (3), by calculating the expecta- Using the explicit form of z0(t) for the MZ interferometer tion value with respect to centered wave function rather than sequence shown in Fig.1 with the initial conditions r0j (ti) = the actual initial state. In the third line we used the mean-value ziδjz and rÛ0j (ti) = 0, we therefore find x (t) theorem of integration [46] to find the number d within 2 6 4 the set where the wave function is nonvanishing, subsequently fφ = 31g vrT /20 − vrgT (14zi + 9vrT)/4 estimated by the maximal size d of the wave function. Thus, the 2 2 2 2 + vrT (vr T + 3vrT zi + 3zi ) 2 3 7 4 frr = v T , fr p = v T , and fpp = v T . r r 6 r 10

[1] M. Kasevich and S. Chu, “Atomic interferometry using stimulated Fermi", Vol. 188, edited by G. M. Tino and M. A. Kasevich (IOS Raman transitions,” Phys. Rev. Lett. 67, 181 (1991). Press, Amsterdam, 2014) p. 171. [2] A. Peters, K. Y. Chung, and S. Chu, “Measurement of gravita- [20] W. P. Schleich, D. M. Greenberger, and E. M. Rasel, “A tional acceleration by dropping atoms,” Nature 400, 849 (1999). representation-free description of the Kasevich-Chu interferom- [3] T. Farah, C. Guerlin, A. Landragin, P. Bouyer, S. Gaffet, F. Pereira eter: A resolution of the redshift controversy,” New J. Phys. 15, Dos Santos, and S. Merlet, “Underground operation at best 013007 (2013). sensitivity of the mobile LNE-SYRTE cold atom gravimeter,” [21] K.-P. Marzlin and J. Audretsch, “State independence in atom Gyroscopy Navig. 5, 266 (2014). interferometry and insensitivity to acceleration and rotation,” [4] M. J. Snadden, J. M. McGuirk, P. Bouyer, K. G. Haritos, and Phys. Rev. A 53, 312 (1996). M. A. Kasevich, “Measurement of the Earth’s gravity gradient [22] S. Kleinert, E. Kajari, A. Roura, and W. P. Schleich, with an atom interferometer-based gravity gradiometer,” Phys. “Representation-free description of light-pulse atom interfer- Rev. Lett. 81, 971 (1998). ometry including non-inertial effects,” Phys. Rep. 605, 1 (2015). [5] G. W. Biedermann, X. Wu, L. Deslauriers, S. Roy, C. Ma- [23] J. M. Hogan, D. M. S. Johnson, and M. A. Kasevich, “Light- hadeswaraswamy, and M. A. Kasevich, “Testing gravity with pulse atom interferometry,” in Atom optics and space physics, cold-atom interferometers,” Phys. Rev. A 91, 033629 (2015). Proceedings of the International School of Physics "Enrico [6] P. Asenbaum, C. Overstreet, T. Kovachy, D. D. Brown, J. M. Fermi", Vol. 168, edited by E. Arimondo, W. Ertmer, W. P. Hogan, and M. A. Kasevich, “Phase shift in an atom interferom- Schleich, and E. M. Rasel (IOS Press, Amsterdam, 2009) p. 411. eter due to spacetime curvature across its wave function,” Phys. [24] A. Roura, W. Zeller, and W. P. Schleich, “Overcoming loss of Rev. Lett. 118, 183602 (2017). contrast in atom interferometry due to gravity gradients,” New J. [7] S. Fray, C. A. Diez, T. W. Hänsch, and M. Weitz, “Atomic inter- Phys. 16, 123012 (2014). ferometer with amplitude gratings of light and its applications to [25] W. Zeller, “The impact of wave-packet dynamics in long-time atom based tests of the equivalence principle,” Phys. Rev. Lett. atom interferometry,” Ph.D. thesis, Universität Ulm (2016). 93, 240404 (2004). [26] W. Magnus, “On the exponential solution of differential equations [8] D. Schlippert, J. Hartwig, H. Albers, L. L. Richardson, C. Schu- for a linear operator,” Commun. Pure Appl. Math 7, 649 (1954). bert, A. Roura, W. P. Schleich, W. Ertmer, and E. M. Rasel, [27] S. Blanes, F. Casas, J. A. Oteo, and J. Ros, “A pedagogical “Quantum test of the universality of free fall,” Phys. Rev. Lett. approach to the Magnus expansion,” Eur. J. Phys. 31, 907 (2010). 112, 203002 (2014). [28] C. Ufrecht, “Theoretical approach to high-precision atom inter- [9] M. G. Tarallo, T. Mazzoni, N. Poli, D. V. Sutyrin, X. Zhang, and ferometry,” Ph.D. thesis, Universität Ulm (2019). G. M. Tino, “Test of Einstein equivalence principle for 0−spin [29] H. Ribeiro, A. Baksic, and A. A. Clerk, “Systematic Magnus- and half-integer-spin atoms: Search for spin-gravity coupling based approach for suppressing leakage and nonadiabatic errors effects,” Phys. Rev. Lett 113, 023005 (2014). in quantum dynamics,” Phys. Rev. X 7, 011021 (2017). [10] L. Zhou, S. Long, B. Tang, X. Chen, F. Gao, W. Peng, W. Duan, [30] H. Cramér, Mathematical