Higher-Order Regularity
of the Free Boundary in
the Inverse First-Passage Problem
Abstract
Consider the inverse first-passage problem: Given a diffusion process on a probability space and a survival probability function on , find a boundary, , such that is the survival probability that does not fall below , i.e., for each , . In earlier work, we analyzed viscosity solutions of a related variational inequality, and showed that they provided the only upper semi-continuous (usc) solutions of the inverse problem. We furthermore proved weak regularity (continuity) of the boundary under additional assumptions on . The purpose of this paper is to study higher-order regularity properties of the solution of the inverse first-passage problem. In particular, we show that when is smooth and has negative slope, the viscosity solution, and therefore also the unique usc solution of the inverse problem, is smooth. Consequently, the viscosity solution furnishes a unique classical solution to the free boundary problem associated with the inverse first-passage problem.
1 Introduction
1.1 The First-Passage Problem and its Inverse
Let be the solution to the stochastic differential equation
where is a standard Brownian motion defined on a probability space and are smooth bounded functions with . The boundary crossing, or first-passage problem for the process concerns the following:
-
1.
The Forward Problem: Given a function , compute the survival probability, , that does not fall below , i.e. evaluate
(1.1) -
2.
The Inverse Problem: Given a survival probability , find a barrier such that .
The forward problem is classical and the subject of a large literature. According to Zucca and Sacerdote (2009), the inverse problem was first suggested by A.N. Shiryaev during a Banach Centre meeting, for the case where is a Brownian motion, and the first-passage distribution is exponential, i.e. for some . Dudley and Gutmann (1977) proved the existence of a stopping time for with a given law. Anulova (1980) demonstrated the existence of a stopping time of the form for a closed set with the properties that if then , and if and , then . Defining , then satisfies the two-sided version of the inverse problem, .
In the 2000’s, the inverse problem became the subject of renewed interest due to applications in financial mathematics. In particular, Avellaneda and Zhu (2001) formulated the one-sided inverse problem given above as a free boundary problem for the Kolmogorov forward equation associated with , and discussed its numerical solution. Other numerical approaches and applications to credit risk were studied by Hull and White (2001), Iscoe and Kreinin (2002), and Huang and Tian (2006). In Cheng et al. (2006), we presented a rigorous mathematical analysis of the free boundary problem (formulated as a variational inequality), first demonstrating the existence of a unique viscosity solution to the problem, and analyzing integral equations satisfied by the boundary , and its asymptotic behaviour for small . In Chen et al. (2011), we proved that the boundary arising from the variational inequality does indeed solve the probabilistic formulation of the inverse problem. We also studied weak regularity properties of the free boundary . In particular, we proved the following (see Chen et al. (2011), Proposition 6):
Proposition 1.
Suppose that is standard Brownian motion, started at 0 (i.e. , , and ), and that is continuous with . Define:
-
1.
If for some positive with , then is continuous on .
-
2.
Assume that for every . Then .
In particular, if , then , yielding continuity of the boundary in the problem posed by Shiryaev.
The purpose of this paper is to study higher-order regularity properties of the boundary in the one-sided inverse problem.
Several other papers have also studied the inverse problem. Ekström and Janson (2016) show that the solution of the inverse first-passage problem is the same as the solution of a related optimal stopping problem, and present an analysis of an associated integral equation for the stopping boundary . Integral equations related to the problem are discussed in Peskir (2002) and Peskir and Shiryaev (2005). Abundo (2006) studied the small-time behaviour of the boundary . A rigorous construction of the boundary based on a discretization procedure was recently presented in Potiron (2021).
Remark 1.1.
Traditionally, the forward problem is studied for upper-semi-continuous (usc) and the conventional survival probability, , is defined in term of the first crossing time, , by
Here, the survival probability, in (1.1), is defined by
It is shown in Chen et al. (2011) that is well-defined for each . In addition, define and by
Then (i) , (ii) when , almost surely, and (iii) when , . Since , it is convenient for the inverse problem to restrict the search to in the class of usc functions, i.e., those that satisfy . In Chen et al. (2011) it is shown that for every , where
(1.2) |
the inverse problem admits a unique usc solution.
1.2 The Free Boundary Problems
We introduce differential operators and defined by
The survival distribution, , and survival density, , are defined by
We denote the distribution of by and its density by :
When is smooth, one can show that satisfies
(1.7) |
and satisfies
(1.12) |
Note that (1.7) and (1.12) are equivalent in the class of smooth functions via the transformation
When is given and regular, say, Lipschitz continuous, the forward problem can be easily handled by first solving the initial–boundary value problem consisting of the first three equations in (1.7) and then evaluating from the last equation in (1.7). For the inverse problem, both (1.7) and (1.12) are free boundary problems since the domain , where the equations, and , are satisfied, is a priori unknown. So far, there is little known concerning the existence, uniqueness, and regularity of classical solutions of the free boundary problems. Here by a classical solution, , of (1.7) we mean that , and each equation in (1.7) is satisfied; similarly, by a classical solution, , of (1.12), we mean that , and each equation in (1.12) is satisfied. In this paper, we investigate the well-posedness of the free boundary problem and the smoothness of the free boundary.
