A free boundary problem for semi-linear elliptic equation and its applications∗
Abstract.
In this paper, we consider a free boundary problem of a semilinear nonhomogeneous elliptic equation with Bernoulli’s type free boundary. The existence and regularity of the solution to the free boundary problem are established by use of the variational approach. In particular, we establish the Lipschitz continuity and non-degeneracy of a minimum, and regularity of the free boundary. As a direct and important application, the well-posedness results on the steady, incompressible inviscid jet and cavitational flow with general vorticity are also obtained in this paper.
1 The Institute of Mathematical Sciences,
The Chinese University of Hong Kong, Hong Kong.
2 Department of Mathematics, Sichuan University,
Chengdu 610064, P. R. China.
2010 Mathematics Subject Classification: 76B10; 76B03; 35Q31; 35J25.
Key words: Semilinear elliptic equation; Free boundary; Regularity; Incompressible flow; General vorticity.
1. Introduction
In this paper, we investigate the free boundary problem of a semilinear nonhomogeneous elliptic equation
(1.1) |
where is a connected open and bounded domain in , is a local Lipschitz graph. The given functions with and with , is so-called vorticity strength function and , .
The semilinear nonhomogeneous equation characterizes the steady, incompressible flow of an ideal fluid in with non-zero variable vorticity. A wide class of vorticity distributions is considered. We are interesting in the existence, regularity and geometric properties of the solution and the free boundary .
On the other hand, physical motivation for our study lies in the proof of minimizing the functional
which is related closely to the jet flow problem with variable vorticity. Here, , is the characteristic function of the set . Here and after, denote for simplicity.
In the remarkable paper [2] by H. Alt and L. Caffarelli, some results on Lipschitz continuity, non-degeneracy lemma of a minimizer , and the analyticity of the free boundary were obtained for the special case . The mathematical results of [2] were used immediately in the study of jet flows [3, 4, 5, 15] and impinging jet flows [12, 14] of inviscid, irrotational and incompressible fluid. For the nonhomogeneous problem (), A. Friedman [24] established the first result on the regularity of the solution and the free boundary for the linear nonhomogeneous case . Based on this result in [24], the well-posedness result of cavitational flow [16, 24] and impinging jet flows [13] of inviscid and incompressible fluid with constant vorticity were obtained. The similar results were extended to the quasilinear homogeneous case
in [6] and nonlinear homogeneous case
in [29].
The first purpose of this paper is to establish the existence, regularity of the solution to the free boundary problem (1.1), and extend the classical results in [2] to the semilinear nonhomogeneous elliptic equation. In particular, the Lipschitz continuity, non-degeneracy lemma of the solution and regularity of the free boundary are obtained (please see Theorem 2.4 for Lipschitz continuity and Theorem 3.15 for the regularity of the free boundary).
On the other hand, there is a large number of literatures on the regularity criteria of the free boundary for linear elliptic problem,
(1.2) |
For the Laplace operator and , L. Caffarelli showed in his pioneer work [9] that Lipschitz free boundary is -smooth, furthermore, he also showed in [10] that ”flat” free boundary is Lipschitz. And higher regularity, such as -smoothness and analyticity of the free boundary follow from the elegant work of D. Kinderlehrer and L. Nirenberg [27]. In the case of the homogeneous case , the regularity criteria results in spirit of works [9, 10] have been subsequently obtained for more general operators, such as concave fully nonlinear uniform elliptic operators of the form in [32, 33] and nonconcave fully nonlinear uniform elliptic operators of the form , and references [11, 21, 22]. Moreover, Silva showed that the Lipschitz free boundary of the problem (1.2) in non-homogenous case is in [30]. Recently, Weiss and Zhang [34] investigated the regularity of the free boundary problem of semilinear elliptic equation , and showed the regularity criteria that the free boundary is a graph implies the -regularity of the free boundary. However, we would like to emphasize that the results in this paper are not the conditional regularity of the free boundary and we do not give any apriori assumptions and topological property on the free boundary and the solution.
An other related interesting problem is the semilinear Dirichlet problem with degenerate gradient on the free boundaries,
(1.3) |
The problem is used for modeling the distribution of a gas with density, , in reaction with a porous catalyst pellet . Alt and Phillips [8] investigated the regularity and the geometry properties of the solution and the free boundaries with some special structural conditions on the force term . The degenerate gradient condition on the free boundaries implies that desired optimal regularity of the solution is in . However, in this paper, the gradient of the solution is non-zero on the free boundaries, thus the optimal regularity desired here is only Lipschitz in . This is the main difference between the free boundary problem (1.1) and the one (1.3).
The second purpose of this paper is to establish the well-posedness theory of the jet flows of inviscid, incompressible fluid with general vorticity issuing from a semi-infinitely long nozzle for symmetric case (see Figure 1). It’s a direct and important application of the mathematical theory of the free boundary problem (1.1).

Before we proceed with the bulk of this paper, we would like to mention the previous results on the well-posedness of the free boundary problem for incompressible inviscid jet flows.
For the irrotational inviscid incompressible flow without surface tension, the existence of an axisymmetric jet flow [5], an asymmetric jet flow [3] and impinging jet flow [12, 14] has been established based on the fundamental work [2]. The main advantage for two-dimensional irrotational flow is that the stream function solves the linear elliptic equation with Bernoulli’s type free boundary. And then the conformal mapping and Green function approach for the linear elliptic equation work for the irrotational flow. Furthermore, for the simplified setting of constant vorticity, which corresponds to a constant vorticity strength in (1.1), the well-posedness of axially symmetric cavitational flow in [24], symmetric impinging jet flow in [13] and symmetric cavitational flow in an infinitely long nozzle in [16] were established, based on the regularity of the free boundary problem with Poisson equation . The assumption of an inviscid incompressible jet with non-zero constant vorticity provides us with the simplest case of a flow that is not irrotational and is attractive for its analytical tractability. However, the simplicity setting is not a mere mathematical convenience, as it is also physical relevant. Indeed, the vorticity remains invariant along the each streamline for inviscid fluid, therefore the fluid possesses constant vorticity as long as we impose the constant vorticity in the upstream.
Generations of scientists working in fluid dynamics has recognized the importance of the vorticity. It has provided a powerful qualitative description for many of the important phenomena of fluid mechanics. From the mathematical point of view, the strong nonlinearity of the equations of vortex motion has made the analysis difficult.
In the present paper, to establish the well-posedness of the incompressible jet issuing from a semi-infinitely long nozzle with general vorticity, we will impose the vorticity in the inlet of the nozzle, and then the vorticity strength function is determined uniquely along the streamlines. To realize this idea, one of the key points is to show the well-posedness of the streamlines. A free boundary problem with semilinear nonhomogeneous elliptic equation as (1.1) is formulated. An associated variational problem is shown to possess a minimizer, which yields a solution of semilinear free boundary problem.
2. Regularity and non-degeneracy of the solutions
To investigate the free boundary of the semilinear elliptic equation (1.1), based on the works [2, 6, 24] by Alt, Caffarelli and Friedman, we study the variational problem in this section, and establish the non-degeneracy lemma and the regularity of the solution.
2.1. The variational problem
Let be a bounded and connected open domain and is a locally Lipschitz graph, is the characteristic function of a set and , which satisfies that
(2.1) |
where is a constant. Denote
It is easy to check that
(2.2) |
Define a function and satisfies
(2.3) |
Consider an energy functional
and an admissible set
where is a given subset of with , and a.e. on , and satisfies
(2.4) |
The variational problem : Find a , such that
Remark 2.1.
For the special case , the variational problem have been studied by Alt and Caffarelli in [2]. They established the Lipschitz continuity and non-degeneracy of the minimizer , and obtained the regularity of the free boundary. Moreover, Friedman in [24] investigated the variational problem for with and , and established the regularity of the free boundary.
As the first step, the existence of the minimizer to the variational problem can be obtained by using the similar arguments in Lemma 1.3 in [2], we omit it here.
2.2. Lipschitz continuity of the minimizer
In this subsection, we will obtain the Lipschitz continuity of the minimizer .
We first focus on the Hölder continuity of the minimizer in , and show that the minimizer satisfies the semilinear elliptic equation in .
Lemma 2.1.
(1) for some
and in .
(2) satisfies that
in | (2.5) |
in the weak sense and
in . | (2.6) |
Furthermore, for any compact subset of .
Proof.
(1) Let and define a function as follows
The maximum principle gives that in , and thus . Then we have
(2.7) |
which implies that
(2.8) |
With the aid of gradient -estimate in (2.8), we can now use the method of Morrey in Theorem 5.3.6 in [28] to deduce the Hölder continuity of the minimizer.
Set for any . It is clear that and in . Furthermore, if and only if in , and one has
(2.9) |
Since for any in (2.2), it follows from (2.9) that
which implies that
(2) For any with in , it follows from the statement (1) that and for any , then we have
which gives (2.5).
