Introduction

Quantum sensing1 and metrology2 are important branches of modern quantum technologies with applications in different areas such as imaging3,4 and spectroscopy5,6,7. In this scenario, atomic-sized sensors encoded in171Yb+8,9,10,11,12 and 40Ca+13 ions provide spatial resolution and sensitivity for the detection of external/target fields. In addition, 171Yb+ ions exhibit negligible emission rates11 and extended coherence times due to the stabilization provided by dynamical decoupling (DD) techniques14,15,16. In particular, DD methods that exploit the multi-level structure in the\({}^{2}{S}_{\frac{1}{2}}\) manifold of the 171Yb+ ion have to lead to the detection of radio frequency (RF) fields with sensitivity close to the standard quantum limit9, while the resulting dressed state qubit was proposed as a robust register for quantum information processing8,10. However, this approach is restricted to a narrow working regime leading to harmonic sensor responses where RF target parameters are encoded. A departure from such a range leads to complex sensor responses where standard inference of the external fields gets challenging. In another vein, machine learning (ML) tools are incorporated to address distinct problems in quantum technologies. In particular, neural networks (NNs) are valuable in distinct quantum sensing scenarios leading to adaptive protocols for phase estimation17,18,19, parameter estimation20,21,22,23,24, and quantum sensors calibration25,26,27.

In this article, we experimentally demonstrate the ability of NNs to decode complex sensor responses, thus significantly extending the operational regime of quantum detectors. In particular, we infer RF target parameters from the measured response of an 171Yb+ atomic sensor in distinct challenging scenarios. These comprise the regime of non-harmonic sensor responses with large shot noise, and the continous interrogation of the sensor (via single-shot measurements) under always-on RF fields. With careful modeling of the sensor-target dynamics, we train NNs to relate intricate responses of the sensor with RF parameters leading to estimations of the latter with high accuracy.

Results

The quantum sensor

The sensor consists of four hyperfine levels (\(\left\vert 0\right\rangle ,\left\vert \acute{0}\right\rangle ,\left\vert 1\right\rangle\), and \(\left\vert -1\right\rangle\)) in the \({}^{2}{S}_{\frac{1}{2}}\) manifold of the 171Yb+ ion, and in a static magnetic field Bz. To enhance the coherence of the sensor, we use two microwave (MW) fields with Rabi frequency Ω leading to the dressed basis set \(\{\left\vert u\right\rangle ,\left\vert d\right\rangle ,\left\vert D\right\rangle ,\left\vert \acute{0}\right\rangle \}\) with \(\left\vert u\right\rangle =(\left\vert B\right\rangle +\left\vert 0\right\rangle )/\sqrt{2}\), \(\left\vert d\right\rangle =(\left\vert B\right\rangle -\left\vert 0\right\rangle )/\sqrt{2}\), \(\left\vert D\right\rangle =(\left\vert -1\right\rangle -\left\vert 1\right\rangle )/\sqrt{2}\), and \(\left\vert B\right\rangle =(\left\vert -1\right\rangle +\left\vert 1\right\rangle )/\sqrt{2}\) which is insensitive to magnetic field fluctuations8,9, see Fig. 1a, b.

Fig. 1: Schematic levels of the 171Yb+ atomic sensor and the experimental setup.
figure 1

a Relevant levels of the 171Yb+ atomic sensor. Two resonant MW fields drive the sensor with a Rabi frequency Ω leading to the configuration in (b) which is defined in the dressed state basis \(\{\left\vert u\right\rangle ,\left\vert d\right\rangle ,\left\vert \acute{0}\right\rangle ,\left\vert D\right\rangle \}\)9. c Schematic configuration of the experimental setup. The ion is trapped in a needle trap which consists of a pair of RF electrodes and four DC electrodes. The MW fields for state manipulation are generated by mixing a 12.4428 GHz signal with the signal from an arbitrary waveform generator (AWG1). The 12.6 GHz signal after a high pass filter (HPF) is amplified by an amplifier (Amp1) and sent to the ion using a MW horn outside the vacuum chamber. The target field around 10 MHz is generated by another arbitrary waveform generator (AWG2) controlled by the computer and broadcast to the ion through a coil after amplification. We place an extra RF switch after AWG2, which can be used to remove the target field from the trap while keeping the signal from AWG2 continuous. A photomultiplier tube (PMT) is used to detect the state-dependent fluorescence. d Scheme of the NN. The response from PMT, i.e., X, is processed by a number of hidden layers (HL1... HLk) leading to the outputs Y.

