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

Optimizing the random search of a finite-lived target by a Lévy flight

Denis Boyer1 boyer@fisica.unam.mx    Gabriel Mercado-Vásquez2 gabrielmv.fisica@gmail.com    Satya N. Majumdar3 satya.majumdar@universite-paris-saclay.fr    Grégory Schehr4 gregory.schehr@u-psud.fr 1Instituto de Física, Universidad Nacional Autónoma de México, Ciudad de México 04510, México 2Pritzker School of Molecular Engineering, University of Chicago, Chicago, IL, 60637, USA 3LPTMS, CNRS, Univ. Paris-Sud, Université Paris-Saclay, 91405 Orsay, France 4 Sorbonne Université, Laboratoire de Physique Théorique et Hautes Energies, CNRS UMR 7589, 4 Place Jussieu, 75252 Paris Cedex 05, France
Abstract

In many random search processes of interest in chemistry, biology or during rescue operations, an entity must find a specific target site before the latter becomes inactive, no longer available for reaction or lost. We present exact results on a minimal model system, a one-dimensional searcher performing a discrete time random walk or Lévy flight. In contrast with the case of a permanent target, the capture probability and the conditional mean first passage time can be optimized. The optimal Lévy index takes a non-trivial value, even in the long lifetime limit, and exhibits an abrupt transition as the initial distance to the target is varied. Depending on the target lifetime, this transition is discontinuous or continuous, separated by a non-conventional tricritical point. These results pave the way to the optimization of search processes under time constraints.

Random search processes are ubiquitous in nature, such as animals searching for food [1, 2], rescue operations looking for survivors after a shipwreck [3, 4] or even searches for a lost object like a key or a wallet. In typical search models, one considers the targets to be “immortal”, i.e., they do not disappear after a certain time. During the last decades, several models of random search of infinitely lived targets have been studied. The most popular among them is the search by a random walker, either diffusive or performing Lévy flights where the jumps are long-ranged. Several strategies have been incorporated to make the search by a random walker optimal. Lévy walks with certain exponent values can maximize the capture rate by a forager of dispersed resources [5, 6, 7, 8, 9, 10, 11, 12]. Another well known strategy is the intermittent search process where short range and long range moves alternate to locate a single target [13, 14]. A popular model that has received much attention in recent years is a resetting random walker, where the walker goes back to its starting point with a finite probability after every step and restarts the search process [15, 16, 17, 18, 19, 20, 21, 22]. In this case, it turns out that the mean time to find an infinitely lived target can be minimized by choosing an optimal resetting probability [15, 16, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 21]. This fact has also been verified in recent experiments in optical traps [33, 34, 35].

Refer to caption
Figure 1: A searcher, performing a Lévy flight in one-dimension, is looking for a non-permanent target (i.e., a ripe fruit) located at the origin. At each time step, the target (in red) stays active with probability a<1a<1, while the searcher performs a random step. If the searcher finds the target in the active state, the search is successful (orange trajectory). In contrast, if the target dies (rots) before being found by the searcher, the search is unsuccessful (blue trajectory).

However, there are many instances where the target has a finite but random lifetime. For instance, ripe fruits in a tree rot in a few days. The lifetime of a fruit is typically random since it depends on the nature of the tree and the weather [36]. Similarly, after a shipwreck, a survivor can last in the water only a finite amount of time, which is random as it depends on the general health of the person and sea conditions [37]. Inside a cell, target sites along the DNA are often blocked for long periods of time, which gives a limited random time to the transcription factors to bind to them [38, 39, 40]. In many examples, the searcher has to capture the target before it disappears or dies. Alternatively, in a dual view, one can consider the target as permanent and the walker with a strong time constraint, as an aerial rescue vehicle having a limited flight time [41]. The termination of the search at a random time also appears in the context of foraging theory, where a searcher abandons a patch at any time with a certain give up probability [42]. For a mortal searcher performing a lattice random walk [43] or Brownian motion [44], the capture probability and conditional mean first-passage time cannot be optimized, or only with an infinite diffusion coefficient. If a resetting mechanism is further implemented, though, a non-zero resetting rate can be optimal provided the mortality rate is not too high [45, 46].

A general question then is: is there any way to optimize the search success for a non-permanent target with a random lifetime? A natural generalization of the Brownian case is to investigate the search by a Lévy flight with a Lévy exponent 0<μ<20<\mu<2. One can then ask whether there is an optimal value of μ\mu that minimizes the conditional search time or, alternatively, maximizes the capture probability of the mortal target. In this Letter, we address this problem for a one-dimensional Lévy flight (see Fig. 1). In our model, the target is fixed at the origin and its lifetime nn is distributed geometrically via p(n)=(1a)anp(n)=(1-a)\,a^{n} where 0<a<10<a<1, i.e., at each discrete step, the target dies with probability 1a1-a and keeps alive with the complementary probability aa. We assume that the Lévy searcher starts from x0>0x_{0}>0 and subsequently evolves in discrete time via

xn=xn1+ηnx_{n}=x_{n-1}+\eta_{n} (1)

where ηn\eta_{n}’s are independent and identically distributed jump variables, each distributed via the probability distribution function f(η)f(\eta), which we assume to be symmetric and continuous with a power-law tail 1/|η|1+μ\propto 1/|\eta|^{1+\mu} where μ(0,2)\mu\in(0,2). Note that both parameters x0x_{0} and aa are given numbers and the searcher has no control in optimizing with respect to them. Thus the only parameter that the searcher has in her disposal to optimize is μ\mu, since it is associated with her motion. The search is successful only if the walker crosses the origin for the first time (takes xn<0x_{n}<0) while the target is still alive. We characterize the search success by two different observables: (i) the capture probability of the target and (ii) the conditional mean first-passage time (CMFPT), i.e., the mean search time conditioned to finding the target alive. We find that, for fixed x0x_{0} and aa, these two quantities can be optimized by varying the Lévy index μ\mu. The two optimal parameters μcap(x0,a)\mu_{cap}^{\star}(x_{0},a) and μFP(x0,a)\mu_{FP}^{\star}(x_{0},a) exhibit very rich phase diagrams in the (x0,a)(x_{0},a) plane.

