Brought to you by:
Paper

Systematic investigations on ion dynamics with noises in Paul trap

, , and

Published 25 October 2023 © 2023 IOP Publishing Ltd
, , Citation Ying-Xiang Wang et al 2023 J. Phys. A: Math. Theor. 56 465302 DOI 10.1088/1751-8121/ad0348

1751-8121/56/46/465302

Abstract

Ions confined in a Paul trap serve as crucial platforms in various research fields, including quantum computing and precision spectroscopy. However, the ion dynamics is inevitably influenced by different types of noise, which require accurate computations and general analytical analysis to facilitate diverse applications based on trapped ions with white or colored noise. In the present work, we investigate the motion of ions in a Paul trap via the Langevin equation using both analytical and numerical methods, systematically studying three different types of noise: the white noise, the colored noise via the Ornstein–Uhlenbeck process and the Wiener process. For the white noise of the case, we provide a recursion method to calculate ion motion for a wide range of parameters. Furthermore, we present an analytical solution to the more realistic stochastic process associated with the colored noise, verified by the Monte Carlo simulation. By comparing the results of the colored noise with those of the white noise, and additionally considering another limit of noise parameters corresponding to the Wiener process, we summarize the effects of different noise types on the ion dynamics.

Export citation and abstract BibTeX RIS

1. Introduction

Trapped ions in a Paul trap experience various types of external noise that can affect their dynamics, making the accurate calculation and general analytical analysis of the ion motion crucial for diverse applications. Due to their controllability and long-range interactions [1], the dynamics of trapped ions have been widely studied, enabling them as useful platforms in various fields such as the precision spectroscopy [2], the atomic clock [3], and the quantum computing [46].

Micromotion near the trapped location requires consideration of excess micromotion, such as second-order Doppler and Stark shifts [7]. To achieve a high accuracy, stability parameters of the Paul trap can be computed via an iterative method using approximate solutions [8]. Monitoring and maintaining the stability of ion motion is essential for a continued system operation. External noise and the ion's micromotion can impact the quality of the trapped ions and its subsequent applications, such as the fidelity of quantum gates [911] in the quantum computing [1215]. Recent studies have re-emphasised the importance of studying the effects of noise for the quantum computing [16].

In order to precisely calculate the micromotion of ions and the effect of noise, the ponderomotive-type approximation has been widely used in previous studies [17]. Moreover, the classical work of Savard et al [18, 19] pointed out the effect of the laser intensity and beam-pointing noise on the threshold of the heating time and the lifetime of the trap. Lu et al [20] also discussed the effect of three typical noises (the phase noise, the amplitude noise, and the wavenumber noise) on the trap and relate them to the Ornstein–Uhlenbeck (OU) noise.

In addition, for the hybrid ion–atom systems [21], one needs to study the kinetic energy of a single ion in the buffer gas and ionic energy distribution [2225], which is much more complex for approximate solutions. DeVoe [26] considered the case in which trapped ions interact with a buffer gas at a finite temperature via hard-sphere collisions. By using Monte Carlo methods, it was found that, after experiencing hard spherical collisions with atoms, the ions will form a non-Gaussian energy distribution with a power-law tail [26]. This type of the non-equilibrium sympathetic cooling has been experimentally demonstrated, e.g. by using the trapped $\mathrm{Sr}^+$ ion immersed in a cloud of Rb atoms [2730].

Mathematically, a single trapped ion is influenced by the noise via the Langevin equation. There exist various methods that theoretically solve this equation, including the Fokker–Planck formalism [31], theWenzel-Kramers-Brillouin (WKB) approximation [32], and the Green's function method [33]. These methods can give approximate solutions and have been verified by numerical methods within certain parameter ranges. Usually, one considers the Langevin equation which has a Gaussian noise with a memory effect. Compared with the white noise, the colored noise is believed to reflect a more realistic stochastic and physical process, which is characterized by the OU process [34, 35]. Previous studies have investigated the effect of the colored noise compared to the white noise on the ion motion [36].

In this work, we investigate the motion of ions in a Paul trap using the Langevin equation through analytical and numerical methods. Specifically, we provide a recursion method for calculating the motion of the ion in a large range of parameters in the case of white noise. Additionally, we derive an analytical solution for the OU process, a more realistic stochastic process associated with colored noise, which we verify using Monte Carlo simulations. We compare the results of the colored noise with those of the white noise and consider the limit of noise parameters that leads to the Wiener process. Finally, we analyze the effects of different types of noise with varying orders of continuity on the ion's dynamics.

The rest of paper is organized as follows. In section 2, we present our model based on the Langevin equation. In section 3, we discuss the case of the white noise with both the recursion method and the WKB method. We show that the recursion method can be used efficiently over a wide range of parameters. In section 4, we consider a more realistic and complex situation in which the noise is the OU-process or the Wiener process by the WKB method. We obtain an analytical solution that explicitly reveals the relationship between the ion motion and parameters, highlighting the different effects on the ion motion under various parameters and discussing the connection between the two types of noise under specific parameters. In addition, we analytically examine the ion motion in the limit of some parameters and demonstrate the effects of the parameter change by numerical simulations. Finally, in section 5, we summarize our main results and present some perspectives.