Once a target field \({\Omega }_{{{{\rm{tg}}}}}\cos ({\omega }_{{{{\rm{tg}}}}}t+{\phi }_{{{{\rm{tg}}}}})\) is applied, one measures the dark state (\(\left\vert D\right\rangle\)) survival probability PD(t) considered here as the sensor response. The standard working regime of the sensor (leading to harmonic responses) implies Ωtg Ω ωtg with ωtg resonant with, e.g., the \(\left\vert \acute{0}\right\rangle \leftrightarrow \left\vert 1\right\rangle\) transition9. This simplifies the total Hamiltonian (see Supplementary Note 1) into

$$H=-\frac{{\Omega }_{{{{\rm{tg}}}}}}{2\sqrt{2}}(\left\vert D\right\rangle \langle \acute{0}| +| \acute{0}\rangle \left\langle D\right\vert ).$$
(1)

From Eq. (1) one finds the harmonic sensor response \({P}_{D}(t)={\cos }^{2}(\pi t/{t}_{{{{\rm{R}}}}})\), with \({t}_{{{{\rm{R}}}}}=2\pi \sqrt{2}/{\Omega }_{{{{\rm{tg}}}}}\) establishing a simple relation between tR and the target field parameter Ωtg. In contrast, a departure from the standard working regime leads to the loss of this type of simple dependencies between sensor responses and target parameters, thus posing serious challenges for reliable field characterization. We demonstrate that our setup combining a quantum sensor and NNs can extract RF fields parameters from complex sensor responses.

Experimental setup

As schematically shown in Fig. 1c, our protocol is executed on a single 171Yb+ ion confined in a Paul trap which is shielded by permalloy to reduce the surrounding magnetic noise28,29. A magnetic field Bz is applied to the ion leading to a Zeeman shift ≈10.0 MHz between \(\left\vert \acute{0}\right\rangle\) and \(\left\vert 1\right\rangle\). Two ~12.6 GHz MW fields \({\Omega }_{j}\cos ({\omega }_{j}t+{\phi }_{j})(j=1,2)\), respectively resonant with transitions \(\left\vert 0\right\rangle \leftrightarrow \left\vert 1\right\rangle\) and \(\left\vert 0\right\rangle \leftrightarrow \left\vert -1\right\rangle\) are used for state manipulation. To generate the dark state \(\left\vert D\right\rangle\), we design pulses with the amplitudes Ω1 and Ω2 evolving in the form of a hyperbolic tangent (details available in Supplementary Note 2). During the sensing window, the amplitudes of dressing fields are kept constant at Ω1 = Ω2 = (2π) × 5.5 kHz. Afterwards the state remaining in \(\left\vert D\right\rangle\) is transferred to \(\left\vert 0\right\rangle\) for detection.

A copper coil placed under the trap generates the target field once an RF current is sent to the coil. As the original RF signal is produced by an arbitrary waveform generator (AWG2), the amplitude, frequency, and the initial phase of the target field can be set by adjusting the parameters of AWG2. An RF switch is placed after the AWG2 output in order to remove interaction between the field and the ion if necessary while keeping the signal source continuously on. After amplification, the RF target field generated by the coil has a maximum amplitude of \({\Omega }_{{{{\rm{tg}}}}}^{\max }=(2\pi )\times 9.0\) kHz. A photomultiplier tube (PMT) is used to detect the state-dependent fluorescence. As schematically shown in Fig. 1d, the response from PMT, i.e., X, is processed through the well-trained NN leading to the outputs Y which approach the targets A. We use data generated from numerical simulations to create the NN. Then, by inputting experimentally collected responses into the NN, we estimate the RF parameters.

Now, we demonstrate the good performance of an 171Yb+-magnetometer assisted by a NN in two scenarios in which input data comprise (i) average values obtained from a reduced number of measurements, and (ii) binary sequences (including only 0, or 1) continuously acquired from single-shot measurements. Both cases are demonstrated in regimes where the sensor delivers complex responses that depart from the standard harmonic regime. Hence, we prove that NNs significantly extends the versatility of quantum sensors.

