This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Active Brownian particle under stochastic position and orientation resetting in a harmonic trap

Amir Shee1,∗ 1Northwestern Institute on Complex Systems and ESAM, Northwestern University, Evanston, IL 60208, United States of America Author to whom any correspondence should be addressed. [email protected]
Abstract

We present an exact analytical study of an Active Brownian Particle (ABP) subject to both position and orientation stochastic resetting in a two-dimensional harmonic trap. Utilizing a Fokker-Planck-based renewal approach, we derive the system’s exact moments, including the mean parallel displacement, mean squared displacement (MSD), and the fourth-order moment of displacement, and compare these with numerical simulations. To capture deviations from Gaussian behavior, we analyze the excess kurtosis, which reveals rich dynamical crossovers over time. These transitions span from Gaussian behavior (zero excess kurtosis) to two distinct non-Gaussian regimes: an activity-dominated regime (negative excess kurtosis) and a resetting-dominated regime (positive excess kurtosis). Furthermore, we quantify the steady-state phase diagrams by varying three key control parameters: activity, resetting rate, and harmonic trap strength, using steady-state excess kurtosis as the primary metric.

Keywords: active Brownian particle, stochastic resetting, harmonic trap, exact moments, excess kurtosis, phase diagrams

1 Introduction

Active particles convert energy into motility and exhibit a wide range of dynamical behaviors [1, 2, 3, 4, 5, 6]. These behaviors are observed across diverse systems, including cytoskeletal filaments in motor protein assays [7, 8, 9, 10], bird flocks [11], fish schools [12], and artificial systems like active colloids and robots. Active particles are inherently out of equilibrium, displaying collective phenomena such as flocking [13, 14, 15], clustering [16, 17, 18], and phase separation [19, 20]. Even at the single-particle level, they exhibit diverse dynamical behaviors, including short-time ballistic motion [21, 22, 23, 24, 25, 26], nonequilibrium steady states in confinement [27, 28, 29, 30], and dynamical transitions in relaxation and first passage properties [31, 32, 33, 34]. Recent studies have demonstrated enhanced control over active agents for decision-making by space dependent rotational diffusion coefficient [35], as well as using external or internal cues, such as magnetic fields [36], chemical gradients [37], and acoustic waves [38, 39], to control their speed and orientation.

Stochastic resetting is a protocol that resets a dynamical system to a specific state, primarily governing the nonequilibrium steady state (NESS) through a continuous influx of probability from resetting. This process can lead to dynamical transitions during the system’s relaxation to the NESS and often results in a non-monotonic mean first passage time [40, 41, 42, 43, 44]. The resetting strategy has broad applications across various systems, including population dynamics and biological processes, particularly in optimizing search problems [45, 46, 47, 48, 49, 50, 51, 52]. The impact of resetting on different diffusive dynamics has also been extensively studied [53, 54, 55, 56, 57, 58, 59, 60, 61, 62, 63, 64, 60]. Resetting has proven to be an optimal navigation strategy for target search, particularly in Brownian motion [40] and under external potentials [65].

The resetting of a free active Brownian agent has been studied recently in [66, 67], with extensions to external confining and absorbing potentials in [68, 69, 70]. Activity introduces new dynamical behavior due to the competition between the internal active timescale and the resetting timescale, first studied in [71]. The first passage properties of active Brownian particle (ABP) and run-and-tumble particle (RTP) under various resetting mechanisms have been explored in [72, 73]. The dynamics of an ABP under position stochastic resetting, orientation stochastic resetting, and combined position and orientation stochastic resetting, focusing on marginal probability distributions, were examined in Kumar et al. [66]. The dynamics under orientation stochastic resetting, using the intermediate scattering function, were studied in [74]. However, the exact dynamics of an ABP under both position and orientation resetting in a harmonic trap, along with an analysis of the steady-state properties, remain unexplored. From the lens of active matter, it is crucial to delve into the intricate interplay between stochastic resetting and activity under the influence of an external force.

In this work, we investigate the exact dynamics of active Brownian particle (ABP) under complete stochastic resetting, involving both position and orientation, in a harmonic trap. We apply the method for exact moment calculations of stiff chains, as described by Hermans et al.[75], to compute the exact moments for the active dynamics case. Using a Fokker-Planck-based approach, we derive the dynamical moment generating equation, which we then use to calculate all dynamical moments. This method has been previously applied to study ABP in a harmonic trap [30], with anisotropic translational noise [76], fluctuating speed [25], and chirality/torque [77], and has also been extended to inertial ABP [26, 78]. Here, we focus on exact moment calculations of ABP under complete stochastic resetting in a harmonic trap using a renewal equation following the Fokker-Planck-based moments calculation. We also characterize the steady-state behavior across various limits and parameter ranges using excess kurtosis.

The remainder of this paper is organized as follows. In Section 2, we introduce the model of an ABP under complete stochastic resetting in a harmonic trap. Using a Fokker-Planck-based approach, we derive the dynamical moment generating equation and apply the final renewal approach to calculate moments under stochastic resetting. This equation is then used to compute the orientation autocorrelation and mean displacement in Section 3. In Section 4, we determine the mean squared displacement and displacement fluctuation. To examine deviations from Gaussian behavior over time, we calculate the fourth-order moment of displacement and the excess kurtosis in Section 5, and characterize the interplay of activity with resetting and trapping in the steady-state phase diagram in Section 6. Finally, we summarize the key results in Section 7.

2 Model and moments generator equation

The standard active Brownian particle in two dimension is described by its position 𝐫=(x,y)\mathbf{r}=(x,y), and its orientation unit vector 𝐮^=(ux,uy){\bf\hat{u}}=(u_{x},u_{y}) where ux=cos(θ)u_{x}=\cos(\theta) and uy=sin(θ)u_{y}=\sin(\theta), evolving over time tt from their initial values (𝐫0,𝐮^0)({\bf r}_{0},{\bf\hat{u}}_{0}). Stochastic resetting imposed on both 𝐫{\bf r} and 𝐮^{\bf\hat{u}} intermittently resets to initial values (𝐫0,𝐮^0)({\bf r}_{0},{\bf\hat{u}}_{0}) with rate rr. The position 𝐫{\bf r} and orientation 𝐮^{\bf\hat{u}} evolves

𝐫˙=v0𝐮^+2D𝝌(t)μk𝐫,\displaystyle\dot{\bf r}=v_{0}{\bf\hat{u}}+\sqrt{2D}\bm{\chi}(t)-\mu k{\bf r}\,, (1)
𝐮^˙=2Drη(t)𝐮^,\displaystyle\dot{\bf\hat{u}}=\sqrt{2D_{r}}\eta(t){\bf\hat{u}}^{\perp}\,, (2)
(𝐫,𝐮^)(𝐫0,𝐮^0).\displaystyle({\bf r},{\bf\hat{u}})\rightarrow({\bf r}_{0},{\bf\hat{u}}_{0})\,. (3)

Here, v0v_{0} is the active speed, DD is the translation diffusion coefficient, μ\mu is the mobility, and kk is the strength of the harmonic trap. DrD_{r} is the orientation diffusion coefficient. The noise terms 𝝌(t)\bm{\chi}(t) and η(t)\eta(t) are modeled as Gaussian white noise with zero mean and variances given by χi(t)χj(t)=δijδ(tt)\langle\chi_{i}(t)\chi_{j}(t^{\prime})\rangle=\delta_{ij}\delta(t-t^{\prime}) and η(t)η(t)=δ(tt)\langle\eta(t)\eta(t^{\prime})\rangle=\delta(t-t^{\prime}), respectively. The interplay of the three timescales Dr1D_{r}^{-1}, r1r^{-1}, and (μk)1(\mu k)^{-1} would lead to rich behavior for ABP under resetting in a harmonic trap.

We rescale the dynamics using timescale τ=Dr1\tau=D_{r}^{-1} and length scale =D/Dr\ell=\sqrt{D/D_{r}}. The dynamics (𝐫𝐫/{\bf r}\to{\bf r}/\ell, 𝐮^𝐮^{\bf\hat{u}}\to{\bf\hat{u}}, tt/τt\to t/\tau) is now controlled by activity strength defined by Péclet Pe=v0τ/\mathrm{Pe}=v_{0}\tau/\ell, resetting rate r=rτr=r\tau, and trapping strength β=μk\beta=\mu k\ell. We investigate the dynamics and steady state behavior with varying three dimensionless parameters activity Pe\mathrm{Pe}, resetting rate rr, and trap strength β\beta.

Refer to caption
Figure 1: First and second order moments. (a) Mean parallel displacement along initial orientation direction in two dimension (equation (8)) for r=0.1(square,),2(circle,),10(triangle,)r=0.1(\mathrm{square},\Box),~{}2(\mathrm{circle},\circ),~{}10(\mathrm{triangle},\triangle) with β=4\beta=4 and Pe=10\mathrm{Pe}=10. (b) Steady state displacement along initial orientation 𝐫rst\langle{\bf r}_{\parallel}\rangle^{st}_{r} (equation (9)) as a function of resetting rate rr for β=1(square,),4(circle,),9(triangle,)\beta=1(\mathrm{square},\Box),4(\mathrm{circle},\circ),9(\mathrm{triangle},\triangle) with Pe=10\mathrm{Pe}=10. (c) Mean squared displacement (MSD) as function of time (equation (12)) correspond to the parameters in (a). The points represent the simulation results, while the lines correspond to the analytical solutions. The initial position is at the origin with the initial orientation along the xx-axis.

Moments generator equation

The probability distribution P(𝐫,𝐮^,t)P({\bf r},{\bf\hat{u}},t) of the position 𝐫{\bf r} and the active orientation 𝐮^{\bf\hat{u}} of the particle follows the Fokker-Planck equation [30]

tP(𝐫,𝐮^,t)\displaystyle\partial_{t}P({\bf r},{\bf\hat{u}},t) =\displaystyle= 2P+𝐮^2PPe𝐮^Pβ𝐫P+2βP,\displaystyle\nabla^{2}P+\nabla_{\bf\hat{u}}^{2}P-\mathrm{Pe}\,{\bf\hat{u}}\cdot\nabla P-\beta{\bf r}\cdot\nabla P+2\beta P\,, (4)

where \nabla is the two-dimensional Laplacian operator, and 𝐮^\nabla_{\bf\hat{u}} is the Laplacian in the orientation space.

Utilizing the Laplace transform P~(𝐫,𝐮^,s)=0𝑑testP(𝐫,𝐮^,t)\tilde{P}({\bf r},{\bf\hat{u}},s)=\int_{0}^{\infty}dt\,e^{-st}\,P({\bf r},{\bf\hat{u}},t) and defining the mean of an observable ψs=𝑑𝐫𝑑𝐮^ψ(𝐫,𝐮^)P~(𝐫,𝐮^,s)\langle\psi\rangle_{s}=\int d{\bf r}\,d{\bf\hat{u}}\,\psi({\bf r},{\bf\hat{u}})\tilde{P}({\bf r},{\bf\hat{u}},s), multiplying by ψ(𝐫,𝐮^)\psi({\bf r},{\bf\hat{u}}) and integrating over all possible (𝐫,𝐮^)({\bf r},{\bf\hat{u}}) we find [30],

ψ0+sψs\displaystyle-\langle\psi\rangle_{0}+s\langle\psi\rangle_{s} =\displaystyle= 2ψs+𝐮^2ψs+Pe𝐮^ψsβ𝐫ψs,\displaystyle\langle\nabla^{2}\psi\rangle_{s}+\langle\nabla_{\bf\hat{u}}^{2}\psi\rangle_{s}+\mathrm{Pe}\langle\,{\bf\hat{u}}\cdot\nabla\psi\rangle_{s}-\beta\langle{\bf r}\cdot\nabla\psi\rangle_{s}\,, (5)

where the initial condition sets ψ0=𝑑𝐫𝑑𝐮^ψ(𝐫,𝐮^)P(𝐫,𝐮^,0)\langle\psi\rangle_{0}=\int d{\bf r}\,d{\bf\hat{u}}\,\psi({\bf r},{\bf\hat{u}})P({\bf r},{\bf\hat{u}},0). Without any loss of generality, we consider the initial condition to follow P(𝐫,𝐮^,0)=δ(𝐫𝐫0)δ(𝐮^𝐮^0)P({\bf r},{\bf\hat{u}},0)=\delta({\bf r}-{\bf r}_{0})\delta({\bf\hat{u}}-{\bf\hat{u}}_{0}), where 𝐫0{\bf r}_{0} and 𝐮^0{\bf\hat{u}}_{0} are the initial position and orientation respectively. We use equation (5) to calculate exact moments as a function of time without resetting, which is already explored in Chaudhuri et al. [30].

The moments of an ABP under stochastic position and orientation resetting in presence of a harmonic trap can be calculated exactly as [66]

ψ(t)r\displaystyle\langle\psi(t)\rangle_{r} =\displaystyle= ertψ(t)+r0t𝑑tertψ(t).\displaystyle e^{-rt}\langle\psi(t)\rangle+r\int_{0}^{t}dt^{\prime}e^{-rt^{\prime}}\langle\psi(t^{\prime})\rangle\,. (6)

We utilize equation (6) to compute the analytic moments exactly under resetting. For comparison with simulations, we perform Euler-Maruyama integration of equations (1) and (2) with stochastic resetting equation (3). The initial position of the particles set at the origin (x,y)=(0,0)(x,y)=(0,0) which corresponds to the minimum of the harmonic potential, with the orientation along the xx-axis i.e., θ=0\theta=0. The positions and orientations of particle stochastically reset to their initial state (x,y,θ)=(0,0,0)(x,y,\theta)=(0,0,0) at rate rr.