1.3 The Weak Formulation
In Cheng et al. (2006), viscosity solutions for the inverse problem, based on the variational inequality
are introduced. It is shown that for any given probability distribution on , there exists a unique viscosity solution. This was followed up in Chen et al. (2011) in which it was shown that the viscosity solution of the variational inequality gives the solution of the (probabilistic) inverse problem. For easy reference, we quote the relevant results.
Definition 1.
Let be given where is as in (1.2). A viscosity solution of the inverse problem associated with is a function defined by
(1.13) |
provided that has the following properties:
-
1.
, ;
-
2.
in and in the set ;
-
3.
If for a smooth , and , the function attains its local minimum on at , then .
One can verify that if (or ) is a classical solution of the free boundary problem (1.7) (or (1.12)), then is a viscosity solution of the inverse problem associated with .
Proposition 2 (Well-posedness of the Inverse Problem Cheng et al. (2006); Chen et al. (2011)).
Let be given.
-
1.
Cheng et al. (2006): There exists a unique viscosity solution, , of the inverse problem associated with .
-
2.
Chen et al. (2011): The viscosity solution is a usc solution of the inverse problem, i.e. and .
-
3.
Chen et al. (2011): There exists a unique usc solution of the inverse problem associated with .
It is clear now that the viscosity solution is the right choice for the inverse problem. For convenience, in the sequel, we shall call , , , or , the solution or the viscosity solution of the inverse problem, where is the viscosity solution boundary, is the viscosity solution for the survival distribution, and is the viscosity solution for the survival density of the inverse problem associated with . We shall also call the curve the free boundary.
1.4 The Main Result: Higher Order Regularity
While the work of Cheng et al. (2006) and Chen et al. (2011) solves the inverse problem, and presents a basic study of weak regularity, here we make a detailed study of the regularity of the free boundary. The main result of this paper is the following, where denotes the integer part of .
Theorem 1 (Regularity of the Free Boundary).
Let be given and be the (viscosity) solution of the inverse problem associated with . Assume that , and for some , either
(1.14) |
- 1.
-
2.
If, in addition, for not an integer, one has , then .
-
3.
Finally, if for not an integer one has , on , , and satisfy all compatibility conditions up to order , including in particular the compatibility condition when , then and .
Remark 1.2.
To derive the compatibility conditions, we consider, for simplicity, the special case and . Set . Then
(1.18) |
The -th order compatibility condition is the -th order derivative of at (with differentiation of in time being replaced by differentiation in space by ):
The -th order derivative of at is obtained by differentiating the equation :
In particular, the first and second order compatibility conditions are
Remark 1.3.
For to be continuous and bounded, it is necessary to assume that is strictly decreasing as in Proposition 1. Indeed, if is a constant in an open interval, then in that interval.
The main tool for the proof of Theorem 1 is the hodograph transformation, defined by the change of variables , the inverse of . Since , we have . Taking to simplify the exposition, and proceeding formally, we can derive that solves quasi-linear pde:
(1.19) |
Assuming it can be shown that is positive on , this system is studied on the set for any fixed and a small positive that depends on . To complete the system, we supply the boundary condition for on by .
The classical approach to the hodograph transformation (see, e.g. Friedman (1982) or Kinderlehrer and Stampacchia (1980)) employs a bootstrapping strategy, assuming some initial degree of smoothness on , and then using the regularity theory for (1.19) to strengthen the regularity of . In particular, standard results for quasi-linear equations (Ladyzhenskaya et al. (1968); Lieberman (1996)) can be used to derive the existence of a unique classical solution of (1.19), and its regularity (including up to the boundary). The assumptions on are sufficient to reverse the hodograph transformation, and transfer the boundary regularity of to (and to show that indeed , where is defined through ). The results achieved through the classical approach are reviewed in Section 4.