Scenario i: Parameter estimation with a reduced number of measurements

We aim to estimate the Rabi frequency of a target field. In this case, the input data string \({{{\bf{X}}}}=\{{P}_{1},{P}_{2},...,{P}_{{N}_{p}}\}\) consist of the average values Pi (i [1, Np]) collected at Np time instants t = ti distributed in a time interval [0, tf] for a specific Ωtg. We use the Hamiltonian in Supplementary Note 1 to numerically compute PD(ti), as each Pi does not follow the expression \({P}_{D}({t}_{i})={\cos }^{2}(\pi {t}_{i}/{t}_{{{{\rm{R}}}}})\) in cases that depart from the standard working regime. In addition, the binary outcome is drawn from a Bernoulli distribution \({z}_{n}^{i} \sim B(1,{P}_{D}({t}_{i}))\in \{0,1\}\) such that each simulated average value is \({P}_{i}=\mathop{\sum }\nolimits_{n = 1}^{{N}_{m}}{z}_{n}^{i}/{N}_{m}\) for a number of shots Nm. We choose Np = 151, Nm = 100 and tf = 2.828 ms (note that this value of tf corresponds to one period of the sensor response for Ωtg = (2π) × 0.5 kHz, i.e., in the harmonic case). The examples (i.e., the data strings X, Y, and A) are computed by selecting 96 values for Ωtg/(2π) in the range [0.5–10] kHz. In addition, as each Pi fluctuates owing to the reduced number of measurements, we perform 100 repetitions for each simulated experimental acquisition (i.e., for each Ωtg). Therefore, our dataset contains 96 × 100 examples of which 70%/15%/15% lead to the training/validation/test datasets. After training the NN (details are available in Methods and Supplementary Note 3) we can estimate the Rabi frequency Ωtg of experimentally collected responses by inputting them into the NN.

In particular, we harvest sensor responses for N = 15 values of Ωtg (while we select other RF parameters as ϕtg = 0 and ωtg = (2π) × 10.56 MHz) which do not belong to the training/validation/test datasets, and for a number of shots Nm = 100, and Nm = 30. In Table 1, we list the estimations y1 from the NN for each target a1 ≡ Ωtg. Each output y1 is obtained after feeding the NN with the experimentally acquired response consisting of average values measured during t [0, tf] with Np = 151. We show the results for a number of shots Nm = 100 and Nm = 30. The regressions of the NN outputs with respect to Ωtg are shown in Fig. 2a, b.

Table 1 Outputs y1 obtained from the NN in Scenario i.
Fig. 2: Estimation results from Scenario i, a reduced number of measurements.
figure 2

a, b Regression of the NN outputs (y1) with respect to the targets a1 ≡ Ωtg for a number of shots Nm = 100 (a) and Nm = 30 (b). The fit (solid-blue) overlaps with the line y1 = a1, while R is the correlation coefficient30 of y1 and a1. This is R = 0.9996 (a), and 0.9994 (b). ce Sensor response (diamonds) for the targets Ωtg = (2π) × 1.1487 kHz (c) and Ωtg = (2π) × 8.9493 kHz (d, e) experimentally obtained for Nm = 30 shots. For comparison, in (c, d) the harmonic response \({P}_{D}(t)={\cos }^{2}(\pi t/{t}_{{{{\rm{R}}}}})\) (solid-blue) is included. The plot in c shows the case with Ωtg = (2π) × 1.1487 kHz lying in the harmonic regime, while in (d) the response for Ωtg = (2π) × 8.9493 kHz significantly deviates from the harmonic behavior. In e, we observe that the curve obtained from numerical simulations (solid-red) fits experimental data for Ωtg = (2π) × 8.9493 kHz. Error bars in ce represent the standard error of the mean.

