Abstract
We present an alternative technique for estimating the response of a cold atom interferometer (CAI). Using data from a conventional inertial measurement unit (IMU) and common strapdown terminology, the position of the atom wave packet is tracked in a newly introduced sensor frame, enabling hybridization of both systems in terms of acceleration and angular rate measurements. The sensor frame allows for an easier mathematical description of the CAI measurement and integration into higher-level navigation systems. The dynamic terms resulting from the transformation of the IMU frame into the CAI sensor frame are evaluated in simulations. The implementation of the method as a prediction model in an extended Kalman filter is explained and demonstrated in realistic simulations, showing improvements of over two orders of magnitude with respect to the conventional IMU strapdown solution. Finally, the implications of these findings for future hybrid quantum navigation systems are discussed.
1 INTRODUCTION
Atom interferometry is a highly precise technique for inertial sensing (Kasevich et al., 1991). By interrogating a free-evolving atom wave packet with a series of laser pulses, information about accelerations and turn rates can be extracted, allowing the calculation of a complete navigation solution (position, velocity, and attitude). Applications of this technique for accelerometers (Barrett et al., 2014), gyroscopes (Gauguet et al., 2009; Schubert et al., 2021), and complete inertial measurement units (IMUs) (Gebbe et al., 2021; Gersemann et al., 2020) based on Bose-Einstein condensates are currently under research. The potential position accuracy reaches 5 m after 1 h of inertial navigation (Jekeli, 2005), which makes atom interferometry a highly promising technique for navigation in global navigation satellite system (GNSS)-denied environments.
Commercial options are already available for static settings, i.e., quantum gravimeters (Vermeulen et al., 2018), and have an accuracy and long-term stability comparable or even superior to those of other high-end conventional sensors (Freier et al., 2016). In the case of changing inertial quantities, some challenges still remain. The atom interferometer itself has a very limited dynamic range when used alone, and conventional measurement models for the interferometer readout are often not suited for highly dynamic trajectories. Thus, in state-of-the-art techniques, hybridization with other sensors, such as conventional accelerometers, is mandatory in order to produce a navigation solution (Cheiney et al., 2018; Richardson et al., 2020; Templier et al., 2022; Zahzam et al., 2022).
In our prior work (Tennstedt & Schön, 2020), a combination of conventional IMUs and atom interferometers in the frame of an extended Kalman filter (EKF) was presented. The phase shift of the cold atom interferometer (CAI) is predicted from the measurement of the conventional IMU, determining the target fringe of the CAI and effectively solving the fringe ambiguity problem. The CAI measurement is then used to correct the prediction model of the IMU, which enables the estimation of a number of IMU error quantities, e.g., sensor bias or misalignment, based on the state vector configuration (Tennstedt & Schön, 2021) This model has been successfully tested with two IMUs of different accuracies, with the higher-class IMU emulating an atom interferometer (Weddig et al., 2021). With the bias of the conventional IMU corrected, the position drift of the combined navigation solution was reduced by a factor of 30. In general, the combined solution benefits from the superior long-term stability of the CAI, as discussed in Tennstedt et al. (2021). This solution was further evaluated for applications on low Earth orbit satellites (HosseiniArani et al., 2022), where the electrostatic accelerometer in use has a lower white noise density than the CAI, but a larger long-term drift.
However, the models introduced thus far have been developed under the assumption that the atom interferometer is aligned with the body frame of the vehicle. In addition, the equations are not scalable, which prevents the integration of one or more atom interferometers in a higher-level navigation system. In this paper, a suitable coordinate system for the sensor frame is introduced in order to allow an easier integration in future high-level navigation and multi-sensor systems. First, a system is developed for a general case with several sensor frames, followed by a special CAI multi-axis scheme based on differential measurements presented in Gersemann et al. (2020). This multi-axis scheme is then used for further evaluation in the scope of this paper.
The remainder of this paper is structured as follows. The measurement principle of CAIs for inertial sensing is briefly summarized in Section 2. The necessity and basic idea for hybridization are explained, and two types of hybridization that are relevant for this paper are described. In Section 3, the mathematical framework for the phase shift prediction, the atom strapdown scheme, is developed. A comparison with a second existing hybridization scheme follows in Section 3.2. Performance studies are carried out in Section 4 based on simulations, where the phase shift is of particular focus. Based on realistic dynamics derived from measurement campaigns, the single phase shift terms are quantified and visualized in order to gain insight and derive requirements for future hybrid sensor designs. Furthermore, the error of the predicted phase shift with respect to the CAI observation is tracked and discussed. The conclusions of this work are presented in Section 5 with a final evaluation.