In order to prove Theorem 1, we wish to employ the same strategy, but with weaker assumptions on the initial regularity of . In doing so, we encounter two main difficulties, the first technical, and the second fundamental. The technical issue is that above the boundary, solves , and when the operator has a zeroth order term. The maximum principle arguments employed in analyzing level sets in our proofs require a differential operator with no zeroth order.111In particular, we need that constant functions satisfy the differential equation. We address these related technical hurdles by considering for an appropriate scaling function , defined as the solution of an auxiliary partial differential equation, such that above the boundary , where for some function . The formal hodograph transformation then leads us to consider the partial differential equation:
(1.20) |
together with the boundary condition , where
(1.21) |
Here, we encounter a fundamental difficulty due to the fact that we do not know a priori that is regular up to the boundary, i.e., we do not have the equation . While we can study the above problem analytically, we have not assumed the requisite regularity to show that , with .222We can, however, show that is well-defined and regular enough inside the domain. It is the regularity for up to the boundary (because of the lack of a priori regularity of ) that is insufficient. This difficulty is surmounted by defining a family of perturbed equations with boundary operators , , and making comparisons with their solutions . In particular, for small , we show that , . Letting , we are then able to obtain that , , and the required regularity of .
Remark 1.4.
For simplicity of exposition, throughout the paper we assume that . This can be done without loss of generality. Indeed, let
where is the inverse of . Then by Itô’s lemma, the process defined by is a diffusion process satisfying . The boundary crossing problem for with barrier is equivalent to the boundary crossing problem for with barrier . In terms of the partial differential equations, this is equivalent to the change of variables
We shall henceforth always assume that .
The remainder of the paper is structured as follows. In the next section, we recall a few properties of the solution of the inverse problem and prove a smoothing property of the diffusion: under condition (ii) of (1.14), condition (i) is satisfied provided that the initial time is shifted to , for any outside a set of measure zero. In Section 3 we provide an interpretation of the free boundary condition for the viscosity solution. In Section 4 we present results that can be derived using the traditional approach to the hodograph transformation . Section 5 presents the proof of Theorem 1, beginning by presenting a required generalization of the Hopf Boundary Point Lemma, then introducing the scaling function and the scaled survival density , and finally analyzing the family of solutions to quasi-linear parabolic equations in order to derive our main results.
2 Regularity Properties of the Viscosity Solutions of the Inverse Problem
In this section, we collect a few results concerning the regularity of the (viscosity) solution of the inverse problem. Recall that and .
Lemma 2.1.
Let be the solution of the inverse problem associated with . Then
(2.1) |
where the inequalities above hold in the sense of distributions.
In addition, for any , the following holds:
-
1.
If , then ; consequently, is a classical solution of the free boundary problem (1.7) on .
-
2.
If and , then and .
-
3.
(Smoothing Property) If on and , then for a.e. ,
Proof. For (2.1), see Cheng et al. (2006), Lemma 2.1, and Chen et al. (2011), Proposition 5 (with Theorems 1 and 3 in this reference summarizing the solution of the inverse boundary crossing problem).
To simplify the presentation, we assume that . The general case is analogous.333Unlike the assumption that , there are points in the paper where this is not the case. Hence, we remark on this explicitly when we take the drift to be zero. The viscosity solution for the survival distribution of the inverse problem satisfies the variational inequality , which, according to Cheng et al. (2006) (see also Friedman (1982)), can be approximated, as , by the solution of
where with . It is worth mentioning that one can further regularize to smooth so that the solution is not only smooth, but also monotonically decreasing in ; see Cheng et al. (2006) for details. We set , and .
1. Since and , the Comparison Principle yields that . Similarly, gives and yields that . From , we have:
(2.2) |
A parabolic estimate (e.g. Krylov (1996), Lemma 8.7.1, page 122) then implies that
(2.3) |
for every where is a constant depending only on .
Let , and let . Then there exists a subsequence such that in . As the limit is shown in Cheng et al. (2006) to be the unique viscosity solution of the inverse problem associated with , the whole sequence in fact converges. In addition, passing to the limit in the estimate (2.3) we see that .
Note that a classical solution of (1.7) requires that the free boundary condition be well-defined. Since is continuous on , the equation for is satisfied in the classical sense. Hence, is a classical solution of (1.7). This proves (1).
3. Using energy estimates, we will show that for each , , and therefore in particular for almost every , from which Sobolov Embedding yields . Since for all and for all , the last assertion of the lemma thus follows.
Suppose on and . Then , so is finite. As above, we have . Using the Comparison Principle as in Cheng et al. (2006), it can be shown that . We then further have , and comparison yields on .
Let be a non-negative smooth function satisfying for , for , and and for . In the sequel, for convenience of notation, we consider as a function of , being evaluated at . Using the differential equation and integration by parts, one can derive the identity
so that
Multiplying both sides of the above equation by , and then integrating over gives:
To control these terms, we make two preliminary estimates. First, since , there exists such that for every . Consequently, on , so that, for every ,
(2.4) |
Secondly, integrating over we obtain:
Integrating this inequality multiplied by over , we get, for every ,
(2.5) |
where we have used the fact that and where .
