Energy saving approximation of Wiener process
under unilateral constraints333The work supported by RSF grant 21-11-00047.
Abstract
We consider the energy saving approximation of a Wiener process under unilateral constraints. We show that, almost surely, on large time intervals the minimal energy necessary for the approximation logarithmically depends on the interval’s length. We also construct an adaptive approximation strategy that is optimal in a class of diffusion strategies and also provides the logarithmic order of energy consumption.
1 Problem setting and main result
Let denote the space of absolutely continuous functions on the interval . For let us call kinetic energy
In the works [1, 3, 7, 8, 9, 10, 11] the approximation of a random process sample path by the function of smallest energy was considered under various constraints on closeness between and the sample path. In particular, in [9] the energy saving approximation was studied for a Wiener process under bilateral uniform constraints.
For let us define the set of admissible approximations as
and let
It was proved in [9] that for every fixed it is true that
where is some absolute constant (the exact value of is unknown), i.e. the optimal approximation energy grows linearly in time.
In this work, we are interested in the behavior of a similar quantity under unilateral constraints, i.e. the set of admissible approximations is
and we are now interested in the behavior of
It is technically more convenient to translate the initial value of the approximating function to the point , so that this function runs above the trajectory of the approximated process . Let
Since the sets of functions and differ by a constant shift, it is easy to see that
Our main result asserts that, when grows, the quantity grows merely logarithmically.
Theorem 1
For every fixed it is true that
(1) |
In Section 2 we establish a connection between the unilateral energy saving approximation of arbitrary continuous function with its minimal concave majorant. In Section 3 we establish the necessary properties of Wiener process minimal concave majorant using the results of Groeneboom [6]. Section 4 contains the proof of Theorem 1.
In Sections 5–6 we consider a class of adaptive Markovian (diffusion) approximation strategies based on the current and past values of . We prove that in this class the optimal strategy is defined by the formula
For this strategy the energy consumption also has the logarithmic order but is two times larger than that for the optimal non-adaptive strategy using the information about the whole trajectory of . Namely,
2 Concave majorants as efficient approximations
It turns out that the optimal energy saving approximation under unilateral constraints may be described in terms of the minimal concave majorant (MCM) of the approximated function. Let be a continuous function. Then the corresponding MCM is the minimal concave function satisfying conditions
Proposition 2
Let . Then the problem under the constraints and
has a unique solution of the following form.
(a) If , then .
(b) If , then is defined differently on three intervals. On the initial interval, is an affine function whose graph contains the point and is a tangent to the graph of . Then coincides with until the first moment when the maximum of is attained. Finally, after that moment, is a constant.
The optimal energy saving majorant is shown in Figure 1.

Proof of the proposition: The problem’s solution exists since for every the set of functions
is compact in the space of continuous functions equipped with the topology of uniform convergence and the functional is lower semi-continuous in this topology.
The uniqueness of the solution follows from the fact that the set of functions satisfying problem’s assumptions is convex and the functional is strictly convex on this set.
Let us describe the solution.
Since case (a) is trivial, we consider case (b).
Let be the solution of our problem. We show first that is a convex non-decreasing function. Consider the function
where is the non-increasing monotone rearrangement of the function . Then is a concave non-decreasing function, and for all . Therefore, satisfies the problem’s constraints. On the other hand,
Due to the problem solution uniqueness we obtain . This equality proves that is concave and non-decreasing.
Since is the smallest concave non-decreasing function satisfying problem’s constraints, we have
Furthermore, let us prove that . Indeed, in case (b) the function
satisfies both problem constraints and ; by the uniqueness of the solution, we obtain . In particular, .
Finally, assume that the strict inequality holds for some . Then, since the function is concave and non-decreasing, there exists a non-decreasing affine function such that
but . However, at the endpoints of the interval the opposite inequality is true, because
Therefore, there exists a non-degenerated interval such that , , .
Since is a constant, while is not a constant on , it follows from Hölder inequality that
We obtain
It follows that the function
satisfies the problem’s constraints and but this is impossible by the definition of . Therefore, the assumption brought us to a contradiction.
3 Minimal concave majorant of a Wiener process
We recall some notation and results from the article [6] that will be used in the sequel. Denote
The function is non-decreasing.