Our results, obtained analytically and numerically, are summarized schematically in Fig. 2 for the capture probability. For any fixed a<a1=2e(152)/11=0.925690a<a_{1}=2\,e\,(\sqrt{15}-2)/11=0.925690\ldots, the index μcap(x0,a)\mu^{\star}_{cap}(x_{0},a) decreases monotonically as a function of x0x_{0}, and jumps to zero abruptly at a critical value x0=xc(a)x_{0}=x_{c}(a). This signals a first-order transition. In contrast, for any a>a1a>a_{1}, μcap(x0,a)\mu^{\star}_{cap}(x_{0},a) again decreases with x0x_{0} but vanishes continuously at xc(a)x_{c}(a), signaling a second-order transition. In the case a>a1a>a_{1}, the critical value xc(a)x_{c}(a) freezes to a constant value xc(a)=xm=0.561459x_{c}(a)=x_{m}=0.561459\ldots. Thus (xm,a1)(x_{m},a_{1}), shown by a red dot in Fig. 2, is a tricritical point that sits at the junction of a 1st1^{\rm st} and 2nd2^{\rm nd}-order transition. The green line x00x_{0}\to 0 is obtained analytically in the Supplemental Material [47]. A qualitatively similar diagram can be drawn for μFP(x0,a)\mu^{\star}_{FP}(x_{0},a), with a tricritical point at a slightly larger value a2=0.973989a_{2}=0.973989\ldots [47].

Refer to caption
Figure 2: Schematic phase diagram of the optimal Lévy index μcap\mu_{cap}^{*} in the (x0,a)(x_{0},a) plane. For fixed aa, as a function of x0x_{0}, the optimal μcap\mu_{cap}^{*} undergoes a first-order transition at x0=xc(a)x_{0}=x_{c}(a) (for a<a1a<a_{1}) which changes to a 2nd2^{\rm nd}-order transition for a>a1a>a_{1}. The critical line xc(a)x_{c}(a) freezes to xm=0.561459x_{m}=0.561459\ldots for a>a1a>a_{1}. The point that separates the first-order and second-order transitions is a tricritical point (shown by the red dot).
Refer to caption
Refer to caption
Refer to caption
Figure 3: (a) Discontinuous transition with short-lived targets (a=0.5a=0.5): numerical Q~μ(x0,a)\widetilde{Q}_{\mu}(x_{0},a) vs. μ\mu for different starting positions close to xmx_{m}. (b) Continuous transition for long-lived targets (aa close to 1): 1aQ~μ(x0,a)\sqrt{1-a}\widetilde{Q}_{\mu}(x_{0},a) as a function of μ\mu and for several x0x_{0} around xmx_{m}. In (a) and (b) the dotted lines represent the concavity approximation (12). (c) Optimal exponent for the CMFPT as a function of x0x_{0} for various aa. Below a2=0.973989a_{2}=0.973989... the transition is discontinuous (a=0.97a=0.97), while it is continuous above (a=0.98a=0.98). The dots correspond to the minima in (b). The index μcap(x0,a)\mu^{\star}_{cap}(x_{0},a) has analogous variations near a1a_{1}.

Both observables, the capture probability and the CMFPT, can be related to one fundamental quantity Qμ(x0,n)Q_{\mu}(x_{0},n) associated with the random walk, denoting the probability that a Lévy walker with index μ\mu, starting at x00x_{0}\geq 0, does not cross 0 up to step nn [48, 49, 50, 51, 52, 53, 54, 55, 56, 57]. Consequently, Qμ(x0,n1)Qμ(x0,n)Q_{\mu}(x_{0},n-1)-Q_{\mu}(x_{0},n) is the probability that the Lévy flight crosses the origin for the first time at the nn-th step, with Qμ(x0,n=0)=1Q_{\mu}(x_{0},n=0)=1. Thus for the target to be captured at the nn-th step, it has to remain alive at least up to step n1n-1, which occurs with probability an1a^{n-1}. Therefore the capture probability Cμ(x0,a)C_{\mu}(x_{0},a), defined as the probability that the searcher starting at x0x_{0} finds the target before the latter becomes inactive, is given by Cμ(x0,a)=n=1an1[Qμ(x0,n1)Qμ(x0,n)]C_{\mu}(x_{0},a)=\sum_{n=1}^{\infty}a^{n-1}\left[Q_{\mu}(x_{0},n-1)-Q_{\mu}(x_{0},n)\right]. This sum can be rewritten as