2 ATOM INTERFEROMETRY FOR INERTIAL SENSING
Atom interferometers use the wave-particle characteristics of an atomic wave packet to deduce inertial quantities. In a basic Mach–Zehnder-type configuration, shown in Figure 1, an atomic wave packet is manipulated by a series of three directed laser pulses in order to split, reverse, and recombine the internal states of the atoms.
During the interaction of the atom wave packet with the beam splitter lasers, the atoms are set in a superposition of two internal energy states, |1, m〉 and |2, m + kħ〉, where m is the momentum of the wave packet, ħ is the reduced Planck constant, and is the effective laser wave number, which itself is related to the wavelength λ, e.g., 780 nm for rubidium. The probability that the atoms are in one of the two states oscillates with a certain Rabi frequency Ωeff and depends on the time of the laser interaction, the pulse time τ. If the atoms are all in state 1 at the beginning of the interaction, they will all switch their state after a certain 2τ pulse into state 2, resembling a mirror pulse. Likewise, if the interaction lasts for a duration of τ, the probability for a transition is 50% and a state-splitter pulse can be realized. During each atom–laser interaction, a phase shift between the two states is accumulated because both states yield different momenta. Under a number of assumptions such as a symmetric pulse sequence (Antoine & Borde, 2003), it is sufficient to describe the position of the center of mass of the atoms at the mid-point line between both states, as shown in Figure 1. The phase shift between the two states then depends on the relative position x of the atoms in the sensor frame, as well as the effective wave number k of the lasers for each atom–laser interaction. The main limitation of this method is that the trajectories of the two atomic states must overlap at the time of the recombination pulse. This requirement might be violated in the presence of rotations, where a loss of contrast attributed to an insufficient closing of the wave packet trajectory occurs (Roura et al., 2014) and a separation phase must be considered.
For the introduced three-pulse Mach–Zehnder configuration and under consideration of the midpoint-line theorem, the final phase shift ϕ(t3) at the time t3 of the last interaction reads as follows:
1
The actual observation after a measurement cycle is the transition probability, or population, of the atoms. This population p can be assessed by counting the number of atoms in one of the two states, N1 or N2. In turn, the population can be related back to the phase shift via the following relation:
2
As a didactic example, the phase shift can now be calculated by the inverse cosine. To retrieve the inertial effects from the phase shift, additional models are needed. Assuming a constant acceleration a, a simple motion equation of is sufficient. The quantity of interest can then be computed by applying the model to Equation (1), plugging the result into Equation (2), and then solving for a. The final equation for the inertial quantity in this example is then:
3
if k = k1 = k2 = k3 (static setting) and T is the interrogation time between the laser pulses, cf., Figure 1.
The approach of using the inverse cosine shows one major drawback of the CAI. The phase shift is only defined in the interval of 0≤ϕ≤π, which is a flank of the cosine function. For a typical setup with an interrogation time of 10 ms and the example above, this range corresponds to an acceleration of only approximately 2 mm/s2. A vibration of the sensor frame may be larger than this value, which leads to ambiguity in the CAI measurement.
This drawback can be circumvented by either compensating for vibrations via a stabilized platform or using different acceleration sensors to measure the vibration and to combine the signals via convolution with the CAI response function, cf., Cheiney et al. (2018) or Richardson et al. (2020), in order to enable measurements with the atom interferometer.
Another way to benefit from the better long-term stability of the CAI is to use a conventional IMU as the main sensor and to correct any drift errors with the atom interferometer measurement, e.g. Tennstedt & Schön (2020), while the solution of the ambiguity problem works in a similar manner. Both techniques will be explained in the next section.
3 HYBRIDIZATION
The idea behind hybridization is driven by two objectives. The first objective is to resolve the fringe ambiguity. This is done by reconstructing the atom interferometer phase with the help of high-rate sensors able to resolve signals that would otherwise exceed the dynamic range of the interferometer. For this step, we propose the atom strapdown technique as a prediction method that utilizes acceleration and angular rate measurements from an IMU as input.
The second objective is to correct the IMU-based prediction model with the observation of the atom interferometer. To this end, an EKF is applied with the bias of the accelerometers and gyroscopes as a state augmentation.
In the following subsections, this formalism is described, compared with another phase reconstruction method that has been previously applied to experimental data, and finally implemented in the framework of an EKF.