methods of statistics (Princeton Uni- J. Zhong, Z. Xiong, J. Wang, et al., “Test of equivalence principle versity Press, Princeton, 1999). at 10−8 level by a dual-species double-diffraction Raman atom [31] E. Wodey, D. Tell, E. M. Rasel, D. Schlippert, R. Baur, interferometer,” Phys. Rev. Lett. 115, 013004 (2015). U. Kissling, B. Kölliker, M. Lorenz, M. Marrer, U. Schläpfer, [11] A. Wicht, J. M. Hensley, E. Sarajlic, and S. Chu, “A prelimi- et al., “A scalable high-performance magnetic shield for nary measurement of the fine structure constant based on atom very long baseline atom interferometry,” e-print arXiv: interferometry,” Phys. Scr. T102, 82 (2002). physics/1911.12320. [12] R. H. Parker, C. Yu, W. Zhong, B. Estey, and H. Müller, [32] C. Ufrecht, F. Di Pumpo, A. Friedrich, A. Roura, C. Schubert, “Measurement of the fine-structure constant as a test of the D. Schlippert, E. M. Rasel, W. P. Schleich, and E. Giese, “An Standard Model,” Science 360, 191 (2018). atom interferometer testing the universality of free fall and the [13] P. W. Graham, J. M. Hogan, M. A. Kasevich, and S. Rajendran, gravitational redshift,” e-print arXiv: quant-ph/2001.09754. “Resonant mode for gravitational wave detectors based on atom [33] A. Bertoldi, F. Minardi, and M. Prevedelli, “Phase shift in atom interferometry,” Phys. Rev. D 94, 104022 (2016). interferometers: Corrections for nonquadratic potentials and [14] M. Sonnleitner, M. Ritsch-Marte, and H. Ritsch, “Attractive finite-duration laser pulses,” Phys. Rev. A 99, 033619 (2019). optical forces from blackbody radiation,” Phys. Rev. Lett. 111, [34] S. Loriani, A. Friedrich, C. Ufrecht, F. Di Pumpo, S. Kleinert, 023601 (2013). S. Abend, N. Gaaloul, C. Meiners, C. Schubert, D. Tell, et al., [15] P. Haslinger, M. Jaffe, V. Xu, O. Schwartz, M. Sonnleitner, “Interference of clocks: A quantum twin paradox,” Sci. Adv. 5, M. Ritsch-Marte, H. Ritsch, and H. Müller, “Attractive force on eaax8966 (2019). atoms due to blackbody radiation,” Nat. Phys. 14, 257 (2018). [35] C. Chiu and L. Stodolsky, “Theorem in matter-wave interferom- [16] P. Storey and C. Cohen-Tannoudji, “The Feynman path integral etry,” Phys. Rev. D 22, 1337 (1980). approach to atomic interferometry. A tutorial,” J. Phys. II France [36] D. M. Greenberger, “The neutron interferometer as a device for 4, 1999 (1994). illustrating the strange behavior of quantum systems,” Rev. Mod. [17] C. Antoine and C. J. Bordé, “Exact phase shifts for atom inter- Phys. 55, 875 (1983). ferometry,” Phys. Lett. A 306, 277 (2003). [37] M. Sanz, E. Solano, and Í. L. Egusquiza, “Beyond adiabatic [18] B. Dubetsky, S. Libby, and P. Berman, “Atom interferometry in elimination: Effective Hamiltonians and singular perturbation,” the presence of an external test mass,” Atoms 4, 14 (2016). in Applications+ Practical Conceptualization+ = [19] E. Giese, S. Kleinert, M. Meister, V. Tamma, A. Roura, and Fruitful Innovation (Springer, Tokyo, 2016). W. P. Schleich, “The interface of gravity and [38] E. Giese, A. Roura, G. Tackmann, E. M. Rasel, and W. P. illuminated by Wigner phase space,” in Atom interferometry, Schleich, “Double Bragg diffraction: A tool for atom optics,” Proceedings of the International School of Physics "Enrico Phys. Rev. A 88, 053608 (2013). 11

[39] L. V.Keldysh, “Diagram technique for nonequilibrium processes,” [44] H. Ammann and N. Christensen, “Delta kick cooling: A new J. Exp. Theor. Phys. 20, 1018 (1965). method for cooling atoms,” Phys. Rev. Lett. 78, 2088 (1997). [40] J. Schwinger, “ of a quantum oscillator,”J. [45] R. Kubo, “Generalized cumulant expansion method,” J. Phys. Math. Phys. 2, 407 (1961). Soc. Jpn. 17, 1100 (1962). [41] A. Peters, K. Y. Chung, and S. Chu, “High-precision gravity [46] P. R. Halmos, Measure theory, Graduate Texts in Mathematics measurements using atom interferometry,” Metrologia 38, 25 (Springer, New York, 2014). (2001). [47] W. Schleich, Quantum Optics in Phase Space (Wiley-VCH, [42] C. J. Bordé, “Theoretical tools for atom optics and interferometry,” Berlin, 2001). C. R. Acad. Sci. 2, 509 (2001). [48] W. B. Case, “Wigner functions and Weyl transforms for pedes- [43] A. Roura, “Circumventing Heisenberg’s uncertainty principle trians,” Am. J. Phys. 76, 937 (2008). in atom interferometry tests of the equivalence principle,” Phys.