According to [6, Corollary 2.1] for every the random variable has the distribution density
where , is a standard normal random variable.
Next, let be the global MCM for the Wiener process . Define a random process as
Our study is essentially based on the following result due to Groeneboom.
Lemma 3
[6, Theorem 3.1]. For every the process , is a pure jump process with independent stationary increments and .
Moreover, there is an explicit description of the Lévy measure of in [6] but we do not need it here. We are only interested in the Kolmogorov’s strong law of large numbers for which asserts that
Using the definition of , letting , and making the variable change , we may reformulate this result as
(2) |
Lemma 4
For every with probability for all sufficiently large it is true that
(3) |
Proof: The lower bound is based upon the inequalities
where are some positive absolute constants.
Let , then the events satisfy
Therefore, by Borel–Cantelli lemma, with probability for all sufficiently large the event does not hold, i.e. with probability for all sufficiently large we have
(4) |
Let be such that (4) holds for and let . Since is non-decreasing, we have
This provides us with a required lower bound for sufficiently large .
In the same way, for the upper bound we have
where are some positive absolute constants.
Consider the sequence . For the events we have
Therefore, by Borel–Cantelli lemma with probability for all sufficiently large the event does not hold, i.e. with probability for all sufficiently large it is true that
(5) |
Let (5) be satisfied for some and let . Since is non-decreasing, we have
This provides us with a required upper bound for all sufficiently large .
The next theorem describes the asymptotic behavior of the energy of the minimal concave majorant for a Wiener process. For some let denote MCM of a Wiener process on the whole real line, starting from the height . Then the majorant is an affine function on some initial interval , its graph contains the point and is a tangent to the graph of , while on the functions and coinсide.
Theorem 5
Let be the global minimal concave majorant of starting from a height . Then, for every fixed it is true that
(6) |
Proof: Compare the quantities
and
They differ by a term (independent of ) corresponding to the initial segment of and by the lower and upper integration limits; notice that the lower integration limits do not depend on in both cases.
By using Lemma 4 for comparing the upper integration limits, we obtain
4 Proof of Theorem 1
Upper bound. The restriction of the global MCM starting from the height onto the interval [0,T] belongs to the set of admissible functions: . We derive from Theorem 5 that
Lower bound. For let denote the local MCM of the Wiener process , starting from the height . Let be the unique solution of the problem we are interested in, . Recall that its structure is described in Proposition 2. Since for large it is true that , for such the assumption of case (b) of that proposition is verified. In particular, it follows that
where
Notice that the function can not take values from the interval . Therefore, if for some it is true that , then it is also true that . In this case we have
It follows that
Let us fix . Let . Then by Lemma 4 we have
a.s. for all sufficiently large . Furthermore, it follows from Theorem 5 that, as ,
By combining these estimates, we obtain
Finally, by letting , we arrive at
as required.
5 Adaptive Markovian approximation
In practice, it is often necessary to arrange an approximation (a pursuit) in real time (adaptively), when the trajectory of the approximated process is known not on the entire time interval but only before the current time instant. In view of the Markov property of Wiener process, a reasonable strategy is to define the speed of a pursuit as a function of current positions of the processes and , without taking past trajectories into account, i.e. let
(7) |
On the qualitative level the function must tend to infinity, as , i.e. when the approximating process approaches the dangerous boundary it accelerates its movement trying to escape from a dangerous position. One has to optimize the function trying to reach the smallest average energy consumption. It is possible to reach the same logarithmic in time order of energy consumption as in the case of non-adaptive approximation but with somewhat larger coefficient. The difference of the coefficients represents the price we pay for not knowing the future of the process we try to approximate.
It is interesting to compare (7) with the form of the optimal adaptive strategy in the case of bilateral constraints [9] where
(8) |
The latter strategy is more simple because the speed is governed only by the distance between the approximated and the approximating processes and does not depend on time.
Let us make a time and space change
Recall that is an Ornstein–Uhlenbeck process and therefore satisfies the equation
(9) |
where is a Wiener process. We have the following expression for the derivative of
(10) |
which yields
(11) |
Let us consider the distance between the approximated and approximating processes
(12) |
We will study time-homogeneous diffusion strategies
(13) |
From equations (9) and (13) it follows that this is equivalent to
which also implies
(14) |
Before proceeding to optimization, let us see how the diffusion strategies act in the initial framework. By (11) and (14) we have
(15) | |||||
(16) |
where . In other words, the form of the strategy is
(17) |
We see that this class of strategies is space-homogeneous but, in general, not time-homogeneous.