The average accuracy in the estimation of the different Ωtg is defined as \(\bar{F}=\frac{1}{N}\mathop{\sum }\nolimits_{j = 1}^{N}{F}_{j}\), with \({F}_{j}=1-| {y}_{1}^{j}-{a}_{1}^{j}| /{a}_{1}^{j}\), and a1 = Ωtg. With the values in Table 1 we find the results \(\bar{F}=98.76 \%\) for Nm = 100 with a standard deviation (SD) of the Fj set SD = 0.7762%. In the case of Nm = 30, we find \(\bar{F}=98.31 \%\) with SD = 1.1483%. We remark that these highly accurate estimations were obtained with examples that comprise values of Ωtg leading to sensor responses that depart from the harmonic case. In particular, in Fig. 2 (c,d,e) we show the cases for Ωtg = (2π) × 1.1487 kHz (c) and Ωtg = (2π) × 8.9493 kHz (d, e), leading to harmonic and non-harmonic responses respectively (both cases comprise Nm = 30). On the one hand, in Fig. 2c, we show that the sensor response (diamonds) follows the harmonic function \({P}_{D}(t)={\cos }^{2}(\pi t/{t}_{{{{\rm{R}}}}})\) (blue-solid line). On the other hand, in Fig. 2d, one can observe that the response (diamonds) deviates from the harmonic case, while Fig. 2e shows that the same response fits to a non-harmonic function (red-solid curve) obtained via numerical simulations of the Hamiltonian (see Supplementary Note 1). By introducing these two responses into our NN, we get the outputs y1 = (2π) × 1.1731 kHz and y1 = (2π) × 8.7825 kHz which result in large accuracies F = y1 − a1/a1 = 97.84%, 98.14%, respectively. In Supplementary Note 4, we show further examples including both Rabi frequency Ωtg and potential detunings ξ between the target frequency and the sensor hyperfine transition.

Scenario ii: Continuous data acquisition

Now, we consider a scenario that comprises single-shot measurements on the ion (i.e., Nm = 1). This scheme is relevant in situations where no reinitialization of the RF field is possible. We demonstrate that the estimation of RF parameters is still feasible in these conditions.

We get one binary value from the atomic sensor (0 or 1) after completing the three experimental stages illustrated in Fig. 3a. More specifically, at stage 1 we cool the ion and prepare the \(\left\vert D\right\rangle\) state in a time t1 = 6.097 ms. At stage 2 the ion is allowed to interact with the target field for a time t2 [0, 6] ms. At stage 3, we readout the ion in a time t3 = 2.447 ms. We repeat these three stages 251 times (leading to a binary string comprising 251 numbers) where t2 varies in the interval [0, 6] ms with a step 0.024 ms. Note that during the three stages the RF source is always on, thus to avoid potential damage on initialization and readout we switch the RF signal to a dummy load at stages 1 and 3.

Fig. 3: Estimation results from Scenario ii, continuous data acquisition.
figure 3

a Schematic configuration of the continuous data acquisition scheme. Each binary value is obtained after completing the preparation, interaction, and measurement stages. bd Regression of the rescaled outputs \({y}_{1}^{r}\) from the NN with respect to the targets \({a}_{1}^{r}={\Omega }_{{{{\rm{tg}}}}}^{r}\). The regression lines \({y}_{1}^{r}=\alpha {a}_{1}^{r}+\beta\) comprise b α = 0.9946, β = 0.0032, c α = 0.9932, β = 0.0040, d α = 0.9925, β = 0.0041, while the correlation coefficients R between the outputs and the targets are all larger than 0.999. All inputs Xr, outputs \({{{{\bf{Y}}}}}^{r}=\{{y}_{1}^{r}\}\), and targets \({{{{\bf{A}}}}}^{r}=\{{a}_{1}^{r}\}\) are rescaled into the range [0, 1].

Via numerical simulations, we train a NN in accordance with the scheme in Fig. 3a. In particular, we use 96 values for Ωtg in the range (2π) × [0.5, 10] kHz (with a step of (2π) × 0.1 kHz) and repeat the data acquisition process 1800 times for each Ωtg. Thus we generate 96 × 1800 = 172,800 examples, of which 70%/15%/15% are used to build the training/validation/test datasets. After training the NN, we find the regression accuracy of the training/validation/test datasets shown in Fig. 3b–d. Note that data in Fig. 3b–d is rescaled into [0, 1] as this is a standardized procedure in NNs.

Now, we experimentally obtain the sensor responses for 8 randomly chosen values of Ωtg in the range (2π) × [0.5, 10] kHz, which do not belong to the training/validation/test datasets (in addition, we select other RF parameters as ϕtg = 0 and ωtg = (2π) × 10.03 MHz). We remark that the obtained responses range from the harmonic shape to those deviating from it. After 251 measurements for each Ωtg, we get binary strings including 251 numbers (0 or 1), where each number is obtained according to the scheme in Fig. 3a. When inputting each string into the trained NN, we get one output y1 from the NN. In order to study the stability of the NN prediction, we have repeated 20 times the data acquisition for each Ωtg. In Table 2, we show the average value \({\bar{y}}_{1}\) of the results from the NN and the SD based on 20 experimentally obtained strings.