3.1 Phase Reconstruction with Atom Strapdown
As clarified in the introduction, the phase shift of an atom interferometer primarily depends on the relative position of the atom wave packet in the sensor frame. The basic equations of a typical strapdown algorithm that is used to compute the kinematic state of a moving platform from accelerometer and gyroscope measurements are adapted in order to estimate the wave packet position and its velocity.
3.1.1 Sensor Frame Phase Shift Representation
To allow a more general three-degree-of-freedom formalism, the position vector pointing to the center of the atom wave packet a in an arbitrary sensor frame s with index j ∈ [x, y, z] will be denoted as . For now, it is assumed that each CAI has only one sensitive direction, defined by its effective wave vector ksj. As an example, for j = x, cf., Figure 2, the vector may read ksx = [k, 0, 0]T, where k is the effective wave number introduced above. The phase shift ϕsj in the corresponding sensor frame can then be calculated by Equation (1). The origin of the s-frame is defined in the exact location at which the atom wave packet is released into free propagation. When the first splitter pulse is applied at t1 = 0, the initial position in the respective sensor frame is zero, and hence, the phase shift according to Equation (1) is reduced to two terms:
4
where T is the interrogation time between the laser pulses and · denotes a scalar product.
If an IMU is used to estimate the position of the atom wave packets in the sensor frame, a framework known from inertial navigation can be utilized: the strapdown algorithm. The differential motion equations for the position and velocity of the respective atom wave packets are as follows:
5
6
Here, are the specific forces, i.e., accelerations, and the term denotes the Coriolis force, as the atoms have a velocity relative to the s-frame. The gravity vector gsj is considered as well. The rotation rates are expressed in the skew-symmetric matrix , where the index i denotes the inertial frame. The index sj with j ∈ [x, y, z] denotes the corresponding sensor frame, cf., Figure 3. As an example, the vector marks the velocity between the sx frame origin and the atom wave packet a in the sx frame, expressed in the sx frame.
Note that the equations are solved in the respective s-frame; hence, the change in the velocity of the atom with respect to the s-frame points in the direction opposite to the acceleration of the s-frame, cf., Equation (6). This does not account for gravity gsj, which affects the motion of the atom wave packet even when the s-frame is not accelerated. Gravity can be considered as an external force. Additional effects, such as centrifugal and Euler accelerations, that depend on the position are not considered in this study, as their effects are comparably small for short interrogation times. For future actual implementations with specified CAI parameters, such as the initial position of the atoms in the s-frame, the differential equation (Equation (6)) can be enhanced accordingly. Furthermore, the recoil velocity caused by the laser–atom interaction is neglected, as the impact on the phase shift is small compared with the effect of the accelerations.
The differential equations are solved by integration to determine the atom positions at times t = T and t = 2T and to calculate the phase via Equation (4). Unless otherwise stated, all parameters are time-dependent.
3.1.2 Consideration of Spatial Displacements Between Sensors
The accelerations and turn rates are now given in a different system b, i.e., the body frame with the origin at the center of the IMU. This arrangement results in a spatial distance to the s-frame in which the quantities must be transformed, i.e., a lever arm , cf., Figure 3. As an example, for a very simplistic and general design of a three-dimensional atomic sensor system, it can be assumed that there are three atomic interferometers with frames sx, sy, and sz, each with a sensitive axis in one of the spatial dimensions. The transformations of specific forces in this case are as follows:
7
The fictitious forces resulting from the transformation are expressed in the centrifugal term and the Euler term , which results from angular accelerations of the frame. The sensor frames and body frame are all connected to a rigid platform; thus, there is no change in the distance , and no Coriolis term needs to be considered. This same rigid platform argument allows us to set the specific force between both frames to zero, . A rotation between the two systems, resembling sensor misalignment, can be included.
The transformation of the turn rates is straightforward:
8
In analogy to the forces, the rigid platform allows no relative rotation between the sensors and body frame, and thus, .
3.1.3 Multi-Axis Model
It is assumed that all three CAI sensors use a single atomic source, as displayed in Figure 3, cf., Gersemann et al. (2020); thus, a single sensor frame s is sufficient. Furthermore, the misalignment between the s and b frames is assumed to be zero, and hence, equals I, the identity matrix.
This multi-axis CAI model has three sensitive axes, which are considered in the model by their respective ks-vector. For example, for the interferometer sequence in the x direction, the vector reads ks = [kx, 0, 0]T, with kx as the respective effective wave number. The coordinate center and the starting position of the atoms in each direction are identical. Note that in this multi-axis CAI, an initial splitter pulse is applied immediately before the Mach–Zehnder sequence, which divides the initial wave packet into two counter-propagating ensembles in order to differentiate between phase shifts based on linear acceleration and those based on rotation. This approach affects the initial velocity and will be important in Section 3.3.