2. Theoretical model

For an ion with charge Q and mass m in the Paul trap, the position r of the ion satisfies the Mathieu equation:

Equation (1)

where Ω is the trap frequency, parameters a and q are related to the dc/ac voltages applied to the trap. The uniform static electric field E can represent the phase difference of the electrodes or the external electric field from other ions when one considers a chain of ions. The field E is usually small and can lead to the excess micromotion of the ion. Using the Floquet's theorem, one can solve the Mathieu equation and calculate the eigenvalues, which can show the motion stability of the ion [37].

In this work, we consider the case of one-dimensional motion along the x-axis. Without loss of generality, we consider both an additional damping term $\gamma \dot{x}$ and a time-dependent noise term F(t). Moreover, we mention that the addition of the two terms can describe various situations, and we provide a possible experimental example in appendix A. Furthermore, we emphasize that this model does not only apply to the specific experimental situation mentioned, also our results may be suitable to any case of the colored noise, which mathematically satisfies our description to the noise. For the one-dimensional case, the equation of the ion motion is thus reduced to the Langevin equation:

Equation (2)

Please note that the noise may have various sources, but usually one assumes its ensemble average '$\langle \rangle $' and variance should satisfy:

Equation (3)

Equation (4)

which means that the noise has a nonvanishing variance but a zero mean value. With equation (3), one actually has neglected the small uniform static electric field in equation (1).

The different forms of f(t) in equation (4) will represent different kinds of noise. We consider two types of noise in this work. Firstly, the white noise, where the stochastic force is modeled as the Gaussian white noise with a constant intensity D, and secondly, the colored noise, where the stochastic force is described by the OU process, which is a Gaussian stationary stochastic process that is widely used to model the dynamics of a particle in a thermal bath.

Our goal is to investigate the effect of the damping coefficient and the noise intensity on the motion of the ion in the Paul trap, and to compare the results obtained from analytical and numerical methods for both types of noise. In the next section, we will discuss the case of the white noise using both the recursion method and the WKB method.

3. White noise

Let us begin to discuss the case where the stochastic force is induced by the white noise, namely:

Equation (5)

where $\delta(x)$ represents the Dirac delta function and D is the strength of the noise. It shows that the stochastic force is absolutely independent of the force at different time and can be simulated using a normal distribution with a mean value of 0 and a variance of D.

3.1. Variance in the long time limit by recursion method

In the long time limit, the mean value of the ion position x and the speed v must be zero. Following [31, 36, 38], we choose to expand the stochastic average of the variance in terms of Fourier series basis set, i.e.

Equation (6)

The Langevin equation can be then transformed to the Fokker–Planck equation [39]. Specifically, equation (2) turns out to be

Equation (7)

in which

Equation (8)

Equation (9)

Equation (10)

In order to further simplify equation (7), we introduce new variables $S^{\pm}_n$, which satisfy:

Equation (11)

Equation (12)

Therefore, we can arrive at a recursion relation of $S^{\pm}_n$:

Equation (13)

and, in turn, the variance $\left\langle x^2 \right\rangle$, after the averaging over time, can be shown to be

Equation (14)

where the time-averaged mean kinetic energy is given by

Equation (15)

Now, the calculations of $\overline{\langle x^2\rangle}$ and $\overline{\langle v^2\rangle}$ can be carried out from equations (14) and (15), which requires us to know the value of $S_0^{+}$ beforehand. It is clear that, for given values of parameters a, q, and γ, $S_0^{+}$ should take a certain specific value since the values of $\overline{\langle x^2\rangle}$ and $\overline{\langle v^2\rangle}$ would otherwise be different for the same set of parameters according to equations (14) and (15). This will in turn lead to different results for the same physical situation of the ion motion. In other words, a fixed value of $S_0^{+}$ should be associated with the trap parameters.

Therefore, one can use the recursion equation (13) to evaluate $S_0^{+}$ with a wide range of parameters by using the numerical method. To do this, we choose an integer $n_{\max}$ as the times of the backward recursion and a random number $S_{n_{\max}}^{+}$ as the starting value of the recursion. By using equation (13), we can calculate the value of any value of $S_n^{+}$ for which $ 0 \unicode{x2A7D} n \lt n_{\max}$, including the particular $S_0^{+}$ that we aim at. It is worth noting that it is unnecessary to calculate the cases for n < 0 because the case of $n = -k$ is exactly equivalent due to $x_k = x_{-k}^*$ if one considers equation (13) for n = k. Hence we replace $S_n^{+}$ with Sn in the following text.