By virtue of the continuity of , we have that is open. For any , it is easy to check that in and for small . Obviously, and we have
which implies that
(2.10) |
It follows from the Schauder interior estimate in [26] that for any compact subset of .
∎
Denote
The dynamic boundary condition on the free boundary can be verified in the following.
Proposition 2.2.
If , for any minimizer , we have
(2.11) |
for any 2-vector , where is the outer normal vector to with . Similarly, if a segment is and on , we have
(2.12) |
Proof.
We define a diffeomorphism
by for any , where is a real number and is suitable small.
Denote
It’s easy to verify that and
where is the identity matrix. Then we have
(2.13) |
Due to the arbitrariness of , the linear term of (2.13) in has to vanish, and this gives that
(2.14) |
Hence, we obtain the conclusion (2.11). To prove (2.12), define for any with on and on , where is suitable small and is the outer normal vector of .
Along the similar arguments in (2.13) and (2.14), we have
(2.15) |
Since , is also up to . Therefore, we complete the proof of (2.12) by using (2.15).
∎
Remark 2.2.
Proposition 2.2 implies that if the free boundary is and is in uniformly up to the free boundary, then
Next, we will give the estimate for the Lipschitz norm for near the free boundary, this is the principal part in establishing the Lipschitz continuity of the minimizer . Denote
Lemma 2.3.
Suppose with , where . Then there exists a positive constant depending only on , such that
where is defined in (2.1).
Proof.
We assume that
(2.16) |
and derive an upper bound on . Denote for simplicity. By scaling and with , we have
in . |
On the other hand, it is easy to check that
Thanks to Harnack’s inequality in [31] and (2.16), we have
(2.17) |
where the constants and depend on .
Let be a point in . Define a function as follows
Since is a minimizer, by virtue of the convexity of , one has
and thus
(2.18) |
It follows from the statement (2) in Lemma 2.1 that in , and the maximum principle gives that
(2.19) |
Then it follows from (2.17) and (2.19) that
Using Harnack’s inequality in [31] again, we have
(2.20) |
We take for simplicity and introduce a function
Then a direct computation gives that
for , and sufficiently large . This gives that
provided that is sufficiently large. It follows from the maximum principle that
which implies that
This together with (2.20) gives that
(2.21) |
With the aid of (2.18) and (2.21), it follows from the similar arguments in Lemma 3.2 in [2] and Lemma 2.2 in [6] that
that is
where the constant depends only on .
∎
With the aid of Lemma 2.3, we will obtain the Lipschitz continuity of , which plays an essential part for constructing a Radon measure. To obtain the Lipschitz continuity, we should use the method of Harnack’s inequality in [31].
Theorem 2.4.
(1). The minimizer is Lipschitz continuous in , namely, .
(2). For any compact
subset of which containing a free boundary point, the
Lipschitz coefficient of in is estimated by
, and the constant depends only on , and .
Proof.
(1). Suppose that , where . Set
By virtue of Lemma 2.3, one has
where the constant depends only on .
Since in , it follows from the elliptic estimate for that
which implies that
Thus, for any compact subset of , is bounded in , where is a small neighborhood of the free boundary. It follows from Lemma 2.1 that in , this together with a.e. in gives that .
(2). Consider any connected domains and contains at least one free boundary point. Let and . Since contains some free boundary points, and we have that is not contained in , then we can find some finite points in ( depends only on and ), such that
and
It follows from Lemma 2.1 and Lemma 2.3 that
By virtue of Harnack’s inequality in [31], we have
and
Inductively, we have
(2.22) |
For any and . We consider the following two cases.
Case 1. , by using elliptic estimate and (2.22), one has
where is a constant depending only on , and .
Hence, we complete the proof of Theorem 2.4.
∎
2.3. Non-degeneracy of the minimizer
As a consequence of Theorem 2.4, we will obtain the non-degeneracy lemma of the minimizer in this subsection.
Lemma 2.5.
Let be a minimizer, for any compact subset of , there exists a positive constant (depending only on and ), such that for any disc ,
(2.23) |
implies that
Proof.
The non-degeneracy lemma will be stated in the following.
Lemma 2.6.
For any , there exists a positive constant , such that for any disc with ,
(2.24) |
implies that
Proof.
Without loss of generality, we assume that . Set and , we have
where . Denote
Since in , it follows from the maximum principle that
for any , which implies that
(2.25) |
Define a function solving the following problem
It is easy to check that
which yields that
(2.26) |
where is the outer normal vector, we have used the fact
due to the fact .
On the other hand, by virtue of the convexity of , we have that
which gives that
(2.27) |
The standard elliptic estimates give that
(2.28) |
Furthermore, with the aid of the trace theorem, it follows from (2.26) - (2.28) that
(2.29) |
where . By virtue of (2.24) and (2.25), one has
This implies that is small enough, provided that is small enough. This together with (2.29) implies that
for sufficiently small . This implies that in , if is small enough.
∎
Remark 2.3.
It should be noted that Lemma 2.5 remains valid, provided that is not contained in and on .
Finally, we give the density estimate of the free boundary point, which gives that for any compact subset of .
Lemma 2.7.
Let be a compact subset of , there exists a positive constant , such that for any disc with and small ,
(2.30) |
where is the two-dimensional Lebesgue measure.
Proof.
Step 1. Without loss of generality, we assume that . It follows from Lemma 2.6 that there exists a point , such that . Set , one has
which implies that
for small , that is
By virtue of non-degeneracy Lemma 2.6, one has
in , |
this gives the left-hand side of (2.30).
Step 2. Define a function solving the following problem
The maximum principle gives that in , and thus , where is defined in (2.3).
Then we have
which together with Poincaré’s inequality and Hölder inequality implies that
(2.31) |
For any point , since with , one has
(2.32) |
It follows from (2.32), Lemma 2.3 and Lemma 2.6 that
(2.33) |
provided that and are small.
Combining (2.31) and (2.33), we have
(2.34) |
for small and small . This gives the right-hand side of (2.30).
∎
3. Regularity of the free boundary
Based on the Lipschitz continuity and non-degeneracy lemma of the minimizer in Section 2, we will establish the regularity of the free boundary in this section.
3.1. Measure estimate of the free boundary
In this subsection, the main objective is to show that the set is finite perimeter locally in .
Set and . First, we will show that is a Radon measure supported on the free boundary .
Lemma 3.1.
is a positive Radon measure with support in . Moreover, is a Radon measure supported on and its singular point is contained in the free boundary.
Proof.
For any with , denote for any . Then we have
(3.1) |
for sufficiently large , where we have used the fact that in for large .
On the other hand, one has
(3.2) |
It follows from (3.1) and (3.2) that
(3.3) |
Taking in (3.3), we conclude that in the sense of distributions. Consequently, there exists a positive Radon measure supported on , such that .
Recalling , there exists a positive Radon measure supported on , such that .
∎
Next, we will give the estimate of the Radon measure .
Lemma 3.2.
Let be a compact subset of . There exist some positive constants and , such that for any disc with and ,
(3.4) |
Furthermore,
(3.5) |
Proof.
For any , taking a test function for a set , it is easy to check that converges to as and
(3.6) |
Taking in (3.6), it follows from Lemma 2.4 that
(3.7) |
where is the one-dimensional Hausdorff measure on and is the outer normal vector.
Let and be the Green function for Laplacian in with pole . If , then the pole is outside the support of the Radon measure , and thus
(3.8) |
Thanks to the non-degeneracy Lemma 2.6, there exists a point , such that for all small . Recalling that is Lipschitz continuous, we have
(3.9) |
for some small constant . Then it follows from (3.8) and (3.9) that
(3.10) |
provided that and are small enough.
On the other hand, we have
(3.11) |
where we have used the fact that . Combining (3.10) and (3.11), one has
It is clear that
this gives the right-hand side of (3.5). On the other hand, we have
which implies that
this gives the left-hand side of (3.5).
∎
We next introduce the representation theorem as follows, which implies that has finite perimeter.
Proposition 3.3.
(Representation Theorem) Let be a minimizer, there holds that:
(1). for any compact subset of .
(2). There exists a Borel function , such that
that is, for any ,
(3). For any compact subset of , with and small ,
for some positive constants independent of and .
Proof.
Thanks to the fact (3.5), and along the similar arguments in the proof of Theorem 4.5 in [2], we can obtain that
Denote , this gives the assertion (1).
Therefore, is absolutely continuous with respect to . It follows from Theorem 2.5.8 in [20] that there exists a Radon-Nikodym derivative , namely,
which together with (3.5) gives the assertion (3).
Then we have
for any , which gives that
due to . This yields the assertion (2).
∎
By virtue of (1) in Proposition 3.3, it follows from Theorem 1 in in [19] that the set has finite perimeter, that is, is Borel measure and the total variation is a Radon measure. Denote the reduced boundary of the set as
where is the unique unit vector with
if such a vector exists, and otherwise. Furthermore, it follows from Theorem 4.5.6 in [20] that
3.2. Blow-up limits
In this subsection, we study some properties of the so-called blow-up limits.