According to Equation (4), the phase shift ϕs depends on the position xs of the atoms in the s-frame. We can state the following:
9
with the single position terms caused by linear acceleration , rotation , and effects of a changing rotation vector during the integration interval now explained in detail.
Acceleration
The position of the atoms in the laser frame s, as measured by a conventional accelerometer in the b-frame, is calculated as follows:
10
In the position shift caused by linear accelerations, three terms are involved in addition to gravity gs: the specific forces measured by the conventional accelerometer, the centrifugal term due to the lever arm, and the Euler term resulting from a change in turn rate.
Rotation
The atom position shift based on a rotation of the s-frame results from the Coriolis term alone:
11
Coning
The presence of turn rates during the measurement interval leads to a change in orientation vector of the s-frame, which must be considered. An additional coning/sculling mixture term appears, cf., Savage (1998), which will subsequently be labeled with the subscript mix. The position of the atoms in the laser frame is then as follows:
12
Note that the same fictitious force terms, namely the centrifugal and Euler terms, apply as in the linear acceleration case. The difference here is that the whole term is rotated by the relative change in orientation .
Higher-Order Terms
The wave packet is potentially affected in the presence of gravity gradients or gravitational waves, which leads to higher-order phase shift terms with respect to the integration time (Antoine & Borde, 2003; Bongs et al., 2006). For the terrestrial navigation case considered herein and the often limited dimensions of CAI vacuum chambers, the interrogation time T is comparably short, typically in the range of 10–25 ms. Thus, higher-order terms only have a diminishing effect in the face of terms resulting from rotation and linear acceleration and, consequently, can be neglected.
This situation changes drastically in space applications (Trimeche et al., 2019), where large interrogation times of several seconds are achievable and gravity gradients as well as higher-order terms can no longer be ignored.
3.2 Evaluation of the Atom Strapdown Phase Prediction
To evaluate the atom strapdown method, this technique is compared with a second method to extract the interferometer phase: the convolution of acceleration sensor signals with the interferometer sensitivity function. This method is a common way to hybridize the signals of conventional sensors with the atom interferometer response and is used by different groups, e.g., Richardson et al. (2020), Cheiney et al. (2018), and Zahzam et al. (2022). First, this method is briefly summarized. Then, the predicted phase shift of the atom strapdown method and the convolution are compared based on an experimental data set Richardson et al. (2020). In that experiment, the acceleration data of the sensor frame are measured by an optomechanical resonator (OMR) attached to the retroreflection mirror of the interferometer. This sensor is a mere prototype, and its characteristics are only briefly stated.
Phase Reconstruction by Convolution
To reconstruct the atom interferometer response that has been distorted because of acting ambient ground vibrations, the OMR signal is convolved with the acceleration sensitivity function f(t) of the atom interferometer, given in Equation (13) and shown in Figure 4. This convolution is then integrated over the interferometer duration to calculate the corresponding phase correction value for that particular atom interferometer shot, as shown in Equation (14):
13
Here, Ωeff is the effective Rabi frequency, τ is the laser pulse duration, and T is the interrogation time:
14
The OMR has a sensitivity on the order of 10−5m/s2 for frequencies above 10 Hz; at lower frequencies, 1/f noise dominates. Therefore, the OMR is capable of suppressing ambient vibrational noise for interferometers with lower integration times, with a maximum limit of 10 ms for good fringe reconstruction. This limitation arises because interferometers with lower T have a higher corner cut-off frequency, which is given by 1/2T; therefore, higher-frequency vibrations will be capable of inducing noise to the interferometer. Beyond an interrogation time of 10 ms, the fringe reconstruction performance obtained via the OMR signal declines because of its poor low-frequency performance.
The acceleration data were acquired at a rate of 100 kHz throughout the interferometer duration, which spans TMZ = 0.02003 s. Consequently, for each shot of the atom interferometer, 2003 acceleration measurement samples are integrated over the interferometer duration in the strapdown in order to calculate the atom position in the sensor frame by using the equations in Section 3.1.3. The positions at the times of the atom–laser interactions are then used in Equation (4) to compute the phase shift. The Earth turn rate at the local latitude, 52.38°, is used as input for .