To validate this method, one should check whether it is possible to obtain the same value of S0 when $n_{\max}$ and $S_{n_{\max}}$ are varied. In the recursion method, S0 is the recursion ending value, while $n_{\max}$ and $S_{n_{\max}}$ is the times and the starting value for the recursion, respectively. The results are shown in figure 1, in which we have set $\Gamma = \frac{\gamma}{\Omega}$ as a dimensionless parameter for the damping coefficient. As can be seen, by varying $n_{\max}$ and $S_{n_{\max}}$, it is indeed possible to obtain the same value of S0 when $n_{\max} \gt 3$ (clearly, values of $n_{\max}$ that are too small cannot be used). In practice, an intermediate value of $n_{\max} = 10$ can be chosen, since larger values are usually not necessary and take longer time to compute. Meanwhile, we find that the choice of $S_{n_{\max}}$ is not important, thus setting $S_{n_{\max}} = 1$ will be sufficient. It seems that different values of $S_{n_{\max}}$ corresponds to the same S0, but we can find the recursion for n from smaller to bigger is rather unstable and only some particular values of S0 can make the whole series converge.

Figure 1.

Figure 1. The recursion ending value S0 calculated by different recursion times $n_{\max}$ and starting values $S_{n_{\max}}$: (a) the real part, (b) the imaginary part. We have taken the trap parameters under a = 0.2, q = 0.8, and $\Gamma = 0.1$. All the curves quickly converge to the same value as $n_{\max}$ increases. The inserts show the convergence speed by showing the logarithmic difference between S0 and its recursion ending value calculated under $n_{\max} = 100$.

Standard image High-resolution image

The above method can produce a converged value of S0 under various parameters a, q, and Γ, but one has to make sure this method can indeed give us the right result. To check whether the S0 calculated by this method is correct, for a given S0, we can further use equations (14) and (15) to calculate $\overline{\langle x^2\rangle}$ and $\overline{\langle v^2\rangle}$ at a wide range of parameters and compare our results with those in literature.

In figure 2, we show the average of the kinetic energy as a function of q for various values of a. We observe curved upward-opening lines, which indicate a rapid divergence of the kinetic energy that can even result in negative values for certain parameters. Consequently, it can be deduced that the ion motion is unstable in during these circumstances. It is easy to find the parameter boundary which corresponds to a finite value by the numerical method. Within the boundaries, the ion motion can be stable and the Mathieu's equation has positive and finite eigenvalues, as shown in figure 3, which agrees with the results of the Mathieu's equation calculated by the conventional methods, such as numerically solving differential equations or using series expansions to solve the eigenvalue problem [40]. When γ increases, the range of the stable area also increases, which implies that the ion motion can be stable more easily. This is consistent with the physical picture. It should be noted that our proposed method is much faster and significantly more accurate than conventional methods [37].

Figure 2.

Figure 2. The average of the variance of speed $\langle \overline{v^2} \rangle$ (represents kinetic energy) under different potential parameters of a and q for a fixed damping coefficient of $\Gamma = 0.1$. Lines of different style represent different parameters of a. For a specific value of a, the kinetic energy tends to be infinite beyond some values of q. A positive and finite the kinetic energy means that the motion of the ion is stable.

Standard image High-resolution image
Figure 3.

Figure 3. The kinetic energy is nearly infinite for some combinations of potential parameters of a and q. Under parameters inside the curve, the motion of the ion is stable. The blue and black lines represent $\Gamma = 10^{-1}$ and 10−5 respectively. The red dots are the results calculated by the conventional method. The choices of the parameters in the rest of this work will be inside the stable region.

Standard image High-resolution image

3.2. Numerical verification to the recursion method

The most convincing and direct way to verify our recursion method is to perform numerical simulations by the molecular dynamics and compare the results with those from equations (14) and (15). To do this, we first convert the equation of motion of the ion into a homogeneous system of first-order differential equations. We introduce the dimensionless parameters:

Equation (16)

Equation (17)

Equation (18)

with l0 being a constant with the dimension of length. There may be a multiplying constant in the noise, but we ignore it and only consider the relative value. We use the difference of the Wiener process to mathematically describe the white noise . With WT representing the Wiener process [41], we have

Equation (19)

Equation (20)

We can use the fourth-order Runge–Kutta (R–K) method to integrate the differential equations. In practice, we compute the ion motion 1000 times with different random forces and add them up for the statistical average of x2 and v2. The implementation of the random force relies on the Python's random number selection. The force satisfies a normal distribution and applies to each ion independently (see appendix B). The results of $\langle x^2 \rangle$ are shown in figure 4.

Figure 4.

Figure 4. The variance of the position $\langle x^2\rangle$ as a function of T, computed by the numerical simulation for potential parameters $a = -0.2$, q = 0.8, the damping coefficient $\Gamma = 0.01$, and the noise strength $D^{\prime} = 0.1$. The insert enlarges the oscillatory feature of the ion motion.

Standard image High-resolution image

To verify our recursion method in the last subsection, we need to consider the limit of the long time and include both the statistical averaging and the time averaging. We need to choose a time range large enough for different parameters. The results are shown in figure 5, from which we can conclude that the results from equations (14) and (15) agree completely with those calculated by the numerical simulations. The small deviations from the numerical simulations are attributed to the fact that we cannot simulate an infinite time for the time averaging. These comparisons show that our recursion method to calculate S0 is undoubtedly correct. It is noteworthy that $\langle x^2\rangle$ starts to increase when the damping parameter Γ is larger than a specific value around 0.6. This can be attributed to the fluctuation–dissipation theorem, which indicates that the fluctuation is proportional to the dissipation. It should be noted that the application of the WKB method below will give similar results in the case of a large Γ.