For any , take two sequences and with and , such that , and as . We call the sequence of functions defined by
as the blow-up sequence with respect to , . It follows from Lemma 2.4 that in any compact set of , provided that is large enough. Since , there exists a blow-up limit , such that for a subsequence ,
(3.12) |
and
Recalling the definition of the Hausdorff distance between two sets and as the infimum of the numbers , such that
Next, we will give the several convergences of the blow-up sequence, the idea borrows from the similar arguments for Laplace equation in Lemma 3.6 in Chapter 3 in [23].
Lemma 3.4.
The following properties hold:
(1)
(2)
(3) If , then .
(4)
(5) If and , then every blow-up limit with respect to is an absolute minimum for the functional in for any , namely,
(3.13) |
where .
Proof.
(1). For any , if , then there exists a small such that with . We next claim that
for sufficiently large . | (3.14) |
In fact, it follows from (3.12) that the claim (3.14) is true, if in .
If in , for any fixed small , there exists a , such that in for any , and one has
which together with the non-degeneracy Lemma 2.6 implies that for , this gives the claim (3.14).
Reversely, for any , if for a subsequence , then for small . Next, we claim that
(3.15) |
If in , we have
which implies that
The strong maximum principle yields that
which gives the claim (3.15).
It is easy to check that the claim (3.15) holds, in the case of in .
Hence, we obtain the convergence of the free boundary in the Hausdorff distance.
(2). For any , it follows from the results in Step 1 that there exists a sequence with , such that . Since , by using Lemma 2.5 and Lemma 2.6 for , we have
(3.16) |
for any , provided that is sufficiently large. Then taking in (3.16) gives that
(3.17) |
which together with Theorem 4.5 in [2] imply that
(3.18) |
Here, is the one-dimensional Hausdorff measure on . Consequently,
where is the two-dimensional Lebesgue measure on .
Let be an -neighborhood of , such that
(3.19) |
Hence, it follows from the results in Step 1 that
for sufficiently large , which together with (3.19) gives that
in . |
(3). If is a free boundary point of the minimizer , it follows from Lemma 2.5 and Lemma 2.6 that
(3.20) |
for any , provided that is sufficiently large. Then taking in (3.20), one has
which gives that .
(4). Let be any compact subset of , thanks to the standard elliptic estimates for , one has
(3.21) |
Next, we claim that
(3.22) |
Since is -measurable, it follows from Corollary 3 of Section 1.7 in [19] that
Denote
We next show that
(3.23) |
In fact, suppose not, we assume that for some with and . With the aid of (3.17), it follows from Theorem 4.3 and Remark 4.4 in [2] that is Lipschitz continuous, which implies that
This gives that has positive density at , which contradicts to . Thus, we obtain the fact (3.23).
With the aid of (3.12) and (3.23), for any , we have
provided that is sufficiently large, that is . It follows from the non-degeneracy Lemma 2.6 that in , which implies that in , and thus is open. Furthermore, one has
in any compact subset of , provided that is sufficiently large. |
This completes the proof of the claim (3.22).
(5). For any and on with , it suffices to show that
(3.24) |
Taking , it is easy to see that and . Set
It is easy to check that
Then we have
(3.25) |
By virtue of the results in the statements (2) and (4), taking in (3.25) gives that
(3.26) |
∎
3.3. Linear growth near the free boundary
In this subsection, we will obtain the gradient estimate of near the free boundary, and show that should grow linearly away from the free boundary. Namely,
(3.27) |
for , where and is the unit vector.
Lemma 3.5.
For any compact subset of , there exist some positive constants and depending only on , and , such that for any disc with , then
(3.28) |
Furthermore, there exists a , such that
(3.29) |
for any disc touching the free boundary with small .
Proof.
Denote for simplicity, and one has
Step 1. In this step, we will show that
(3.30) |
Denote , it suffices to prove that . In view of definition of , there exists a sequence with and . Let be the nearest point to and denote . Let be a blow-up limit of a sequence with respect to , without loss of generality, we assume that
By virtue of the statement (5) in Lemma 3.4, we have that is a minimizer. It follows from the similar arguments in Lemma 2.2 and Lemma 2.4 in [2] that is subharmonic in and in . Furthermore, and
(3.31) |
this gives that . Define , where . It is easy to check that is harmonic in . Moreover, it follows from (3.31) that
The strong maximum principle gives that
which together with implies that
Since in , we have that . Due to the uniqueness of the Cauchy problem for the Laplace equation, one has
Next, we claim that
(3.32) |
Suppose that the claim (3.32) is not true. Define
Since is a local minimizer, it follows from Corollary 3.3 in [2] that is Lipschitz continuous, and thus . Suppose , and let
Choose a blow-up sequence with respect to , let be the blow-up limit. Using above arguments again, we have
Since is a minimizer for (3.13) and is a free boundary point of , we can show that the set has density zero at any point , which contradicts to Lemma 3.7 in [2].
Therefore, we obtain that and as . For any , one has
provided that is small enough. It follows from the non-degeneracy Lemma 3.4 in [2] that in some strip . Thus, the proof of the claim (3.32) is done.
Finally, noting that is a minimizer to the variational problem (3.13), by means of Theorem 2.5 in [2], we can conclude that on the free boundary of , and thus .
Step 2. In this step, we will show that there exists a constant , such that
(3.33) |
for any disc . Denote and for simplicity.
Define for any , where . It is easy to check that
And thus is subharmonic function in . It follows from (3.30) that in a small neighborhood of the free boundary . We extend by and set
Then is superharmonic in . Furthermore, in and in . It follows from Theorem 8.26 in [26] that
(3.34) |
where we have used the fact in Lemma 2.7. Taking in (3.34), we have
which implies that
Thus we have
Step 3. For any disc touching the free boundary, there exists a free boundary point , such that . Since , the gradient estimate (3.28) gives that
(3.35) |
where , provided that is small.
∎
To obtain the linear growth (3.27) of near the free boundary, we next show that
Lemma 3.6.
For any compact subset of and disc with , then
(3.36) |
Proof.
Without loss of generality, we assume that , and denote for simplicity. For any with , define a function
Since , we have that , namely,
This gives that
(3.37) |
Taking , the Lipschitz continuity of gives that
in . |
Taking and
It follows from (3.37) that
(3.38) |
Since in , it is easy to check that
(3.39) |
By virtue of (3.28), (3.38) and (3.39), we have
where , which gives that
(3.40) |
∎
With the aid of Lemma 3.6, we have
Lemma 3.7.
For any blow-up limit of at , is a half plane solution with slope . Furthermore, the linear growth (3.27) holds.
Proof.
Let be a blow-up sequence with respect to for any , and with , it follows from Lemma 3.5 and Lemma 3.6 that
and
as . Those imply that for any blow-up limit , we have
On the other hand, is harmonic in , thus must be invariant in each connected component of . In fact, applying the operator and to , respectively, we have
This together with the fact gives that
where the vector and matrix as
Furthermore, one has
which implies
Thus, is a linear function, there exist a unit vector , and two numbers and , such that one has
(3.41) |
Since is a local minimizer, it follows from Lemma 3.7 in [2] that the set has a positive measure, this gives that and . In view of (3.41), one has
which implies the linear growth (3.27). ∎
3.4. Flatness of the free boundary
In this subsection, we will study the regularity of the free boundary , and obtain some flatness property of the free boundary point.
First, we introduce the relevant flatness class of the free boundary point (See the definition 5.1 in [2]).
Definition 3.1.
(Flat free boundary points) Let and . We say that is of the flatness class in with a unit vector (see Figure 2) if
(i) is a minimizer to the variational problem .
(ii) and
and
(iii) in and .

We next show that the flatness from above implies the flatness from below.
Lemma 3.8.
There exists a positive constant , such that for any small , if in and , then in (see Figure 3).

Proof.
We divide two steps to show this lemma.
Step 1. In this step, we will show that
(3.42) |
Denote , it suffices to show that . Taking a sequence with , such that , and
It follows from the non-degeneracy Lemma 2.6 that . Consider a blow-up sequence with respect to , where and . Taking a subsequence with a blow-up limit , such that
It is easy to check that . We next claim that
(3.43) |
In fact, the definition of gives that
Since is harmonic in and , the strong maximum principle gives the claim (3.43).
Step 2. Without loss of generality, we assume that and . Set , and , one has
(3.44) |
Let
and choose be the maximum with the property
Thus there exists a point . It is easy to check that , due to . Define a function solving the problem
Noticing that on , the maximum principle gives that
(3.45) |
Next, we will show that
(3.46) |
To see this, define a function as follows
where and the constants and depending on will be determined later. It is easy to check that
(3.47) |
Therefore, it follows from (3.47) and that
provided that
On the other hand, take , is sufficiently large and is small, one has
which together with the maximum principle gives that
Recalling that , we have
In view of , one has that on . The maximum principle gives that in .
