One of the most important PDEs is the Laplace equation
The corresponding inhomogeneous PDE is Poisson’s equation
Both equations are linear PDEs of second order with the unknown function . A function that solves Laplace’s equation is called harmonic. As is typical with linear inhomogeneous equations, the sum of a solution of Poisson’s equation and a harmonic function is again a solution to Poisson’s equation.
These equations show up in many situations. In physics they describe for example the potential (also called the voltage) of an electric field in the vacuum with some distribution of charges . To give some more detail, perhaps you are familiar with Coulomb’s law: if we have a particle with charge at the origin and another particle with charge at , then the force on the second particle is
where is an empirical constant. (If you haven’t seen this before, it is very much like Newton’s equation for gravity.) If the charges have the same sign the force pushes the second particle in the direction (repulsion); if the opposite sign the force is in the direction (attraction). We interpret the bracket as the electric field of the first particle. Then this same rule could be stated that a positively-charged particle moves in the direction of the electric field and a negatively-charged particle in the opposite direction. And in fact this vector field is a gradient
For historical reasons, the potential is defined . So then we could say that a positively-charged particle tries to decrease the electric potential, like a ball rolling down a hill. The steeper the change in potential, the stronger the force. We will use this example of electric potential to give an interpretation of some of our results. Indeed, much of this theory was developed first by physicists and some techniques seem strange if one does not know the physics motivation!
The Laplace equation is invariant with respect to all rotations and translations of the Euclidean space . Therefore we first look for solutions which are invariant with respect to all rotations. These solutions depend only on the length of the position vector . For such functions we calculate:
Hence the Laplace equation simplifies to an ODE
Let us solve this ODE:
We see two things here. The space of solutions is two dimensional, with one solution being just the constant solution . The other solution is not a solution on all of because is has a singularity at the origin. Never-the-less these are important ‘solutions’ to consider!
Definition 3.1. Let be the following solutions of the Laplace equation:
Here denotes the volume of the unit ball in Euclidean space . We call these fundamental solutions of the Laplace equation.
This solution lies in the space of radially symmetric solutions. And notice that for it is the electric potential of a single particle. We have chosen , which makes the solution tend to zero for large . The constant is chosen in such a way that the following theorem holds:
Theorem 3.2. For a solution of Poisson’s equations is given by
Moreover, the distribution corresponding to the fundamental solution obeys .
Proof. We see that the function is twice continuously differentiable since is twice continuously differentiable and because it has compact support we can differentiate under the integral sign. We calculate
In particular, . We decompose this integral in the sum of an integral nearby the singularity of and an integral away from this singularity:
We use and and estimate the first integral for :
In the integral, because is second order, we can change to without changing signs. Then integration by parts yields
We are able to apply integration by parts because has compact support; we can restrict to some large ball without changing the integral. The second term converges in the limit to zero:
Another integration by parts of the first term yields
Here we used that is harmonic for . The gradient of is equal to . The outer normal of on points towards the origin and is given by the expression . Together . As we will prove rigorously in Lemma 3.3, the limit of as is . We can understand this intuitively by observing that for small and by continuity . Therefore
Putting these three limits together
Because the left hand side is independent of , we conclude that it must have been equal to all along.
It remains to prove the claim about distributions. For any test function we have per the definition of distribution derivative
But then we can see this as the calculation above with . The conclusion is that the value of the integral is . Moving the minus sign around we arrive at . But this is the definition of the delta distribution. □
In general, a fundamental solution of a constant coefficient linear PDE has the property that in the sense of distribution. We make these assumptions on so that is just the real-linear combination of partial derivatives, and so interacts well with convolution. In particular, if we apply to the convolution of and the fundamental solution
This shows that the convolution solves the inhomogeneous PDE as long as it is well defined and the derivative rule for convolutions holds.