Figure 5.

Figure 5. (a) The average of the variance of the position $\overline{\langle x^2\rangle}$; (b) The ratio of the average of the variance of the position and the kinetic energy $ \overline{\langle x^2\rangle}/\overline{\langle v^2\rangle}$. The results are calculated by two different methods as a function of the damping coefficient Γ. The parameters have been taken as potential parameters $a = -0.2$, q = 1, and noise strength $D^{\prime} = 0.1$. In (a), the results have been normalized to 1 at the damping coefficient $\Gamma = 0$.

Standard image High-resolution image

The above discussions mainly focus on S0 and the average of the variance. Actually, we can also verify equations (13) for other terms in the Fourier series of Sn , and the time evolution of x2 can be also examined. When one calculates S0 by the recursion method, at the same time we obtain Sn (obviously, for a given n, we should take a large enough $n_\textrm{{max}}$ to ensure a converged Sn ). Meanwhile, in the long time limit, $\langle x^2\rangle$ has the following Fourier decomposition,

Equation (21)

where $X_n = x_n+x_n^* = 2x_{n}^{r}$ ($x_{n}^{r}$ is the real part of xn ) due to the fact that $\langle x^2 \rangle$ is a real number (cf equation (6)). As an example, we verify the factors of the cosine terms. According to equations (11) and (12), we have

Equation (22)

Following equation (21), one can make a discrete Fourier transform to $\langle x^2 \rangle$ in the long time limit:

Equation (23)

As x0 has been proven to be accurate and reliable, we can compare $2x_n/x_0$ using the recursion method and $X_n/x_0$ using the numerical method. The results are listed in table 1, which corresponds to the parameters given in the caption. Note that the errors arising from the fact that the discrete Fourier transform itself has a relative error of about $(\Delta t)^2\sim10^{-4}$ since we have chosen $\Delta t = 0.01$. As can be seen from table 1, the difference between the recursion method and the numerical simulation is around 10−4. One does not need to further verify the case for a bigger $n~(n\gt4)$ since $X_n/x_0$ will be smaller than this error limit.

Table 1. Fourier series of different orders for trap parameters $a = -0.2$ and q = 0.8, where two cases of the damping coefficient $\Gamma = 0.01$ and 0.1 have been considered. For each case, the first row is for the theoretical method and the second row for the numerical method a total time T = 1000 and a stepsize $ \Delta T = 0.01$. The data are directly shown in the table with the same number of valid digits, regardless of the error.

Γ n 01234
  ${2x_{n}^{r}}/{x_0}$ 1.0000−3.9000 $\times 10^{-1}$ 5.6683 $\times 10^{-2}$ −4.1249 $\times 10^{-3}$ 1.7897 $\times 10^{-4}$
0.01 ${X_n}/{x_0}$ 1.0000−3.9003 $\times 10^{-1}$ 5.6694 $\times 10^{-2}$ −4.0472 $\times 10^{-3}$ 2.6688 $\times 10^{-4}$
 Δ03 $\times 10^{-5}$ 1 $\times 10^{-5}$ 8 $\times 10^{-5}$ 9 $\times 10^{-5}$
  ${2x_{n}^{r}}/{x_0}$ 1.0000−3.8372 $\times 10^{-1}$ 5.4872 $\times 10^{-2}$ −3.9330 $\times 10^{-3}$ 1.6782 $\times 10^{-4}$
0.1 ${X_n}/{x_0}$ 1.0000−3.8389 $\times 10^{-1}$ 5.4880 $\times 10^{-2}$ −3.8444 $\times 10^{-3}$ 2.6091 $\times 10^{-4}$
 Δ017 $\times 10^{-5}$ 1 $\times 10^{-5}$ 9 $\times 10^{-5}$ 10 $\times 10^{-5}$

3.3. WKB method

As was recently proposed by Conangla et al [32], one can apply the WKB method to calculate the variance of the position x(t). The advantage of the WKB method is that one can examine the ion motion at any timescale, not restricted to the case of the long time limit. Furthermore, the analytical expressions can tell us how the ion motion will change in response to variations in the trap and noise parameters. In particular, we consider the overdamped situation where the dc term is neglected, i.e. one takes a = 0 in equation (2).

To simplify the subsequent derivation, we denote $T = \Omega t$, $\Gamma = \frac{\gamma}{\Omega}$, $D^{\prime} = \sigma^2$, and $\lambda_1 = -\frac{q^2}{8\Gamma^3}$. Following [32], at time T, one can obtain an approximate solution to equation (2):

Equation (24)

According to the Itô isometry [42], the variance of the approximate solution equals to

Equation (25)

where E stands for the ensemble average, with the same meaning of the symbol '$\langle \rangle$'.