Cμ(x0,a)=1(1a)Q~μ(x0,s=a)a,C_{\mu}(x_{0},a)=\frac{1-(1-a)\widetilde{Q}_{\mu}(x_{0},s=a)}{a}, (2)

where Q~μ(x0,s)n=0snQμ(x0,n)\widetilde{Q}_{\mu}(x_{0},s)\equiv\sum_{n=0}^{\infty}s^{n}Q_{\mu}(x_{0},n) is the generating function of Qμ(x0,n)Q_{\mu}(x_{0},n). Similarly, the CMFPT Tμ(x0,a)T_{\mu}(x_{0},a), the mean time taken by the successful trajectories to locate the target [44], can be expressed as Tμ(x0,a)=n=1nan1[Qμ(x0,n1)Qμ(x0,n)]/Cμ(x0,a)T_{\mu}(x_{0},a)=\sum_{n=1}^{\infty}na^{n-1}\left[Q_{\mu}(x_{0},n-1)-Q_{\mu}(x_{0},n)\right]/C_{\mu}(x_{0},a), where Cμ(x0,a)C_{\mu}(x_{0},a) acts as a normalization factor. This can also be rewritten again in terms of the generating function of the survival probability

Tμ(x0,a)=aaln[1(1a)Q~μ(x0,s=a)].T_{\mu}(x_{0},a)=a\frac{\partial}{\partial a}\ln\left[1-(1-a)\widetilde{Q}_{\mu}(x_{0},s=a)\right]. (3)

Thus to analyze either Cμ(x0,a)C_{\mu}(x_{0},a) or Tμ(x0,a)T_{\mu}(x_{0},a), we need the generating function Q~μ(x0,s)\widetilde{Q}_{\mu}(x_{0},s) for Lévy flights. Unfortunately, there is no simple expression for Q~μ(x0,s)\widetilde{Q}_{\mu}(x_{0},s). However its Laplace transform with respect to x0x_{0} is given by the exact Pollaczek-Spitzer formula [48, 50],

0Q~μ(x0,s)eλx0𝑑x0=1λ1sφ(λ,s)\displaystyle\int_{0}^{\infty}\widetilde{Q}_{\mu}(x_{0},s)\,e^{-\lambda x_{0}}dx_{0}=\frac{1}{\lambda\sqrt{1-s}}\varphi(\lambda,s) (4)
withφ(λ,s)=exp[λπ0ln[1sf^(k)]λ2+k2𝑑k],\displaystyle{\rm with}\quad\varphi(\lambda,s)=\exp\left[-\frac{\lambda}{\pi}\int_{0}^{\infty}\frac{\ln[1-s\hat{f}(k)]}{\lambda^{2}+k^{2}}dk\right]\,,\quad (5)

where f^(k)=f(η)eikη𝑑η\hat{f}(k)=\int_{-\infty}^{\infty}f(\eta)\ e^{{\rm i}k\eta}d\eta is the Fourier transform of the step distribution. Here we will focus on Lévy stable jump distribution, with f^(k)=e|k|μ\hat{f}(k)=e^{-|k|^{\mu}} with 0<μ20<\mu\leq 2.

With an infinite-lived target (a=1a=1), recall that Cμ=1C_{\mu}=1, owing to the recurrence property of 1d1d random walks, while Tμ=T_{\mu}=\infty, independently of x0x_{0} and f(η)f(\eta) [58]. Hence, there is no option of optimizing them by varying μ\mu. However, for a finite-lived target where a<1a<1, both quantities become nontrivial functions of μ\mu and can be optimized by choosing μ\mu appropriately with optimal values μcap(x0,a)\mu^{\star}_{cap}(x_{0},a) and μFP(x0,a)\mu^{\star}_{FP}(x_{0},a). One finds that, even for short-lived targets, CμC_{\mu} at optimality can be larger than the maximal value 1/21/2 that could be achieved by a naive ballistic strategy (see [47]).

In order to maximize the capture probability in Eq. (2) by varying μ\mu, for fixed x0x_{0} and aa, it turns out that we need to minimize Q~μ(x0,s=a)\widetilde{Q}_{\mu}(x_{0},s=a) with respect to μ\mu. We will study the exact relation in Eq. (4), both analytically in certain limits and numerically by inverting the Laplace transform in Eq. (4) using the Gaver-Stehfest method [59, 60], which we explain in [47].

We start by plotting the numerically obtained Q~μ(x0,a)\widetilde{Q}_{\mu}(x_{0},a) as a function of μ\mu, for fixed x0x_{0} and aa. In Fig. 3a we show the data for a=0.5a=0.5 and four different values of x0x_{0}. For small x0x_{0}, the curve has a single minimum at a nonzero value of μcap(x0,a)\mu^{\star}_{cap}(x_{0},a), while there is a local maximum at μ=0\mu=0. As x0x_{0} increases to some value xmx_{m}, the derivative of Q~μ(x0,a)\widetilde{Q}_{\mu}(x_{0},a) with respect to μ\mu at μ=0+\mu=0^{+} 111Here we consider the μ0+\mu\to 0^{+} limit (and not strictly μ=0\mu=0). In the limit μ0+\mu\to 0^{+} the jump distribution is normalizable but not when μ=0\mu=0 exactly. Hence we restrict only to the case μ0+\mu\to 0^{+}. vanishes, i.e., μQ~μ(xm,a)|μ=0=0.\partial_{\mu}\widetilde{Q}_{\mu}(x_{m},a)\Big{|}_{\mu=0}=0\;. This value of xmx_{m} can be determined analytically [see Eq. (7) below] and is given by xm=eγE=0.561459x_{m}=e^{-\gamma_{E}}=0.561459\ldots, where γE\gamma_{E} is the Euler constant. When x0x_{0} slightly exceeds xmx_{m}, the curve has two minima: one at μ=0+\mu=0^{+} and one at μ=μcap(x0,a)\mu=\mu^{\star}_{cap}(x_{0},a), but the value at μ=0+\mu=0^{+} is higher. This situation persists for xm<x0<xc(a)x_{m}<x_{0}<x_{c}(a). When x0x_{0} exceeds xc(a)x_{c}(a), the local minimum at μ=0+\mu=0^{+} becomes the global one and μcap(x0,a)\mu^{\star}_{cap}(x_{0},a) drops to 0+0^{+}, triggering a first-order transition. The point xc(a)x_{c}(a) is thus determined by