Returning to the estimation of , (2.4) immediately gives:
(2.6) |
For , we integrate by parts:
(2.7) |
The second term on the right is bounded by (2.4), and the fact that . Now, using that :
(2.8) |
To control , we again use , so that:
(2.9) |
As above:
(2.10) |
Furthermore, from (2.5), we have:
(2.11) |
Putting things together, we have that:
where is a constant depending on and . Sending we see that the above estimate also holds for . Finally, since on , we have . For each taking we obtain
Since can be arbitrarily small, we conclude that
This completes the proof. ∎
Remark 2.1.
We remark that for to be a classical solution of the free boundary problem (1.12), we need the existence of the limit of , as , for each . Here the conclusion of the third assertion in the previous lemma is not sufficient for to be a classical solution. Thus, from an analytical viewpoint, finding classical solutions of the free boundary problem (1.12) is much harder than that of the free boundary problem (1.7).
Remark 2.2.
Taking to be Brownian motion with , we have the following:
-
1.
If , then:
where is the probability density function of a standard normal random variable. Clearly (1) applies (, and is a classical solution of (1.7), as can be verified by direct computation). However, , so that (2) and (3) do not apply (which is not surprising since otherwise we would have , contradicting the definition of ).
-
2.
For the exponential survival function for some , and , so that both (2) and (3) apply, and in particular , and the smoothing property in (3) of the lemma holds.
3 The Free Boundary Condition For Viscosity Solutions
We interpret the free-boundary condition for the viscosity solution as follows.
Lemma 3.1.
Let and be the unique viscosity solution of the inverse problem associated with . Suppose and is continuous at . Then for any function that satisfies , we have:
(3.1) |
Remark 3.1.
Proof. Suppose the first inequality in (3.1) does not hold. Then the first limit in (3.1) is strictly bigger than so there exist small constants and such that
where . Consequently, for each and ,
(3.2) |
Now for any sufficiently small positive , consider the smooth function
We compare and in the set
We claim that the minimum of on is negative and is attained at some point . First of all, and
Next we show that on the parabolic boundary of ,
When , and . For small enough , on the remainder of the parabolic boundary of , we can verify that so that we can use (3.2) to derive
On the lower part of the parabolic boundary of , we have so and . Finally, on the right lateral boundary of , we have and so
Thus, on the parabolic boundary of provided that is sufficiently small.
Hence, for every small positive , there exists such that attains at the minimum of over . Now by the definition of as a viscosity solution (see, e.g. Cheng et al. (2006)), we have . However, we can calculate (with )
The last quantity is positive if we take sufficiently small. Thus we obtain a contradiction, and that the first inequality in (3.1) holds.
Now we prove the second inequality in (3.1). Since , the second inequality in (3.1) is trivially true when . Hence, we consider the case . Suppose the second inequality in (3.1) does not hold. Then the second limit in (3.1) is strictly less than so there exist small constants and such that
Since and in , the above inequality implies that . Hence, we must have , for every . Consequently, for each and ,
(3.3) |
Now for any sufficiently small positive , consider the smooth function
We compare and in the set
We claim that the maximum of on is positive and is attained at some point .
First of all, and
Next we show that on the parabolic boundary of . On the left lateral boundary of , , so and . For the remainder of the parabolic boundary we can use (3.3) to derive
On the lower side of the parabolic boundary of , so and . Finally, on the right lateral boundary of , we have and so
Thus, on the parabolic boundary of if is sufficiently small. Consequently, there exists such that attains at the positive maximum of over .
We claim that . Indeed, if , then which contradicts the fact that the maximum of on is positive. Thus . It then implies that is smooth in a neighborhood of . Consequently, . However,
The last quantity is negative if we take sufficiently small. Thus we obtain a contradiction, and the second inequality in (3.1) holds. This completes the proof. ∎
4 The Traditional Hodograph Transformation
The hodograph transformation considers the inverse, , of the function , so that for each fixed , the curve is the -level set of . A bootstrapping procedure is applied. By beginning with a weak regularity assumption on the free boundary , and applying the regularity theory for the partial differential equation satisfied by , one can obtain higher-order regularity of . In this section, we present two results derived using this traditional approach, one for the viscosity solution and the other for the classical solution of the free boundary problem.
Proposition 3.
Let be the solution of the inverse problem associated with . Assume that for some interval , , and , where is not an integer. Then .
Proof.
As is already , we need only consider the case , so is continuously differentiable. Since , by working on the function on the fixed domain with the boundary condition and then translating the regularity of back to , one can show that for every ,
In addition, since and , the Hopf Lemma (Protter and Weinberger (1967), Theorem 3.3, pages 170-171) implies that . Consequently, as , , and with , we can use Lemma 3.1 and Remark 3.1 to derive that .