3 Mean orientation and mean displacement

We consider the initial orientation of ABP along 𝐮^=𝐮^0\langle{\bf\hat{u}}\rangle={\bf\hat{u}}_{0} and proceed to calculate mean orientation. To calculate 𝐮^\langle{\bf\hat{u}}\rangle, we use ψ=𝐮^\psi={\bf\hat{u}} in the equation (5), leads to 𝐮^s=𝐮^0/(1+s)\langle{\bf\hat{u}}\rangle_{s}={\bf\hat{u}}_{0}/(1+s). Inverse Laplace transform leads to the average orientation without stochastic resetting 𝐮^(t)=et𝐮^0\langle{\bf\hat{u}}(t)\rangle=e^{-t}{\bf\hat{u}}_{0}. In the presence of stochastic resetting using renewal approach in equation (6) and taking the dot product with the initial orientation, we get the orientation autocorrelation,

𝐮^(t)𝐮^(0)r\displaystyle\langle{\bf\hat{u}}(t)\cdot{\bf\hat{u}}(0)\rangle_{r} =\displaystyle= r+e(1+r)t(1+r).\displaystyle\frac{r+e^{-(1+r)t}}{(1+r)}\,. (7)

The plot of equation (7) is compared with simulations, as shown in A, figure 5(a). As (tt\to\infty), the orientation autocorrelation saturates to r/(1+r)r/(1+r). In the steady state 𝐮^(t)𝐮^(0)r\langle{\bf\hat{u}}(t)\cdot{\bf\hat{u}}(0)\rangle_{r} varies from 0 as r0r\to 0 to 11 as rr\to\infty.

Now, we proceed to calculate 𝐫s\langle{\bf r}\rangle_{s} we consider the initial position at 𝐫0=0\langle{\bf r}\rangle_{0}=0, using the equation (5) leads to 𝐫s=Pe𝐮^s/(s+β)\langle{\bf r}\rangle_{s}=\mathrm{Pe}\langle{\bf\hat{u}}\rangle_{s}/(s+\beta). Substituting 𝐮^s\langle{\bf\hat{u}}\rangle_{s} and then inverse Laplace transform gives the mean displacement in the absence of resetting 𝐫(t)\langle{\bf r}(t)\rangle [30]. The parallel and perpendicular component of the displacement vector to the initial orientation defined as 𝐮^0{\bf\hat{u}}_{0} as 𝐫=(𝐫𝐮^0)𝐮^0{\bf r}_{\parallel}=({\bf r}\cdot{\bf\hat{u}}_{0}){\bf\hat{u}}_{0} and 𝐫=𝐫𝐫{\bf r}_{\perp}={\bf r}-{\bf r}_{\parallel}. In the presence of stochastic resetting using renewal approach in equation (6), we get

𝐫(t)r=Pe𝐮^0(e(β+r)te(1+r)t)1βrPe𝐮^01β[1e(1+r)t1+r1e(β+r)tβ+r].\displaystyle\langle{\bf r}_{\parallel}(t)\rangle_{r}=\frac{\mathrm{Pe}{\bf\hat{u}}_{0}(e^{-(\beta+r)t}-e^{-(1+r)t})}{1-\beta}-\frac{r\mathrm{Pe}{\bf\hat{u}}_{0}}{1-\beta}\left[\frac{1-e^{-(1+r)t}}{1+r}-\frac{1-e^{-(\beta+r)t}}{\beta+r}\right]\,. (8)

In the absence of harmonic trap (β=0\beta=0), equation (8) simplifies to ABP under stochastic resetting [66], see A, figure 5(b). In figure 1(a), we plot equation (8) as solid lines, showing excellent agreement with the simulation results represented by the points. For a small resetting rate (rr), 𝐫(t)r\langle{\bf r}_{\parallel}(t)\rangle_{r} exhibits non-monotonic behavior: it starts with a small value in the short-time, thermally dominated diffusion regime, reaches a maximum at intermediate times in the activity-dominated regime, and then decays back to smaller values due to trapping, see squares (\Box) for r=0.1r=0.1 in figure 1(a). At long times, 𝐫(t)r\langle{\bf r}_{\parallel}(t)\rangle_{r} decays to zero in the absence of stochastic resetting [30], exhibiting the same behavior as for low resetting rates. For intermediate values of the resetting rate (rr), 𝐫(t)r\langle{\bf r}_{\parallel}(t)\rangle_{r} initially shows a small value in the short-time, thermally dominated diffusion regime, reaches a maximum at intermediate times in the activity-dominated regime, and then does not decay back to smaller values due to the cancellation of trapping effects by stochastic resetting, see circles (\circ) for r=2r=2 in figure 1(a). For large values of the resetting rate (rr), 𝐫(t)r\langle{\bf r}_{\parallel}(t)\rangle_{r} initially shows a small value in the short-time, thermally dominated diffusion regime, slightly increases in the activity-dominated regime, and then saturates due to the dominance of stochastic resetting, see triangles (\triangle) for r=20r=20 in figure 1(a).

At, long times tt\to\infty, mean parallel displacement 𝐫rst=limt𝐫(t)r\langle{\bf r}_{\parallel}\rangle^{st}_{r}=\lim_{t\to\infty}\langle{\bf r}_{\parallel}(t)\rangle_{r}

𝐫rst\displaystyle\langle{\bf r}_{\parallel}\rangle^{st}_{r} =\displaystyle= rPe𝐮^0(r+β)(1+r).\displaystyle\frac{r\mathrm{Pe}{\bf\hat{u}}_{0}}{(r+\beta)(1+r)}\,. (9)

The mean parallel displacement at steady state show a local maximum at resetting rate rmax=βr_{\mathrm{max}}=\sqrt{\beta}. In figure 1(b), we plot equation (9) (solid lines) as a function of resetting rates (rr), comparing it with simulation results (points). The maximum mean parallel displacement occurs at resetting rates rmax=β=1,2,3r_{\mathrm{max}}=\sqrt{\beta}=1,~{}2,~{}3 for trap strengths β=1(squares,),4(circles,),9(triangles,)\beta=1(\mathrm{squares},\Box),~{}4(\mathrm{circles},\circ),~{}9(\mathrm{triangles},\triangle), respectively. It is important to compute the second-order moments to further quantify the fluctuating dynamics.

Refer to caption
Figure 2: Time evolution of the fourth-order moment and excess kurtosis. (a), (b), (c) Fourth-order moment of displacement 𝐫4r\langle{\bf r}^{4}\rangle_{r} (equation (81)); (d), (e), (f) Kurtosis 𝒦r\mathcal{K}_{r} as a function of time tt for different conditions: resetting rates r=0.1,2,20r=0.1,~{}2,~{}20 with Pe=10\mathrm{Pe}=10 and β=4\beta=4 in (a) and (d); trapping strength β=0.5,4,50\beta=0.5,~{}4,~{}50 with Pe=10\mathrm{Pe}=10 and r=1r=1 in (b) and (e); and activity Pe=1,5,10\mathrm{Pe}=1,~{}5,~{}10 with r=1r=1 and β=4\beta=4 in (c) and (f). The solid lines represent analytic predictions, while points denote simulation results. The initial position is at the origin, with the initial orientation along the xx-axis. The region where 𝒦r<0\mathcal{K}_{r}<0 (red) represents the activity-dominated regime, and 𝒦r>0\mathcal{K}_{r}>0 (green) represents the resetting-dominated regime.

4 Mean-squared displacement (MSD)

We proceed to compute mean-squared displacement 𝐫2s\langle{\bf r}^{2}\rangle_{s} defining observable ψ=𝐫2\psi={\bf r}^{2}, utilizing equation (5) with initial position at origin 𝐫20=0\langle{\bf r}^{2}\rangle_{0}=0 leads to the mean-squared displacement in Laplace space 𝐫2s\langle{\bf r}^{2}\rangle_{s}

𝐫2s\displaystyle\langle{\bf r}^{2}\rangle_{s} =\displaystyle= 1s+2β[4s+2Pe𝐫𝐮^s],\displaystyle\frac{1}{s+2\beta}\left[\frac{4}{s}+2\mathrm{Pe}\langle{\bf r}\cdot{\bf\hat{u}}\rangle_{s}\right]~{}, (10)

where the second term 𝐫𝐮^s\langle{\bf r}\cdot{\bf\hat{u}}\rangle_{s} calculated assuming ψ=𝐫𝐮^\psi={\bf r}\cdot{\bf\hat{u}} in equation (5) as

𝐫𝐮^s=Pes(s+β+1).\displaystyle\langle{\bf r}\cdot{\bf\hat{u}}\rangle_{s}=\frac{\mathrm{Pe}}{s(s+\beta+1)}~{}. (11)

Inverse Laplace transform of equation (10) gives mean-squared displacement 𝐫2(t)\langle{\bf r}^{2}(t)\rangle without stochastic resetting (see B), which previously obtained in Chaudhuri et al. [30]. Finally, using renewal approach in equation (6), we get

𝐫2(t)r=ert[2β(1e2βt)+Pe2β(1+β)2Pe2eβt1β[eβt2βet1+β]]\displaystyle\langle{\bf r}^{2}(t)\rangle_{r}=e^{-rt}\left[\frac{2}{\beta}\left(1-e^{-2\beta t}\right)+\frac{\mathrm{Pe}^{2}}{\beta(1+\beta)}-\frac{2\mathrm{Pe}^{2}e^{-\beta t}}{1-\beta}\left[\frac{e^{-\beta t}}{2\beta}-\frac{e^{-t}}{1+\beta}\right]\right]
+(2+2β+Pe2)(1ert)β(1+β)+2rβPe2[1e(1+β+r)t](1+β+r)(1β2)+r(Pe2+2β2)(1e(2β+r)t)β(2β+r)(1β).\displaystyle+\frac{(2+2\beta+\mathrm{Pe}^{2})(1-e^{-rt})}{\beta(1+\beta)}+\frac{2r\beta\mathrm{Pe}^{2}[1-e^{-(1+\beta+r)t}]}{(1+\beta+r)(1-\beta^{2})}+\frac{r(\mathrm{Pe}^{2}+2\beta-2)(1-e^{-(2\beta+r)t})}{\beta(2\beta+r)(1-\beta)}\,.
(12)

In figure 1(c), we compare the equation (12) represented by solid lines, with simulations depicted by points, showing excellent agreement. The various limiting cases of equation (12) depending on parameters rr, Pe\mathrm{Pe}, and β\beta discussed in B.

At small time (t0t\to 0), equation (12) leading to

𝐫2r4t+(Pe24β2r)t2+13[8β2Pe23βPe2+8βr2Pe2r+2r2]t3+𝒪(t4).\displaystyle\langle{\bf r}^{2}\rangle_{r}\simeq 4t+(\mathrm{Pe}^{2}-4\beta-2r)t^{2}+\frac{1}{3}\left[8\beta^{2}-\mathrm{Pe}^{2}-3\beta\mathrm{Pe}^{2}+8\beta r-2\mathrm{Pe}^{2}r+2r^{2}\right]t^{3}+\mathcal{O}(t^{4})\,.

The MSD at small time exhibits diffusive behavior 𝐫2r=4t\langle{\bf r}^{2}\rangle_{r}=4t, which crosses over to ballistic behavior 𝐫2rt2\langle{\bf r}^{2}\rangle_{r}\sim t^{2} when Pe2>2r+4β\mathrm{Pe}^{2}>2r+4\beta at t=4/(Pe22r4β)t^{*}=4/(\mathrm{Pe}^{2}-2r-4\beta). In the steady state, MSD 𝐫2rst=𝐫2r(t)\langle{\bf r}^{2}\rangle^{st}_{r}=\langle{\bf r}^{2}\rangle_{r}(t\to\infty) gives

𝐫2rst\displaystyle\langle{\bf r}^{2}\rangle^{st}_{r} =\displaystyle= 42β+r+2Pe2(1+β+r)(2β+r).\displaystyle\frac{4}{2\beta+r}+\frac{2\mathrm{Pe}^{2}}{(1+\beta+r)(2\beta+r)}~{}. (14)

The standard deviation is given by σ=𝐫2rst/2\sigma=\sqrt{\langle{\bf r}^{2}\rangle^{st}_{r}/2}, the Gaussian probability distribution is P(|𝐫|)=e𝐫2/2σ2/2πσ2P(|{\bf r}|)=e^{-{\bf r}^{2}/2\sigma^{2}}/2\pi\sigma^{2}. These are represented by solid lines in figures 3(d), 3(e), and 3(f).