Consider the points with and , and define a function solving the following problem
Hopf’s Lemma gives that
(3.48) |
Suppose that
for a positive constant . It should be noted that in , and the maximum principle gives that
Therefore, we have
where is the outer normal vector and we have used (3.48). This leads a contradiction, provided that is sufficiently large.
Hence, there exists some point , such that
(3.49) |
In the following, denote , we will investigate the blow-up limit. Thanks to Lemma 7.3 and Corollary 7.4 in [2], we have the following non-homogeneous blow-up limit and we omit the proof here.
Lemma 3.9.
(Non-homogenous blow up)Let in with , , and . For any , set
and
then for a subsequence,
Furthermore, and uniformly in , and is a continuous function.
Furthermore, we can show that the limit function is a convex function.
Lemma 3.10.
is convex with respect to .
Proof.
Set , and , one has
If the assertion is not true, there exists a and a linear function in , such that
Denote
and
for continuous function , where . Define a test function for a set , it is easy to check that converges to the characteristic function as .
Taking in (3.51), one has
which gives that
(3.52) |
Denote
It is easy to check that has finite perimeter in and
(3.53) |
By using the similar estimate in Page 136-Page 137 in [2], one has
(3.54) |
It follows from (3.52)-(3.54) that
(3.55) |
that is
which contradicts to the facts and .
∎
Lemma 3.11.
There exists a constant , such that for any ,
where .
Proof.
Set , and , one has
(3.56) |
It is easy to check that be the class of in , provided that is sufficiently large. Therefore, it suffices to show the lemma for , that is
(3.57) |
Set
Recalling the flatness condition for , the free boundary of lies in the strip . Since and , we have
in . |
The flatness condition implies that in . Therefore, one has
(3.58) |
It is easy to check that
(3.59) |
Thanks to the flatness conditions on and Lemma 3.7, there exists a subsequence still labeled by , such that
In view of (3.58), we can conclude that
(3.60) |
By virtue of , we have that for any , there exists a , such that
(3.62) |
Then it gives that
(3.63) |
Next, we will show that
(3.64) |
To obtain the fact (3.64), we first show that for any small and any large constant , we have
(3.65) |
as .
On another hand, to see this, thanks to the non-homogeneous blow-up Lemma 3.9, it suffices to show that
(3.66) |
It follows from (3.62) that for any , one has
(3.67) |
as , where we have used the fact
Taking any sequence with and , and consider in , where is the free boundary point and is an arbitrary large constant. Denote with . It follows from Lemma 3.9 that . Next, we claim that
(3.68) |
where is a constant depending on and , provided that is sufficiently large. In fact, the definition of gives that
Since in and we have that
and
where is a constant depending on and . Therefore, we obtain the claim (3.68).
For any , let be the solution of the Dirichlet problem
It follows from (3.65) that
(3.70) |
for any large constant (independent of and ), provided that is sufficiently large (depending on and ). Moreover,
provided that is sufficiently large. In view of (3.70), the maximum principe gives that
(3.71) |
Taking in (3.71), we have
Consequently,
(3.72) |
With the aid of Lemma 3.11, it follows from Lemma 7.7 - Lemma 7.9 in [2] that the flatness from below implies the flatness from above (see Figure 4).

Lemma 3.12.
For any , there exist a large constant , and a small , such that if
then
Lemma 3.13.
Let be a monotone increasing and continuous function with . For any , there exists a small , such that if
and for , then there exist a large constant and , such that
for some and , where and .
Proof.
Step 1. By virtue of Lemma 3.8, one has
Then for some small to be determined later, it follows from Lemma 3.12 that
where
To improve the value of , we define
It is easy to check that
For subharmonic function , we can use the similar arguments in P141 in [2] and obtain that there exists a , such that
which implies that
Denote
Taking be sufficiently small, such that and , then the continuity of gives that
Furthermore,
and
Thus we have
On the other hand, we have
Notice that
due to and .
Step 2. Repeating the arguments in Step 1, and choosing sufficiently small , we obtain
for some and , where
Similarly, we can repeat the arguments in Step 1 for a finite number , and we choose the constant being small enough in the statement for each step, thus
(3.74) |
for some and , where
(3.75) |
Applying Lemma 3.8 again, we have
3.5. The regularity of the free boundary
In Lemma 3.13, we show that the flatness in is improved in a smaller disc . Based on the results in the previous subsection, we will investigate the regularity of the free boundary in this subsection.
By virtue of Theorem 8.1 in [2], the flatness of free boundary implies immediately the -smoothness of the free boundary, which is the main result in this paper. We omit the proof and present the result as follows.
Theorem 3.14.
For any compact subset of , there exists and , such that if
(3.77) |
where , and
then
Namely, a graph in direction of a function, for any and on this curve,
In particular, if , then can be written in the form ; Furthermore,
Moreover, the free boundary possesses higher regularity, provided that the functions and the vorticity strength function possess higher regularity. With the aid of Lemma 3.7 and Theorem 3.14, by using the similar arguments in Theorem 3.12 and Corollary 3.13 in [24], we have
Theorem 3.15.
(1) If is and is , for ,
then the free boundary is locally in
with .
(2) If is and is , for
, then the free boundary is
locally in with .
(3) If and are analytic, then the free boundary
is locally analytic in .
Proof.
For any , let be a blow up sequence with respect to the discs , it follows from Lemma 3.7 that there exists a unit vector , such that
With the aid of Lemma 3.4, we have
which implies that there exists a sequence with , such that
provided that is sufficiently large. Applying Theorem 3.14, we have that is .
Next, we can take a transformation to flatten the free boundary. Then reflect to the full neighborhood of the free boundary, applying the Schauder estimates for the elliptic equations in divergence form in Section 9 in [1], we can obtain the regularity of up to the free boundary.
It follows from Proposition 2.2 that
Since is up to the free boundary and on the free boundary, we have
Without loss of generality, we assume that the outward normal direction to is in the direction of the positive -axis. Extend and as functions into a full neighborhood of . In view of , we have that
(3.78) |
Define a mapping as follows,
In view of (3.78), it is easy to check that
And thus the mapping is a local diffeomorphism near .
Construct a function as follows,
Therefore, the free boundary is transformed into , and we have
Consequently, one has
(3.79) |
It follows from (3.79) that
It is easy to check that is a quasilinear elliptic equation in a neighborhood of . Furthermore, satisfies the Neumann type boundary condition as follows,
(3.80) |
where .
Noting that is in near , which implies that the coefficients of the operator are in . By using the elliptic regularity in Section 9 in [1], we obtain that is in near . Furthermore, the free boundary can be denoted by , and thus the free boundary is near .
Applying the Schauder estimates for elliptic equations in [1], we can obtain the regularity of up to the free boundary . Using the above arguments again, we can conclude that the free boundary is near , provided that and .
By bootstrap argument, we can obtain the higher regularity of the free boundary .
Finally, if is analytic and is analytic, by virtue of the results of Section 6.7 in [28], we can conclude that is analytic. Hence, we obtain the analyticity of the free boundary .
∎
4. Applications: steady ideal jet flow and cavitational flow with general vorticity
As an important application of the mathematical theory of the free boundary problem (1.1), we will investigate the well-posedness of the steady incompressible inviscid fluid with free streamline. There are at least two classical hydrodynamical problems of two-dimensional steady flows with non-trivial vorticity can be described mathematically as a free boundary problem (1.1) for a semilinear elliptic equation, the one is the two-dimensional incompressible inviscid jet flow problem and another one is the two-dimensional incompressible inviscid cavitational flow problem. For a classical example, we will investigate briefly the existence and uniqueness of the two-dimensional incompressible inviscid jet flow issuing from a given semi-infinitely long nozzle in this section, and state the similar results on the incompressible cavitational flow problem.
In the previous sections, the technique of variational method has provided some solutions to the free boundary problem (1.1). The regularity of the free boundary and the regularity of the derivative to the solution up to it, follows from Theorem 3.15. However, the common topological properties of the free boundary between and are still unknown. Of course, it is an important and natural question to ask if the free boundary is smooth enough to provide a classical solution of the steady incompressible jet flow problem under consideration. The main aim of this section is to establish the existence and uniqueness of the steady incompressible jet flow issuing from symmetric semi-infinitely long nozzle via the mathematical theory on the free boundary problem established before.
4.1. Statement of the physical problem and main results
The problem we address is that the flow of an incompressible, inviscid fluid issues from a given symmetric semi-infinitely long nozzle and emerges as a jet with two symmetric free boundaries (see Figure 5). The flow is assumed to be both steady and irrotational. The free boundary initiates smoothly at the end point of the nozzle wall and extends to infinity in downstream, where the jet flow tends to some uniform flow.