μQ~μ(xc,a)|μcap(xc)=0,Q~μ(xc,a)|μcap(xc)=q0,\left.\partial_{\mu}\widetilde{Q}_{\mu}(x_{c},a)\right|_{\mu^{\star}_{cap}(x_{c})}=0,\quad\left.\widetilde{Q}_{\mu}(x_{c},a)\right|_{\mu^{\star}_{cap}(x_{c})}=q_{0}, (6)

where q0Q~μ=0(xc,a)q_{0}\equiv\widetilde{Q}_{\mu=0}(x_{c},a). From Eq. (4), q0=1/(1a)(1ae1)q_{0}=1/\sqrt{(1-a)(1-ae^{-1})}, independent of the position (see [47]). This scenario presented above for a=0.5a=0.5 continues to hold up to a=a10.926a=a_{1}\approx 0.926.

For a>a1a>a_{1}, a different scenario occurs as displayed in Fig. 3b where again Q~μ(x0,a)\widetilde{Q}_{\mu}(x_{0},a) is plotted as a function of μ\mu for different values of x0x_{0}. In contrast to Fig. 3a, the curves always have a single minimum at μ=μcap(x0,a)\mu=\mu^{\star}_{cap}(x_{0},a) that decreases continuously to 0+0^{+} as x0x_{0} approaches a critical value xc(a)=xmx_{c}(a)=x_{m}, signaling a second-order phase transition. Thus the first and second-order phase transitions merge at a=a1a=a_{1}, making it a tricritical point. These numerical observations lead to the phase diagram presented in Fig. 2.

The CMFPT exhibits the same qualitative features as above, with a tricritical point now located at a=a20.974a=a_{2}\approx 0.974... In Fig. 3c, we plot μFP(x0,a)\mu^{\star}_{FP}(x_{0},a) as a function of x0x_{0} for four different values of aa close to a2a_{2}. The jump discontinuity at x0=xc(a)x_{0}=x_{c}(a) is finite for a<a2a<a_{2} while it vanishes for aa2a\geq a_{2}, confirming indeed that (x0=xm,a=a2)(x_{0}=x_{m},a=a_{2}) is a tricritical point for μFP(x0,a)\mu^{\star}_{FP}(x_{0},a) in the (x0,a)(x_{0},a) plane.

We show how a1a_{1} and a2a_{2} can be computed analytically using a standard Landau-like expansion well known in critical phenomena. There, by expanding the free energy in powers of the order parameter, the Landau theory gives access to the phase diagram close to a continuous critical/tricritical point. Here we follow the same spirit with μ\mu playing the role of the “order parameter”. We then expand Q~μ\widetilde{Q}_{\mu} in powers of μ\mu near μ=0+\mu=0^{+}: Q~μ(x0,a)=q0+q1μ+q2μ2/2!+q3μ3/3!+q4μ4/4!+\widetilde{Q}_{\mu}(x_{0},a)=q_{0}+q_{1}\mu+q_{2}\mu^{2}/2!+q_{3}\mu^{3}/3!+q_{4}\mu^{4}/4!+\ldots, where the dependence of the qiq_{i}’s on x0x_{0} and aa is implicit. Depending on these parameters, the signs of qiq_{i}’s in this expansion may change, leading either to a first or second order transition and also to the possibility of a tricritical point. In the standard Landau’s theory with a positive order parameter it is enough to keep terms up to order O(μ3)O(\mu^{3}) and a tricritical point occurs when q1=q2=0q_{1}=q_{2}=0 while q3>0q_{3}>0 [62] (see also [63] in the context of stochastic resetting). However, in our case, the dependence of qiq_{i}’s on x0x_{0} and aa are such that this standard scenario is not realized and one needs to keep terms up to order O(μ4)O(\mu^{4}). From Eq. (4), we show that [47]

q1=ae121a(1ae1)3/2(lnx0+γE)\displaystyle q_{1}=\frac{ae^{-1}}{2\sqrt{1-a}(1-ae^{-1})^{3/2}}(\ln x_{0}+\gamma_{E}) (7)
q2=3ea241a(ea)5/2(lnx0+γE)2.\displaystyle q_{2}=\frac{3\sqrt{e}a^{2}}{4\sqrt{1-a}(e-a)^{5/2}}(\ln x_{0}+\gamma_{E})^{2}\;. (8)