The effective diffusion coefficient Deff=(β+r)𝐫2st/2D_{\rm eff}=(\beta+r)\langle{\bf r}^{2}\rangle_{\rm st}/2

Deff\displaystyle D_{\rm eff} =\displaystyle= 2(β+r)2β+r+(β+r)Pe2(1+β+r)(2β+r).\displaystyle\frac{2(\beta+r)}{2\beta+r}+\frac{(\beta+r)\mathrm{Pe}^{2}}{(1+\beta+r)(2\beta+r)}\,. (15)

Setting resetting r=0r=0 and activity Pe=0\mathrm{Pe}=0, Deff=1D_{\rm eff}=1 holds the fluctuation dissipation relation (FDR). Excess diffusion coefficient due to non-equilibrium nature, calculated as Dex=Deff1D_{\rm ex}=D_{\rm eff}-1 reads

Dex\displaystyle D_{\rm ex} =\displaystyle= r2β+r+(β+r)Pe2(1+β+r)(2β+r).\displaystyle\frac{r}{2\beta+r}+\frac{(\beta+r)\mathrm{Pe}^{2}}{(1+\beta+r)(2\beta+r)}\,. (16)

when Dex=0D_{\rm ex}=0, the fluctuation-dissipation relation holds for r=Pe=0r=\mathrm{Pe}=0. Both activity (Pe\mathrm{Pe}) and resetting (rr) breaks this relation and resulting in Dex>0D_{\rm ex}>0. However, it is not possible to precisely distinguish between the activity-dominated and resetting-dominated regimes by calculating DexD_{\rm ex} when varying the parameters r,β,Per,\beta,\mathrm{Pe}. We distinguish them by calculating the excess kurtosis in the next section.

In the absence of stochastic resetting (r=0r=0), we get ABP in a harmonic trap [30]. In the absence of activity (Pe=0\mathrm{Pe}=0), Brownian particle in two dimensions under resetting gives 𝐫2rst=4/(2β+r)\langle{\bf r}^{2}\rangle^{st}_{r}=4/(2\beta+r), Deff=2(β+r)/(2β+r)D_{\rm eff}=2(\beta+r)/(2\beta+r), and Dex=r/(2β+r)D_{\rm ex}=r/(2\beta+r).

Moreover, the displacement fluctuations and their limiting cases are discussed in C. The components of the second-order moment, such as those along the initial orientation direction, are also discussed in D.

5 Excess kurtosis: signature of non-Gaussian behavior

To calculate the excess kurtosis, we first need to calculate the fourth-order moment of the displacement. In the similar method, we calculated the fourth order moment of displacement under stochastic position and orientation resetting, detailed derivation and final result of 𝐫4(t)r\langle{\bf r}^{4}(t)\rangle_{r} in equation (81), see detailed derivation in E. The small time expansion (t0t\to 0), the fourth moment of displacement 𝐫4r\langle{\bf r}^{4}\rangle_{r}

𝐫4r32t2+163(3Pe24r12β)t3\displaystyle\langle{\bf r}^{4}\rangle_{r}\simeq 32t^{2}+\frac{16}{3}\left(3\mathrm{Pe}^{2}-4r-12\beta\right)t^{3}
+13[224β2+3Pe4+24r2+144βr96βPe216Pe236rPe2]t4+𝒪(t5),\displaystyle+\frac{1}{3}\left[224\beta^{2}+3\mathrm{Pe}^{4}+24r^{2}+144\beta r-96\beta\mathrm{Pe}^{2}-16\mathrm{Pe}^{2}-36r\mathrm{Pe}^{2}\right]t^{4}+\mathcal{O}(t^{5})\,, (17)

exhibits small time diffusive behavior 𝐫4r=32t2\langle{\bf r}^{4}\rangle_{r}=32t^{2}, which crosses over to ballistic behavior 𝐫4rt3\langle{\bf r}^{4}\rangle_{r}\sim t^{3} at t=6/(3Pe24r12β)t^{*}=6/(3\mathrm{Pe}^{2}-4r-12\beta) with Pe2>4(r+3β)/3\mathrm{Pe}^{2}>4(r+3\beta)/3. In figures 2(a), 2(b), and 2(c), we have shown the 𝐫4r\langle{\bf r}^{4}\rangle_{r} as a function of time tt for resetting rates r=0.1,2,20r=0.1,~{}2,~{}20 with Pe=10\mathrm{Pe}=10 and β=4\beta=4 in (a), for trap strength β=0.5,4,50\beta=0.5,~{}4,~{}50 with Pe=10\mathrm{Pe}=10 and r=1r=1 in (b), and for activity Pe=1,5,10\mathrm{Pe}=1,~{}5,~{}10 with β=4\beta=4 and r=1r=1 in (c). The solid lines are the plot of equation (81) which excellently agrees with the simulation results depicted by points. In figures 2(a) and 2(b), we can see the increase of resetting rates rr and trapping strengths β\beta decreases the 𝐫4r\langle{\bf r}^{4}\rangle_{r}. On the other hand, the increase of activity parameter Pe\mathrm{Pe} increases the 𝐫4r\langle{\bf r}^{4}\rangle_{r}. In the long time limit (tt\to\infty), 𝐫4r\langle{\bf r}^{4}\rangle_{r} reaches a steady state due to the effect of the harmonic trap and/or stochastic resetting. The steady state fourth moment of displacement 𝐫4rst\langle{\bf r}^{4}\rangle^{st}_{r}

𝐫4rst=8(1+β+r)(2β+r)(4+2β+r)(1+3β+r)(4β+r)×\displaystyle\langle{\bf r}^{4}\rangle^{st}_{r}=\frac{8}{(1+\beta+r)(2\beta+r)(4+2\beta+r)(1+3\beta+r)(4\beta+r)}\times
[48β3+8(1+r)2(4+r)+Pe4(8+3r)+8β2(20+6Pe2+11r)\displaystyle\left[48\beta^{3}+8(1+r)^{2}(4+r)+\mathrm{Pe}^{4}(8+3r)+8\beta^{2}(20+6\mathrm{Pe}^{2}+11r)\right.
+4Pe2(8+14r+3r2)+2β[3Pe4+8Pe2(7+3r)+24(3+4r+r2)]].\displaystyle\left.+4\mathrm{Pe}^{2}(8+14r+3r^{2})+2\beta\left[3\mathrm{Pe}^{4}+8\mathrm{Pe}^{2}(7+3r)+24(3+4r+r^{2})\right]\right]\,. (18)

In order to quantify the impact of these parameters on position distributions at different times as well as in the steady state, we calculate excess kurtosis which define the deviation from a Gaussian process.

In the absence of activity (Pe=0\mathrm{Pe}=0) and stochastic resetting (r=0r=0), Brownian particle in two dimensions in a harmonic trap gives zero kurtosis which signifies the Gaussian process. We calculate the kurtosis in presence of activity and stochastic position and orientation resetting which deviates from zero. Thus, the non-equilibrium nature of the system leading to the deviation from Gaussian nature given by excess kurtosis [30],

𝒦r\displaystyle\mathcal{K}_{r} =\displaystyle= 𝐫4r2𝐫2r21.\displaystyle\frac{\langle{\bf r}^{4}\rangle_{r}}{2\langle{\bf r}^{2}\rangle_{r}^{2}}-1\,. (19)

The fourth order moment of displacement 𝐫4r\langle{\bf r}^{4}\rangle_{r} and mean-squared displacement 𝐫2r\langle{\bf r}^{2}\rangle_{r} have already been calculated as shown in equation (81) and equation (12), respectively. We expand the kurtosis in the small time limit (t0t\to 0),

𝒦r\displaystyle\mathcal{K}_{r} \displaystyle\simeq rt3(3Pe4+16βr4rPe2)96t2+𝒪(t3).\displaystyle\frac{rt}{3}-\frac{(3\mathrm{Pe}^{4}+16\beta r-4r\mathrm{Pe}^{2})}{96}t^{2}+\mathcal{O}(t^{3})~{}. (20)

The next order term 𝒪(t3)\mathcal{O}(t^{3}) is shown in equation (86). It has shown the deviation of kurtosis towards positive or negative values in the small time controlled by rr, Pe\mathrm{Pe}, and β\beta. The positive deviation holds for small activity (Pe0\mathrm{Pe}\to 0) but it deviates towards negative values with increase of activity. For high activity, the kurtosis deviates towards negative values at t=32r/(3Pe4+16βr4rPe2)t^{*}=32r/(3\mathrm{Pe}^{4}+16\beta r-4r\mathrm{Pe}^{2}).

In figures 2(d), 2(e), and 2(f), we have shown the 𝒦r\mathcal{K}_{r} as a function of time tt for resetting rates r=0.1,2,20r=0.1,~{}2,~{}20 with Pe=10\mathrm{Pe}=10 and β=4\beta=4 in (d), for trapping strength β=0.5,4,50\beta=0.5,~{}4,~{}50 with Pe=10\mathrm{Pe}=10 and r=1r=1 in (e), and for Péclet values Pe=1,5,10\mathrm{Pe}=1,~{}5,~{}10 with β=4\beta=4 and r=1r=1 in (f). The solid lines are the plot of equation (19) which excellently agrees with the simulation results depicted by points. We characterize the activity- and resetting-dominated time regimes using excess kurtosis (𝒦r\mathcal{K}_{r}). At short times, thermal diffusion leads to near-Gaussian behavior (𝒦r0\mathcal{K}_{r}\sim 0). Negative 𝒦r\mathcal{K}_{r} indicates a flat, light-tailed with non-zero peak at Pe/β\mathrm{Pe}/\beta position distribution dominated by activity (Pe\mathrm{Pe}), while positive 𝒦r\mathcal{K}_{r} suggests a heavy-tailed with peak at zero position distribution dominated by resetting (rr). In figure 2(d), we clearly observe activity-dominated behavior (𝒦r<0\mathcal{K}_{r}<0) at low resetting rates (rr) and resetting-dominated behavior (𝒦r>0\mathcal{K}_{r}>0) at high resetting rates (rr). In figure 2(e), we observe non-monotonic behavior at small trap strength (β\beta). At short times, 𝒦r\mathcal{K}_{r} shifts to negative values, indicating activity-dominated behavior, crosses zero at intermediate times, and eventually saturates to positive values, reflecting long-term resetting-dominated behavior. The transitions from small to long times are: Gaussian (𝒦r=0\mathcal{K}_{r}=0) to light-tail (𝒦r<0\mathcal{K}_{r}<0), back to Gaussian (𝒦r=0\mathcal{K}_{r}=0), and finally to heavy-tail (𝒦r>0\mathcal{K}_{r}>0) position distributions. In figure 2(f), we observe activity-dominated behavior (𝒦r<0\mathcal{K}_{r}<0) at high activity (Pe\mathrm{Pe}) and resetting-dominated behavior (𝒦r>0\mathcal{K}_{r}>0) at low activity (Pe\mathrm{Pe}).

To quantify the steady-state properties of position distributions and capture the transitions between activity- and resetting-dominated regions, we calculate the steady-state excess kurtosis and analyze the numerically obtained position distributions.

Refer to caption
Refer to caption
Figure 3: Steady-state kurtosis, 𝒦rst\mathcal{K}^{st}_{r}, as a function of rr with Pe=10\mathrm{Pe}=10 and β=4\beta=4 in (a), β\beta with Pe=10\mathrm{Pe}=10 and r=1r=1 in (b), and Pe\mathrm{Pe} with r=1r=1 and β=4\beta=4 in (c). The blue solid line represents the exact analytical solution for steady-state kurtosis as given by equation (21). The symbols (blue open circles) are from simulation results. The regions where 𝒦rst<0\mathcal{K}^{st}_{r}<0 (red) correspond to the activity (A)-dominated regime, while regions where 𝒦rst>0\mathcal{K}^{st}_{r}>0 (green) correspond to the resetting (R)-dominated regime. The critical points where 𝒦rst=0\mathcal{K}^{st}_{r}=0 (black solid symbols) occur at rc=3.93r_{c}=3.93 in (a), βcI=0.77\beta^{I}_{c}=0.77 and βcII=4804\beta^{II}_{c}=4804 in (b), and Pec=2.90\mathrm{Pe}_{c}=2.90 in (c). The radial probability distributions for three resetting rates in (d) correspond to (a), four harmonic trap strengths in (e) to (b), and three activity values in (f) to (c).

6 Steady state excess kurtosis and phase diagrams

Refer to caption
Figure 4: Phase Diagrams. Steady-state kurtosis, 𝒦rst\mathcal{K}^{st}_{r}, is shown in the Per\mathrm{Pe}-r plane with β=1\beta=1 in (a), in the Peβ\mathrm{Pe}-\beta plane with r=1r=1 in (b), and in the rβr-\beta plane with Pe=10\mathrm{Pe}=10 in (c). The black dashed line corresponds to 𝒦rst=0\mathcal{K}^{st}_{r}=0. The region where 𝒦rst<0\mathcal{K}^{st}_{r}<0 is denoted as the activity-dominated regime (A), and the region where 𝒦rst>0\mathcal{K}^{st}_{r}>0 is denoted as the resetting-dominated regime (R).

In the steady state(tt\to\infty), equation (19) results

𝒦rst=2(4+2β+r)(1+3β+r)(4β+r)(2+2β+Pe2+2r)2×\displaystyle\mathcal{K}^{st}_{r}=\frac{2}{(4+2\beta+r)(1+3\beta+r)(4\beta+r)(2+2\beta+\mathrm{Pe}^{2}+2r)^{2}}\times
[12β4rβ3(6Pe420Pe2r2r(26+17r))\displaystyle\left[12\beta^{4}r-\beta^{3}(6\mathrm{Pe}^{4}-20\mathrm{Pe}^{2}r-2r(26+17r))\right.
+r(1+r)(Pe4(2+r)+2(1+r)2(4+r)+2Pe2(2r2+9r+4))\displaystyle\left.+r(1+r)(\mathrm{Pe}^{4}(2+r)+2(1+r)^{2}(4+r)+2\mathrm{Pe}^{2}(2r^{2}+9r+4))\right.
+β2(Pe4(14+r)+2Pe2r(32+19r)+2r(17r2+55r+38))\displaystyle\left.+\beta^{2}(-\mathrm{Pe}^{4}(14+r)+2\mathrm{Pe}^{2}r(32+19r)+2r(17r^{2}+55r+38))\right.
+βr(Pe4(1+3r)+2(1+r2)(22+7r)+Pe2(22r2+80r+52))].\displaystyle\left.+\beta r(\mathrm{Pe}^{4}(1+3r)+2(1+r^{2})(22+7r)+\mathrm{Pe}^{2}(22r^{2}+80r+52))\right]\,. (21)

We now calculate various limiting cases to explore the properties of excess kurtosis. I. Brownian particles in a harmonic trap : In the absence of activity (Pe=0\mathrm{Pe}=0) and stochastic resetting (r=0r=0), Brownian particle in two dimensions in a harmonic trap gives 𝒦rst=0\mathcal{K}_{r}^{st}=0. It signifies the Gaussian process, and we calculate the steady state kurtosis to quantify the deviation from 𝒦rst=0\mathcal{K}_{r}^{st}=0.