Once we know the continuity of and the positivity of , we can define the inverse of for and where . Then implicit differentiation gives:444Recall that we take to simplify the exposition.
Also, where and for . It then follows from the local regularity theory for parabolic equations (see Ladyzhenskaya et al. (1968), Theorem IV.5.3, pages 320-322) that when where is not an integer, since we have . Consequently, . This completes the proof. ∎
Proposition 4.
Suppose on and where is not an integer. Assume that is a classical solution of the free boundary problem (1.12) satisfying
(4.1) |
Then .
Proof. Let be arbitrarily fixed. The conditions (4.1) and imply that the set is compact, so that there exists such that is bounded and uniformly positive in Consequently, the inverse of is well-defined for . Setting we have that , is uniformly positive and bounded in and
Since satisfies in , as above local regularity then implies that defined on can be extended onto such that . Hence . Sending and we conclude that .∎
5 Proof of The Main Result
In this section, we prove our main result, Theorem 1. This provides regularity of the viscosity solution of the inverse problem without the a priori regularity assumptions used in the previous section. We begin with a technical result - a generalization of Hopf’s Lemma in the one-dimensional case that is needed in our later arguments. We then introduce the scaling function , and the new hodograph transformation for the scaled function . Finally, we prove Theorem 1 by analyzing a family of perturbed equations related to the PDE satisfied by .
5.1 A Generalization of Hopf’s Lemma in the One-Dimensional Case
In order to apply the weak formulation of the free boundary condition, we need the following extension of the classical Hopf Lemma.
Lemma 5.1 (Generalized Hopf’s Lemma).
Let where , and are bounded functions and . Assume that in where and are Lipschitz continuous functions. Also assume that in . Then for every , there exists such that
Moreover, if in addition , then ; similarly, if , then
Proof. Without loss of generality, we can assume that for all . By modifying and in by and we can further assume that is uniformly positive on ; i.e., there exists such that for all . Furthermore, by approximating by smooth functions from below and by smooth functions from above, with the same Lipschitz constants of and , we can assume that both and are smooth functions.
Now for large positive constants and to be determined, consider the function
Direct calculation gives
where
First taking and then taking a suitably large we see that . Thus, by comparison, we have in and obtain the assertion of the Lemma.∎
5.2 Scaling the Survival Density
In making comparison arguments in the proof of Theorem 1, we would like to have that constant functions are solutions of the partial differential equation under consideration. Unfortunately, while this is true for the operator , it may fail for . To resolve this technical difficulty, we introduce a suitable scaling of the function .
To this end, let be the bounded solution of the initial value problem:
(5.1) |
Then is smooth, uniformly positive, and bounded in for any . Now write
It is easy to verify that
where
In particular, note that constant functions satisfy the equation .
5.3 The New Hodograph Transformation
As we have seen, the traditional hodograph transformation is defined as the inverse of . In order to work with the scaled survival density and the operator introduced above, we instead define as the inverse of :
If and , then by the Implicit Function Theorem, locally the above equation defines a unique smooth for near that satisfies . In addition, by implicit differentiation,
Thus, implies that satisfies the following quasi-linear partial differential equation of parabolic type:
The free boundary is given by on which we can derive the boundary condition as follows. Since and , we see that . Hence, we have the non-linear boundary condition
This is a standard boundary condition for the quasi-linear parabolic equation for .
5.4 The Basic Assumption
Let and be the unique viscosity solution of the inverse problem associated with . To prove Theorem 1, we need only establish the assertions of the theorem in a finite time interval for any fixed positive . Hence, in the sequel, we assume, for some and , that
(5.2) |
These conditions are satisfied by the general assumptions made in Theorem 1 if (i) in (1.14) is imposed. If instead of (i), condition (ii) in (1.14) is imposed, we can apply the third assertion of Lemma 2.1 to shift the initial time by considering first the solution for at which , and then sending .
Under (5.2), we will show that and that is a classical solution of the free boundary problem (1.12) on .
First of all, from Lemma 2.1 (1) and (2), we see that
5.5 The Level Sets.
The hodograph transformation we are going to use is based on the inverse, , of where . That is, for each , is the -level set of . For to be well-defined, we need to consider in the set where is positive.
We begin by investigating the initial value specified by Assume that satisfies (5.2). Then the function , admits a unique inverse, , which satisfies:
Next we consider for small . Fix . There exists such that and for all . Also, since , there exists such that on and on . Finally, from in we see that is continuous and uniformly positive and on for some . Hence for each , the equation , for , admits a unique solution which we denoted by . This solution has the property that . In addition, . Hence, .