For x0<xm=eγEx_{0}<x_{m}=e^{-\gamma_{E}}, we have q1<0q_{1}<0 and q2>0q_{2}>0. In contrast, for x0>xmx_{0}>x_{m}, we have both q1,q2>0q_{1},q_{2}>0 and both of them vanish simultaneously at x0=xmx_{0}=x_{m}, for any aa. The tricritical point thus occurs when q3(xm,a)q_{3}(x_{m},a) changes sign. We have [47]

q3(xm,a)=aeK81a(ea)7/2(11a2+8ea4e2),q_{3}(x_{m},a)=\frac{a\sqrt{e}K}{8\sqrt{1-a}(e-a)^{7/2}}(11a^{2}+8ea-4e^{2})\;, (9)

where K=2ζ(3)=2.40411K=2\zeta(3)=2.40411.... Thus q3(xm,a)<0q_{3}(x_{m},a)<0 for a<a1a<a_{1} where a1=2e(152)/11a_{1}=2e(\sqrt{15}-2)/11 is the unique root of 11a2+8ea4e2=011a^{2}+8ea-4e^{2}=0 in (0,1)(0,1). At the transition point x0=xc(a)x_{0}=x_{c}(a) and for a<a1a<a_{1}, since q3<0q_{3}<0, we need to keep terms up to order O(μ4)O(\mu^{4}) (assuming that q4>0q_{4}>0 in the Landau expansion). From Eqs. (6), the first-order jump discontinuity Δ(a)μcap(xc(a),a)\Delta(a)\equiv\mu_{cap}^{\star}(x_{c}(a),a) is given by [47]

Δ(a)=23q4(2|q3|+4q329q2q4)|x0=xc(a),\displaystyle\Delta(a)=\frac{2}{3\,q_{4}}\left(2|q_{3}|+\sqrt{4q_{3}^{2}-9q_{2}q_{4}}\right)\Big{|}_{x_{0}=x_{c}(a)}\;, (10)

This discontinuity vanishes when q30q_{3}\to 0 and q20q_{2}\to 0 which occurs at the point (x0=xm,a=a1)(x_{0}=x_{m},a=a_{1}), indicating that this is a tricritical point. If a>a1a>a_{1} then q2>0q_{2}>0 and q3>0q_{3}>0 : when q1q_{1} changes sign (always at x0=xmx_{0}=x_{m}), a 2nd2^{\rm nd} order transition occurs. Hence xc(a)x_{c}(a) freezes to xmx_{m} for a>a1a>a_{1}. A similar Landau-like expansion can be carried out exactly for the CMFPT, which leads to the same qualitative conclusions, with a2=0.973989a_{2}=0.973989\ldots [47].

As mentioned before, for a permanent target (a=1a=1), there is no optimal strategy since the capture probability is 11 and the CMFPT infinite, irrespective of μ\mu. However, surprisingly, for long-lived targets, there is a nontrivial optimal strategy characterized by the same μcap=μFP\mu_{cap}^{*}=\mu_{FP}^{*} for both observables. As a1a\to 1, Eqs. (4) and (3) imply Q~μ(x0,a)gμ(x0)/1a\widetilde{Q}_{\mu}(x_{0},a)\approx g_{\mu}(x_{0})/\sqrt{1-a} and Tμ(x0,a)gμ(x0)/(21aT_{\mu(x_{0},a)}\approx g_{\mu}(x_{0})/(2\sqrt{1-a}), where gμ(x0)g_{\mu}(x_{0}) is independent of aa. Hence, both the capture probability and the CMFPT are optimized by minimizing gμ(x0)g_{\mu}(x_{0}) with respect to μ\mu. Since the expression of gμ(x0)g_{\mu}(x_{0}) is complicated, it is hard to obtain the full functional form of μcap=μFP\mu_{cap}^{*}=\mu_{FP}^{*} for all x0x_{0}. However, close to the transition point xmx_{m}, where μcap\mu_{cap}^{*} is expected to be small due to the continuous transition, gμg_{\mu} directly follows from the small μ\mu expansion of QμQ_{\mu} above. Using Eqs. (7) and (9), we obtain exactly to leading order for small (xmx0)/xm(x_{m}-x_{0})/x_{m}

μcap=μFPA(xmx0xm)1/2,x0<xm,\displaystyle\mu_{cap}^{*}=\mu_{FP}^{\star}\approx A\left(\frac{x_{m}-x_{0}}{x_{m}}\right)^{1/2}\;,\;x_{0}<x_{m}\;, (11)

where A=2(e1)/ζ(3)(11+8e4e2)=1.7549A=2(e-1)/\sqrt{\zeta(3)(11+8e-4e^{2})}=1.7549\ldots (see SM [47] for more details). This shows that the limit a1a\to 1 does allow an optimization with respect to μ\mu.

So far, we have analyzed the exact formula in Eq. (4) in the small μ\mu limit. When a1a\to 1 and x00x_{0}\to 0, far from xmx_{m}, a small x0x_{0} expansion in [47] gives μcap0.905954\mu^{\star}_{cap}\to 0.905954..., as indicated in Fig. 2. But to obtain analytically the full curves in Figs. 3a and 3b, as a function of μ\mu from Eq. (4) for any (x0,a)(x_{0},a) is extremely hard. Yet, we have found a concavity approximation allowing a very accurate analytical estimate of Q~μ(x0,a)\widetilde{Q}_{\mu}(x_{0},a). Starting from the concavity of the logarithm, we approximate iwiln(ri)ln(iwiri)\sum_{i}w_{i}\ln(r_{i})\approx\ln(\sum_{i}w_{i}r_{i}) for any set of positive reals rir_{i} and normalized weights iwi=1\sum_{i}w_{i}=1. With this, one can perform the inverse Laplace transform in Eq. (4) and deduce the general expression