II. Brownian particles under stochastic resetting : In the absence of activity (Pe=0\mathrm{Pe}=0) and harmonic trap (β=0\beta=0), Brownian particle in two dimensions under stochastic position and orientation resetting gives 𝒦rst=1\mathcal{K}^{st}_{r}=1. The positive kurtosis signifies the heavy tailed position distribution at long times.

III. Brownian particles under stochastic resetting in a harmonic trap : In the absence of activity (Pe=0\mathrm{Pe}=0), Brownian particles in two dimensions under stochastic position and orientation resetting in a harmonic trap gives 𝒦rst=r/(4β+r)\mathcal{K}^{st}_{r}=r/(4\beta+r). It suggests the decrease of kurtosis with the increase of trapping strength β\beta. For very high resetting rates, 𝒦rst\mathcal{K}^{st}_{r} reaches a maximum value of 11.

IV. ABP under stochastic resetting : In the absence of harmonic trap (β=0\beta=0), ABP under complete stochastic resetting, equation (21) gives,

𝒦rst=[4(1+r)2(4+r)+4(2r2+9r+4)Pe2+2(2+r)Pe4](4+r)(2(1+r)+Pe2)2.\displaystyle\mathcal{K}^{st}_{r}=\frac{\left[4(1+r)^{2}(4+r)+4(2r^{2}+9r+4)\mathrm{Pe}^{2}+2(2+r)\mathrm{Pe}^{4}\right]}{(4+r)(2(1+r)+\mathrm{Pe}^{2})^{2}}~{}. (22)

Remarkably, in the absence of a harmonic trap (β=0\beta=0), equation (22) yields only positive values, indicating heavy-tailed position distributions (see G).

V. ABP in a harmonic trap : In the absence of stochastic resetting (r=0r=0), ABP in a harmonic trap gives the steady state kurtosis 𝒦st\mathcal{K}^{st} already explored in [30].

In figures 3(a), 3(b), and 3(c), the steady state kurtosis 𝒦rst\mathcal{K}^{st}_{r} from equation (21) is shown as blue solid lines, compared with simulation results represented by blue open points. The black solid points denote the Gaussian behavior at 𝒦rst=0\mathcal{K}^{st}_{r}=0. The negative kurtosis 𝒦rst<0\mathcal{K}^{st}_{r}<0 indicates persistent active motion, leading to a flat, light-tailed position distribution with a non-zero peak. This is referred to as the activity (A) dominated regime. Positive kurtosis, 𝒦rst>0\mathcal{K}^{st}_{r}>0, indicates passive motion, resulting in a narrow, heavy-tailed position distribution peaking at zero. This is referred to as the resetting (R) dominated regime. In figures 3(d), 3(e), and 3(f), we plot the radial distribution from simulations (points) alongside the corresponding Gaussian distribution (solid lines) using the second-order moment.

Figure 3(a) shows a monotonic transition from the activity (A) dominated regime at small resetting rates (rr) to the resetting (R) dominated regime at larger resetting rates, with a critical resetting rate (rcr_{c}) marking Gaussian behavior. Interestingly, in the resetting dominated regime, the steady-state kurtosis 𝒦rst\mathcal{K}^{st}_{r} exhibits a non-monotonic transition, reaching its maximum positive value at intermediate rr. This is purely due to the interplay between activity (Pe\mathrm{Pe}) and resetting (rr), which occurs even in the absence of a harmonic trap (see G). In figure 3(d), we present the radial distribution for three resetting rates: the activity-dominated regime at r<rcr<r_{c}, Gaussian behavior at r=rcr=r_{c}, and the resetting-dominated regime at r>rcr>r_{c}.

Figure 3(b) shows a non-monotonic transition as a function of trap strength β\beta. Initially, the resetting (R) dominated regime occurs at weak trap strength, transitioning to the activity (A) dominated regime at stronger trap strengths, with a critical trap strength (βcI\beta^{I}_{c}) indicating Gaussian behavior. Subsequently, the activity (A) dominated regime transitions back to the resetting regime at very strong trap strengths, marked by another critical trap strength (βcII\beta^{II}_{c}). In figure 3(e), we present the radial distribution for four trap strengths: the resetting-dominated regime at β<βc\beta<\beta_{c}, Gaussian behavior at β=βcI\beta=\beta^{I}_{c}, the activity-dominated regime at β>βc\beta>\beta_{c}, Gaussian behavior at β=βcII\beta=\beta^{II}_{c}. In the resetting-dominated regime for β>βcII\beta>\beta^{II}_{c}, the steady-state kurtosis 𝒦rst\mathcal{K}^{st}_{r} takes very small positive values, making the heavy-tail visualization nearly impossible.

Figure 3(c) shows a monotonic transition from the resetting (R) dominated regime at small activity (Pe\mathrm{Pe}) to the activity (A) dominated regime at larger activity, with a critical resetting rate (Pec\mathrm{Pe}_{c}) marking Gaussian behavior. In figure 3(f), we present the radial distribution for three activity (Pe\mathrm{Pe}) values: the resetting-dominated regime at Pe<Pec\mathrm{Pe}<\mathrm{Pe}_{c}, Gaussian behavior at Pe=Pec\mathrm{Pe}=\mathrm{Pe}_{c}, and the activity-dominated regime at Pe>Pec\mathrm{Pe}>\mathrm{Pe}_{c}.

In figures 4(a), 4(b), and 4(c), we plot the phase diagram considering 𝒦rst\mathcal{K}^{st}_{r} as order parameter to show the activity (A) and resetting (R) dominated regime separating by Gaussian line 𝒦rst=0\mathcal{K}^{st}_{r}=0 depicted by black dashed line. Figure 4(a) shows the phase diagram in the Pe\mathrm{Pe}-rr plane, highlighting two key points: at small resetting rates (rr), the transition progresses from weak resetting (R) to Gaussian, then to the activity (A) dominated regime. At large resetting rates, increasing activity enhances the resetting (R) dominated regime, resulting in a heavier tail in the distribution. Figure 4(b) shows the phase diagram in the Pe\mathrm{Pe}-β\beta plane, highlighting a re-entrant transition with increasing trap strength (β\beta), progressing from resetting (R) to Gaussian, then to activity (A), back to Gaussian, and finally returning to resetting-dominated behavior. The suppression of resetting behavior with increasing trap strength in figure 4(b) is also illustrated in the rr-β\beta plane in figure 4(c).

7 Conclusions

In this work, we computed the exact analytical moments of a two-dimensional Active Brownian Particle (ABP) subject to complete stochastic resetting (both position and orientation) in a harmonic trap. These analytical results were validated against numerical simulations, demonstrating excellent agreement, and the nonequilibrium steady-state behavior was thoroughly analyzed.

In the steady state, the orientation autocorrelation decays to a non-zero value in the presence of resetting, increasing with the resetting rate and approaching unity as the resetting rate becomes large. The steady state mean displacement exhibits non-monotonic behavior with resetting rate, initially increasing from zero as the resetting rate rises, peaking, and then returning to near-zero at very high resetting rates. We showed that ballistic dynamics emerge in the mean squared displacement (MSD) at intermediate times with increasing activity, though these are suppressed when resetting dominates. Additionally, the MSD reveals that the fluctuation-dissipation relation (FDR) holds in the absence of both activity and resetting, but breaks down, yielding a non-zero excess diffusion coefficient, when either is present.

To distinguish between the activity-dominated (A) and resetting-dominated (R) regimes, we computed the fourth moment of displacement and the corresponding excess kurtosis, providing a complete characterization of the system. In the steady state, excess kurtosis captures the transition between the activity-dominated regime (negative excess kurtosis) and the resetting-dominated regime (positive excess kurtosis), with the Gaussian regime represented by zero excess kurtosis. We anticipate that this rich steady-state behavior can be experimentally validated in active colloids or robotic systems, as shown in [79]. In the future, it will be crucial to explore Active Brownian Particles (ABPs) under two additional stochastic resetting protocols, only position resetting and only orientation resetting in a harmonic trap, to gain valuable insights into optimal resetting strategies in external potentials.

Acknowledgements

AS acknowledges partial financial support from the John Templeton Foundation, Grant 62213.

Appendix A Orientation autocorrelation and mean displacement without harmonic trap

Refer to caption
Figure 5: (a) Orientation autocorrelation (equation (7)) and (b) mean parallel displacement (equation (8)) as a function of time in two dimension for r=0.1(square,),1(triangle,),10(diamond,),100(circle,)r=0.1(\mathrm{square},\Box),~{}1(\mathrm{triangle},\triangle),~{}10(\mathrm{diamond},\Diamond),~{}100(\mathrm{circle},\circ) with β=0\beta=0 and Pe=10\mathrm{Pe}=10. Points are from simulation results and solid lines are analytic plot. The initial position is at the origin with the initial orientation along the xx-axis.

In figure 5(a), we plot the orientation autocorrelation from equation (7) for varying resetting rates (rr). As the resetting rate increases, the orientation autocorrelation saturates at higher, non-zero values. In figure 5(b), we plot the mean parallel displacement as a function of time from equation (8) for varying resetting rates (rr) with activity Pe=10\mathrm{Pe}=10 in the absence of a harmonic trap (β=0\beta=0).

Appendix B Limiting cases of mean-squared displacement (MSD)

B.1 ABP in a harmonic trap

Inverse Laplace transform of equation (10) gives mean-squared displacement 𝐫2(t)\langle{\bf r}^{2}(t)\rangle for an ABP in a harmonic trap without stochastic resetting (or by setting r=0r=0 in equation (12)), previously calculated by Chaudhuri et al. [30],

𝐫2(t)\displaystyle\langle{\bf r}^{2}(t)\rangle =\displaystyle= 2β(1e2βt)+Pe2β(1+β)2Pe2eβt1β[eβt2βet1+β].\displaystyle\frac{2}{\beta}\left(1-e^{-2\beta t}\right)+\frac{\mathrm{Pe}^{2}}{\beta(1+\beta)}-\frac{2\mathrm{Pe}^{2}e^{-\beta t}}{1-\beta}\left[\frac{e^{-\beta t}}{2\beta}-\frac{e^{-t}}{1+\beta}\right]\,. (23)

At small times (t0t\to 0), the expansion of equation (23) leads to

𝐫2\displaystyle\langle{\bf r}^{2}\rangle \displaystyle\simeq 4t+(Pe24β)t2+13[8β2Pe23βPe2]t3+𝒪(t4).\displaystyle 4t+(\mathrm{Pe}^{2}-4\beta)t^{2}+\frac{1}{3}\left[8\beta^{2}-\mathrm{Pe}^{2}-3\beta\mathrm{Pe}^{2}\right]t^{3}+\mathcal{O}(t^{4})\,. (24)

In the steady state (tt\to\infty), equation (23) gives

𝐫2st\displaystyle\langle{\bf r}^{2}\rangle^{st} =\displaystyle= 2β+Pe2β(1+β).\displaystyle\frac{2}{\beta}+\frac{\mathrm{Pe}^{2}}{\beta(1+\beta)}\,. (25)

The effective diffusion coefficient Deff=β𝐫2st/2D_{\rm eff}=\beta\langle{\bf r}^{2}\rangle_{\rm st}/2

Deff=1+Pe22(1+β),\displaystyle D_{\rm eff}=1+\frac{\mathrm{Pe}^{2}}{2(1+\beta)}\,, (26)

which enhanced with increase of activity Pe\mathrm{Pe} and suppressed with increase of harmonic stiffness β\beta.