Evaluation
The resulting phase shifts from both methods show good agreement, cf., Figure 5. However, a relatively constant offset of approximately 0.04% is visible, as well as occasional spikes of up to 0.186%. The largest difference during the whole experiment with over 17000 CAI shots is approximately 2%.
One possibility for the deviation lies in the beam splitter pulse durations, which are not yet considered in the atom strapdown method. The two state splitter pulses each have a duration of 7.5 μs, while the duration of the mirror pulse is twice as long, leading to a total of 30 μs. During these pulses, the sensitivity of the interferometer regarding the inertial quantity changes, following a sinusoidal pattern, cf., Equation (13). Because of the 100-kHz accelerometer data rate, approximately 3–5 of the 2003 data points are integrated differently in the two methods.
Finally, we include some general remarks on the two methods: The convolution technique uses the sensitivity of the atom interferometer with respect to changes in the phase shift, which is then further related to the sensitivity to the acceleration of the sensor frame.
In the atom strapdown method, the position of the atoms is calculated explicitly. However, the phase shift in Equation (1), which uses the atom position as input, originates from the sensitivity of the atom interferometer response. Thus, the convolution method can be seen as a more specialized and slim way of combining signals of different sensors with the CAI measurement, whereas the atom strapdown method is a more general approach, allowing the consideration of arbitrary target frames.
The new method can be further used to include different motion models and dynamics in the differential equation system. Furthermore, different sensors could be utilized to predict the atom position, for example, odometry from car wheels or visual sensors.
3.3 Implementation of Phase Prediction in a Hybrid Filter
Thus far, the atom strapdown method has been introduced as a way to predict the phase shift of an atom interferometer in an arbitrary sensor frame. The general steps for implementing the phase shift prediction in a filter framework (see Figure 6) have been sketched in prior works, e.g., Tennstedt & Schön (2021). Some important aspects are now repeated and explained, including the adaption of the measurement model, as the new dynamic equations need some special care.
It is worth stressing the clear differentiation between the kinematic state of the moving platform calculated in the IMU strapdown, i.e., the position and velocity of the body frame (index b) with respect to the Earth frame (index e) as well as the attitude in a target navigation frame (superscript n), from the kinematic state of the atom wave packet in the sensor frame (atom strapdown). The parameters that are used in both systems are the bias of the accelerations ba and angular rate measurements bg. The kinematic state of the wave packet changes with each new measurement cycle; thus, it is not necessary to keep track of these states beyond the corresponding measurement interval. Consequently, the state vector of the filter inherits the kinematic state of the platform with the IMU bias as a state augmentation:
15
With this design decision, the filter innovation, which is based on the atom phase shift model, can also be used to correct the kinematic state of the platform, due to coupling with the bias in the system model. The corresponding transition matrix F over the course of a system time step Δt is as follows:
16
The uncertainties for the accelerometers wa and gyroscopes wg are modeled as white noise with respective variances and random walk processes. The variance of the latter is modeled by an integrator with white noise wrw,in as input, sampled from a Gaussian distribution with variance . For each individual axis and sensor, the process reads as follows:
17
The processes are combined in the system perturbation vector w = [wa, wg]T and affect the bias of the accelerations and angular rates. Hence, the perturbation input B matrix reads as follows:
18
where the covariance matrix is given as follows:
19
The discrete time process noise matrix with subscript k indicating the time step t → t + Δt can then be assembled:
20
The remaining equations follow that of a regular EKF with error states, as described in Tennstedt et al. (2021). Thus, at the filter step at time step k, the total state is updated with the estimated error state :
21
The superscript – denotes a predicted state, whereas + denotes a filtered state. At each filter update, the predicted error state is always initialized as zero; hence, the filtered error state is essentially the difference between the CAI observation pk and the predicted observation , weighted by the Kalman gain Kk:
22
The observation model h of the CAI is based on the phase prediction obtained from the IMU data. The acceleration measurement of the IMU, , must be adapted in order to include a potential deviation ba, i.e., bias, from the true acceleration value :
23
The analytical CAI observation equation per axis is the transition probability p:
24
where ϕ0 is the steered laser phase and ϕ is the phase shift prediction from the IMU data. Readout contrast and offset of the actual signal are not considered in the scope of this paper. In fact, only the contrast affects the sensitivity in a real application.
The steered phase ϕ0 is set such that the sum of ϕ0 and ϕ equals in order to maximize the sensitivity with respect to the additional bias-induced phase shift δϕ.