Now we can consider two limiting cases. Firstly, in the long time limit, one has

Equation (26)

which indicates that the variance is a constant Y0 independent on T. Secondly, in the limit of $ T \rightarrow 0$, one gets

Equation (27)

According to equation (26), one immediately arrives at the conclusion that $E\left[x(T)^2\right]$ has an upper limit Y0 which is proportional to the damping parameter Γ. This fact can be verified by the recursion method discussed in the previous subsections. In doing this, for the overdamped case where a = 0, we calculate $E\left[x(T)^2\right]$ at different Γ and for various values of q. The results are presented in figure 6, which clearly shows that, above a certain value of Γ, $E\left[x(T)^2\right]$ increases as Γ grows, which are consistent with our assumptions (overdamped). In particular, for all the curves, $E\left[x(T)^2\right]$ indeed shows a linear relationship with Γ when it is sufficiently large. At this point, we have further confirmed the correctness of our recursion method. In addition, we mention that the WKB method can be also used in the case of the colored noise, as ill be shown below.

Figure 6.

Figure 6. For the potential parameter a = 0, the average of the variance of the position $\overline{\langle x^2\rangle}$ is shown as a function of the damping coefficient Γ for different values of q, related to the ac voltage.

Standard image High-resolution image

4. Colored noise

In the previous section, we discussed the white noise. The ideal white noise is easy to be simulated through the theoretical sampling, but the colored noises are more realistic in most circumstances. In practice, the noise that appears in the engineering and the measurement data is usually colored noise. Therefore, it is more important to discuss the effects of the colored noise on the dynamics of the trapped ion.

In this section, we will turn to examine the colored noise, which in some cases is equivalent to the OU process. We will consider two cases for different parameter scales of the OU-process, in which the noises have different properties. In one case, we provide an analytical solution to show the relationship between the ion motion and the parameters. In the other case, we focus on the Wiener process, which has a higher order of continuity than the white noise and interconnects them.

4.1. The case of OU process

The colored noise is no longer Markovian. The type of noise chosen in the current work is a classical one with a Gaussian correlation function, i.e.

Equation (28)

where $\lambda^{\prime}$ represents the rate of the mean reversion and σ is the noise intensity. In this case, the colored noise is equivalent to the OU-process with the stochastic differential equation (SDE) given by [34]

Equation (29)

in which Wt stands for the Wiener process and thus $\mathrm dW_t$ means a white noise.

We can now consider the influence of using different types of noises. In the overdamped case, if we ignore the difference of the constant term F(t) in equation (2), the function is the same as that discussed in the previous section. Thus, we can continue to use the WKB method, with appropriate changes. Following [32], we can get an approximate solution to x(T) in the integral form:

Equation (30)

To carry out the above integral, we make a slight change to equation (29), i.e.

Equation (31)

which can be inserted into equation (30), yielding

Equation (32)

where we have assumed $X_0 = 0$ because the starting of the OU-process is zero. Since now we have an equation having x(T) on both sides, we can make transposition and solve the dynamics of the ion with parameters and the noise function as follows:

Equation (33)

Then, we can calculate the variance $E\left[x(T)^2\right]$ by the statistical averaging:

Equation (34)

in which there are three terms. As has been shown in the previous section, the first term can be calculated by the Itô isometry [42] to be

Equation (35)

Remembering that $E[X_t] = 0$ is considered in this work, we can thus drop the second term $E_2(T)$. For the third term, we know that the variance of the OU-process is given by

Equation (36)

from which we consider the case when the noise becomes stable. In this case, since $\lambda^{\prime} \rightarrow +\infty$, one thus has $E[X_T^2] = \frac{\sigma^2}{2\lambda^{^{\prime}}}$. After substituting these results into equation (34), we finally arrive at

Equation (37)

Now, let us first consider the case of the long time limit, which also means $\left|\lambda_1\right| T \rightarrow +\infty$. Similarly, under these conditions, it is easy to show that $E\left[x(T)^2\right]$ can also reach a finite limit, i.e.

Equation (38)

We further assume $\lambda^{^{\prime}}T \gg |\lambda_1|T$, which means that the noise becomes stable much faster than the ion motion reaches the above limit. Thus, we finally get the following results:

Equation (39)

We approximate the analytical solution in order to show the relationship between the variance and the parameters distinctly, which can be verified by numerical simulations. In principle, we can still use the R–K method to integrate the motion of the ion, but we first need to generate the colored noise. We calculate the value of the colored noise by solving the SDE for the OU-process according to equation (29). Considering that the function itself contains the white noise, one judges that the error from the algorithm of the numerical solution to the SDE has little influence on our results. Hence, we choose to use the forward Euler method:

Equation (40)

where $\eta(T)$ is the white noise. For the convenience of calculations, we assume that $\eta(T)$ is independent of time and subject to a standard normal distribution. The stepsize $\Delta T$ is chosen to be the same as that we use to calculate the motion of the ion.