B.2 ABP under complete stochastic resetting

In the absence of harmonic trap (β=0\beta=0), equation (12) simplifies to

𝐫2(t)r\displaystyle\langle{\bf r}^{2}(t)\rangle_{r} =\displaystyle= 4(1+r)+2Pe2r(1+r)(4+2Pe2)ertr+2Pe2e(1+r)t(1+r).\displaystyle\frac{4(1+r)+2\mathrm{Pe}^{2}}{r(1+r)}-\frac{(4+2\mathrm{Pe}^{2})e^{-rt}}{r}+\frac{2\mathrm{Pe}^{2}e^{-(1+r)t}}{(1+r)}\,. (27)

At small times (t0t\to 0), the expansion of equation (27) leads to

𝐫2(t)r\displaystyle\langle{\bf r}^{2}(t)\rangle_{r} \displaystyle\simeq 4t+(Pe22r)t213[Pe2(1+2r)2r2]t3+𝒪(t4).\displaystyle 4t+(\mathrm{Pe}^{2}-2r)t^{2}-\frac{1}{3}\left[\mathrm{Pe}^{2}(1+2r)-2r^{2}\right]t^{3}+\mathcal{O}(t^{4})\,. (28)

The MSD at small time exhibits diffusive behavior 𝐫2r=4t\langle{\bf r}^{2}\rangle_{r}=4t, which crosses over to ballistic behavior 𝐫2rt2\langle{\bf r}^{2}\rangle_{r}\sim t^{2} when Pe2>2r\mathrm{Pe}^{2}>2r at t=4/(Pe22r)t^{*}=4/(\mathrm{Pe}^{2}-2r). At long times (tt\to\infty), 𝐫2r\langle{\bf r}^{2}\rangle_{r} reaches a steady state,

𝐫2rst\displaystyle\langle{\bf r}^{2}\rangle^{st}_{r} =\displaystyle= 4r+2Pe2r(1+r).\displaystyle\frac{4}{r}+\frac{2\mathrm{Pe}^{2}}{r(1+r)}~{}. (29)

It reaches steady state at tst=𝐫2rst/4t^{st}=\langle{\bf r}^{2}\rangle^{st}_{r}/4 for Pe22r\mathrm{Pe}^{2}\leq 2r and tst=[𝐫2rst/(Pe22r)]1/2t^{st}=[\langle{\bf r}^{2}\rangle^{st}_{r}/(\mathrm{Pe}^{2}-2r)]^{1/2} for Pe2>2r\mathrm{Pe}^{2}>2r.

Note, the steady-state MSD in equations (29) and (25) have same form, with 𝐫2rst=2𝐫2st\langle{\bf r}^{2}\rangle^{st}_{r}=2\langle{\bf r}^{2}\rangle^{st}. Thus, in the steady state, the impact of stochastic resetting rate and trap stiffness on ABP is equivalent.

The effective diffusion coefficient Deff=r𝐫2st/2D_{\rm eff}=r\langle{\bf r}^{2}\rangle_{\rm st}/2

Deff=2[1+Pe22(1+r)].\displaystyle D_{\rm eff}=2\left[1+\frac{\mathrm{Pe}^{2}}{2(1+r)}\right]\,. (30)

This is twice the effective diffusion coefficient of ABP in a harmonic trap. Similar to ABP in a harmonic trap, the effective diffusion coefficient here is enhanced with an increase in activity Pe\mathrm{Pe} and suppressed with increase of resetting rate rr.

B.3 Brownian particle under stochastic resetting in harmonic trap

In the absence of activity (Pe=0\mathrm{Pe}=0), equation (12) simplifies to

𝐫2(t)r=4(1e(2β+r)t)2β+r.\displaystyle\langle{\bf r}^{2}(t)\rangle_{r}=\frac{4\left(1-e^{-(2\beta+r)t}\right)}{2\beta+r}\,. (31)

At small times (t0t\to 0), the expansion of equation (31) leads to

𝐫2r4t2(2β+r)t2+23[4β2+4βr+r2]t3+𝒪(t4).\displaystyle\langle{\bf r}^{2}\rangle_{r}\simeq 4t-2(2\beta+r)t^{2}+\frac{2}{3}\left[4\beta^{2}+4\beta r+r^{2}\right]t^{3}+\mathcal{O}(t^{4})\,. (32)

The MSD at small time exhibits diffusive behavior 𝐫2r=4t\langle{\bf r}^{2}\rangle_{r}=4t, which reaches steady state (tt\to\infty), equation (31) gives

𝐫2rst=4(2β+r).\displaystyle\langle{\bf r}^{2}\rangle^{st}_{r}=\frac{4}{(2\beta+r)}\,. (33)

The effective diffusion coefficient Deff=(β+r)𝐫2st/2D_{\rm eff}=(\beta+r)\langle{\bf r}^{2}\rangle_{\rm st}/2

Deff=2(β+r)(2β+r).\displaystyle D_{\rm eff}=\frac{2(\beta+r)}{(2\beta+r)}\,. (34)

The effective diffusion coefficient increases with increase of resetting rate rr.

B.4 Brownian particle under stochastic resetting

In the absence of activity (Pe=0\mathrm{Pe}=0) and harmonic trap (β=0\beta=0), equation (12) simplifies to

𝐫2(t)r=4(1ert)r.\displaystyle\langle{\bf r}^{2}(t)\rangle_{r}=\frac{4\left(1-e^{-rt}\right)}{r}\,. (35)

At small times (t0t\to 0), the expansion of equation (35) leads to

𝐫2r4t2rt2+23r2t3+𝒪(t4).\displaystyle\langle{\bf r}^{2}\rangle_{r}\simeq 4t-2rt^{2}+\frac{2}{3}r^{2}t^{3}+\mathcal{O}(t^{4})\,. (36)

In the steady state (tt\to\infty), equation (35) gives

𝐫2rst=4r.\displaystyle\langle{\bf r}^{2}\rangle^{st}_{r}=\frac{4}{r}\,. (37)

The effective diffusion coefficient Deff=r𝐫2st/2D_{\rm eff}=r\langle{\bf r}^{2}\rangle_{\rm st}/2

Deff=2.\displaystyle D_{\rm eff}=2\,. (38)

The effective diffusion coefficient is constant and twice of the thermal diffusion coefficient.

Appendix C Displacement fluctuations and its limiting cases

The displacement fluctuation δ𝐫2r=𝐫2r𝐫r2\langle\delta{\bf r}^{2}\rangle_{r}=\langle{\bf r}^{2}\rangle_{r}-\langle{\bf r}\rangle_{r}^{2}. At small time (t0t\to 0),

δ𝐫2r\displaystyle\langle\delta{\bf r}^{2}\rangle_{r} \displaystyle\simeq 4t2(2β+r)t2+23[4β2+Pe2+4βr+r2]t3+𝒪(t4).\displaystyle 4t-2(2\beta+r)t^{2}+\frac{2}{3}\left[4\beta^{2}+\mathrm{Pe}^{2}+4\beta r+r^{2}\right]t^{3}+\mathcal{O}(t^{4})\,. (39)

In the steady state,

δ𝐫2rst\displaystyle\langle\delta{\bf r}^{2}\rangle^{st}_{r} =\displaystyle= 4[2+Pe2+3r+r2+2β+βr](2+r)(1+β+r)(2β+r).\displaystyle\frac{4[2+\mathrm{Pe}^{2}+3r+r^{2}+2\beta+\beta r]}{(2+r)(1+\beta+r)(2\beta+r)}\,. (40)

The limiting cases are discussed below.

C.1 ABP in a harmonic trap

The displacement fluctuation δ𝐫2=𝐫2𝐫2\langle\delta{\bf r}^{2}\rangle=\langle{\bf r}^{2}\rangle-\langle{\bf r}\rangle^{2}. Here, we set resetting rate r=0r=0.

At small times (t0t\to 0),

δ𝐫2\displaystyle\langle\delta{\bf r}^{2}\rangle \displaystyle\simeq 4t4βt2+23[4β2+Pe2]t3+𝒪(t4).\displaystyle 4t-4\beta t^{2}+\frac{2}{3}\left[4\beta^{2}+\mathrm{Pe}^{2}\right]t^{3}+\mathcal{O}(t^{4})\,. (41)

In the steady state (tt\to\infty),

δ𝐫2st\displaystyle\langle\delta{\bf r}^{2}\rangle^{st} =\displaystyle= 2β+Pe2β(1+β).\displaystyle\frac{2}{\beta}+\frac{\mathrm{Pe}^{2}}{\beta(1+\beta)}\,. (42)

C.2 ABP under complete stochastic resetting

The displacement fluctuation δ𝐫2r=𝐫2r𝐫r2\langle\delta{\bf r}^{2}\rangle_{r}=\langle{\bf r}^{2}\rangle_{r}-\langle{\bf r}\rangle_{r}^{2}. Here, we set trap strength β=0\beta=0.

At small times (t0t\to 0),

δ𝐫2r\displaystyle\langle\delta{\bf r}^{2}\rangle_{r} \displaystyle\simeq 4t2rt2+23(Pe2+r2)t3+𝒪(t4).\displaystyle 4t-2rt^{2}+\frac{2}{3}\left(\mathrm{Pe}^{2}+r^{2}\right)t^{3}+\mathcal{O}(t^{4})\,. (43)

In the steady state (tt\to\infty),

δ𝐫2rst\displaystyle\langle\delta{\bf r}^{2}\rangle^{st}_{r} =\displaystyle= 4r+4Pe2r(1+r)(2+r).\displaystyle\frac{4}{r}+\frac{4\mathrm{Pe}^{2}}{r(1+r)(2+r)}\,. (44)

C.3 Brownian particles under stochastic resetting in a harmonic trap

Here, we set activity Pe=0\mathrm{Pe}=0. The displacement fluctuation δ𝐫2r=𝐫2r𝐫r2=𝐫2r\langle\delta{\bf r}^{2}\rangle_{r}=\langle{\bf r}^{2}\rangle_{r}-\langle{\bf r}\rangle_{r}^{2}=\langle{\bf r}^{2}\rangle_{r}.

C.4 Brownian particles under stochastic resetting

Here, we set activity Pe=0\mathrm{Pe}=0 and trap strength β=0\beta=0. The displacement fluctuation δ𝐫2r=𝐫2r𝐫r2=𝐫2r\langle\delta{\bf r}^{2}\rangle_{r}=\langle{\bf r}^{2}\rangle_{r}-\langle{\bf r}\rangle_{r}^{2}=\langle{\bf r}^{2}\rangle_{r}.

Appendix D Second order moment and fluctuations along initial orientation direction

D.1 ABP in a harmonic trap

The initial orientation of the ABP along xx-direction 𝐮^0=x^{\bf\hat{u}}_{0}=\hat{x}. Using ψ=r2=x2\psi=r_{\parallel}^{2}=x^{2}, we get using equation (5),

(s+2β)r2s\displaystyle(s+2\beta)\langle r_{\parallel}^{2}\rangle_{s} =\displaystyle= 2/s+2Pexuxs.\displaystyle 2/s+2\mathrm{Pe}\langle xu_{x}\rangle_{s}\,. (45)

Now, we proceed to calculate the second term xuxs\langle xu_{x}\rangle_{s} assuming ψ=xux\psi=xu_{x}, we get using equation (5),

xux\displaystyle\langle xu_{x}\rangle =\displaystyle= Peux2ss+1+β.\displaystyle\frac{\mathrm{Pe}\langle u_{x}^{2}\rangle_{s}}{s+1+\beta}\,. (46)

Again, we need to calculate ux2s\langle u_{x}^{2}\rangle_{s}, substituting ψ=ux2\psi=u_{x}^{2} in equation (5), we get

ux2s\displaystyle\langle u_{x}^{2}\rangle_{s} =\displaystyle= s+2s(s+4).\displaystyle\frac{s+2}{s(s+4)}\,. (47)

Finally,

r2s\displaystyle\langle r_{\parallel}^{2}\rangle_{s} =\displaystyle= 1(s+2β)[2s+2Pe2(s+2)s(s+4)(s+1+β)].\displaystyle\frac{1}{(s+2\beta)}\left[\frac{2}{s}+\frac{2\mathrm{Pe}^{2}(s+2)}{s(s+4)(s+1+\beta)}\right]\,. (48)

Inverse Laplace transformation of the above equation leads to the r2(t)\langle r_{\parallel}^{2}(t)\rangle for ABP in a harmonic trap in the absence of stochastic resetting

r2(t)=e4tPe22(3+β)(2+β)2e(1+β)tPe2(3+β)(1+β)+e2βt(2β+Pe2)(2+β)β+2+2β+Pe22β(1+β).\displaystyle\langle r_{\parallel}^{2}(t)\rangle=\frac{e^{-4t}\mathrm{Pe}^{2}}{2(-3+\beta)(-2+\beta)}-\frac{2e^{-(1+\beta)t}\mathrm{Pe}^{2}}{(-3+\beta)(1+\beta)}+\frac{e^{-2\beta t}(2-\beta+\mathrm{Pe}^{2})}{(-2+\beta)\beta}+\frac{2+2\beta+\mathrm{Pe}^{2}}{2\beta(1+\beta)}\,.

In the small time limit (t0t\to 0), r2(t)\langle r_{\parallel}^{2}(t)\rangle gives

r2(t)2t+(2β+Pe2)t2+13(4β23Pe23βPe2)t3+𝒪(t4).\displaystyle\langle r_{\parallel}^{2}(t)\rangle\simeq 2t+(-2\beta+\mathrm{Pe}^{2})t^{2}+\frac{1}{3}\left(4\beta^{2}-3\mathrm{Pe}^{2}-3\beta\mathrm{Pe}^{2}\right)t^{3}+\mathcal{O}(t^{4})\,. (50)

In the steady state (tt\to\infty), we get

r2st\displaystyle\langle r_{\parallel}^{2}\rangle^{st} =\displaystyle= 2(1+β)+Pe22β(1+β).\displaystyle\frac{2(1+\beta)+\mathrm{Pe}^{2}}{2\beta(1+\beta)}\,. (51)

Displacement fluctuations along parallel direction δr2r=r2rrr2\langle\delta r_{\parallel}^{2}\rangle_{r}=\langle r_{\parallel}^{2}\rangle_{r}-\langle r_{\parallel}\rangle^{2}_{r}.