Denote the upper nozzle wall of the semi-infinitely long plane symmetric nozzle, which satisfies that
, and . | (4.1) |
Let be the symmetric axis, and be the end point of the nozzle wall.
The problem of incompressible jet flow with general vorticity can be formulated as the free boundary problem of finding a domain in -plane, whose boundary consists of the upper nozzle wall , the symmetric axis , and a priori unknown curve expressed by for with
(4.2) |
The condition (4.2) is the so-called continuous fit condition and smooth fit condition, respectively.
A vector representing the horizontal velocity, the vertical velocity and the pressure of the flow in , which belongs to and satisfies the steady incompressible Euler equations
(4.3) |
and the slip boundary condition
(4.4) |
where is the normal direction of the boundary.
The free boundary is assumed to be a material surface of the incompressible fluid, and then the velocity still satisfies the slip boundary condition (4.4) on . Moreover, the classical assumption (neglecting the effects of surface tension) is constant pressure condition, namely,
Here, we denote the constant atmosphere pressure.
In the upstream, we assume that
(4.5) |
and denote the constant champer pressure in the inlet of the nozzle. Here, the variation and amplitude of the horizontal velocity are arbitrary, and it implies that the vorticity of the jet flow is arbitrary in the upstream.
Remark 4.1.
Once we impose the horizontal velocity in the inlet of the nozzle, the mass flux of incoming flow is determined by
Before we state the well-posedness results on the incompressible jet problem, it should be noted that there are two invariant quantities for the steady incompressible inviscid fluid, i.e.,
(4.6) |
and
(4.7) |
where denotes the vorticity of the fluid in two dimensions. In particular, the relation (4.7) gives that quantity called Bernoulli’s function remains invariant along the each streamline, which tells us two facts,
(1) the speed remains a constant denoted as along the free boundary .
(2) along the upper nozzle wall and the free boundary ,
as long as the continuous fit condition of the free boundary holds, where denotes the uniform pressure in the upstream. Here, the quantity is nothing but the pressure difference between the upstream and the downstream. It should be noted that the quantity is an undermined parameter here, and we will show that the appropriate choice is guaranteed by the continuous fit condition.
The incompressible jet problem is stated as follows.
The incompressible jet flow problem. Given a symmetric semi-infinitely long nozzle , a horizontal velocity in the inlet and a constant atmosphere pressure on the free surface, does there exist a unique symmetric incompressible inviscid jet flow issuing from the nozzle , and the free boundary initiates smoothly at the endpoint .
Moreover, we introduce the definition of a solution to the incompressible jet flow problem in the following.
Definition 4.1.
(A solution to the incompressible jet flow problem).
Given an atmosphere pressure and an incoming velocity
, a vector is called a solution to the incompressible jet
flow problem, provided that
(1) The free boundary can be expressed by a -smooth
function for any , and there exists an appropriate pressure difference , such that satisfies the continuous and smooth fit condition (4.2).
(2)
solves the Euler system (4.3), and satisfies the boundary
condition (4.4).
(3) There exists a unique positive constant such that
where is the asymptotic height of the free boundary in downstream.
(4) on .
(5) uniformly for , as .
The main results on the existence and uniqueness of the incompressible jet flow problem read as follows.
Theorem 4.1.
Suppose that the nozzle wall satisfies the assumption condition (4.1). Assume that the horizontal velocity in upstream satisfies that
(4.8) |
Then, there exist a unique difference pressure and a unique
solution
to the incompressible jet flow problem, such that
(1) The jet flow satisfies the following asymptotic behavior in the
far fields,
uniformly in any compact subset of , as , and
uniformly in any compact subset of , as
where , and and are uniquely
determined by , and .
(2) in .
(3) for any .
Remark 4.2.
The assumption in (4.8) follows from the symmetry of the incompressible jet flow. However, if we consider an asymmetric incompressible jet flow, the condition can be instead of .
Remark 4.3.
To obtain the continuous fit condition, we choose the difference pressure as a parameter, and then show that there exists a unique , such that the free boundary satisfies the continuous fit condition in (4.2).
Similarly, we can obtain the well-posedness results on the incompressible cavitational flow problem. We will give the statement of the physical problem and the well-posedness results as follows.

The incompressible cavitational flow problem. Given a two-dimensional symmetric obstacle , atmosphere pressure and the horizontal velocity in the upstream, does there exist a unique incompressible symmetric inviscid cavitational flow past the given obstacle, and the free boundary initiates smoothly at the corner of the obstacle?
Definition 4.2.
(A solution to the incompressible cavitational flow problem).
A vector is called a solution to the incompressible cavitational flow
problem, provided that
(1) The free boundary can be expressed by a -smooth
function for any , such that is , namely,
(4.10) |
(2)
solves the Euler system (4.3), and satisfies the boundary
condition (4.6), where
is the flow field bounded by and .
(3) There exists a positive constant such that
where is the asymptotic height of the free boundary in downstream.
(4) on .
(5) uniformly for , as .
We give the results as follows.
Theorem 4.2.
Suppose that the solid wall satisfies the assumption (4.9). Given an atmosphere pressure and , which satisfies that
(4.11) |
Then, there exist a unique difference pressure
and a unique solution
to the cavitational flow problem, where denotes the pressure in the inlet of the channel. Furthermore,
(1) The cavitational flow satisfies the following asymptotic
behavior in the far fields,
uniformly in any compact subset of , as , and
uniformly in any compact subset of , as
where and are uniquely
determined by , and .
(2) in .
(3) for any .
4.2. Reformulation of the free boundary problem
In the following, we will reformulate the original physical problem into a free boundary problem of a semilinear elliptic equation. The similar idea has been adapted in the compressible subsonic flows with general vorticity in an infinitely long nozzle in [17, 18, 35].
First, it follows from the continuity equation in the incompressible Euler system (4.3) that there exists a stream function , such that
(4.12) |
Second, suppose that the streamlines are well-defined in the whole flow fluid, and thus the incompressible jet flow problem can be solved along the each streamline as follows. On another hand, the positivity of the horizontal velocity of the jet will be verified later, which gives the well-definedness of the streamlines in the whole flow fluid.

Denote the possible flow field (see Figure 7) and the flow field. For any point , it can be pulled back along one streamline to the point in the inlet. Thus
which implies that
(4.13) |
It’s easy to see that the function can be solved uniquely by (4.13) provided that is -smooth. Meanwhile, it’s clear that . Due to the fact that the vorticity is invariable on each streamline, we obtain the governing equation to the stream function,
Furthermore, it is not difficult to check the following facts
(4.14) |
provided that satisfies the assumption (4.8).
On another hand, we can impose the Dirichlet boundary conditions,
where . Moreover, the free boundary can be defined as
(4.15) |
The constant pressure condition on the free boundary together with the Bernoulli’s law gives that the speed remains a constant on , denote the constant speed, i.e.,
Clearly, the constant is determined uniquely by .
Therefore, we formulate the following free boundary problem of the stream function that
(4.16) |
4.3. Uniqueness of asymptotic behavior in downstream
Next, we will show that the asymptotic behavior of the jet flow in downstream can be determined by the state of the incoming flow and the difference pressure .
For any initiate point in the inlet, the well-definedness of the streamlines implies that there exists a unique point denoted in the downstream for any (see Figure 8), which can be pulled back to the imposed initiate point along a streamline.

Denote the horizontal velocity in the downstream, the conservation of mass and Bernoulli’s law gives that
and
Here, the difference pressure between the inlet and the downstream is regarded as a parameter. These also give the initial value problem to the function as follows,
(4.17) |
for any and . Thus,
(4.18) |
It is easy to check that is strictly decreasing with respect to the parameter . In particular, the asymptotic height of the jet in downstream is in fact .
Noting that for any and , thus there exists an inverse function , such that for any . Then the horizontal velocity in downstream can be solved uniquely by
(4.19) |
once is solved by the initiate value problem (4.17) for any .
Remark 4.4.
For any , it is easy to check that
Moreover, the asymptotic height is strictly increasing with respect to , and .
4.4. The variational approach
To solve the free boundary problem (4.16), we introduce a variational problem with a parameter . For any , denote the truncated domain (see Figure 9),
and
and

Since, a priorily, one does not know whether the stream function satisfies that . Hence, we need extend the function as follows
(4.20) |
Set
(4.21) |
For , define a function , which satisfies that
By virtue of (4.14), it is easy to check that
Firstly, we give the following functional that
where the admissible set is defined as follows
and
(4.22) |
Truncated variational problem : For any and , find a such that
Remark 4.5.
Proposition 4.3.
For any and , there exists a minimizer to the variational problem with in .
Proof.
The existence of the minimizer to the variational problem follows from Lemma 2.1, and denote the minimizer to the variational problem .