To give the physics explanation, the fundamental solution is the potential of a single particle with unit charge. The charge of a particle is described by the delta distribution because it is only at a point but the total amount is finite. Consider the situation with two particles . This formula (pretending that is a function) says that their potential is
The interpretation is that if you have charges described by , then treat them as a sum (or integral) of particles. Each particle produces an electric potential , and the total potential is the sum (or integral).
Fundamental solutions are not usually unique however. Consider the present case of the Laplace equation. If we have any harmonic function then shows that is also a fundamental solution. The difference between two fundamental solutions solves the Laplace equation, so this is the only possibility for other fundamental solutions. Different fundamental solutions can produce different solutions to the PDE. We shall see that the fundamental solution we have chosen is the only one that vanishes at infinity, which makes it in some sense the best one.
The difference between the first and second claim of the theorem is the assumption of regularity of : twice continuously differentiable or smooth respectively. In fact it is possible to generalise this theorem further: the convolution of with is defined for continuous functions and belongs to . In this case the result of the convolution may not be differentiable but it is a solution of Poisson’s equation in the sense of distributions. However, if one assumes that is Lipschitz continuous and belongs to then is twice differentiable (in the usual sense) and solves the PDE. This situation is typical of the delicate questions of regularity of the solution.
In the previous section we constructed a solution to the inhomogeneous equation. Any other solution must differ from the constructed one by a harmonic function. We should therefore understand harmonic functions in order to understand the space of solutions. In this section we shall prove the following property of a harmonic function on an open domain : the value of at the center of any ball with compact closure in is equal to the mean of on the boundary of the ball. Conversely, if this holds for all balls with compact closure in , then is harmonic. This relation is called mean value property and has many important consequences.
Let us introduce some notation. Given a function let
be its spherical mean. Here denotes the volume of the unit ball in Euclidean space and equality follows from Lemma 2.9(iv) using . We write when the function and center point are clear.
The ball mean or of on the ball is
using the co-area formula, Lemma 2.11. Many statements can therefore be made either in terms of ball means or spherical means.
The spherical mean, and means generally, have several nice properties. First note that the normalisation constant in the definition ensures that and likewise for any other constant. The mean is real-linear in the function: , which just follows from linearity of the integral. Likewise it follows from the monotonicity of the integral that if then . From these basic properties follows continuity at the center:
Proof. By the definition of continuity for all there is a radius such that for all points we know . For any it follows that
But this is the definition that . □
Particularly important is the relationship between the spherical mean and the Laplacian of . Differentiating the spherical mean with respect to the radius and using the divergence theorem gives
Therefore if is harmonic then is constant. With these important properties of means prepared, we are ready to fully prove our claim.
Theorem 3.5 (Mean Value Property). Let on an open domain . We say that has the mean value property if
for all balls with . A twice continuously differentiable function has the mean value property if and only if it is harmonic. Additionally, the same result holds if ball means are used in place of spherical means.
Proof. We have just calculated that if is harmonic then is constant. From the previous lemma we then conclude that for all applicable . Conversely, if , then by the continuity of there is a ball where is strictly positive (or negative). For this ball and any ball contained in it the right hand side of equation (3.4) is strictly positive (or negative) and the spherical mean is strictly monotonic. Therefore it is not constant.
To show the statement about ball means relate it to the spherical means:
Thus if is constant and equal to , so is the ball mean. If the ball mean is constant and equal to then we differentiate both sides with respect to
Therefore too. □
Keeping with our theme of distributions, we might wonder how we can reinterpret the mean value property for distributions. As is typical for extending definitions to distributions, we first develop a formula for regular distributions. Suppose that is continuous and . For each point , we view the spherical mean as a function on . Therefore . For any test function we compute
Therefore we make the following definition for any distribution . For any there is a ball . The spherical mean of around is the distribution on with the formula
This is well-defined for two reasons. First, the support of excludes , so is identically zero on a neighborhood of . In particular, dividing by does not produce a singularity. And second, the support of is contained in . This shows that it is a test function on .