Table 2 Estimation results from the NN in Scenario ii.

In addition, the regression of the outputs \({\bar{y}}_{1}\) with respect to the targets is in Fig. 4a. Finally, in Fig. 4b, we illustrate for a1 = (2π) × 2.1572 kHz the histogram of the NN outputs y1 obtained after feeding the NN with 20 strings of experimental data. The above analysis illustrates the ability of NNs to achieve accurate estimations in scenarios involving single-shot measurements (thus, when the RF field is not controllable), leading to highly versatile quantum sensors.

Fig. 4: Accuracy analysis for continuous data acquisition.
figure 4

a Regression of \({\bar{y}}_{1}\) obtained from the NN with respect to each target. In addition, the fit line (blue) and the correlation coefficient R are shown. b Histogram of 20 outputs y1 from the NN for the case a1 = (2π) × 2.1572 kHz denoted with a vertical red line.

Discussion

One can resort to other estimators for predicting parameters, e.g., using Bayesian inference. Following the well-known Bayes theorem, one may compute the posterior distribution pX) p(XΘ)p(Θ) where p(Θ), p(XΘ) denote the prior and likelihood, respectively, while X refers to the data obtained by interrogating the quantum sensor at different time instances, and Θ = {θ1, . . . , θk} denotes k unknown parameters which we aim to estimate by our quantum sensor. For a Bayesian estimator, an accurate microscopic model is needed in order to calculate the likelihood p(XΘ). More specifically, this is

$$\begin{array}{rcl}&&p({{{\bf{X}}}}| \Theta )={\Pi }_{i = 1}^{{N}_{p}}f({X}_{i},{N}_{m},{\tilde{P}}_{i}({t}_{i};\Theta ))\\ &{{{\rm{with}}}}&f(x,n,p)=\frac{n!}{x!(n-x)!{p}^{x}{(1-p)}^{(n-x)}},\end{array}$$
(2)

where the function f(x, n, p) refers to the probability of observing x success outcomes through n trials from the Binomial distribution with success probability p. \({\tilde{P}}_{i}({t}_{i};\Theta )\) is the survival probability PD computed using the total Hamiltonian H (see Supplementary Note 1: Eq. (3)) at time ti whose obtention requires to load a string of values for the Θ parameters in the microscopic model and then compute its dynamical evolution. Note this is a procedure that has to be repeated for each value of the Θ parameters. Finally, via the marginal distribution p(θjX) = ∫ΠijdθipX), one could derive the average value and SD as

$$\begin{array}{c}{\theta }_{j}^{{{{\rm{est}}}}}=\int\,d{\theta }_{j}{\theta }_{j}p({\theta }_{j}| {{{\bf{X}}}}),\\ {(\delta {\theta }_{j}^{{{{\rm{est}}}}})}^{2}=\int\,d{\theta }_{j}{({\theta }_{j}-{\theta }_{j}^{{{{\rm{est}}}}})}^{2}p({\theta }_{j}| {{{\bf{X}}}}).\end{array}$$
(3)

In contrast, less prior knowledge of the microscopic model is needed when using NNs. This owes to NNs learning the input–output relation from the training/validation/test datasets that can be obtained from numerical simulations, or directly from experiments. The latter is especially useful when numerical simulation of the system dynamics becomes challenging (for instance, in a sensor that consists of several entangled ions).

In this manner, our work demonstrates the good performance in parameter estimation that results from an appropriate hybridization of ML tools with quantum sensing techniques. This is a strategy that can be easily extended to other quantum platforms, such as, e.g., nitrogen-vacancy (NV) centers in diamond, to decipher complex NV responses emerging from dense nuclear samples comprising nuclear spins which are strongly coupled to the sensor and/or among them.

Methods

NN-based magnetometer

A NN enables to find the relation between Np measured data \({{{\bf{X}}}}=\{{x}_{1},{x}_{2},...,{x}_{{N}_{p}}\}\), and n output data Y = {y1, y2, . . . yn} that approach the targets A = {a1, a2, . . . , an}. During the training stage of the NN, the following cost function

$$C=\frac{1}{nN}\mathop{\sum }\limits_{j=1}^{N}\mathop{\sum }\limits_{i=1}^{n}{({y}_{i}^{j}-{a}_{i}^{j})}^{2}$$
(4)

is minimized for a training set that comprises N examples. This is done by using gradient descent methods such that the NN parameters (i.e., weights and biases) are adjusted to satisfy F(X) = Y ≈ A.