Q~μ,approx(x0,s)=11se1π0ln[1sf^(k)]sin(kx0)k𝑑k,\widetilde{Q}_{\mu,approx}(x_{0},s)=\frac{1}{\sqrt{1-s}}e^{-\frac{1}{\pi}\int_{0}^{\infty}\ln[1-s\hat{f}(k)]\frac{\sin(kx_{0})}{k}dk}\;, (12)

where we have used the identity 1[k/(λ2+k2)]=sin(kx0){\cal L}^{-1}[k/(\lambda^{2}+k^{2})]=\sin(kx_{0}) for x0>0x_{0}>0 (see also [47]). Eq. (12) is easy to evaluate numerically. Interestingly, the first two terms of its small μ\mu expansion coincide with the exact expressions q0q_{0} and q1q_{1} above, as well as the first terms of its small x0x_{0} expansion [47]. Consequently, Eq. (12) gives the correct slope-change point xmx_{m} and also captures qualitatively the order of the transitions (see the dashed lines in Figs. 3a and 3b), along with the existence of a tricritical point.

We conclude with the remark that this problem of a finite-lived target is reminiscent of a Lévy flight subject to resetting with a probability rr to its initial position. The mean first-passage time (MFPT) to find a permanent target at the origin was computed for the resetting Lévy flight [56] where the walker has two parameters μ\mu and rr that can be used to optimize the MFPT (see also [64] for a related problem). Indeed, the optimal pair (μ,r)(\mu^{*},r^{*}) was computed and found to undergo a first-order transition at a critical value of the initial distance x0x_{0} from the target. This is rather different from our problem where the Lévy flight has only a single parameter μ\mu, which it can vary to optimize the MFPT. In our model, the walker has no control over the parameter aa associated with the lifetime of the target. Hence, here we optimize the search strategy by varying only μ\mu for fixed aa, which leads to a new phase diagram with a tricritical point.

In summary, we have studied a simple model of a Lévy flight of index μ\mu in one-dimension searching for a finite-lived target at the origin with mean lifetime 1/(1a)1/(1-a). We have shown that the capture probability of the target can be maximized by choosing an optimal μcap\mu^{\star}_{cap} for fixed aa and x0x_{0} (where x0x_{0} denotes the initial distance from the target). The presence of a finite life-time leads to a very rich and nontrivial phase diagram for μcap\mu^{\star}_{cap} in the (x0,a)(x_{0},a) plane. This work opens up many interesting possibilities for future works. For instance, it would be interesting to find the optimal strategy in higher dimensions, for multiple Lévy flights and for the case where the distribution of the target lifetime is non-exponential.

DB acknowledges support from the LPTMS at the Université Paris-Saclay (France) and from CONACYT (Mexico) grant Ciencia de Frontera 2019/10872. We thank Lya Naranjo for illustration support in Fig. 1.