The derivative of Equation (24) with respect to the accelerometer bias reads as follows:
25
In the limit of δϕ → 0, we obtain the following:
26
The observation matrix Ha for the acceleration bias is then written as follows:
27
Only the acceleration and mixed term, cf., Equation (10) and Equation (12), respectively, are sensitive for an acceleration bias. Because of the skew-symmetric matrix , the second term is non-zero only if the angular rate in the direction of the respective perpendicular axes is non-zero. This result is in line with the prior statement that the phase shift on a particular sensitive CAI axis is only affected by rotations perpendicular to the axis.
The observation matrix Hg for the gyroscope bias follows a similar pattern. The IMU gyroscope measurement consists of the true angular rates and the bias bg:
28
The derivative of Equation (24) with respect to the gyroscope bias is as follows:
29
As before, the limit of δϕ → 0 gives the following:
30
The observation matrix Hg for the gyroscope bias results in a rather extensive expression:
31
The first two terms come from the derivative of the centrifugal term with respect to the gyroscope bias. The term resembles the Coriolis term observed for the atoms in a rotating s-frame. This term scales with the atom velocity. The fourth term results from the partial derivative of the mixed term, which scales with the linear accelerations. Terms five to seven are produced by the product rule when differentiating the centrifugal term between the b- and s-frame. The final Euler term including the angular acceleration is not considered for the EKF measurement linearization in the frame of this work because of the inaccessibility of these values. The angular accelerations can be acquired by either numerical differentiation of the turn rates or state vector augmentation when using at least three conventional IMUs. However, the first method is rather inaccurate because of the differentiation of gyroscope white noise in the signal, whereas the second method may be evaluated in a future study.
The complete observation matrix H can now be constructed:
32
Thus far, the observation equation and matrices in the s-frame have been stated for an arbitrary CAI observation on a single sensitive axis as defined by the wave vector ks. In the six-degree-of-freedom case, there are now two measurements per spatial axis at a time in order to differentiate between angular rates and linear accelerations. In this case, the observation vector reads , where the superscripts A and B refer to anti-parallel propagating wave packets after the initial separation split. The individual phase shifts are predicted for each observation independently; thus, for three sensitive CAI axes with two counter-propagating wave packets, there is a total of six different predicted atom wave packet states.
For now, it is assumed that the interferometers on all axes perform measurements at the same time. Thus, the observation matrices are individually computed according to Equations (27) and (31) with the respective ks and and are finally added together to produce a single update step.
4 RESULTS AND DISCUSSION
4.1 Impact of Dynamics on Phase Shift Terms
Setting
We now consider a hybrid solution between a conventional IMU and the CAI according to the multi-axis scheme. The previous equations revealed a number of fictitious force terms when the IMU measurement is transformed from the b-frame to the sensor frame. To evaluate the impact of different dynamics on the single phase shift terms, a trajectory with a constant velocity of 1 m/s in the driving direction s1 as well as one period of a sinusoidal changing yaw rate with an amplitude of 0.05 rad/s is simulated. With these settings, it is possible to apply an angular rate and angular acceleration on s3 and an accelerations in the s2 direction and thus excite all phase shift terms.
The goal is to determine to which order of magnitude the single terms affect the phase shift of the interferometer. Afterwards, the effect of the most dominant terms on the error phase shift in the case of an IMU bias is evaluated depending on the sensor distance .
Results
The results in Figure 7 show that in the forward direction, the rotation term is dominant. Whereas the Euler and centrifugal terms of the acceleration as well as the mixed term generally scale with the distance between the b- and s-frames, the largest effect arises from the centrifugal term. In the right direction, the largest impact results from the lateral accelerations of the trajectory. However, the angular accelerations are scaled by the lever arm as well, which leads to an increased Euler term that surpasses even the lateral accelerations in parts of the simulation when the lever arm is set to 1 m, e.g., at t = 25 s. Regarding the mixed terms, the effect is generally quite small. In fact, the mixed term resembles the transformation terms of the linear accelerations, scaled by the integrated angular rates. Thus, for most terrestrial applications, these terms are smaller than the centrifugal and Euler terms from the transformation of the linear acceleration alone.
Next, we focus on the effect of the lever arm on the dynamic terms. In contrast to the prior example, the error of the predicted phase in the case of a bias in the accelerations and angular rates is now examined. Instead of a time-dependent trajectory, only a slice of the trajectory with the dynamics at an exact time is given. The phase error in this situation is evaluated based on the size of the lever arm, where is varied in the interval ℓ ∈ [0,1] m on an equidistant scale.