The mean value property is that the spherical means of the function are constant in the radius. Hence the corresponding property of distributions should require to be ‘constant’ in a suitable sense. We will prove in an exercise that a distribution corresponds to a constant function if and only if
Together we have
Definition 3.6 (Weak Mean Value Property). Let be a distribution on an open domain . It is called harmonic if in the sense of distributions. We say that has the weak mean value property if for each the respective spherical mean is a constant distribution. More explicitly, this means that for each ball with and each with the distribution vanishes on the test function .
What is the relationship of the weak mean value property to the (strong) mean value property? Suppose for a continuous function . If has the mean value property, then we observe that
In other words, for each the distribution corresponds to the constant function . Thus has the weak mean value property. Conversely, suppose that has the weak mean value property: For each point there is a constant such that . But we may use the fundamental lemma of the calculus of variations, Lemma 2.15, to conclude that . Hence the spherical mean of is constant in the radius. In summary:
The functions may look a little scary, but in fact they are actually friendly once you get to know them. They are smooth functions characterised by two properties:
they are radially symmetric around , and
they have compact support in .
It is clear that any has these two properties. If a smooth function has Property 1, then it is a function of the distance . Another way to state Property 2 is to say that the support is contained in an annulus centered at . Because it vanishes in a neighborhood of , there are no issues with the non-smoothness of at . So define to get the function with the relation .
These functions also behave well under convolution, so long as its the convolution of a ‘big annulus’ with a ‘little annulus’. By this we mean the following. Consider . Further suppose that is identically zero on and the support of lies in for . Then also obeys Property 1 and 2. Let us demonstrate this now. First, due to Lemma 2.13 we know that is rotationally symmetric around . Second, the convolution has compact support in by the addition formula for supports. It remains to show that it vanishes in a neighbourhood of . But this too follows from the addition formula for the support of a convolution, since .
There is a final point to be made about the total integral of these functions. Recall the formula . We apply this to the function , which has the mean value property, to get . Writing this out as integrals shows
In particular, the integral of is zero if and only if the integral of is zero. And as a reminder, when we introduced convolutions we noted that the integral of is the product of the integral of each function. Important to the proof below is that if has total integral zero, so too does . In particular, the weak mean value property applies to it.
Now we ready to prove that a distribution has the weak mean value property if and only if it is a harmonic distribution. This should be seen as a generalisation of Theorem 3.5. Something stronger comes out of this proof, a famous result known as Weyl’s lemma. It tells us that weak solutions of the Laplace equations coincide with the strong solutions, and all solutions are smooth.
Weyl’s Lemma 3.8. On an open domain , a distribution is harmonic if and only if it has the weak mean value property. For each harmonic distribution there exists a harmonic function with .
Proof. The steps of the proof are as follows:
We show that harmonic distributions have the weak mean value property.
For any distribution with the weak mean value property, we can define a function through spherical means. This function is smooth and harmonic.
We show that corresponds to the original distribution . So every distribution with the weak mean value property is a harmonic distribution.
Step 1. Suppose that is a harmonic distribution. Choose any point and suppose . For every with integral , we will show that there exists a test function with . This is sufficient to prove that has the weak mean value property because then .
By the assumption on that the total integral is zero we can define a test function through with . Then we define
This function depends only on . Because one end of the integral is set at and has compact support, has compact support in . Similarly it is constant on for some . For near therefore, . And for we can reuse the calculation of the Laplacian for radial function from the search for the fundamental solution:
Note
which implies
This concludes Step 1.
In Step 2, we assume that has the weak mean value property and construct a smooth harmonic function . For any open subset there is a radius such that . For all choose any with and define
Due to Lemma 2.16, is smooth. But we need to check that this definition is independent of the choice of . We can unwind the definitions of the convolution
Now suppose that is another choice. Then is a test function on with total integral zero (it is equal to the integral of minus the integral of , both of which are ). The weak mean value property now implies
Next we prove that the distribution has the weak mean value property. How does act on a test function ? Again this is answered by Lemma 2.16, . This formula simplifies a little due to being a radial function. Let be any function from the definition of the weak mean value property. Then we must show that . The trick is to use the freedom definition of to choose a suitable . We know that there is an such that vanishes on . We can choose such that its support lies inside the ball . Then by the discussion above we know that is again a function of the form considered in the weak mean value property. Therefore . In other words has the weak mean value property. By Lemma 3.7, has the mean value property and further by Theorem 3.5, is harmonic.