Now we extend the inverse, , of to where is an arbitrarily fixed positive constant and is a small positive constant that depends on and on the viscosity solution . More precisely, we prove the following.
Lemma 5.2.
Let be given and be the unique viscosity solution of the inverse problem associated with . Assume that (5.2) holds. Let be defined in (5.1) and
Then there exists such that the function
(5.3) |
is well-defined and satisfies
Proof.
We believe that the idea behind this proof may have appeared previously in the literature, but are not aware of a precise reference. For completeness, and possible other applications, we present a full proof. We consider the level sets of in . We call a critical value of if there exists such that and either or is not differentiable at . Hence, if is not a critical value, then by the Implicit Function Theorem, the level set consists of smooth curves, each of which either does not have boundary (i.e. lies completely inside ) or has boundary on . Since and for all , by Sard’s Theorem (e.g. Guillemin and Pollack (1974), pages 39-45) the set of all critical values of has measure zero.
As before, we denote by the inverse of .
Since and , we can find a smooth function such that and for all . We define
Then in (5.3) is well-defined and for all and . In addition, for all and for all .
Next, let be a non-critical value of and let be the smooth curve in that connects to . We first claim that is a simple curve, i.e., it cannot form a loop. Indeed, if it forms a loop, then the loop is in so the differential equation in implies that inside the loop which is impossible, since is not a critical value. Hence, is a simple curve. We parameterize by its arc-length parameter, , in the form , with . It is not difficult to show that , so stays in a bounded region and we must have and . In addition, by the earlier discussion of for small positive we see that and for all small positive .
When , we can differentiate to obtain
(5.4) | |||
(5.5) |
Now we define . Since , we must have and . Consequently, . Evaluating (5.4) at , we obtain . Consequently, evaluating (5.5) at and using we obtain . Since is not a critical value of , we must have , so that . This implies that for all bigger than and close to .
Next we define We claim that . Indeed, suppose . Then we must have and . As above, first evaluating (5.4) at we obtain and then evaluating (5.5) at and using we get that . However, this is impossible since and for all Thus, we must have . In summary, we have
Now we claim that . Suppose not. Then for all . Since , we see that cannot touch . That is, for all . Thus, we must have . However, this would imply forms a loop which is impossible. Hence, .
Let be the number such that . Denote the inverse of , by . We can apply the Maximum Principle (Protter and Weinberger (1967), Theorem 3.2, recalling that on ) for on the domain to conclude that for every and . Hence, we must have . Since is not a critical value and in , we derive from (5.4) that for every . Hence, .
Finally, let be any two non-critical values of that satisfy . Then for , is a smooth function and for . Note that satisfies the equation in , the Maximum Principle then implies that on . By the Implicit Function Theorem, we then know that . Finally, sending and along non-critical values of , we conclude that . This completes the proof of the Lemma.∎
Remark 5.1.
Here we have used the Maximum Principle for . The Maximum Principle may not hold for since the equation for is where the non-homogeneous term may cause difficulties. Of course, we also need to know that and on so that the Maximum Principle can be applied to .
Using the same idea as in the proof, one can derive the following which maybe useful in qualitative and/or quantitative studies of the free boundary.
Proposition 5.
Let be the solution of the inverse problem associated with . Assume that and on . Let where is defined in (5.1).
-
1.
Denote by the number of roots (without counting multiplicity) of in . Then is a decreasing function. In particular, if , i.e. changes sign only once in , then for all ; that is, changes sign only once in .
-
2.
Suppose is the Delta function (i.e. a.s.). Then there exists such that and in and in for every .
The first assertion can be proven by a variation of our proof. The second assertion follows from the fact that the Delta function can be approximated by a sequence of the bell-shaped positive functions, each of which has only one local maximum and no local minimum; see Chen et al. (2008). We omit the details.
5.6 The Initial Boundary Value Problem
In the sequel, is defined by (5.3) and is a fixed small positive constant. We consider the quasi-linear parabolic initial boundary value problem, for the unknown function
(5.10) |
We know from the theory of quasi-linear equations that this problem admits a unique classical solution for small enough, which is also smooth up to the boundary, so long as for and for are uniformly positive. The main difficulty is to show that .555Once we have , and , it follows from classical theory that is only less differentiable than . We know that satifies all of (5.10), except for the third equation (i.e., we don’t know a priori that , because we don’t have smoothness of up to the boundary).
To do this, we consider for each , the following initial boundary value problem, for ,
(5.15) |
where is a family that has the following properties:
5.7 Well-Posedness of the Perturbed Problems
We now show that for each , (5.15) admits a unique classical solution. Since (5.10) and (5.15) belong to the same type of initial-boundary value problem, we state our result in terms of (5.10). The conclusion for (5.15) is analogous.