Denote and for any . Thus, and . Moreover, if and only if in , and we have
due to for , which implies that
Similarly, we can show that
∎
Since in , we can remove the truncation of the function in (4.20). Denote the free boundary of as
Proposition 4.4.
The minimizer satisfies that
(1) .
(2) in
and in in the weak sense. Furthermore, in and
for any compact subset of
,
.
(3) The free boundary of the minimizer is .
(4) on the free boundary .
Moreover, on the segment and on
the segment .
The Statement (1) in Proposition 4.4 gives the Lipschitz continuity of the minimizer in the interior of . Next, we will show that the minimizer is also Lipschitz continuous near the boundary of .
Lemma 4.5.
is Lipschitz continuous in every compact subsect of that does not contain the point .
Proof.
We first consider the Lipschitz continuity of near the boundary . Denote and and for any . We consider the following two cases.
Case 1. , we can obtain the Lipschitz continuity of at by using the similar arguments in the proof Theorem 2.4.
Case 2. . Without loss of generality, we assume that for and . Set and . Let be a solution to the following boundary problem,
Since on , the maximum principle gives that
If , we have
Denote for , it follows from (4.23) that
By using the elliptic estimates in [26], we have
If , the elliptic estimate in [26] gives that .
Since in , the Lipschitz continuity of near can be obtained by using elliptic regularity.
∎
4.5. The uniqueness, monotonicity and free boundary of the minimizer
We first give the uniqueness of the minimizer for any given and , and show that is monotone increasing with respect to .
Once we have the regularity of the free boundary at hand, we will establish some topological properties of the free boundary, provided that some special geometric conditions on the solid boundaries are assumed. For example, it’s desired that the free boundary is -graph, as long as we assume that the nozzle wall is a -graph and the horizontal velocity of the incoming flow is positive.
To obtain this topological property of the free boundary, we will establish the monotonicity of the minimizer with respect to first.
Lemma 4.6.
The minimizer of the truncated variational problem is unique and is monotone increasing with respect to .
Proof.
Assume that and are two minimizers to the truncated variational problem . Set in for any .
Noticing that is a minimizer to the functional
and the admissible set , where . Extend in and in .
Next, we will show that
(4.25) |
Since and are minimizers, it suffices to show that
(4.26) |
Next, we claim that
(4.28) |
for small . Suppose that the claim (4.28) is not true, the continuity of and give that and in for small , then we have that is not a solution of in . In fact, if not, the maximum principle gives that in , due to . This contradicts to our assumptions.
Let be the solution of the following boundary value problem
Thus in , it is easy to check that
(4.29) |
Extend in , it follows from (4.25) and (4.29) that
which contradicts to the minimality of .
Since near , it follows from (4.28) that in the connected component of which contains an -neighborhood of . In view of the boundary value of , we have that is a connected arc. The maximum principle gives that any component of must touch the boundary . Therefore, the domain is connected and
(4.30) |
In particular, taking in (4.30), we have that the minimizer is monotone increasing with respect to .
∎
The monotonicity of with respect to implies that the free boundary is a -graph, and then there exists a function for , such that
To obtain the continuity of , we first give the following non-oscillation lemma.
Lemma 4.7.
(Non-oscillation Lemma) Suppose that there exist some constants with and a domain with dist for some , which is bounded by two disjointed arcs (, the lines and . Denote and the endpoints of for . Then there exists a constant , such that
Proof.
Denote and . Since in , we have
which implies that
(4.32) |
where we have used the Lipschitz continuity of due to Lemma 4.5.
∎
With the aid of Lemma 4.7, we have
Lemma 4.8.
is a continuous function in .
Proof.
By using the non-oscillation Lemma 4.7, it follows from the similar arguments in Lemma 5.4 in [5] that has at most one limit point as and . Moreover, has most one limit point as or as for any . It suffices to show that
If not, without loss of generality, we assume that there exists a point , such that . Denote with . The monotonicity of with respect to and the Lipschitz continuity of give that
where for small . Thus, is a part of the free boundary , it follows from (4) in Proposition 4.4 that
(4.33) |
It is easy to check that
The strong maximum principle gives that in . In view of on , thanks to Hopf’s lemma, we have
(4.34) |
On the other hand, it follows from (4.33) that
which contradicts to (4.34).
Hence, we obtain the continuity of the function .
∎
Next, we will introduce the bounded gradient lemma for near the boundary of .
Lemma 4.9.
For any , if for any , then for small , there exists a constant independent of and , such that
(4.35) |
Proof.
We first consider the case . It suffices to show that
for any small .
Denote , and . It follows from Proposition 4.4 that
Denote , it is easy to check that
Obviously, . Since the boundary is , and on , then the Harnack’s inequality is still valid up to the boundary . By using the similar arguments in Theorem 2.4 and Lemma 4.5, we can obtain that
where is a constant independent of . This implies the estimate (4.35).
For the other case with for any , we can show that the estimate (4.35) is still valid by using the above arguments.
∎
For the point , there is a possible case that and for any , and thus for the -regularity of and at , here we can not use the smooth fit condition in Theorem 6.1 and Lemma 6.4 in [7] directly. (see Figure 10)
Therefore, we will estimate the -regularity of and near the free boundary as follows.

Proposition 4.10.
For any , if for any , then we have
where is the outer normal vector to at . Moreover, is -smooth at .
Proof.
Without loss of generality, we assume and .
Suppose with , and is a blow-up limit of , where and for any . By virtue of the results of Subsection 3.2, there exists a blow-up limit , such that
(4.36) |
It is easy to check that , we next claim that
(4.37) |
Consider the complex -plane with . Let be a straight line in -plane with the direction , and passing through the origin. Set
Since the blow-up limit is still a harmonic function, we can use the similar arguments in the proof of Lemma 6.2 in [7] to show that
(4.38) |
Next, we will show that
(4.39) |
Suppose not, it follows from (4.38) that
and for some . | (4.40) |
For the harmonic function , one has
if const in , then is an open mapping in |
for any compact subsect of . This together with (4.40) implies that
(4.41) |
Since for any , it follows from (4.41) that is linear in respect to , and is a straight line of the form with . Thus, there exists a constant , such that
We next consider the following two cases for .
Case 1. . In view of Lemma 3.4, is a local minimizer in and , and Theorem 2.5 in [2] gives that , which contradicts to .
Case 2. . Since uniformly in any compact subset of , for any and , there exists a , such that
Noticing that there are free boundary points of in , the assumption implies that . Without loss of generality, we assume that . Therefore,
Introduce a function as follows
where
for a sequence . Denote the domain , and choose a largest , such that in and contains a free boundary point of . The can be chosen such that and .
Let be the solution of the following Dirichlet problem
where . Obviously, on and in , the strong maximum principle gives that in . Thanks to Hopf’s Lemma, one has
(4.42) |
where is the inner normal vector to at .
The definition of gives that , where the constant is independent of and . Applying the regularity theory for the semilinear elliptic equation yields that
where the constants are independent of and , which together with (4.42) implies that
This leads a contradiction to the assumption , provided that is small enough and is sufficiently large.
Hence, we complete the proof of (4.39).
Next, we will show the claim (4.37). Since is harmonic in , it follows from Lemma 7.2 in [7] that
(4.43) |
where , is the outer normal vector and the curvature if the streamline is concave to the fluid. Due to and on the free boundary , it follows from (4.43) that is convex to the domain , and the set is convex.
If in . Denote , it is easy to see that is harmonic in . The fact on implies that . In view of , one has
Therefore, thanks to Phragmén-Lindelöf Theorem (see also Theorem 1.1 in [25]), we have
which implies that is a function of and
It follows from Theorem 2.5 in [2] that , and thus the claim (4.37) holds.
If there exists a free boundary point of with , denote the maximal free boundary arc containing . We next consider the following two cases.
Case 1. The free boundary is a straight line . The fact in implies that the straight line must be parallel to the -axis. Moreover, below , due to that is monotone decreasing with respect to . Thus one has
Case 2. The free boundary is not a straight line. We next consider the following two subcases (see Figure 11).

Subcase 2.1. above . Since is monotone decreasing with respect to , we can rule out this subcase.
Subcase 2.2. below . The convexity of the free boundary implies that must intersect the -axis, denote the intersection point. Moreover, the convexity of and monotonicity of with respect to imply that the angle between and -axis at point lies in . Applying the similar arguments in the proof of (2.30) in [17], one has
(4.44) |
for small . Denote , using the non-degeneracy Lemma 3.4 and Remark 3.5 in [3], we have
(4.45) |
On the other hand, it follows from (4.44) and (4.45) that
which leads a contradiction, provided that is small enough.
Hence, the proof of the claim (4.37) is done.
With the aid of the claim (4.37), we will show that
(4.46) |
For any sequence of points with , let be the nearest point to . Denote
Next, we consider the following two cases.