Lastly, we have Step 3, where we prove . The functions have support and total integral . Thus the corresponding functions are a smooth mollifier on . We again use the freedom in the choice of to see that for every . Now Lemma 2.12 implies . □
To conclude this section we show that the mean value property leads to a growth estimate.
Corollary 3.9. Let be a harmonic function on an open domain and a ball with compact closure in . For all multi-indices we have the estimate
Proof. We have just seen in Weyl’s lemma that all harmonic functions are smooth and thus all partial derivatives of a harmonic function are harmonic. The mean value property and integration by parts (the divergence theorem version) yield for
The inductive application gives first , and using the induction hypothesis
the relation . The given is the solution. □
Proof. The foregoing corollary shows that is bounded by for each and . In the limit the first partial derivatives vanish identically. Therefore is constant. □
We have already mentioned the intuition that if a harmonic function is increasing in some direction then it must decreasing in another. This would imply that a harmonic function cannot have a local extremum, and this is indeed the case. Suppose a harmonic function has a maximum at a point of an open connected domain . The mean value property implies on all balls
By the fundamental lemma of the calculus of variations (or a standard argument from continuity), we conclude that for all . Hence takes the maximum on all these balls . This shows that the set is open. But it is also the preimage of a single value, and therefore closed. It is non-empty since by assumption does have a maximum. By the definition of connected, this set must be all of .
Strong Maximum Principle 3.11. If a harmonic function has on a connected open domain a maximum, then is constant. □
There is a more geometric proof in the case that is path connected. We again begin with showing that takes its maximum on every ball centered at in the domain. Since is path-connected every other point is connected with by a continuous path with and . The compact image is covered by finitely many balls with and . Supplementing the balls if necessary, we can assume that the center of each ball belongs to the previous ball. Then repeating the argument Inductively, is constantly on all these balls too and hence along , and on since this is true for all .
A practical consequence is the following
Weak Maximum Principle 3.12. Let the harmonic function on a bounded open domain extend continuously to the boundary . The maximum of is taken on the boundary .
Proof. By Heine Borel the closure is compact and the continuous function takes on a maximum. If it does not belong to , then is constant on the corresponding connected component and the maximum is also taken on . □
Since the negative of a harmonic function is harmonic the same conclusion holds for minima.
The triumph of the Maximum Principle is that it generalises to many elliptic operators (Definition 2.1), unlike the mean value property. It really goes to the heart of ellipticity.
Theorem 3.13. Let be an elliptic operator on a bounded open domain whose coefficients and extend continuously and elliptic to , and . Every twice differentiable solution of which extends continuously to takes its maximum on .
Proof. Let us first show that is uniform elliptic, i.e. there exists with
The continuous function attains on the compact set a minimum . Hence is uniform elliptic.
Next we use a trick to move to the case where of the function is strictly positive. For with we conclude
The continuous coefficients are bounded on the compact set . Therefore there exists with . By linearity of we obtain on for all .
Now we show that the continuous functions cannot attain a maximum on even though they must attain a maximum on . At any such interior maximum the first derivative of the function which is twice differentiable on vanishes and the Hessian is negative semi-definite. At this point we need a little bit of linear algebra to explain the connection between the Hessian and the Laplacian. The Hessian is a real symmetric matrix, so it is diagonalizable by an orthogonal matrix , that is . is a diagonal matrix whose entries are the eigenvalues of . Because is negative semidefinite, all the eigenvalues are negative or zero. In symbols . The Laplacian is the trace of the Hessian. Therefore
Similarly, for any elliptic operator
Because the eigenvalues are non-positive, we define . Continuing with the calculation
and this contradicts . Therefore for all the maximum of belongs to the boundary. Finally, we use the following comparison between and to reach the conclusion.