Lemma 5.3.
Let be smooth and bounded functions on . Assume that and
Then problem (5.10) admits a unique classical solution that satisfies
If in addition where is not an integer and , then so that .
If , , is not an integer, and all compatibility conditions at up to order are satisfied, then so
Proof.
According to the general theory of quasi-linear partial differential equations of parabolic type (see Ladyzhenskaya et al. (1968)), to show that (5.10) admits a unique classical solution, it suffices to establish an a priori estimate for an upper bound and a positive lower bound for . For this purpose, suppose we have a classical solution of (5.10). Then using a local analysis we have . Set . Then we can differentiate the first two equations in (5.10) to obtain:
Now denote
Then by comparison, we have
An an example, we demonstrate the derivation of the first order compatibility condition:
We remark that from our definition of , the compatibility of the initial and boundary data at for (5.10) is automatically satisfied. For (5.15), since , the first order compatibility condition at is also satisfied, so .
Finally, to demonstrate continuous dependence, we integrate over the difference of the differential equations in (5.10) and (5.15) multiplied by and use integration by parts to obtain
Upon using the boundedness of and , Cauchy’s inequality, and the Sobolev embedding
and we find that there exists a positive constant such that
Gronwall’s inequality then yields the estimate
This estimate in turn implies the convergence of to as . By Sobolev embedding and the boundedness of , this convergence also implies the convergence:
To show that , it suffices to show the following:
Lemma 5.4.
For every , and on .
The proof will be given in the next three subsections.
5.8 The Inverse Hodograph Transformation
To show that and , we define as the inverse function of :
Since , the inverse is well-defined. We record the key equation for future reference:
(5.17) |
By implicit differentiation, we find that
Finally, setting we see that
When , using , we see from (5.17) that This implies that , so using with we obtain or
Next, substituting in (5.17) we obtain
The boundary condition then implies that
Finally, substituting in (5.17) we have
The boundary condition then gives
In summary, we see that has the following properties:
Note that
5.9 The Proof that and for
Let be arbitrarily fixed. We define
We know that and are all continuous and that is upper-semi-continuous (Chen et al. (2011)). Also, , , and . Thus, we have that is well-defined and .
We claim that and on . Suppose the claim is not true. Then we must have have and for all and
We shall show that neither of the above can happen. For this we compare and in the set
where
Since for all , we have in . Also, on the left lateral parabolic boundary of , , we have . On the right lateral boundary of , , we have . Thus on the parabolic boundary of if . Finally, if , we have for all . Thus, on the parabolic boundary of . Since , we cannot have , so the Strong Maximum Principle implies that in .
Now consider case (i): . Then as is smooth and , the Hopf Lemma implies that . This is impossible since . Thus case (i) cannot happen.
Next, we consider case (ii): . Since is smooth, the generalized Hopf Lemma 5.1 implies that there exist and such that
However, since , the above inequality implies
But this contradicts the first inequality in (3.1) [with ] of Lemma 3.1. Hence, case (ii) also cannot happen.
In conclusion, when , we must have and on .
5.10 The Proof that and for
Here, we use the facts that and is continuous on , proven in Chen et al. (2011), and repeated here as Proposition 1. Also, we need the fact that in .
Let be arbitrary. We define
Since are all continuous and and , we see that is well-defined and .
We claim that and on . Suppose the claim is not true. Then and for all and
To show that none of the above can happen, we compare and as before in the set
Then, by the definition of , we have on the parabolic boundary of and in . The Maximum Principle then implies that in .
Now consider case (i): . Then as is smooth and , the Hopf Lemma implies that . This is impossible since . Thus case (i) cannot happen.
Next we consider case (ii): . Again, since is smooth, the generalized Hopf Lemma 5.1 implies that there exist and such that
However, since , the above inequality implies
This contradicts the second inequality in (3.1) [with ] of Lemma 3.1. Hence, case (ii) also cannot happen. Thus, when , we have and on .
5.11 Proof of Theorem 1
Once we know that for and , we can send to conclude that . Consequently, .
As we know , one can show that , so
Using and , we then obtain
Thus, is a classical solution of (1.12) on . If, in addition, for some that is not an integer, then by local regularity, . If we further have for some that is not an integer, and all compatibility conditions up to the order are satisfied, then so .
Finally, upon noting that can be arbitrarily large, we also obtain the assertion of Theorem 1, which completes the proof of this theorem.