δr2(t)r\displaystyle\langle\delta r_{\parallel}^{2}(t)\rangle_{r} =12(e4tPe265β+β24e(1+β)tPe2(3+β)(1+β)2(eteβt)2Pe2(1+β)2\displaystyle=\frac{1}{2}\left(\frac{e^{-4t}\mathrm{Pe}^{2}}{6-5\beta+\beta^{2}}-\frac{4e^{-(1+\beta)t}\mathrm{Pe}^{2}}{(-3+\beta)(1+\beta)}-\frac{2(e^{-t}-e^{-\beta t})^{2}\mathrm{Pe}^{2}}{(-1+\beta)^{2}}\right. (52)
+2e2βt(2β+Pe2)(2+β)β+2+2β+Pe2β+β2).\displaystyle\left.+\frac{2e^{-2\beta t}(2-\beta+\mathrm{Pe}^{2})}{(-2+\beta)\beta}+\frac{2+2\beta+\mathrm{Pe}^{2}}{\beta+\beta^{2}}\right)\,.

In the small time limit (t0t\to 0), r2(t)r\langle r_{\parallel}^{2}(t)\rangle_{r} gives

δr2(t)r\displaystyle\langle\delta r_{\parallel}^{2}(t)\rangle_{r} \displaystyle\simeq 2t2βt2+4β2t33+𝒪(t4).\displaystyle 2t-2\beta t^{2}+\frac{4\beta^{2}t^{3}}{3}+\mathcal{O}(t^{4})\,. (53)

In the steady state (tt\to\infty), we get

δr2rst=r2st=2(1+β)+Pe22β(1+β).\displaystyle\langle\delta r_{\parallel}^{2}\rangle_{r}^{st}=\langle r_{\parallel}^{2}\rangle^{st}=\frac{2(1+\beta)+\mathrm{Pe}^{2}}{2\beta(1+\beta)}\,. (54)

Now the perpendicular component of displacement fluctuation δr2r=δ𝐫2rδr2r\langle\delta r_{\perp}^{2}\rangle_{r}=\langle\delta{\bf r}^{2}\rangle_{r}-\langle\delta r_{\parallel}^{2}\rangle_{r}.

δr2(t)r\displaystyle\langle\delta r_{\perp}^{2}(t)\rangle_{r} =e4tPe22(65β+β2)+4e(1+β)tPe23β3β2+β3+2+2β+Pe22β+2β2\displaystyle=-\frac{e^{-4t}\mathrm{Pe}^{2}}{2(6-5\beta+\beta^{2})}+\frac{4e^{-(1+\beta)t}\mathrm{Pe}^{2}}{3-\beta-3\beta^{2}+\beta^{3}}+\frac{2+2\beta+\mathrm{Pe}^{2}}{2\beta+2\beta^{2}} (55)
e2βt(23β+β2+Pe2)β(23β+β2).\displaystyle-\frac{e^{-2\beta t}(2-3\beta+\beta^{2}+\mathrm{Pe}^{2})}{\beta(2-3\beta+\beta^{2})}\,.

In the small time limit (t0t\to 0), r2(t)r\langle r_{\parallel}^{2}(t)\rangle_{r} gives

δr2(t)2t2βt2+23(2β2+Pe2)t316[4β3+5Pe2+3βPe2]t4+𝒪(t5).\displaystyle\langle\delta r_{\perp}^{2}(t)\rangle\simeq 2t-2\beta t^{2}+\frac{2}{3}(2\beta^{2}+\mathrm{Pe}^{2})t^{3}-\frac{1}{6}\left[4\beta^{3}+5\mathrm{Pe}^{2}+3\beta\mathrm{Pe}^{2}\right]t^{4}+\mathcal{O}(t^{5})\,. (56)

In the steady state (tt\to\infty), we get

δr2rst=δr2rst=r2st=2(1+β)+Pe22β(1+β).\displaystyle\langle\delta r_{\perp}^{2}\rangle_{r}^{st}=\langle\delta r_{\parallel}^{2}\rangle^{st}_{r}=\langle r_{\parallel}^{2}\rangle^{st}=\frac{2(1+\beta)+\mathrm{Pe}^{2}}{2\beta(1+\beta)}\,. (57)

D.2 ABP under stochastic resetting in a harmonic trap

Now, we get the r2(t)\langle r_{\parallel}^{2}(t)\rangle under stochastic position and orientation resetting using equation (6)

r2(t)r=ert(e4tPe22(3+β)(2+β)2e(1+β)tPe2(3+β)(1+β)+e2βt(2β+Pe2)(2+β)β+2+2β+Pe22β(1+β))\displaystyle\langle r_{\parallel}^{2}(t)\rangle_{r}=e^{-rt}\left(\frac{e^{-4t}\mathrm{Pe}^{2}}{2(-3+\beta)(-2+\beta)}-\frac{2e^{-(1+\beta)t}\mathrm{Pe}^{2}}{(-3+\beta)(1+\beta)}+\frac{e^{-2\beta t}(2-\beta+\mathrm{Pe}^{2})}{(-2+\beta)\beta}+\frac{2+2\beta+\mathrm{Pe}^{2}}{2\beta(1+\beta)}\right)
+β(22ert)e(2β+r)t(1+e2βt)rβ(2β+r)+rPe22[e(4+r)t(65β+β2)(4+r)+4e(1+β+r)t(3+β)(1+β)(1+β+r)\displaystyle+\frac{\beta(2-2e^{-rt})-e^{-(2\beta+r)t}(-1+e^{2\beta t})r}{\beta(2\beta+r)}+\frac{r\mathrm{Pe}^{2}}{2}\left[-\frac{e^{-(4+r)t}}{(6-5\beta+\beta^{2})(4+r)}+\frac{4e^{-(1+\beta+r)t}}{(-3+\beta)(1+\beta)(1+\beta+r)}\right.
2e(2β+r)t(2+β)β(2β+r)+4(2+r)r(4+r)(1+β+r)(2β+r)ertβr+β2r].\displaystyle\left.-\frac{2e^{-(2\beta+r)t}}{(-2+\beta)\beta(2\beta+r)}+\frac{4(2+r)}{r(4+r)(1+\beta+r)(2\beta+r)}-\frac{e^{-rt}}{\beta r+\beta^{2}r}\right]\,. (58)

In the small time limit (t0t\to 0), r2(t)r\langle r_{\parallel}^{2}(t)\rangle_{r} gives

r2(t)2t+(2β+Pe2r)t2+13(4β23Pe23βPe2+4βr2Pe2r+r2)t3\displaystyle\langle r_{\parallel}^{2}(t)\rangle\simeq 2t+(-2\beta+\mathrm{Pe}^{2}-r)t^{2}+\frac{1}{3}\left(4\beta^{2}-3\mathrm{Pe}^{2}-3\beta\mathrm{Pe}^{2}+4\beta r-2\mathrm{Pe}^{2}r+r^{2}\right)t^{3}
+𝒪(t4).\displaystyle+\mathcal{O}(t^{4})\,. (59)

In the steady state (tt\to\infty), we get

r2rst\displaystyle\langle r_{\parallel}^{2}\rangle_{r}^{st} =\displaystyle= 2[(1+r)(4+r)+Pe2(2+r)+β(4+r)](4+r)(1+β+r)(2β+r).\displaystyle\frac{2[(1+r)(4+r)+\mathrm{Pe}^{2}(2+r)+\beta(4+r)]}{(4+r)(1+\beta+r)(2\beta+r)}\,. (60)

Displacement fluctuations along parallel direction δr2r=r2rrr2\langle\delta r_{\parallel}^{2}\rangle_{r}=\langle r_{\parallel}^{2}\rangle_{r}-\langle r_{\parallel}\rangle^{2}_{r}.

δr2(t)r=12[ert(e4tPe265β+β24e(1+β)tPe2(3+β)(1+β)+2e2βt(2β+Pe2)(2+β)β+2+2β+Pe2β+β2)\displaystyle\langle\delta r_{\parallel}^{2}(t)\rangle_{r}=\frac{1}{2}\left[e^{-rt}\left(\frac{e^{-4t}\mathrm{Pe}^{2}}{6-5\beta+\beta^{2}}-\frac{4e^{-(1+\beta)t}\mathrm{Pe}^{2}}{(-3+\beta)(1+\beta)}+\frac{2e^{-2\beta t}(2-\beta+\mathrm{Pe}^{2})}{(-2+\beta)\beta}+\frac{2+2\beta+\mathrm{Pe}^{2}}{\beta+\beta^{2}}\right)\right.
2Pe2(rβr+βe(β+r)t(1+r)e(1+r)t(β+r))2(1+β)2(1+r)2(β+r)2\displaystyle\left.-\frac{2\mathrm{Pe}^{2}(r-\beta r+\beta e^{-(\beta+r)t}(1+r)-e^{-(1+r)t}(\beta+r))^{2}}{(-1+\beta)^{2}(1+r)^{2}(\beta+r)^{2}}\right.
+Pe2r(e(4+r)t(65β+β2)(4+r)+4e(1+β+r)t(3+β)(1+β)(1+β+r)2e(2β+r)t(2+β)β(2β+r)\displaystyle\left.+\mathrm{Pe}^{2}r\left(-\frac{e^{-(4+r)t}}{(6-5\beta+\beta^{2})(4+r)}+\frac{4e^{-(1+\beta+r)t}}{(-3+\beta)(1+\beta)(1+\beta+r)}-\frac{2e^{-(2\beta+r)t}}{(-2+\beta)\beta(2\beta+r)}\right.\right.
+4(2+r)r(4+r)(1+β+r)(2β+r)ertβr+β2r)+β(44ert)4e(β+r)trsinh(βt)β(2β+r)].\displaystyle\left.\left.+\frac{4(2+r)}{r(4+r)(1+\beta+r)(2\beta+r)}-\frac{e^{-rt}}{\beta r+\beta^{2}r}\right)+\frac{\beta(4-4e^{-rt})-4e^{-(\beta+r)t}r\sinh(\beta t)}{\beta(2\beta+r)}\right]\,. (61)

In the small time limit (t0t\to 0), r2(t)r\langle r_{\parallel}^{2}(t)\rangle_{r} gives

δr2(t)2t(2β+r)t2+13(4β2+4βr+Pe2r+r2)t3+𝒪(t4).\displaystyle\langle\delta r_{\parallel}^{2}(t)\rangle\simeq 2t-(2\beta+r)t^{2}+\frac{1}{3}\left(4\beta^{2}+4\beta r+\mathrm{Pe}^{2}r+r^{2}\right)t^{3}+\mathcal{O}(t^{4})\,. (62)

In the steady state (tt\to\infty), we get

δr2rst=Pe2r2(1+r)2(β+r)2+22β+r+2Pe2(2+r)(4+r)(1+β+r)(2β+r).\displaystyle\langle\delta r_{\parallel}^{2}\rangle_{r}^{st}=-\frac{\mathrm{Pe}^{2}r^{2}}{(1+r)^{2}(\beta+r)^{2}}+\frac{2}{2\beta+r}+\frac{2\mathrm{Pe}^{2}(2+r)}{(4+r)(1+\beta+r)(2\beta+r)}\,. (63)

D.3 ABP under stochastic resetting

r2(t)r=e(4+r)t3r(1+r)(4+r)(Pe2r(1+r)+2e3tPe2r(4+r)3e4t(2+Pe2)(1+r)(4+r))\displaystyle\langle r_{\parallel}^{2}(t)\rangle_{r}=\frac{e^{-(4+r)t}}{3r(1+r)(4+r)}\left(\mathrm{Pe}^{2}r(1+r)+2e^{3t}\mathrm{Pe}^{2}r(4+r)-3e^{4t}(2+\mathrm{Pe}^{2})(1+r)(4+r)\right)
+2[Pe2(2+r)+(1+r)(4+r)]r(1+r)(4+r).\displaystyle+\frac{2[\mathrm{Pe}^{2}(2+r)+(1+r)(4+r)]}{r(1+r)(4+r)}\,. (64)

In the small time limit (t0t\to 0), r2(t)r\langle r_{\parallel}^{2}(t)\rangle_{r} gives

r2(t)\displaystyle\langle r_{\parallel}^{2}(t)\rangle \displaystyle\simeq 2t+(Pe2r)t2+13(3Pe22Pe2r+r2)t3+𝒪(t4).\displaystyle 2t+(\mathrm{Pe}^{2}-r)t^{2}+\frac{1}{3}\left(-3\mathrm{Pe}^{2}-2\mathrm{Pe}^{2}r+r^{2}\right)t^{3}+\mathcal{O}(t^{4})\,. (65)

In the steady state (tt\to\infty), we get

r2rst\displaystyle\langle r_{\parallel}^{2}\rangle_{r}^{st} =\displaystyle= 2[Pe2(2+r)+(1+r)(4+r)]r(1+r)(4+r).\displaystyle\frac{2[\mathrm{Pe}^{2}(2+r)+(1+r)(4+r)]}{r(1+r)(4+r)}\,. (66)

Displacement fluctuations along parallel direction δr2r=r2rrr2\langle\delta r_{\parallel}^{2}\rangle_{r}=\langle r_{\parallel}^{2}\rangle_{r}-\langle r_{\parallel}\rangle^{2}_{r}.