Because this holds for all the boundedness of on implies the theorem. □
The negative of the functions in the theorem obey and take a minimum on the boundary. In particular, the solutions of take the maximum and the minimum on the boundary.
Now let us see why maximum principles are so important. We consider the following very natural boundary value problem:
Dirichlet Problem 3.14. For a given function on a bounded open domain and on we look for a solution of on which extends continuously to and coincides there with .
The condition that extends continuously to the boundary is necessary for the boundary value problem to be meaningful. Otherwise the values on the boundary could be complete unrelated to the rest of the function. We say that a function is times continuously differentiable on the closure of an domain, if it is times continuously differentiable on and all partial derivatives of order at most extend continuously to .
Let be an open and bounded domain and suppose that there are two solutions and to the Dirichlet problem for the Poisson equation with inhomogeneous term and boundary value . Then the difference solves the homogeneous problem, i.e. it is harmonic, and on . Therefore by the weak maximum principle we know that both the maximum and minimum of on every connected component of is . The only possibility is that on all of . This shows that solutions to the Dirichlet problem are unique.
Putting this another way, we can uniquely determine a harmonic function if we know its values on the boundary of its domain. This gives us a way to understand the space of harmonic functions.
We just saw that the solution to the Dirichlet problem is unique, if a solution exists. In this section we try to find some conditions which ensure the existence.
First we prepare some well known formulas, which hopefully you have already proved as an exercise. In first formula we apply the Divergence Theorem to :
Green’s First Formula 3.15. Let the Divergence Theorem hold on the open and bounded domain . Then for two functions we have
If we subtract the formula for interchanged and , then we obtain:
Green’s Second Formula 3.16. Let the Divergence Theorem hold on the open and bounded domain . Then for two functions we have
The significance of these formulas becomes apparent when we apply them to the fundamental solution . This function is harmonic for , so we need to exclude a small ball . We apply Green’s second formula on the domain . The left hand side becomes
As argued in Theorem 3.2 (the part with ) this integral is well defined in the limit . For the right hand side of Green’s second formula, there are two boundary components to consider, namely and . The integrals over are of a type and respectively. We have in the limit
For the other integral, we must be very careful of signs. As required by the divergence theorem, let be the unit normal vector to that points towards . It can be expressed as . Therefore is the unit normal vector to pointing away from the origin. This is the opposite sign as the in Theorem 3.2. We have
Rearranging the terms gives
Green’s Representation Theorem 3.17. Let the Divergence Theorem hold on the open and bounded domain . Then for and a function we have
This representation formula allows us to reconstruct a function from its Laplacian and the values of and the normal derivative on . But the Weak Maximum Principle implies the function is already uniquely determined by its Laplacian and boundary values, the normal derivatives on the boundary are redundant information. The question is, how can we calculate the normal derivatives from the other two pieces of information? If the domain admits a function of the following type, then there is a clean formula.
Green’s Function 3.18. A function is called Green’s function for the bounded open domain , if it has the following two properties:
For the function extends to a harmonic function on .
For the function extends continuously to and vanishes on .
From the physics perspective, a Green’s function tells us the potential at of a single particle at if the potential is forced to be zero on the boundary. This is the case if the boundary is a metal cage (a Faraday cage). The first condition can also be expressed as in the sense of distributions, where is the delta distribution centered at . We could imagine expanding the definition of a Green’s function so that unbounded domains were allowed, but the potential has to go to zero ‘at infinity’ in the second condition. The shifted fundamental solution would then be a Green’s function of .
Let’s put them to use. We apply Green’s Second Formula to the function . It is a harmonic function on all of so there is no need to exclude a ball this time. Further, because we know the integrals with are well defined, so therefore are the ones with . We have
Now Green’s Representation Theorem implies
We should think of this as an improved version of Green’s representation formula, enabled by the existence of a Green’s function. We will shortly prove that conversely that if functions and have sufficient regularity, then
defines a function that solves the Dirichlet Problem. Therefore the Dirichlet Problem reduces to the search of the Green’s Function.