Now we proceed to the optimization of the shift coefficient determining the pursuit strategy. Let us use some basic facts about one-dimensional time-homogeneous diffusion, cf. [2, Ch.IV.11] and [5, Ch.2]. Let
(18) | |||||
(19) |
Assume that condition
(20) |
is verified. Then, in Feller classification, the point is the entrance-boundary and not an exit-boundary for diffusion (13). This means that the diffusion remains forever in . Moreover, the function
(21) |
where , is the density of the unique stationary distribution for . For the energy, by using (11) and (14), we obtain (a.s., as )
Taking into account condition (20), it remains to solve the variational problem
over the class of densities concentrated on . After the variable change
the variational problem transforms into
We show in the next section that this minimum equals ; it is attained at the function
It follows that the asymptotic energy behavior for the optimal strategy is
i.e. the optimal choice of the shift in the adaptive setting leads to two times larger energy consumption than for the optimal strategy in the non-adaptive setting.
In order to find the optimal shift, write
Note that the density indeed satisfies the necessary condition (20).
Returning to the initial problem, we obtain the shift , thus the strategy (17) takes the form
Curiously, the optimal diffusion strategy is not only space-homogeneous but also time-homogeneous, unlike arbitrary strategies of this class.
6 Solution of the variational problem
6.1 Quantum harmonic oscillator
Consider the Sturm–Liouville problem on the eigenvalues of a differential operator
It represents a special case of the quantum harmonic oscillator equation, extensively studied by physicists, see [4, §23]. Its solution is well known. Usually one considers this equation on the entire real line. When performing the restriction to , one should take into account the boundary condition , hence, to keep the restrictions to of odd solutions on and multiply them by in order to keep the normalization. We arrive at the orthonormal base that consists of the functions , given by
where are Hermite polynomials; these functions satisfy
where .
In particular, the minimal eigenvalue is , , while the corresponding eigenfunction is .
6.2 Minimization
Consider quadratic form
For twice differentiable functions satisfying additional assumption integration by parts yields
In particular,
since is an orthonormal base.
If , then
and for the equality is attained in this chain. Therefore,
The authors are grateful to A.I. Nazarov for useful advice.
References
- [1] D.I. Blinova, M.A. Lifshits. Energy of taut strings accompanying Wiener process and random walk in a band of variable width. J. Math. Sci., 2022, 268, No. 5, 573–588.
- [2] A.N. Borodin. Stochastic processes. Birkhäuser, 2017.
- [3] Z.A. Kabluchko, M.A. Lifshits. Adaptive energy saving approximation for stationary processes. Izvestia: Mathematics, 2019, 83, no. 5, 932–956.
- [4] L.D. Landau, E.M. Lifshitz. Quantum Mechanics: Non-Relativistic Theory. Course of Theoretical Physics, Vol.3. 3rd Edition. Elsevier, 2013.
- [5] A.N. Borodin, P. Salminen. Handbook of Brownian motion. Facts and Formulae. Birkhäuser, Basel, 1996.
- [6] P. Groeneboom. The concave majorant of Brownian motion. Ann. Probab., 1983, 11, no.4, 1016–-1027.
- [7] I. Ibragimov, Z. Kabluchko, M. Lifshits. Some extensions of linear approximation and prediction problems for stationary processes. Stoch. Proc. Appl., 2019, 129, 2758–2782.
- [8] Z. Kabluchko, M. Lifshits. Least energy approximations for processes with stationary increments. J. Theor. Probab., 2017, 30, no. 1, 268–296.
- [9] M. Lifshits, E. Setterqvist. Energy of taut string accompanying Wiener process. Stoch. Proc. Appl., 2015, 125, 401–427.
- [10] M.A. Lifshits, A.A. Siuniaev. Energy of taut strings accompanying random walk. Probab. Math. Stat., 2021, 41, no.1, 9–23.
- [11] E. Schertzer. Renewal structure of the Brownian taut string. Stoch. Proc. Appl., 2018, 128, 487–504.