In figure 7, we demonstrate the value of the colored noise under two different values of rate of the mean reversion $\lambda^{\prime}$. One finds that for a larger $\lambda^{\prime}$, as shown in figure 7(a), the volume $\lambda^{\prime} T$ quickly rises, which means that the noise becomes to be stable so fast that we cannot observe any apparent change of the process over time. The behavior is very similar to the case of the white noise. On the contrary, when $\lambda^{\prime} T$ becomes small, as shown in figure 7(b), it resembles the Wiener process, as expected.

Figure 7.

Figure 7. Colored noise of the OU-process Xt calculated by the numerical simulation as a function time: (a) $\lambda^{\prime} = 1$; (b) $\lambda^{\prime} = 0.01$, where $\lambda^{\prime}$, as an important parameter for the colored noise, represents the rate of the mean reversion. The noise appears drastically different in which (a) looks like the white noise while (b) is similar to the Wiener process.

Standard image High-resolution image

Now, we consider the case where $\lambda^{\prime} $ is much larger than $\frac{1}{T}$. We first calculate the noise XT and supply it to the R–K integration algorithm to the following equations:

Equation (41)

Equation (42)

In figure 8(a), we show our result from the numerical calculation, which appears similar to the case of the white noise. This is not surprising, as can be seen from equation (39), one has a factor of $\frac{\sigma^2}{2 \Gamma^2\left|\lambda_1\right|}$, which is the same as the case of the white noise. The second factor $\frac{1}{\lambda^{^{\prime} 2}}$ only contributes a multiple constant. The last factor $\left[1-\frac{\lambda_1}{\lambda}+o\left(\frac{\lambda_1}{\lambda}\right)\right]$ approaches to 1 and thus does not largely change our result when $\lambda^{^{\prime}} \gg |\lambda_1|$. Overall, the result should resemble the case of the white noise and may only increase slightly due to the last term since $\lambda_1\lt0$.

Figure 8.

Figure 8. Different verification to our analytical solutions. (a) The variance of position for the ion $\langle x^2\rangle$ as a function of time when the noise has the rate of the mean reversion $\lambda^{\prime} = 1$ and the noise intensity σ = 0.1, where the trap potential parameters a = 0, q = 0.7, and the damping coefficient $\Gamma = 1$. The orange dashed line is the value calculated by equation (37). (b) The same with (a) for a = 0, q = 0.7, $\Gamma = 1$, and σ = 0.1. The blue, orange, green, red line respectively represents different rate of the mean reversion of the noise for $\lambda^{^{\prime}} = 1, 2, 3, 4$. (c) $\langle x^2 \rangle \times \lambda^{^{\prime} 2}$ as a function of time for a = 0, q = 0.7, $\Gamma = 1$, and σ = 0.1, verifying how our results is related to $\lambda^{\prime}$. (d) $\langle x^2 \rangle / \Gamma$ as a function of time with $\lambda^{\prime} = 1 $ for a = 0, q = 1, and σ = 0.1. The blue, orange, green, red line respectively represents different damping coefficient of $\Gamma = 5.0, 5.5, 6.0, 6.5$.

Standard image High-resolution image

In figure 8(b), we compare the results under various values of $\lambda^{\prime}$. As can be seen, one observes that the average of the variance of the ion position gradually becomes smaller when one increases the value of $\lambda^{\prime}$. To verify the effect of the factor $\frac{1}{\lambda^{^{\prime} 2}}$ in equation (39) for the case of $\lambda^{^{\prime}}\gg |\lambda_1|$, we calculate $\langle x^2 \rangle \times \lambda^{^{\prime} 2}$ as a function of time. As shown in figure 8(c), we find that the four curves tend to overlap with each other as $T\rightarrow \infty$, which confirms that our results for $\frac{1}{\lambda^{^{\prime} 2}}$ are reliable.

Then, we move to examine the motion of the ion under different damping strengths by changing the damping coefficient Γ. We have chosen four different values of Γ around 6.0 but with small differences, because the computation time for a converged result increases rapidly as Γ becomes large due to the fact that the volume is $\lambda_1 T$ with $\lambda_1 = -\frac{2q^2}{\Gamma^3}$. The results are shown in figure 8(d), from which we find that the four curves converge to similar values when $T\rightarrow \infty$. Again, please note that the time needed for the convergence increases with the increase of Γ. These observations are similar to those made for the case of the white noise.

Finally, we consider the case of $T\rightarrow 0$, and it is unsuitable to use equation (37) to make approximations because equation (36) may result in entirely different outcomes. Instead, we are able to simulate this case by choosing a smaller $\lambda^{\prime}$, and we find that the noise appears to be distinct, as shown in figure 7(b). Then, we calculate the variance of the position with $\lambda^{\prime} = 0.01$, and the result is shown in figure 9. Significantly, the behavior of the outcome is markedly different from those previously discussed, as different tendencies are observed in three distinct regions. Specifically, when $T\rightarrow \infty$, the curve indeed gradually converges, because the noise becomes stable so that the previous approximation remains valid. In the case of $T\rightarrow 0$, the variance $\langle x^2 \rangle$ increases nonlinearly, while it is approximately proportional to the time when T is in an intermediate range.