6 Conclusion
In earlier work, we studied the inverse first-passage problem for a one-dimensional diffusion process by relating it to a variational inequality. We investigated existence and uniqueness, as well as the asymptotic behaviour of the boundary for small times, and weak regularity of the boundary. In this paper, we studied higher-order regularity of the free boundary in the inverse first-passage problem. The main tool used was the hodograph transformation. The traditional approach to the transformation begins with some a priori regularity assumptions, and then uses a bootstrap argument to obtain higher regularity. We presented the results of this approach, but then went further, studying the regularity of the free boundary under weaker assumptions. In order to do so, we needed to perform the hodograph transformation on a carefully chosen scaling of the survival density, and to analyze the behaviour of a related family of quasi-linear parabolic equations. We expect that the method presented here can be applied to other parabolic obstacle problems.
References
- Abundo [2006] M. Abundo. Limit at zero of the first-passage time density and the inverse problem for one-dimensional diffusions. Stochastic Analysis and Applications, 24:1119–1145, 2006.
- Anulova [1980] S.V. Anulova. Markov times with given distribution for a Wiener process. Theory Probab. Appl., 25(2):362–366, 1980.
- Avellaneda and Zhu [2001] M. Avellaneda and J. Zhu. Modeling the distance-to-default process of a firm. Risk, 14(12):125–129, 2001.
- Chen et al. [2008] X. Chen, J. Chadam, L. Jiang, and W. Zheng. Convexity of the exercise boundary of the American put option on a zero dividend asset. Mathematical Finance, 18(1):185–197, 2008.
- Chen et al. [2011] X. Chen, L. Cheng, J. Chadam, and D. Saunders. Existence and uniqueness of solutions to the inverse boundary crossing problem for diffusions. Annals of Applied Probability, 21(5):1663–1693, 2011.
- Cheng et al. [2006] L. Cheng, X. Chen, J. Chadam, and D. Saunders. Analysis of an inverse first passage problem from risk management. SIAM Journal on Mathematical Analysis, 38(3):845–873, 2006.
- Dudley and Gutmann [1977] R.M. Dudley and S. Gutmann. Stopping times with given laws. In Sém. de Probab. XI (Strasbourg 1975/76), volume 581 of Lecture Notes in Math., pages 51–58. Springer, 1977.
- Ekström and Janson [2016] E. Ekström and S. Janson. The inverse first-passage problem and optimal stopping. Annals of Applied Probability, 26(5):3154–3177, 2016.
- Friedman [1982] A. Friedman. Variational Principles and Free-Boundary Problems. John Wiley & Sons, 1982.
- Guillemin and Pollack [1974] V. Guillemin and A. Pollack. Differential Topology. Prentice-Hall, 1974.
- Huang and Tian [2006] H. Huang and W. Tian. Constructing default boundaries. Banques & Marchés, pages 21–28, Jan.-Feb. 2006.
- Hull and White [2001] J. Hull and A. White. Valuing credit default swaps II: Modeling default correlations. Journal of Derivatives, 8(3):12–21, 2001.
- Iscoe and Kreinin [2002] I. Iscoe and A. Kreinin. Default boundary problem. Algorithmics Inc., Internal Paper, 2002.
- Kinderlehrer and Stampacchia [1980] D. Kinderlehrer and G. Stampacchia. An Introduction to Variational Inequalities and their Applications. Academic Press, New York and London, 1980.
- Krylov [1996] N.V. Krylov. Lectures on Elliptic and Parabolic Equations in Hölder Spaces. American Mathematical Society, Providence, Rhode Island, 1996.
- Ladyzhenskaya et al. [1968] O.A. Ladyzhenskaya, V.A. Solonnikov, and N.N. Ural’tseva. Linear and Quasi-Linear Equations of Parabolic Type, volume 23 of Translations of Mathematical Monographs. American Mathematical Society, Providence, Rhode Island, 1968.
- Lieberman [1996] G.M. Lieberman. Second Order Parabolic Differential Equations. World Scientific, Singapore, 1996.
- Peskir [2002] G. Peskir. On integral equations arising in the first-passage problem for Brownian motion. Integral Equations Appl., 14(4):397–423, 2002.
- Peskir and Shiryaev [2005] G. Peskir and A.N. Shiryaev. Optimal Stopping and Free Boundary Problems. Birkhäuser, 2005.
- Potiron [2021] Y. Potiron. Existence in the inverse Shiryaev problem. Available on arxiv.org, 2021.
- Protter and Weinberger [1967] M.H. Protter and H.F. Weinberger. Maximum Principles in Differential Equations. Springer, 1967.
- Zucca and Sacerdote [2009] C. Zucca and L. Sacerdote. On the inverse first-passage-time problem for a Wiener process. Annals of Applied Probability, 19(4):1319–1346, 2009.