The parameters are set as follows, according to the maximum values that can be expected for typical maneuvers. A constant angular rate rad/s and acceleration m/s2 represent the dynamic situation. The angular rate and angular acceleration for the mixture Euler term are scaled down by a factor of to account for being normally phase-shifted by 90 degrees with respect to each other, as Figure 7 demonstrates in the right direction for the example of the Euler and acceleration terms.
The constant biases of the acceleration and angular rate signals are defined as ba = 10−4 m/s2 and bg = 3 · 10−5 rad/s on all axes.
The interrogation time and wave numbers are T = 25 ms and kx,y,z = 4π/780 nm, respectively. The initial atom velocity m/s is summed with the integrated linear accelerations in order to emulate the largest atom velocity to be expected.
As shown in Figure 8, the linear acceleration and the rotation in the corresponding direction are naturally independent of the lever arm. The acceleration has a relatively large impact, which is caused by the comparably large accelerometer bias. The two terms that show the greatest scaling with the magnitude of the lever arm are the centrifugal and Euler terms. This trend matches the results in Figure 7, but also shows that even a small distance of 0.4 m in any direction can surpass the dynamic range defined by the atom interferometer measurement equation.
4.2 Impact of Dynamics on Bias Estimates
Thus far, the new algorithms have been introduced and the adaption of a hybrid filter presented. The results have indicated that the spatial displacement between the CAI and the IMU might lead to difficulties for the bias estimate. The correction of this bias is a crucial step in improving the hybrid navigation solution. Based on another trajectory derived from a number of different dynamic experiments (Weddig et al., 2022), the effect of the dynamics in the s-frame on the bias estimate is further investigated.
Setting
A trajectory that involves most of the maneuvers identified in the analysis of different data sets collected with IMUs during real-world measurement campaigns is simulated, cf., Figure 9. After an initial static phase, an accelerated phase follows with some minor vibrations on all axes. This phase is followed by a sliding right turn, with only the angular rate changing. Afterwards, a left turn is applied with accelerations in the b1 and b2 directions, as well as a spike in both signals at t = 26 s. The trajectory is concluded with another accelerated phase with vibrations and a small turn rate on the b3 axis.
The simulated IMU acceleration signal is affected by a discrete white noise process with a standard deviation of 5 · 10−5 m/s2. Additionally, a random walk process based on integrated white noise with variance 9 · 10−10 m2/s5 is applied in order to introduce a significant drift of the signal. The gyroscope measurement consists of the angular rates with additive white noise samples with a standard deviation of 1.4544 · 10−5 rad/s and a random walk process with input white noise sampled from a variance of 4 · 10−14 rad2/s3. This measurement corresponds to the performance of a navigation-grade ringlaser gyroscope. The data rate of the simulated trajectory and the IMU is 200 Hz. The CAI is modeled with a discrete white noise process with a variance of 1 · 10−3 a.u. added to the transition probability. The interrogation time is set to T = 25 ms, corresponding to a filter rate of Δt = 2T = 50 ms, with no dead or readout times considered. The uncertainty of the simulated CAI translates to a phase uncertainty of approximately 63 mrad or, in conjunction with the interrogation time, an acceleration uncertainty with a standard deviation of 6.3·10−6 m/s2 per shot.
A spatial displacement of m is applied, which was selected based on the results from the previous section, where the values for and are based on the maximum amplitude of the oscillations and the bias is equal to the deviation of approximately 2σ for the respective white noise process. This distance may resemble an IMU placed directly over the atom interferometer frame. The simulation was processed in 100 different runs, in which all noise processes were sampled independently. Importantly, this number allows for the intended qualitative discussion, but is not sufficient for a statistically proven Monte-Carlo analysis, as twelve independent variables are sampled in total.
Note that in the trajectory at approximately 26 s, when the large spike in accelerations in the x and y direction occurs, the oscillation amplitude in the acceleration signal in the z direction also reaches its maximum. At the same time, the amplitude of the angular rates in x and y peaks, while a large turn rate in the z direction is present as well. According to the experiment in the previous section, it can be expected that the phase shift error resulting from this situation might be similar to or even larger than the dynamic range of the atom interferometer and thus produce an error in the filtered solution.