In our case, we deal with an 171Yb+-magnetometer where we aim to estimate RF parameters from experimentally collected responses by inputting them into the NN. The input data contains, in Scenario i, average values obtained from a reduced number of measurements and, in Scenario ii, a sequence containing binary values continuously acquired from single-shot measurements. For the first case, the input data string \({{{\bf{X}}}}=\{{P}_{1},{P}_{2},...,{P}_{{N}_{p}}\}\) consist on the average values Pi (i [1, Np]) collected in a time interval [0, tf] for a specific set of targets A. The simulated average value is \({P}_{i}=\mathop{\sum }\nolimits_{n = 1}^{{N}_{m}}{z}_{n}^{i}/{N}_{m}\) for a number of shots Nm, where the binary outcome is drawn from a Bernoulli distribution \({z}_{n}^{i} \sim B(1,{P}_{D}({t}_{i}))\in \{0,1\}\). In the second scenario that comprises continuous data acquisition, the input data string X is made of binary numbers, 0 and 1, which are obtained according to the scheme in Fig. 3a. Repeating this procedure N times, we achieve the whole dataset that comprises N examples. In both cases, the examples with the data strings X, Y, and A are computed by selecting a number of values of the targets in, and beyond, the regime leading to harmonic sensor responses. Among all the examples of the total datasets, 70%/15%/15% lead to the training/validation/test sets. A number of repetitions for each data acquisition are repeated such that the NN learns the statistical fluctuations resulting from a reduced number of measurements.

Experimental timing sequence

As shown in Fig. 5, in each cycle (each period to obtain one value Pi (i [1, NP]) of a response / input data string), the ion is cooled down approximately in the Doppler limit by a red-detuning 369.5 nm laser starting from stage 1. After that, the state of the ion is initialized to \(\left\vert 0\right\rangle\) by an optical pumping process. After a 0.5 μs ‘MW trigger’ signal, a MW π-pulse resonant with \(\left\vert 0\right\rangle \leftrightarrow \left\vert 1\right\rangle\) transfers the state from \(\left\vert 0\right\rangle\) to \(\left\vert 1\right\rangle\). Subsequently, the STIRAP pulses drive the system from the state \(\left\vert 1\right\rangle\) to the dark state \(\left\vert D\right\rangle\). At stage 2, the amplitudes of MW fields are held at a constant Ω within the fixed time interval [0, tD]. Simultaneously, the target RF field is applied for the time interval [0, t2], where t2 ≤ tD. For the scheme of a finite number of measurements (Scenario i), after the time instant t2, the RF field is removed by turning off the AWG2 output. The RF target field is restarted at the next cycle such that the repetition of measurement can be done. For the scheme of single-shot measurement, Scenario ii, the RF signal is switched to a dummy load after the time instant t2 while the origin RF source is always on. At stage 3, the STIRAP pulses transfer the rest of population of \(\left\vert D\right\rangle\) back to \(\left\vert 1\right\rangle\). Another π-pulse transfers the population of state \(\left\vert 1\right\rangle\) to \(\left\vert 0\right\rangle\). Thus, we can get the response of \(\left\vert D\right\rangle\) by measuring the probability of \(\left\vert 0\right\rangle\), and a state dependent fluorescence detection could be used to determine it. Before and after the cycle, there are extra ‘idle’ steps (1 μs) needed to end this cycle and start the next cycle. In our experiments, the timings are all controlled by a TTL pulse generator (Spincore PB24-100-4k-PCIe) see Table 3 for the specific values of the time intervals invested in each process.

Fig. 5: The sequence diagram of the experiment.
figure 5

The whole cycle could be divided into three stages. Stage 1 consists of cooling, initialization, and preparation of the state \(\left\vert 1\right\rangle\) (the first blue zone) and the state \(\left\vert D\right\rangle\) (the first green zone). At Stage 2 (the pink zone of length tD), we allow the ion interacts with the target field at the interval [0, t2] (the yellow zone), where t2 ≤ tD. In stage 3, we transfer the state \(\left\vert D\right\rangle\) back to \(\left\vert 0\right\rangle\) and detect the probability of \(\left\vert 0\right\rangle\).

Table 3 Experimental timings for two scenarios (‘Idle’ and ‘MW trigger’ are not shown).