δr2(t)r=e2(1+r)t(1+e(1+r)t)2Pe2(1+r)2+ert(2t+112Pe2(9+e4t+8et+12t))\displaystyle\langle\delta r_{\parallel}^{2}(t)\rangle_{r}=-\frac{e^{-2(1+r)t}(-1+e^{(1+r)t})^{2}\mathrm{Pe}^{2}}{(1+r)^{2}}+e^{-rt}\left(2t+\frac{1}{12}\mathrm{Pe}^{2}\left(-9+e^{-4t}+8e^{-t}+12t\right)\right)
+e(4+r)t12r(1+r)(4+r)[Pe2r2(1+r)8e3tPe2r2(4+r)\displaystyle+\frac{e^{-(4+r)t}}{12r(1+r)(4+r)}\left[-\mathrm{Pe}^{2}r^{2}(1+r)-8e^{3t}\mathrm{Pe}^{2}r^{2}(4+r)\right.
+24e(4+r)t(Pe2(2+r)+(1+r)(4+r))3e4t(1+r)(4+r)(8+8rt+Pe2(43r+4rt))].\displaystyle\left.+24e^{(4+r)t}(\mathrm{Pe}^{2}(2+r)+(1+r)(4+r))-3e^{4t}(1+r)(4+r)(8+8rt+\mathrm{Pe}^{2}(4-3r+4rt))\right]\,.

In the small time limit (t0t\to 0), r2(t)r\langle r_{\parallel}^{2}(t)\rangle_{r} gives

δr2(t)\displaystyle\langle\delta r_{\parallel}^{2}(t)\rangle \displaystyle\simeq 2trt2+r3(Pe2+r)t3+𝒪(t4).\displaystyle 2t-rt^{2}+\frac{r}{3}\left(\mathrm{Pe}^{2}+r\right)t^{3}+\mathcal{O}(t^{4})\,. (68)

In the steady state (tt\to\infty), we get

δr2rst\displaystyle\langle\delta r_{\parallel}^{2}\rangle_{r}^{st} =\displaystyle= 2(1+r)2(4+r)+Pe2(4+2r+r2)r(1+r)2(4+r).\displaystyle\frac{2(1+r)^{2}(4+r)+\mathrm{Pe}^{2}(4+2r+r^{2})}{r(1+r)^{2}(4+r)}\,. (69)

D.4 Brownian particle under stochastic resetting in a harmonic trap

In the absence of activity (Pe=0\mathrm{Pe}=0), simplifies to Brownian particle under stochastic resetting

r2(t)r\displaystyle\langle r_{\parallel}^{2}(t)\rangle_{r} =\displaystyle= 22β+r(1e(2β+r)t).\displaystyle\frac{2}{2\beta+r}\left(1-e^{-(2\beta+r)t}\right)\,. (70)

In the small time limit (t0t\to 0), r2(t)r\langle r_{\parallel}^{2}(t)\rangle_{r} gives the expansion for Brownian particle under stochastic resetting

r2(t)r\displaystyle\langle r_{\parallel}^{2}(t)\rangle_{r} \displaystyle\simeq 2t(2β+r)t2+13(4β2+4βr+r2)t3+𝒪(t4).\displaystyle 2t-(2\beta+r)t^{2}+\frac{1}{3}\left(4\beta^{2}+4\beta r+r^{2}\right)t^{3}+\mathcal{O}(t^{4})\,. (71)

In the steady state (tt\to\infty), we get the expression for Brownian particle under stochastic resetting

r2rst\displaystyle\langle r_{\parallel}^{2}\rangle^{st}_{r} =\displaystyle= 22β+r.\displaystyle\frac{2}{2\beta+r}\,. (72)

Displacement fluctuations along parallel direction δr2r=r2rrr2=r2r\langle\delta r_{\parallel}^{2}\rangle_{r}=\langle r_{\parallel}^{2}\rangle_{r}-\langle r_{\parallel}\rangle^{2}_{r}=\langle r_{\parallel}^{2}\rangle_{r}.

D.5 Brownian particle under stochastic resetting

In the absence of activity (Pe=0\mathrm{Pe}=0), simplifies to Brownian particle under stochastic resetting

r2(t)r\displaystyle\langle r_{\parallel}^{2}(t)\rangle_{r} =\displaystyle= 2r(1ert).\displaystyle\frac{2}{r}\left(1-e^{-rt}\right)\,. (73)

In the small time limit (t0t\to 0), r2(t)r\langle r_{\parallel}^{2}(t)\rangle_{r} gives the expansion for Brownian particle under stochastic resetting

r2(t)r\displaystyle\langle r_{\parallel}^{2}(t)\rangle_{r} \displaystyle\simeq 2trt2+r2t33+𝒪(t4).\displaystyle 2t-rt^{2}+\frac{r^{2}t^{3}}{3}+\mathcal{O}(t^{4})\,. (74)

In the steady state (tt\to\infty), we get the expression for Brownian particle under stochastic resetting

r2rst\displaystyle\langle r_{\parallel}^{2}\rangle^{st}_{r} =\displaystyle= 2r.\displaystyle\frac{2}{r}\,. (75)

Displacement fluctuations along parallel direction δr2r=r2rrr2=r2r\langle\delta r_{\parallel}^{2}\rangle_{r}=\langle r_{\parallel}^{2}\rangle_{r}-\langle r_{\parallel}\rangle^{2}_{r}=\langle r_{\parallel}^{2}\rangle_{r}.

Appendix E Detailed derivation of fourth order moment of displacement

Here, we show the detailed derivation of fourth moment of displacement under stochastic position and orientation resetting.

We get the fourth order moment of displacement in Laplace space 𝐫4s\langle{\bf r}^{4}\rangle_{s} in the absence of stochastic resetting (r=0r=0) with initial position at origin 𝐫0=0{\bf r}_{0}=0 using equation (5), gives

𝐫4s\displaystyle\langle{\bf r}^{4}\rangle_{s} =\displaystyle= 1s+4β[16𝐫2s+4Pe(𝐮^𝐫)𝐫2s].\displaystyle\frac{1}{s+4\beta}\left[16\langle{\bf r}^{2}\rangle_{s}+4\mathrm{Pe}\langle({\bf\hat{u}}\cdot{\bf r}){\bf r}^{2}\rangle_{s}\right]\,. (76)

Further, we proceed to calculate (𝐮^𝐫)𝐫2s\langle({\bf\hat{u}}\cdot{\bf r}){\bf r}^{2}\rangle_{s},

(𝐮^𝐫)𝐫2s\displaystyle\langle({\bf\hat{u}}\cdot{\bf r}){\bf r}^{2}\rangle_{s} =\displaystyle= [8𝐮^𝐫s+Pe𝐫2s+2Pe(𝐮^𝐫)2s](s+1+3β).\displaystyle\frac{\left[8\langle{\bf\hat{u}}\cdot{\bf r}\rangle_{s}+\mathrm{Pe}\langle{\bf r}^{2}\rangle_{s}+2\mathrm{Pe}\langle({\bf\hat{u}}\cdot{\bf r})^{2}\rangle_{s}\right]}{(s+1+3\beta)}\,. (77)

The first two quantities 𝐮^𝐫s\langle{\bf\hat{u}}\cdot{\bf r}\rangle_{s} and 𝐫2s\langle{\bf r}^{2}\rangle_{s} already calculated. The third term (𝐮^𝐫)2s\langle({\bf\hat{u}}\cdot{\bf r})^{2}\rangle_{s} calculated as

(𝐮^𝐫)2s\displaystyle\langle({\bf\hat{u}}\cdot{\bf r})^{2}\rangle_{s} =\displaystyle= [2/s+2𝐫2s+2Pe𝐮^𝐫s](s+4+2β).\displaystyle\frac{\left[2/s+2\langle{\bf r}^{2}\rangle_{s}+2\mathrm{Pe}\langle{\bf\hat{u}}\cdot{\bf r}\rangle_{s}\right]}{(s+4+2\beta)}\,. (78)

Finally, we get 𝐫4s\langle{\bf r}^{4}\rangle_{s} in Laplace space

𝐫4s\displaystyle\langle{\bf r}^{4}\rangle_{s} =\displaystyle= 8s3[8D2+Dv024(3s+2Dr)(s+Dr)2+v043s+8Dr(s+Dr)2(s+4Dr)].\displaystyle\frac{8}{s^{3}}\left[8D^{2}+Dv_{0}^{2}\frac{4(3s+2D_{r})}{(s+D_{r})^{2}}+v_{0}^{4}\frac{3s+8D_{r}}{(s+D_{r})^{2}(s+4D_{r})}\right]\,. (79)

Inverse Laplace transformation of equation (79) leads to 𝐫4(t)\langle{\bf r}^{4}(t)\rangle for ABP in a harmonic trap without stochastic resetting already explored in [30]

𝐫4(t)=2Pe4e2(2+β)t(3+β)(2+β)(2+β)(3+β)4e2βt[4+4β24Pe2Pe4](1+β)β2(1+β)\displaystyle\langle{\bf r}^{4}(t)\rangle=\frac{2\mathrm{Pe}^{4}e^{-2(2+\beta)t}}{(-3+\beta)(-2+\beta)(2+\beta)(3+\beta)}-\frac{4e^{-2\beta t}[-4+4\beta^{2}-4\mathrm{Pe}^{2}-\mathrm{Pe}^{4}]}{(-1+\beta)\beta^{2}(1+\beta)}
+4e(1+3β)t[12Pe232βPe2+12β2Pe2+5Pe43βPe4](3+β)(1+β)β(1+β)(1+3β)\displaystyle+\frac{4e^{-(1+3\beta)t}[-12\mathrm{Pe}^{2}-32\beta\mathrm{Pe}^{2}+12\beta^{2}\mathrm{Pe}^{2}+5\mathrm{Pe}^{4}-3\beta\mathrm{Pe}^{4}]}{(-3+\beta)(-1+\beta)\beta(1+\beta)(1+3\beta)}
+e4βt[16+72β80β2+24β316Pe2+56βPe224β2Pe24Pe4+3βPe4](2+β)(1+β)β2(1+3β)\displaystyle+\frac{e^{-4\beta t}[-16+72\beta-80\beta^{2}+24\beta^{3}-16\mathrm{Pe}^{2}+56\beta\mathrm{Pe}^{2}-24\beta^{2}\mathrm{Pe}^{2}-4\mathrm{Pe}^{4}+3\beta\mathrm{Pe}^{4}]}{(-2+\beta)(-1+\beta)\beta^{2}(-1+3\beta)}
+16+72β+80β2+24β3+16Pe2+56βPe2+24β2Pe2+4Pe4+3βPe4β2(1+β)(2+β)(1+3β)\displaystyle+\frac{16+72\beta+80\beta^{2}+24\beta^{3}+16\mathrm{Pe}^{2}+56\beta\mathrm{Pe}^{2}+24\beta^{2}\mathrm{Pe}^{2}+4\mathrm{Pe}^{4}+3\beta\mathrm{Pe}^{4}}{\beta^{2}(1+\beta)(2+\beta)(1+3\beta)}
4e(1+β)t[12Pe2+32βPe2+12β2Pe2+5Pe4+3βPe4](1+β)β(1+β)(3+β)(1+3β).\displaystyle-\frac{4e^{-(1+\beta)t}[-12\mathrm{Pe}^{2}+32\beta\mathrm{Pe}^{2}+12\beta^{2}\mathrm{Pe}^{2}+5\mathrm{Pe}^{4}+3\beta\mathrm{Pe}^{4}]}{(-1+\beta)\beta(1+\beta)(3+\beta)(-1+3\beta)}\,. (80)

In presence of resetting, substituting ψ=𝐫4\psi={\bf r}^{4} in equation (6) and using equation (80), we get