References

  • Bartumeus and Catalan [2009] F. Bartumeus and J. Catalan, Optimal search behavior and classic foraging theory, Journal of Physics A: Mathematical and Theoretical 42, 434002 (2009).
  • Viswanathan et al. [2011] G. M. Viswanathan, M. G. da Luz, E. P. Raposo, and H. E. Stanley, The physics of foraging: an introduction to random searches and biological encounters (Cambridge University Press, 2011).
  • Serra et al. [2020] M. Serra, P. Sathe, I. Rypina, A. Kirincich, S. D. Ross, P. Lermusiaux, A. Allen, T. Peacock, and G. Haller, Search and rescue at sea aided by hidden flow structures, Nature Communications 11, 2525 (2020).
  • Kosmas et al. [2022] V. Kosmas, M. Acciaro, and M. Besiou, Saving migrants’ lives at sea: Improving search and rescue operations, Production and Operations Management 31, 1872 (2022).
  • Shlesinger and Klafter [1986] M. F. Shlesinger and J. Klafter, Lévy walks versus Lévy flights, in On growth and form: Fractal and non-fractal patterns in physics (Springer, 1986) pp. 279–283.
  • Viswanathan et al. [1999] G. M. Viswanathan, S. V. Buldyrev, S. Havlin, M. G. E. da Luz, E. P. Raposo, and H. E. Stanley, Optimizing the success of random searches, Nature 401, 911 (1999).
  • Bartumeus and Levin [2008] F. Bartumeus and S. A. Levin, Fractal reorientation clocks: Linking animal behavior to statistical patterns of search, Proceedings of the National Academy of Sciences 105, 19072 (2008).
  • Levernier et al. [2020] N. Levernier, J. Textor, O. Bénichou, and R. Voituriez, Inverse square Lévy walks are not optimal search strategies for d2d\geq 2, Physical Review Letters 124, 080601 (2020).
  • Buldyrev et al. [2021] S. V. Buldyrev, E. P. Raposo, F. Bartumeus, S. Havlin, F. R. Rusch, M. G. E. da Luz, and G. M. Viswanathan, Comment on “Inverse square Lévy walks are not optimal search strategies for d2d\geq 2”, Physical Review Letters 126, 048901 (2021).
  • Levernier et al. [2021] N. Levernier, J. Textor, O. Bénichou, and R. Voituriez, Reply to “Comment on ‘Inverse square Lévy walks are not optimal search strategies for d2d\geq 2”’, Physical Review Letters 126, 048902 (2021).
  • Guinard and Korman [2021] B. Guinard and A. Korman, Intermittent inverse-square Lévy walks are optimal for finding targets of all sizes, Science Advances 7, eabe8211 (2021).
  • Clementi et al. [2021] A. Clementi, F. d’Amore, G. Giakkoupis, and E. Natale, Search via parallel Lévy walks on 2\mathbb{Z}^{2}, in Proceedings of the 2021 ACM Symposium on Principles of Distributed Computing (2021) pp. 81–91.
  • Oshanin et al. [2007] G. Oshanin, H. Wio, K. Lindenberg, and S. Burlatsky, Intermittent random walks for an optimal search strategy: one-dimensional case, Journal of Physics: Condensed Matter 19, 065142 (2007).
  • Bénichou et al. [2011] O. Bénichou, C. Loverdo, M. Moreau, and R. Voituriez, Intermittent search strategies, Reviews of Modern Physics 83, 81 (2011).
  • Evans and Majumdar [2011a] M. R. Evans and S. N. Majumdar, Diffusion with stochastic resetting, Physical Review Letters 106, 160601 (2011a).
  • Evans and Majumdar [2011b] M. R. Evans and S. N. Majumdar, Diffusion with optimal resetting, Journal of Physics A: Mathematical and Theoretical 44, 435001 (2011b).
  • Reuveni et al. [2014] S. Reuveni, M. Urbakh, and J. Klafter, Role of substrate unbinding in Michaelis–Menten enzymatic reactions, Proceedings of the National Academy of Sciences 111, 4391 (2014).
  • Roldán et al. [2016] E. Roldán, A. Lisica, D. Sánchez-Taltavull, and S. W. Grill, Stochastic resetting in backtrack recovery by RNA polymerases, Physical Review E 93, 062411 (2016).
  • Pal and Reuveni [2017] A. Pal and S. Reuveni, First passage under restart, Physical Review Letters 118, 030603 (2017).
  • Falcón-Cortés et al. [2017] A. Falcón-Cortés, D. Boyer, L. Giuggioli, and S. N. Majumdar, Localization transition induced by learning in random searches, Physical Review Letters 119, 140603 (2017).
  • Evans et al. [2020] M. R. Evans, S. N. Majumdar, and G. Schehr, Stochastic resetting and applications, Journal of Physics A: Mathematical and Theoretical 53, 193001 (2020).
  • Pal et al. [2022] A. Pal, S. Kostinski, and S. Reuveni, The inspection paradox in stochastic resetting, Journal of Physics A: Mathematical and Theoretical 55, 021001 (2022).
  • Evans and Majumdar [2014] M. R. Evans and S. N. Majumdar, Diffusion with resetting in arbitrary spatial dimension, Journal of Physics A: Mathematical and Theoretical 47, 285001 (2014).
  • Pal et al. [2016] A. Pal, A. Kundu, and M. R. Evans, Diffusion under time-dependent resetting, Journal of Physics A: Mathematical and Theoretical 49, 225001 (2016).
  • Reuveni [2016] S. Reuveni, Optimal stochastic restart renders fluctuations in first passage times universal, Physical Review Letters 116, 170601 (2016).
  • Montero and Villarroel [2016] M. Montero and J. Villarroel, Directed random walk with random restarts: The Sisyphus random walk, Physical Review E 94, 032132 (2016).
  • Bhat et al. [2016] U. Bhat, C. De Bacco, and S. Redner, Stochastic search with Poisson and deterministic resetting, Journal of Statistical Mechanics: Theory and Experiment 2016, 083401 (2016).
  • Mercado-Vásquez et al. [2020] G. Mercado-Vásquez, D. Boyer, S. N. Majumdar, and G. Schehr, Intermittent resetting potentials, Journal of Statistical Mechanics: Theory and Experiment 2020, 113203 (2020).
  • Bressloff [2020] P. C. Bressloff, Diffusive search for a stochastically-gated target with resetting, Journal of Physics A: Mathematical and Theoretical 53, 425001 (2020).
  • De Bruyne et al. [2022] B. De Bruyne, S. N. Majumdar, and G. Schehr, Optimal resetting Brownian bridges via enhanced fluctuations, Physical Review Letters 128, 200603 (2022).
  • Mercado-Vásquez et al. [2022] G. Mercado-Vásquez, D. Boyer, and S. N. Majumdar, Reducing mean first passage times with intermittent confining potentials: a realization of resetting processes, Journal of Statistical Mechanics: Theory and Experiment 2022, 093202 (2022).