Results
The acceleration bias estimates, cf., Figure 10, are able to follow the noise process quite well, and the effects of the dynamics of the trajectory can hardly be seen. The increase in uncertainty over time, as reflected in the empirical standard deviation, is explained by the random walk that was added to the base signals. In the standard deviation of the gyroscope bias, two distinct peaks can be seen at approximately 26 s and 33–35 s, prominently in the x and y direction. This expected decrease in performance is also visible in the velocity state, cf., Figure 11. Although the velocity estimate of the hybrid solution is largely improved during the first 25 s with a factor of approximately 100 at best, there is a sudden increase in the error at approximately 26 s. The acceleration spikes in the trajectory scale with the accelerometer and gyroscope errors in the mixed term, leading to a comparably large phase error. At 33 s, a similar case occurs when the magnitude of the trajectory signals on all axes is at a local maximum.
In total, the error of the predicted phase was larger than π/2 in approximately 14% of the CAI measurements. This result does not necessarily imply that the target fringe is lost and the EKF update fails, as convergence to the true trajectory is generally still possible in the range of +π. Nevertheless, a larger linearization error occurs when the measurement sensitivities are computed in these cases, leading to a non-optimal update of the filter despite the fact that the trajectory is resolved by the IMU and the innovation in Equation (22) is independent of the linearization.
For comparison, the results of a smooth slalom trajectory without any spikes in the signals are depicted in Figure 12. The CAI-aided EKF solution allows an improvement of over two orders of magnitude as compared with the IMU strapdown solution alone.
5 CONCLUSION
In this paper, the atom strapdown method was introduced and explained, as well as its application for state prediction in an EKF. This hybridization technique based on a classical IMU and a CAI has two main qualities: First, it allows the phase shift of an atom interferometer to be predicted based on data from a conventional IMU. Equations were developed in a newly introduced sensor frame for the CAI. Second, this method allows one to combine measurements of an atom interferometer with conventional IMU measurements, in order to correct for IMU biases.
A comparison between the strapdown method and the hybridization technique based on convolution, developed and used in the physics community, shows that both methods deliver comparable results in the predicted CAI phase shift, while the atom strapdown method is a more general concept that allows the consideration of different dynamics and sensor systems. The methods were compared based on a static data set. Further kinematic experiments, including a one-axis CAI with controlled time-varying accelerations as well as a full six-degree-of-freedom setup, are currently planned. More detailed CAI models that include the effects of dynamics and temperature on the atoms will be investigated as well.
Taking the dynamics from real car trajectories obtained from IMU measurements, our simulations showed that it is possible to produce a valid hybrid solution of a conventional IMU and a six-axis CAI, as the trajectory states and biases are reliably estimated. Comparison with an IMU-only solution showed performance improvements of about two orders of magnitude. In the case of large signal spikes on several axes at the same time, the solution loses accuracy due to cross-coupling of different axes and due to a scale effect induced by the transformation from the b- to s-frame. Methods for identifying those situations on the fly in order to discard the CAI measurement if necessary will be investigated in future studies.
In addition, two conclusions for practical implementation can be drawn: First, the lever arm, especially between the accelerometer and the CAI, should be kept small. The actual limit is strongly related to the IMU noise characteristics and the dynamics that are expected in individual applications. The exemplary study in Section 4.1 revealed that a distance of 40 cm between the two sensors was the maximum for the dynamic situation considered. Similar assessments could be performed for further applications. The distance is less important for the gyroscope, as the transformations from the b- to s-frame in Equation (8) show that the angular rate measurement is not affected by the lever arm. Solutions such as the OMR prototype applied directly to the retroreflection mirror of the CAI seem to provide a valid option for future hybrid accelerometer implementations, while at the same time, a high-precision gyroscope with a larger distance to the CAI could be used.
Second, the best situations for using a CAI in conjunction with an IMU might be in navigation tasks in which no spikes in the trajectory signals occur, e.g., in maneuvers with nearly constant acceleration and turn rates. Even if the EKF can cope with those situations, the best performance is achieved in smooth maneuvers. More specific applications must be examined in the future.
CONFLICT OF INTEREST
The authors declare no potential conflict of interests.
HOW TO CITE THIS ARTICLE
Tennstedt, B., Rajagopalan, A., Weddig, N.B., Abend, S., Schön, S., & Rasel, E. M. (2023). Atom strapdown: Toward integrated quantum inertial navigation systems. NAVIGATION, 70(4). https://doi.org/10.33012/navi.604
ACKNOWLEDGMENTS
This research was funded by the Federal Ministry for Economic Affairs and Climate Action (BMWK) due to an enactment of the German Bundestag under Grant 50RK1957 (QGyro) and 50NA2106 (QGyro+). The team also acknowledges funding from the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy—EXC-2123 QuantumFrontiers—Project-ID 390837967.
This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the original work is properly cited.