A Green’s function is unique. If there are two Green’s functions on , then their difference is harmonic for all :
and vanishes for . By the weak maximum principle, this difference must be zero. As an aside, if we return to the generalised case where , then the difference between two Green’s functions is a harmonic function that goes to zero at infinity. Therefore it is bounded and Liouville’s theorem tells us it is constant (and thus constantly zero). Therefore the shifted fundamental solution is the unique Green’s function for .
Further
Theorem 3.19 (Symmetry of the Green’s Function). If there is a Green’s Function for the bounded domain , then holds for all .
Proof. For let be sufficiently small, such that both balls and are disjoint subsets of . Green’s Second Formula implies for the domain and the functions and
For the estimate for in the proof of Theorem 3.2 shows that both second terms converge to zero. The calculation of in the proof of Theorem 3.2 carries over and shows that the first terms converge to and , respectively. □
Finding a Green’s function for an arbitrary domain can be difficult, and they do not even exist for all domains. However it is feasible for highly symmetric domains, and the advantage is that then the solution has a concrete formula. We shall calculate Green’s function for all balls in . Let us first restrict to the unit ball . The key is to try and add a harmonic function to that equals it on the boundary. We may use the inversion in the unit sphere . It maps the inside of the unit ball to the outside and vice versa, fixing the boundary.
Proof. Fix . There are two properties that we must satisfy. First the function should extend to a harmonic function on all . Observe that is a point outside unit ball, so is never zero and thus this function is well-defined for all . Moreover, we have proved in a exercise that composing a harmonic function with rescaling, reflection or translation of its domain creates another harmonic function.
For the vanishing on the boundary, note that there is no problem extending for , because and are not in . To show that it’s zero, we need some geometry. For we have
Because is a function that only depends on the length of its argument, and are equal on the boundary . □
Although the definition of appears to treat and differently, in fact from the above proof, which does not use on , shows that the it is symmetric as expected.
The affine map is a diffeomorphism from onto and a homeomorphism from onto . We can use this coordinate change to transform a Dirichlet problem on the ball to one on . If solves on and then solves on and for . The same is true in reverse. Thus the ability to solve the Dirichlet on one ball confers the ability to solve the Dirichlet problem on every ball (and the same for other domains related by similarity).
We can use this insight to give the Green’s function for a general ball. We use an equivalent characterisation of the Green’s function: for every the harmonic difference is a solution to the Dirichlet problem
(This gives an alternative proof of uniqueness.) For and a point the related Dirichlet problem on the unit ball is with on and
for . By linearity and since constant functions are harmonic, we can write down the unique solution on by inspection:
Putting this all together gives
It remains to prove therefore that taking the Green’s representation formula and inserting and with sufficient regularity does indeed define a solution to the Dirichlet problem. We do this only for the specific example of the unit ball, but by the above discussion an analogous result will hold for any ball.
Poisson’s Representation Formula 3.21. For , and the unique solution of the Dirichlet Problem on is given by
Proof. It suffices to consider the two cases and separately.
Consider first. The essential point is the symmetry of the Green’s function, so whatever properties hold in the second variable also hold in the first. From Theorem 3.2 we have function that satisfies . Their difference has the formula
But the bracketed expression is harmonic in and therefore is harmonic. This shows that . Moreover, we know that is zero for and hence so too is .
The case is the new part. We define the Poisson kernel . By the Symmetry of the Green’s Function the function is harmonic. Hence for the given function is harmonic. It remains to show
extends continuously to and coincides there with . The issue is that the integral is over so their is a singularity in the integration in this limit. We compute for and (the reader should check this same formula holds for too):
This clearly shows the singularity at but that for all other it is zero. We observe
the integral kernel is positive for .
The following formula, which follows from Green’s Representation Formula for the function on the domain :
For all , , and there is the bound Therefore the family of functions converge uniformly to zero for on .