Figure 9.

Figure 9. The variance of position for the ion $\langle x^2\rangle$ as a function of time T when the noise has the rate of the mean reversion $\lambda^{\prime} = 0.01 $ and the noise intensity σ = 0.1, where the trap potential parameters a = 0, q = 0.1, and the damping coefficient $\Gamma = 0.5$. The results are computed from equation (42) using the R–K method for total time T = 1000 and stepsize $\mathrm dT = 0.05$.

Standard image High-resolution image

4.2. The case of Wiener process

Now, we will discuss the limit case of $\lambda^{\prime} \rightarrow 0$ for the OU-process, in which case figure 7(b) is found to be very similar to the Wiener process. Hence, we can make approximation to equation (26) and the noise becomes a Wiener process. In particular, we denote it as Wt [41]:

Equation (43)

We have used the inverse of the Wiener process to mathematically describe the white noise. Therefore, the Wiener process has a higher order of continuity than the white noise, which will simulate the case where the noise is disturbed and keeps being accumulated.

As the Wiener process is the limit case of the OU-process, we can find the ion motion using the method in the preceding subsection. Specifically, we use the WKB method and make similar procedures as those to equation (30). By integrating by parts, we have

Equation (44)

Again, we use the Itô isometry [42] to calculate the variance of the position of the ion:

Equation (45)

Apparently, by taking different limit of T, one can arrive at the following results:

Equation (46)

For the numerical simulation, we can calculate the Wiener process as we have done in the previous subsection. According to the SDE of the Wiener process, we can use the forward Euler method and arrive at

Equation (47)

The result is shown in figure 10, which is very similar to that in figure 7(b). Then, using the noise, we can calculate the motion of the ion by the R–K method for the following equations:

Equation (48)

Equation (49)

Figure 10.

Figure 10. The value of the Wiener process Xt as a function of time T with the noise intensity of the noise σ = 1, where we have taken T = 1000 and dT = 0.01.

Standard image High-resolution image

The result of $E\left[x(T)^2\right]$ is shown in figure 11, from which we find that the curve is partially similar to the front part of figure 9. It is easy to see that $E\left[x(T)^2\right]$ is proportional to the time when T is sufficiently large, but it endlessly increases over time and has no limit.

Figure 11.

Figure 11. The variance of position for the ion $\langle x^2 \rangle$ where the trap potential parameters a = 0, q = 0.1, and the damping coefficient $\Gamma = 0.1$. We have chosen the Wiener process using the noise intensity σ = 0.01.

Standard image High-resolution image

For $T \rightarrow 0$, $E\left[x(T)^2\right]$ increases very slowly and the curve appears to be flat. In order to examine the exponent over T when $T \rightarrow 0$, we calculate $\ln{\langle x^2 \rangle}$ as a function of $\ln{T}$. As shown in the insert of figure 12, the natural logarithm of $\langle x^2 \rangle$ is proportional to $\ln{T}$ when T is larger than 1 ($\ln{T}\gt0$). This confirms that $\langle x^2 \rangle$ has a power law relation with T. To be more visually clear, we can calculate the slope of the line. As can be seen from the main figure of figure 12, it is evident that the value of the slope is only slightly larger than 3, which is apparently due to the effect of higher power terms.

Figure 12.

Figure 12. Logarithmic relationship of physical quantities at smaller time. The insert shows $\ln{\langle x^2 \rangle}$ as a function of $\ln{T}$. The curve is linear when T is not too small. The main figure shows the slope for the curve in the insert without considering the case where T is small. The value of the slope is only slightly larger than 3, due to contributions of higher order terms.

Standard image High-resolution image

With the above discussions in mind, we can come back to the results of the OU-process as shown in figure 9. Obviously, the first and the second part of figure 9 agree with the case of the Wiener process. The first part requires $\lvert \lambda_1 \rvert T\ll 1$ and $\lambda^{\prime} T \ll 1$. The variance of x increases slowly and follows the power law. The second part requires $\lvert \lambda_1 \rvert T\gg 1$ and $\lambda^{\prime} T \ll 1$ when the noise has not been stable and the variance of x is proportional to the time. Finally, the third part requires $\lambda^{\prime} T \gg 1$ when the noise has been stable and the result is the same as the white noise and the colored noise of the OU-process when $\lambda^{\prime} \gg \lvert \lambda_1 \rvert$. If $\lambda^{\prime} \sim \lvert \lambda_1 \rvert$, the second part may not be distinguishable. The first part will be hard to be identified if $\lambda^{\prime} T \gg 1$.

In summary, we have investigated the motion of a single trapped ion subject to a colored noise environment by using an analytical approximation technique. Our theoretical results are consistent with numerical simulations, which indicates the accuracy of our approximation. The approach presented here can be extended to study other systems influenced by the colored noise.