  • Biroli et al. [2023] M. Biroli, S. N. Majumdar, and G. Schehr, Critical number of walkers for diffusive search processes with resetting, Physical Review E 107, 064141 (2023).
  • Tal-Friedman et al. [2020] O. Tal-Friedman, A. Pal, A. Sekhon, S. Reuveni, and Y. Roichman, Experimental realization of diffusion with stochastic resetting, The Journal of Physical Chemistry Letters 11, 7350 (2020).
  • Besga et al. [2020] B. Besga, A. Bovon, A. Petrosyan, S. N. Majumdar, and S. Ciliberto, Optimal mean first-passage time for a Brownian searcher subjected to resetting: experimental and theoretical results, Physical Review Research 2, 032029(R) (2020).
  • Faisant et al. [2021] F. Faisant, B. Besga, A. Petrosyan, S. Ciliberto, and S. N. Majumdar, Optimal mean first-passage time of a Brownian searcher with resetting in one and two dimensions: experiments, theory and numerical tests, Journal of Statistical Mechanics: Theory and Experiment 2021, 113203 (2021).
  • Janmaat et al. [2006] K. R. Janmaat, R. W. Byrne, and K. Zuberbühler, Primates take weather into account when searching for fruits, Current Biology 16, 1232 (2006).
  • Tikuisis [1997] P. Tikuisis, Prediction of survival time at sea based on observed body cooling rates, Aviation, Space, and Environmental Medicine 68, 441 (1997).
  • Golding et al. [2005] I. Golding, J. Paulsson, S. M. Zawilski, and E. C. Cox, Real-time kinetics of gene activity in individual bacteria, Cell 123, 1025 (2005).
  • Wong et al. [2008] O. K. Wong, M. Guthold, D. A. Erie, and J. Gelles, Interconvertible lac repressor–DNA loops revealed by single-molecule experiments, PLoS biology 6, e232 (2008).
  • Chen et al. [2014] Y.-J. Chen, S. Johnson, P. Mulligan, A. J. Spakowitz, and R. Phillips, Modulation of DNA loop lifetimes by the free energy of loop formation, Proceedings of the National Academy of Sciences 111, 17396 (2014).
  • Waharte and Trigoni [2010] S. Waharte and N. Trigoni, Supporting search and rescue operations with UAVs, in 2010 international conference on emerging security technologies (IEEE, 2010) pp. 142–147.
  • Charnov [1976] E. L. Charnov, Optimal foraging, the marginal value theorem, Theoretical Population Biology 9, 129 (1976).
  • Yuste et al. [2013] S. B. Yuste, E. Abad, and K. Lindenberg, Exploration and trapping of mortal random walkers, Physical Review Letters 110, 220603 (2013).
  • Meerson and Redner [2015] B. Meerson and S. Redner, Mortality, redundancy, and diversity in stochastic search, Physical Review Letters 114, 198101 (2015).
  • Belan [2018] S. Belan, Restart could optimize the probability of success in a bernoulli trial, Phys. Rev. Lett. 120, 080601 (2018).
  • Radice [2023] M. Radice, Effects of mortality on stochastic search processes with resetting, Phys. Rev. E 107, 024136 (2023).
  • Boyer et al. [2023] D. Boyer, G. Mercado-Vásquez, S. N. Majumdar, and G. Schehr, Supplemental Material,     (2023).
  • Pollaczek [1952] F. Pollaczek, Fonctions caractéristiques de certaines répartitions définies au moyen de la notion d’ordre. Application à la théories des attentes, C. R. Acad. Sci. Paris 234, 2334 (1952).
  • Andersen [1954] E. S. Andersen, On the fluctuations of sums of random variables II, Mathematica Scandinavica 2, 195 (1954).
  • Spitzer [1956] F. Spitzer, A combinatorial lemma and its application to probability theory, Transactions of the American Mathematical Society 82, 323 (1956).
  • Redner [2001] S. Redner, A guide to first-passage processes (Cambridge University Press, 2001).
  • Comtet and Majumdar [2005] A. Comtet and S. N. Majumdar, Precise asymptotics for a random walkerÍs maximum, Journal of Statistical Mechanics: Theory and Experiment 2005, P06013 (2005).
  • Majumdar [2010] S. N. Majumdar, Universal first-passage properties of discrete-time random walks and Lévy flights on a line: Statistics of the global maximum and records, Physica A: Statistical Mechanics and its Applications 389, 4299 (2010).
  • Bray et al. [2013] A. J. Bray, S. N. Majumdar, and G. Schehr, Persistence and first-passage properties in nonequilibrium systems, Advances in Physics 62, 225 (2013).
  • Majumdar et al. [2017] S. N. Majumdar, P. Mounaix, and G. Schehr, Survival probability of random walks and Lévy flights on a semi-infinite line, Journal of Physics A: Mathematical and Theoretical 50, 465002 (2017).
  • Kusmierz et al. [2014] L. Kusmierz, S. N. Majumdar, S. Sabhapandit, and G. Schehr, First order transition for the optimal search time of Lévy flights with resetting, Physical Review Letters 113, 220602 (2014).
  • Klinger et al. [2022] J. Klinger, R. Voituriez, and O. Bénichou, Splitting probabilities of symmetric jump processes, Physical Review Letters 129, 140603 (2022).
  • Feller [2008] W. Feller, An introduction to probability theory and its applications, vol 1 (John Wiley & Sons, 2008).
  • Stehfest [1970] H. Stehfest, Algorithm 368: Numerical inversion of Laplace transforms [d5], Commun. ACM 13, 47 (1970).
  • Kuhlman [2013] K. L. Kuhlman, Review of inverse Laplace transform algorithms for Laplace-space numerical approaches, Numerical Algorithms 63, 339 (2013).
  • Note [1] Here we consider the μ0+\mu\to 0^{+} limit (and not strictly μ=0\mu=0). In the limit μ0+\mu\to 0^{+} the jump distribution is normalizable but not when μ=0\mu=0 exactly. Hence we restrict only to the case μ0+\mu\to 0^{+}.
  • Kincaid and Cohen [1975] J. M. Kincaid and E. G. D. Cohen, Phase diagrams of liquid helium mixtures and metamagnets: experiment and mean field theory, Physics Reports 22, 57 (1975).
  • Pal and Prasad [2019] A. Pal and V. V. Prasad, Landau-like expansion for phase transitions in stochastic resetting, Physical Review Research 1, 032001(R) (2019).
  • Campos and Méndez [2015] D. Campos and V. Méndez, Phase transitions in optimal search times: How random walkers should combine resetting and flight scales, Physical Review E 92, 062115 (2015).