Case 1. There exists a constant , such that . Define a blow-up sequence
It is easy to check that contains some free boundary points of and , which does not intersect the free boundary of , where . Moreover,
(4.47) |
There exist a blow-up limit and a point with , such that
Due to the fact that does not intersect the free boundary of and is -smooth near , one has
Then we have
(4.48) |
Case 2. . Define a blow-up sequence
Obviously, contains the free boundary points of for sufficiently large . Then there exist a blow-up limit and a point with , such that
where . Since the free boundary locally in the Hausdorff distance (see Lemma 3.4) and is a straight line, we can conclude that the free boundary of satisfies the flatness condition at in the direction , for sufficiently large . By virtue of Theorem 3.15, one has the free boundary of is uniformly in . The elliptic regularity gives that is uniformly in for some . Thus we have
4.6. The continuous fit condition for free boundary
In this subsection, we will check the continuous fit condition of the free boundary at , namely, for any , there exists a , such that
(4.49) |
To obtain the continuous fit condition (4.49), we first establish the continuous dependence of and with respect to the parameter .
Lemma 4.11.
If with , then
and
(4.50) |
Proof.
It follows from the similar arguments in the proof of Lemma 3.4, there exists a subsequence and a , such that
Furthermore, is a minimizer to the variational problem . The uniqueness of the minimizer to the variational problem gives that .
Suppose that the assertion (4.50) is not true, then there exists a sequence , such that
We consider the following three cases and derive a contradiction.
Case 1. . The monotonicity of with respect to gives that .
Next, we claim that
(4.51) |
where with .
It follows from the statement (4) in Proposition 4.4 that
(4.52) |
To obtain the claim (4.51), it suffices to show that
(4.53) |
Denote , and
for small , Proposition 4.4 gives that the free boundary is , then we have
in for some . |
For any fixed , then there exists a sequence with , such that as . In fact, suppose that there exists a small , such that and in , which gives that in and . The maximum principle implies that in , which contradicts to .
Take a small with , such that
(4.54) |
where as . Define a function as follows
where
(4.55) |
Denote the domain . It is easy to check that . Let to be the largest one, such that in , and . Furthermore,
Let be the solution of the following Dirichlet problem
where . It is clear that on , which gives that
(4.56) |
where is the inner normal vector to at .
Taking implies that
Thanks to the standard estimates of the solutions of the semilinear elliptic equation, we conclude that in converges to in in -sense. Denote It follows from (4.56) that
(4.57) |
Taking yields that and
which together with (4.53) gives that the claim (4.51) holds.
With the aid of the fact (4.51), we can obtain a contradiction by using the similar arguments in the proof of Lemma 4.8.
Case 2. and . Similar to the proof of the claim (4.51) in Case 1, one has
(4.58) |
where with , and we can obtain a contradiction.
Case 3. and . Since on , it follows from the statement (4) in Proposition 4.4 that
for sufficiently large , where .
Let be a domain bounded by , , and . Moreover,
Thanks to the non-oscillation Lemma 4.7 for in , there exists a constant independent of , such that
which gives a contradiction for sufficiently large .
∎
Next, we will study the relation between the initial point and for any .
Lemma 4.12.
For any , the initial point satisfies that
(1) there exists a constant (independent of ), such that
for any .
(2) there exists a (independent of ), such
that for any .
Proof.
(1) Suppose that there exist a free boundary point and a sufficiently large , such that . Then there exist a and a disc with (independent of ), such that . It follows from the non-degeneracy Lemma 2.6 and Remark 2.3 that
which leads to a contradiction for sufficiently large .
Similarly, we can obtain a contradiction if there is no free boundary point, taking with and using the non-degeneracy Lemma 2.5 for the boundary point (see Remark 2.3).
(2) Suppose not, then for any , there exists a , such that . In view of Remark 4.4, we have
the asymptotic height in downstream lies in , | (4.59) |
provided that is small. Let for , choosing to be the smallest one, such that
for . It follows from (4.59) that . We first show that
(4.60) |
Suppose not, there exists a point , such that . Then there exists a small such that
On the other hand,
The strong maximum principle gives that in , due to . Applying the strong maximum principle again, then in , which leads a contradiction.
The boundary value of gives that . Next, we claim that
(4.61) |
In fact, it follows from Proposition 4.10 that the free boundary is at and its tangent is in the direction of the positive -axis, the definition of implies the claim (4.61).
Thus, let be a free boundary point of . Since the free boundary is at , it follows from Hopf’s lemma that
where is the outer normal vector of at , which leads a contradiction.
∎
Moreover, set
(4.62) |
Proposition 4.13.
For any , there exists a as in (4.62), such that the free boundary satisfies the continuous fit condition and the smooth fit condition
(4.64) |
Moreover,
(4.65) |
and
(4.66) |
Proof.
Step 1. The definition of and the continuous dependence of with respect to imply that
which together with Proposition 4.10 gives that . Thus, the continuous fit condition and the smooth fit condition (4.64) hold.
Step 2. Suppose that the assertion (4.65) is not true, then there exists an , such that
(4.67) |
On the other hand, the fact implies that the asymptotic height . Thus we have
(4.68) |
For and is small, define
where is defined as in Subsection 4.3. Denote for any . Let be the smallest one, such that
for some . It follows from (4.68) that . The strong maximum principle implies that , and thus . Similar to the proof of Lemma 4.12, we can show that . We next consider the following three cases.
Case 1. . Since the free boundary is at , the Hopf’s lemma gives that
where is the outer normal vector of at , which contradicts to the assumption .
Case 2. and for small . In view of that the boundary is at , it follows from Hopf’s lemma that
(4.69) |
On the other hand, the statement (4) in Proposition 4.4 gives that
which contradicts to (4.69).
Step 3. In this step, we will show the assertion (4.66). Denote for any , and choosing the smallest , such that
for some . It suffices to show that
If not, suppose that . By virtue of the strong maximum principle and the boundary value of , we can choose to be the free boundary point of . Thus, we can derive a contradiction by using Hopf’s lemma.
∎
4.7. The existence of the incompressible jet flow
In previous subsections, we show that there exist a and a minimizer for any , such that the free boundary satisfies the continuous fit condition (4.64). Taking , we will obtain the existence of the solution to the jet flow problem in this subsection.
First, we consider the following variational problem.
The variational problem : For any , find a such that
for any with on , the admissible set
By using the similar arguments in Lemma 4.11, taking a sequence with , such that
as , and is a minimizer to the variational problem . It follows from (4.63) that
Lemma 4.6 gives that is monotone increasing with respect to , and the free boundary of is -graph. Moreover, can be described by a continuous function for , and
Theorem 3.15 and Proposition 2.2 give that the free boundary is and on , where is the outer normal vector.
Next, we will obtain the positivity of horizontal velocity.
Lemma 4.14.
in .
Proof.
Denote , one has
where . The monotonicity of with respect to implies that in , the strong maximum principle gives that
Since in and on , it follows from and Hopf’s lemma that
Similarly, we have
Next, we will show that on . Suppose that there exists , such that on . Since the free boundary is -smooth at , the normal outer vector is at and on . Thanks to Hopf’s lemma, we have
(4.72) |
By virtue of the boundedness of , we have at .
The positivity of horizontal velocity implies that the function is for any . ∎
Next, we will obtain the asymptotic height of the free boundary .
Lemma 4.15.
, where is the asymptotic height of the free boundary and .
Proof.
Suppose that the limit does not exist. It follows from the non-degeneracy 2.6 and (4.70) that
which implies that there exist two sequence and , such that
(4.73) |
with .
By virtue of Remark 4.4, there exist with , such that
(4.74) |
Set and the free boundary . Define a curve
where the function is defined as in (4.55). Let be the largest one, such that the curve touches the free boundary at some points , and thus
and . | (4.75) |
Let be the solution to the Dirichlet problem
(4.76) |
where domain is bounded by , , and .
Since in and , the maximum principle implies that in and
(4.77) |
where is the outer normal vector of .
By virtue of (4.75), we can choose a subsequence , such that
(4.78) |
After a translation in the direction, such that the free boundary point lies on the -axis, then we have
(4.79) |
and uniformly in any compact subset of . Moreover, satisfies that and
(4.80) |
By using the similar arguments in [35], we can show that the Dirichlet problem (4.80) has a unique solution
where with , and is defined as in Subsection 4.3.
Next, we will show that
(4.82) |
.
Define , as the free boundary of and a curve
Let be the largest one, such that the curve touches the free boundary at a point , and thus
and . |
Denote a domain , which is bounded by , , and . Let be the solutions to the Dirichlet problem (4.76), replacing and by and . Obviously, in , it follows from the maximum principle that in and
(4.83) |
where is the outer normal vector of . By using the previous arguments, we can show that there exists a subsequence , such that
and uniformly in any compact subset of and
Thus, there exists an asymptotic height of the free boundary , such that
Next, we claim that
If not, without loss of generality, we assume that , and there exists a , such that . Similar to the proof of (4.81) and (4.82), we can show that
which contradicts to the assumption .