𝐫4(t)r=𝐫4(t)ert+24β3+4(2+Pe2)2+8β2(10+3Pe2)+β(72+56Pe2+3Pe4)β2(1+β)(2+β)(1+3β)\displaystyle\langle{\bf r}^{4}(t)\rangle_{r}=\langle{\bf r}^{4}(t)\rangle e^{-rt}+\frac{24\beta^{3}+4(2+\mathrm{Pe}^{2})^{2}+8\beta^{2}(10+3\mathrm{Pe}^{2})+\beta(72+56\mathrm{Pe}^{2}+3\mathrm{Pe}^{4})}{\beta^{2}(1+\beta)(2+\beta)(1+3\beta)}
4rPe2[12+12β2+5Pe2+β(32+3Pe2)](1+β)β(1+β)(3+β)(1+3β)(1+β+r)+4r[4β2+(2+Pe2)2](1+β)β2(1+β)(2β+r)\displaystyle-\frac{4r\mathrm{Pe}^{2}[-12+12\beta^{2}+5\mathrm{Pe}^{2}+\beta(32+3\mathrm{Pe}^{2})]}{(-1+\beta)\beta(1+\beta)(3+\beta)(-1+3\beta)(1+\beta+r)}+\frac{4r[-4\beta^{2}+(2+\mathrm{Pe}^{2})^{2}]}{(-1+\beta)\beta^{2}(1+\beta)(2\beta+r)}
+2rPe4(3613β2+β4)(4+2β+r)+4rPe2[12+12β2+5Pe2β(32+3Pe2)](3+β)(1+β)β(1+β)(1+3β)(1+3β+r)\displaystyle+\frac{2r\mathrm{Pe}^{4}}{(36-13\beta^{2}+\beta^{4})(4+2\beta+r)}+\frac{4r\mathrm{Pe}^{2}[-12+12\beta^{2}+5\mathrm{Pe}^{2}-\beta(32+3\mathrm{Pe}^{2})]}{(-3+\beta)(-1+\beta)\beta(1+\beta)(1+3\beta)(1+3\beta+r)}
+24rβ34(2+Pe2)28β2(10+3Pe2)+β(72+56Pe2+3Pe4)(2+β)(1+β)β2(1+3β)(4β+r)\displaystyle+\frac{24r\beta^{3}-4(2+\mathrm{Pe}^{2})^{2}-8\beta^{2}(10+3\mathrm{Pe}^{2})+\beta(72+56\mathrm{Pe}^{2}+3\mathrm{Pe}^{4})}{(-2+\beta)(-1+\beta)\beta^{2}(-1+3\beta)(4\beta+r)}
[24β3+4(2+Pe2)2+8β2(10+3Pe2)+β(72+56Pe2+3Pe4)]ertβ2(1+β)(2+β)(1+3β)\displaystyle-\frac{[24\beta^{3}+4(2+\mathrm{Pe}^{2})^{2}+8\beta^{2}(10+3\mathrm{Pe}^{2})+\beta(72+56\mathrm{Pe}^{2}+3\mathrm{Pe}^{4})]e^{-rt}}{\beta^{2}(1+\beta)(2+\beta)(1+3\beta)}
+4rPe2e(1+r+β)t[12+12β2+5Pe2+β(32+3Pe2)](1+β)β(1+β)(3+β)(1+3β)(1+β+r)\displaystyle+\frac{4r\mathrm{Pe}^{2}e^{-(1+r+\beta)t}[-12+12\beta^{2}+5\mathrm{Pe}^{2}+\beta(32+3\mathrm{Pe}^{2})]}{(-1+\beta)\beta(1+\beta)(3+\beta)(-1+3\beta)(1+\beta+r)}
+4re(r+2β)t[4β2(2+Pe2)2](1+β)β2(1+β)(2β+r)2rPe4e[2(2+β)+r]t(3613β2+β4)(4+2β+r)\displaystyle+\frac{4re^{-(r+2\beta)t}[4\beta^{2}-(2+\mathrm{Pe}^{2})^{2}]}{(-1+\beta)\beta^{2}(1+\beta)(2\beta+r)}-\frac{2r\mathrm{Pe}^{4}e^{-[2(2+\beta)+r]t}}{(36-13\beta^{2}+\beta^{4})(4+2\beta+r)}
+4rPe2e(1+r+3β)t[1212β25Pe2+β(32+3Pe2)](3+β)(1+β)β(1+β)(1+3β)(1+3β+r)\displaystyle+\frac{4r\mathrm{Pe}^{2}e^{-(1+r+3\beta)t}[12-12\beta^{2}-5\mathrm{Pe}^{2}+\beta(32+3\mathrm{Pe}^{2})]}{(-3+\beta)(-1+\beta)\beta(1+\beta)(1+3\beta)(1+3\beta+r)}
+re(r+4β)t[24β3+4(2+Pe2)2+8β2(10+3Pe2)β(72+56Pe2+3Pe4)](2+β)(1+β)β2(1+3β)(4β+r).\displaystyle+\frac{re^{-(r+4\beta)t}[-24\beta^{3}+4(2+\mathrm{Pe}^{2})^{2}+8\beta^{2}(10+3\mathrm{Pe}^{2})-\beta(72+56\mathrm{Pe}^{2}+3\mathrm{Pe}^{4})]}{(-2+\beta)(-1+\beta)\beta^{2}(-1+3\beta)(4\beta+r)}\,. (81)

At small time (t0t\to 0),

𝐫4r32t2+163(3Pe24r12β)t3\displaystyle\langle{\bf r}^{4}\rangle_{r}\simeq 32t^{2}+\frac{16}{3}\left(3\mathrm{Pe}^{2}-4r-12\beta\right)t^{3}
+13[224β2+3Pe4+24r2+144βr96βPe216Pe236rPe2]t4+𝒪(t5).\displaystyle+\frac{1}{3}\left[224\beta^{2}+3\mathrm{Pe}^{4}+24r^{2}+144\beta r-96\beta\mathrm{Pe}^{2}-16\mathrm{Pe}^{2}-36r\mathrm{Pe}^{2}\right]t^{4}+\mathcal{O}(t^{5})\,. (82)

In the absence of activity by substituting Pe=0\mathrm{Pe}=0 in equation (82), we get small time behavior of a Brownian particle in two dimensions under resetting

𝐫4\displaystyle\langle{\bf r}^{4}\rangle \displaystyle\simeq 2t264(r+3β)t33+[224β2+24r2+144βr]t43+𝒪(t5).\displaystyle 2t^{2}-\frac{64\left(r+3\beta\right)t^{3}}{3}+\frac{\left[224\beta^{2}+24r^{2}+144\beta r\right]t^{4}}{3}+\mathcal{O}(t^{5})\,. (83)

In the steady state(tt\to\infty),

𝐫4rst=8(1+β+r)(2β+r)(4+2β+r)(1+3β+r)(4β+r)×\displaystyle\langle{\bf r}^{4}\rangle^{st}_{r}=\frac{8}{(1+\beta+r)(2\beta+r)(4+2\beta+r)(1+3\beta+r)(4\beta+r)}\times
[48β3+8(1+r)2(4+r)+Pe4(8+3r)+8β2(20+6Pe2+11r)\displaystyle\left[48\beta^{3}+8(1+r)^{2}(4+r)+\mathrm{Pe}^{4}(8+3r)+8\beta^{2}(20+6\mathrm{Pe}^{2}+11r)\right.
+4Pe2(8+14r+3r2)+2β[3Pe4+8Pe2(7+3r)+24(3+4r+r2)]].\displaystyle\left.+4\mathrm{Pe}^{2}(8+14r+3r^{2})+2\beta\left[3\mathrm{Pe}^{4}+8\mathrm{Pe}^{2}(7+3r)+24(3+4r+r^{2})\right]\right]\,. (84)

In the absence of activity (Pe=0\mathrm{Pe}=0), Brownian particle in two dimensions under resetting in a harmonic trap

𝐫4rst=8[48β3+8(1+r)2(4+r)+8β2(20+11r)+48β(3+4r+r2)](1+β+r)(2β+r)(4+2β+r)(1+3β+r)(4β+r).\displaystyle\langle{\bf r}^{4}\rangle^{st}_{r}=\frac{8\left[48\beta^{3}+8(1+r)^{2}(4+r)+8\beta^{2}(20+11r)+48\beta(3+4r+r^{2})\right]}{(1+\beta+r)(2\beta+r)(4+2\beta+r)(1+3\beta+r)(4\beta+r)}\,. (85)

In the absence of activity (Pe=0\mathrm{Pe}=0) and harmonic trap (β=0\beta=0), Brownian particle in two dimensions under resetting, 𝐫4rst=64/r2\langle{\bf r}^{4}\rangle^{st}_{r}=64/r^{2}.

Appendix F Limiting cases of excess kurtosis

At small time (t0t\to 0), the kurtosis in equation (19) results

𝒦rrt3(3Pe4+16βr4rPe2)96t2+12880[45Pe6+60Pe4+32r2Pe2+64β2r\displaystyle\mathcal{K}_{r}\simeq\frac{rt}{3}-\frac{(3\mathrm{Pe}^{4}+16\beta r-4r\mathrm{Pe}^{2})}{96}t^{2}+\frac{1}{2880}\left[45\mathrm{Pe}^{6}+60\mathrm{Pe}^{4}+32r^{2}\mathrm{Pe}^{2}+64\beta^{2}r\right.
+24βrPe242rPe456rPe232r3128βr2]t3+𝒪(t4).\displaystyle\left.+24\beta r\mathrm{Pe}^{2}-42r\mathrm{Pe}^{4}-56r\mathrm{Pe}^{2}-32r^{3}-128\beta r^{2}\right]t^{3}+\mathcal{O}(t^{4})\,. (86)

In the absence of activity (Pe=0\mathrm{Pe}=0), simplifies to Brownian particle under stochastic resetting in two dimensions,

𝒦rrt3βrt26+r[2β2r24βr]90t3+βr[4β2+12βr+3r2]360+𝒪(t5).\displaystyle\mathcal{K}_{r}\simeq\frac{rt}{3}-\frac{\beta rt^{2}}{6}+\frac{r\left[2\beta^{2}-r^{2}-4\beta r\right]}{90}t^{3}+\frac{\beta r\left[4\beta^{2}+12\beta r+3r^{2}\right]}{360}+\mathcal{O}(t^{5})\,. (87)

In the absence of resetting rate (r=0r=0), simplifies to ABP in a harmonic trap in two dimensions,

𝒦rPe4t232+Pe4(4+3Pe2)t3192Pe4[136120β2+360Pe2+135Pe4]23040t4\displaystyle\mathcal{K}_{r}\simeq-\frac{\mathrm{Pe}^{4}t^{2}}{32}+\frac{\mathrm{Pe}^{4}(4+3\mathrm{Pe}^{2})t^{3}}{192}-\frac{\mathrm{Pe}^{4}\left[136-120\beta^{2}+360\mathrm{Pe}^{2}+135\mathrm{Pe}^{4}\right]}{23040}t^{4}
+𝒪(t5).\displaystyle+\mathcal{O}(t^{5})\,. (88)
𝒦rst(Pe0)rr+4β+Pe2(r2+2βr)(1+β+r)(1+3β+r)(4β+r)+𝒪(Pe3).\displaystyle\mathcal{K}^{st}_{r}(\mathrm{Pe}\to 0)\simeq\frac{r}{r+4\beta}+\frac{\mathrm{Pe}^{2}(r^{2}+2\beta r)}{(1+\beta+r)(1+3\beta+r)(4\beta+r)}+\mathcal{O}(\mathrm{Pe}^{3})\,. (89)

Appendix G ABP under stochastic resetting without harmonic trap

Refer to caption
Figure 6: ABP under stochastic resetting without harmonic trap (β=0\beta=0). (a) Fourth order moment of displacement 𝐫4r\langle{\bf r}^{4}\rangle_{r} (equation (81)) and (b) Kurtosis 𝒦r\mathcal{K}_{r} as a function of time tt for varying resetting rates r=1(squares,),10(triangles,),100(circles,)r=1(\mathrm{squares},\Box),10(\mathrm{triangles},\triangle),100(\mathrm{circles},\circ) with Pe=10\mathrm{Pe}=10. The solids lines are analytic predictions and symbols are from numerical simulations. The initial position is at the origin with the initial orientation along the xx-axis.

Figure 6 presents the fourth-order displacement moment in (a) and the excess kurtosis in (b) as a function of time of ABP under stochastic resetting without a harmonic trap (β=0\beta=0). In figure 6(a), the analytic fourth-order moment of displacement, 𝐫4r\langle{\bf r}^{4}\rangle_{r}, is shown as solid lines, compared to simulation results (points) for resetting rates r=1,10,100r=1,~{}10,~{}100 with Pe=10\mathrm{Pe}=10. For low resetting rates (r=1r=1), ballistic behavior 𝐫4rt3\langle{\bf r}^{4}\rangle_{r}\sim t^{3} is observed in the intermediate time regime. In figure 6(b), the analytic excess kurtosis, 𝒦r\mathcal{K}_{r}, is shown as solid lines, compared to simulation results (points) for the same resetting rates. Negative kurtosis is observed in the intermediate time regime for r=1r=1.

Refer to caption
Refer to caption
Refer to caption
Figure 7: Steady state behavior of ABP under stochastic resetting without harmonic trap (β=0\beta=0). Steady state excess kurtosis 𝒦rst\mathcal{K}^{st}_{r} (equation (22)) as a function of resetting rates (rr) for activities Pe=0.1,1,10\mathrm{Pe}=0.1,~{}1,~{}10 in (a) and as a function of Pe\mathrm{Pe} for r=0.1,1,10r=0.1,~{}1,~{}10 in (b). (c) Plot of 𝒦rst\mathcal{K}^{st}_{r} in rPer-\mathrm{Pe} plane.

Figure 7 presents the steady-state quantification of ABP under stochastic resetting without a harmonic trap (β=0\beta=0). In figure 7(a), we show steady state kurtosis 𝒦rst\mathcal{K}^{st}_{r} (equation (22)) as a function of rr for Pe=0.1,1,10\mathrm{Pe}=0.1,~{}1,~{}10. In figure 7(b), we show the 𝒦rst\mathcal{K}^{st}_{r} (equation (22)) as a function of Pe\mathrm{Pe} for r=0.1,1,10r=0.1,~{}1,~{}10. We see the non-monotonic behavior of 𝒦rst\mathcal{K}^{st}_{r} as a function of rr with high activity Pe\mathrm{Pe} (figure 7(a)). It demonstrates the presence of an optimal resetting rate where 𝒦rst\mathcal{K}^{st}_{r} is maximized for fixed activity, suggesting a maximum heavy tail in the position distribution, which increases the likelihood of finding the particle at a very long distance. We can also observe non-monotonic in 𝒦rst\mathcal{K}^{st}_{r} as a function of Pe\mathrm{Pe} for constant rr, which peaks at intermediate Pe\mathrm{Pe} (figure 7(b)). In figure 7(c), we show the 𝒦rst\mathcal{K}^{st}_{r} in rPer-\mathrm{Pe} plane, exhibits this non-monotonic behavior in both parameters rr and Pe\mathrm{Pe}.

We found that at intermediate timescales, high activity (Pe\mathrm{Pe}) and low resetting rate (rr) result in persistent behavior, characterized by bimodal distributions and negative kurtosis. At longer times, the system reaches a nonequilibrium steady state (NESS), where moments stabilize. The steady-state kurtosis shows non-monotonic behavior, deviating from the positive kurtosis value of 11 seen in Brownian particles under resetting, across different values of both activity (Pe\mathrm{Pe}) and resetting rate (rr).

References

References