5. Conclusion

To conclude, we have discussed the dynamics of a single ion in the Paul trap with a damping and a noise via the Langevin equation. We have proposed a recursion method, which allows us to compare the analytical solutions with those from the Monte Carlo simulations. Our recursion method consumes much less time to calculate the variance of the position of the ion and is applicable in a wide range of trap parameters of a, q and the damping coefficient Γ.

In addition, we have applied the WKB method to solve the Langevin equation and discussed two limiting cases. We use the WKB method to derive an analytical formula for the case of the OU-process and the Wiener process. We examined the OU-process under different parameters and related it to the Wiener process. Based on some approximations, we can analytically derive the relationship between the parameters and the variance.

Our work provides a comprehensive analysis of the ion motion in the Paul trap under the influence of damping and noise. We have developed novel analytical solutions using the WKB method and a recursion method, which are both applicable to a wide range of parameters. Our results have shown that one can accurately estimate and optimize the particle position according to the trap, damping and noise parameters, which would be particularly important for experiments that critically depend on the particle signal collection.

Overall, our results provide insight into the behavior of an ion in a trapped quantum system subjected to OU noise. The dependence of the variance on $\lambda^{\prime}$ and Γ offers useful information for the control and manipulation of experimental parameters in ion trap systems. In particular, equation (39) also shows the relation between $\langle x^2 \rangle$ and $\lambda^{\prime}$ ($\lambda^{\prime}$ represents the reciprocal of correlation time). Our results conclude that the variance is proportional to the square of the correlation time while correlation time is an essential parameter being considered in the noise controlling [20]. Our model provides a recursion method and expands the WKB method for investigating the dynamics of trapped ions under different experimental conditions, and has the potential to enhance our understanding of ion traps and their applications.

Finally, considering that the OU noise is strongly associated with concrete noises such as the phase noise, the amplitude noise, or the wavenumber noise (which may be caused by an extensive range of experiment settings [20]), we believe that our finding allows for better control of noise in the fields of the quantum computing and quantum error correction. Our research can be useful for these areas to investigate the effect of specific and practical noise on the particle signal collection and the logic gate fidelity.

Acknowledgments

This work is supported by National Natural Science Foundation of China under Grant Nos. 12234002 and 92250303.

Data availability statement

The data cannot be made publicly available upon publication because they are not available in a format that is sufficiently accessible or reusable by other researchers. The data that support the findings of this study are available upon reasonable request from the authors.

Appendix A: Possible relationship with experiments

In this section, we discuss how we choose the parameters and their links to relevant experiments. The parameters a and q in equation (2) are trap parameters, which respectively mean the strength of the dc/ac voltages and satisfy [7]:

Equation (A.1)

where R and Z0 are the parameters that describe the scale of the trap, κ is a geometrical factor, V0 is the amplitude of ac voltage and U0 represents the dc voltage. In the section of the recursion method, our method can be used under arbitrary a and q , as we have shown in figure 3. One should make sure the ion dynamics is stable under such parameters when we need to calculate the dynamics of the ion motion in figure 4. However, the excess micromotion are negligible for $\left|a\right| \gg q^2 $ [7], and people generally study the case that $\left|a_i\right| \ll q_i^2 \ll 1$ [7, 36]. Hence, we calculate the case of a = 0 by the WKB method. We neglect $ q_i^2 \ll 1$, although it is not necessary in our study. The damping term and the noise term can be attributed to the interactions with the buffer gas, in particular the collisions with the gas particles. We give an example of the OU-process in the experiment. The process Xt satisfies the following [36]:

Equation (A.2)

where D symbolizes the strength of the noise, Wt represents the white noise, and τ is the correlation time. The correlation time is related to the density of the bath, the Langevin rate for atom-ion collisions [43]. These physical quantities can be measured in experiments through some physical quantities [36]. The colored noise parameters that we have chosen contain two scales in order to simulate any probable cases for discussions.

Appendix B: Numerical simulation

We give some details of the numerical simulation used in the main text. The white noise Wt in equation (20) satisfies the nature of the Wiener process [41]: For any $0 \unicode{x2A7D} s \lt t$, the random variable $W_t - W_s$ is normally distributed with a mean of 0 and a variance of t − s, i.e. for any a < b, one has

Equation (B.1)

Therefore, $\mathrm dW_t = \sigma \eta_t \sqrt{\mathrm d t}$ where ηt is the standard normal distribution. We simulate this by the random function in Python. For the colored noise, we use the numerical method to solve the SDE and get the list Xt . Then the differential of the noise Xt can be calculated as the following:

Equation (B.2)

To obtain the statistical average, we calculate the motion of N = 1000 ions, or equivalently, calculate in the same way for 1000 times. To improve the computation efficiency, we perform simultaneous calculations by parallel operations. The noise is calculated by 1000 different paths at the same time while noises on different paths are irrelevant, so each ion follows a different noise for each path, while there do not exist ion-ion interactions. Then, we have

Equation (B.3)

Please note that this method consumes a large amount of memory, and it is hard to choose an even larger N on our PC. All our results are calculated by Python, through which it is easy to generate random numbers.

Please wait… references are loading.
10.1088/1751-8121/ad0348