Finally, it follows from (4.71) that
∎
Finally, the asymptotic behavior of the jet flow will be obtained in the following.
Proposition 4.16.
The rotational jet flow satisfies the following asymptotic behavior in the far fields,
uniformly in any compact subset of , as , and
uniformly in any compact subset of , as where and are uniquely determined by , and as in Remark 4.4.
Proof.
For any sequence , by using the uniform elliptic estimate, there exists a subsequence , such that
where is any compact subsect of , and
(4.84) |
Then boundary value problem (4.84) possesses a unique solution
which implies that
uniformly in any compact subset of , as . The Bernoulli’s law gives that
uniformly in any compact subset of , as .
For any sequence with and , by virtue of Lemma 4.15, one has
(4.85) |
For any large and small , it follows from (4.85) that there exists a , such that the free boundary of lies in a -neighborhood of the line in , and thus the free boundary of satisfies the flatness condition in , provided that is sufficiently large. It follows from Theorem 3.15 that the free boundary of convergence to the line in norm, namely,
Applying Theorem 3.15 again, one has
By using the uniform elliptic estimate, there exists a subsequence , such that
and
(4.86) |
where .
∎
4.8. The uniqueness of the incompressible jet flow
In this subsection, we will obtain the uniqueness of and the solution to the jet flow problem.
Proposition 4.17.
and the solution to the jet flow problem are unique.
Proof.
Suppose that and are two solutions to the jet flow problem, which satisfy the conditions in Definition 4.1.
Step 1. In this step, we will show that . Suppose not, without loss of generality, we assume that . In view of the asymptotic behaviors of and in Proposition 4.16, one has
(4.88) |
Consider a function for and is the free boundary of , choosing the smallest such that
(4.89) |
Similar to the proof of (4.60), the strong maximum principle implies that . We next consider the following two cases.
Case 1. , we can choose . Then one has
where is the outer normal vector. Since in , which yields that
which leads a contradiction to the assumption .
Case 2. , it follows from (4.88) that . Then we can take be the free boundary point of and , and the strong maximum principle gives that
Since the free boundaries and are at , it follows from Hopf’s lemma that
where is the outer normal vector to at . This leads a contradiction to the assumption .
Hence, we obtain the uniqueness of , and denote .
Step 2. In this step, we will show that . By virtue of the asymptotic behaviors of and in Proposition 4.16, we have
Without loss of generality, we assume that there exists some , such that
(4.90) |
Denote for , and let to be the smallest one, such that
(4.91) |
It follows from assumption (4.90) that and .
Similar to the proof of Case 1 in Step 1, we can take be the free boundary point of and , applying Hopf’s lemma yields that
where is the outer normal vector to at . This leads a contradiction. ∎
References
- [1] S. Agmon, A. Douglis, L. Nirenberg, Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions. I, Comm. Pure Appl. Math., 12, 623-727, (1959).
- [2] H. W. Alt, L. A. Caffarelli, Existence and regularity for a minimum problem with free boundary, J. Reine Angew. Math., 325, 105-144, (1981).
- [3] H. W. Alt, L. A. Caffarelli, A. Friedman, Asymmetric jet flows, Comm. Pure Appl. Math., 35, 29-68, (1982).
- [4] H. W. Alt, L. A. Caffarelli, A. Friedman, Jet flows with gravity, J. Reine Angew. Math., 35, 58-103, (1982).
- [5] H. W. Alt, L. A. Caffarelli, A. Friedman, Axially symmetric jet flows, Arch. Ration. Mech. Anal., 81, 97-149, (1983).
- [6] H. W. Alt, L. A. Caffarelli, A. Friedman, A free boundary problem for quasilinear elliptic equations, Ann. Scuola Norm. Sup. Pisa Cl. Sci., 11, 1-44, (1984).
- [7] H. W. Alt, L. A. Caffarelli, A. Friedman, Compressible flows of jets and cavities, J. Differential Equations, 56, 82-141, (1985).
- [8] H. W. Alt, D. Phillips, A free boundary problem for semilinear elliptic equations, J. Reine Angew. Math., 368, 63-107, (1986).
- [9] L. A. Caffarelli, A Harnack inequality approach to the regularity of free boundaries, I. Lipschitz free boundaries are , Rev. Mat. Iberoamericana, 3, 139-162, (1987).
- [10] L. A. Caffarelli, A Harnack inequality approach to the regularity of free boundaries. II. Flat free boundaries are Lipschitz, Comm. Pure Appl. Math., 42, 55-78, (1989).
- [11] M. C. Cerutti, F. Ferrari, S. Salsa, Two-phase problems for linear elliptic operators with variable coefficients: Lipschitz free boundaries are , Arch. Ration. Mech. Anal., 171, 329-348, (2004).
- [12] J. F. Cheng, L. L. Du, Hydrodynamic jet incident on an uneven wall, Math. Models Methods Appl. Sci., 28, 771-827, (2018).
- [13] J. F. Cheng, L. L. Du, Y, F, Wang, Two-dimensional impinging jets in hydrodynamic rotational flows, Ann. Inst. H. Poincaré Anal. Non Linéaire, 34, 1355-1386, (2017).
- [14] J. F. Cheng, L. L. Du, Y, F, Wang, On incompressible oblique impinging jet flows, J. Differential Equations, 265, 4687-4748, (2018).
- [15] J. F. Cheng, L. L. Du, W, Xiang, Incompressible jet flows in a de Laval nozzle with smooth detachment, Arch. Ration. Mech. Anal., 232, 1031-1072, (2019).
- [16] J. F. Cheng, L. L. Du, Q. Zhang, Two-dimensional cavity flow in an infinitely long channel with non-zero vorticity, J. Differential Equations, 263, 4126-4155, (2017).
- [17] L. L. Du, C. J. Xie, On subsonic Euler flows with stagnation points in two dimensional nozzles, Indiana Univ. Math. J., 63, 1499-1523, (2014).
- [18] L. L. Du, C. J. Xie, Z. P. Xin, Steady subsonic ideal flows through an infinitely long nozzle with large vorticity, Comm. Math. Phys., 328, 327-354, (2014).
- [19] L. C. Evans, Measure theory and fine properties of functions, Studies in Advanced Mathematics, American Mathematical Society, Providence, 1992.
- [20] H. Federer, Geometric measure theory, Berlin-Heidelberg, New York, 1969.
- [21] F. Ferrari, S. Salsa, Regularity of the free boundary in two-phase problems for linear elliptic operators, Adv. Math., 214, 288-322, (2007).
- [22] F. Ferrari, S. Salsa, Subsolutions of elliptic operators in divergence form and application to two-phase free boundary problems, Bound. Value Probl., Article 57049, 21 pp., (2007).
- [23] A. Friedman, Variational principles and free-boundary problems, Pure and Applied Mathematics, John Wiley Sons, Inc., New York, 1982.
- [24] A. Friedman, Axially symmetric cavities in rotational flows, Comm. Partial Differential Equations, 8, 949-997, (1983).
- [25] D. Gilbarg, The Phragmn-Lindelöf theorem for elliptic partical differential equations, J. Rational Mech. Anal., 1, 411-417, (1952).
- [26] D. Gilbarg, N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, Classics in Mathematics, Springer-Verlag, Berlin, 2001.
- [27] D. Kinderlehrer, L. Nirenberg, The smoothness of the free boundary in the one phase Stefan problem, Comm. Pure Appl. Math., 31, 257-282, (1978).
- [28] C. B. Morrey Jr, Multiple integrals in the calculus of variations, Springer-Verlag, Berlin, 1966.
- [29] S. Omata, A free boundary problem for nonlinear elliptic equations, Proc. Japan Acad. Ser. A Math. Sci., 66, 281-286, (1990).
- [30] D. De Silva, Free boundary regularity for a problem with right hand side, Interfaces Free Bound., 13, 223-238, (2011).
- [31] N. S. Trudinger, On Harnack type inequalities and their application to quasilinear elliptic equations, Comm. Pure Appl. Math., 20, 721-747, (1967).
- [32] P. Y. Wang, Regularity of free boundaries of two-phase problems for fully nonlinear elliptic equations of second order. I. Lipschitz free boundaries are , Comm. Pure Appl. Math., 53, 799-810, (2000).
- [33] P. Y. Wang, Regularity of free boundaries of two-phase problems for fully nonlinear elliptic equations of second order. II. Flat free boundaries are Lipschitz, Comm. Partial Differential Equations, 27, 1497-1514, (2002).
- [34] G. Weiss, G. H. Zhang, A free boundary approach to two-dimensional steady capillary gravity water waves, Arch. Ration. Mech. Anal., 203, 747-768, (2012).
- [35] C. J. Xie, Z. P. Xin, Existence of global steady subsonic Euler flows through infinitely long nozzle, SIAM J. Math. Anal., 42, 751-784, (2010).