We will now prove that for continuous the properties (i)-(iii) ensure that in the limit the family of functions converge on uniformly to . For any , , and we have estimate
Therefore for any and we have the uniform estimate
Taking the limit we see that the limit is bounded by the second term for any , since the first term tends to zero. But the second term can be arbitrarily small, and therefore the uniform limit must be zero. This proves the claim. □
A harmonic function on which extends continuously to obeys
Like the Weak Maximum Principle, this shows that is completely determined by the values on , except here the result is constructive. One can also integrate this formula in over a ball, and after interchanging the integral and using some geometry, arrive at the Mean Value property.
One new consequence of this formula is an additional regularity result for harmonic functions. The dependence on in the formula is well-behaved for with , because is bounded away from its singularity. Therefore partial derivatives of with respect to can be expressed with similar formulas depending only on the values of on a fixed ball . For all the Taylor series of in converges uniformly to . This implies:
Another regularity result, which speaks to the connection between harmonic functions and holomorphic functions (if you know some complex analysis), is the so called ‘removable singularities’ theorem:
Lemma 3.23. Let be an open neighbourhood of and a bounded harmonic function on . Then extends as a harmonic function to .
Proof. On a ball with compact closure in , Theorem 3.21 gives a harmonic function which coincides on with . The family of harmonic functions on vanish on . If for any the function takes on a negative value, then due to the boundedness of and and the unboundedness of the harmonic function has a negative minimum on . This contradicts the Strong Maximum Principle. Hence is non-negative. Analogously us for negative non-positive. Otherwise would have a positive maximum in . In both limits and vanishes identically on and is a harmonic extension of to . □
The proof shows a slightly stronger statement. Each harmonic function on whose absolute value is for all bounded by on with sufficiently small depending on has an harmonic extension to .
In this optional section we deliver on the promise in Section 2.2 to give a PDE without any solutions. The key is the following lemma, which shows that no (nontrivial) function has a Laplacian that grows negatively faster than the function grows. This should be compared to Liouville’s theorem 3.10, in which a growth bound is used to show that a harmonic function (a solution to ) is constant. Then we only need to construct a PDE which implies this property but that does not solve. The idea and lemma come from the paper “Nonexistence of weak solutions for some degenerate elliptic and parabolic problems on Rn” (Mitidieri and Pohozaev, 2001).
Proof. The trick is to choose a particular family of test functions and use them to derive decreasing bounds on the integral of that can only be satisfied by . Choose a smooth bump function on that has the value for , the value for , and is monotonic increasing/decreasing for . We define
Because they are radially symmetric, it is easy to describe their supports:
So is positive on the open annulus . Likewise on the closed annulus .
We will bound the integral of on . Because we are working with non-negative functions we can increase the domain of the integral:
Now we apply Green’s second formula on . The test function and all its derivatives vanish on the boundary , the result is to transfer the Laplacian to .
The introduction of these strange factors will become clear in a moment. In the next step we use a result you might not know. You should be familiar with the Cauchy-Schwarz inequality for vectors, which says for . But it holds for all inner products, including the inner product on functions.
Now we see that the choice of factors has created another on the right hand side. We can manipulate the inequality, by dividing across and squaring:
This bound is useful because does not appear on the right hand side, it is solely in terms of .
In the next phase of the proof we use the specific form of (until now, we have only used that Green’s formula applies to ). Recall that the Laplacian in polar coordinates is
We use the chain rule:
We substitute this into the integral and then make the change of variable (which implies and ):
Now we have an integral that doesn’t even depend on . Of course the precise value of the integral depends on the choice of , but it is possible to choose one such that the integral is finite. Therefore we have a bound
Finally we can prove the statement of the lemma. Choose any point and let be such that . Consider . For all we have since the integrand is positive and this is expanding the domain. But this implies for all . The only possibility is . But this implies on . Therefore . □
With this lemma it is easy to construct a PDE with no solutions, even before we impose any boundary conditions, namely . Any solution has the property
and therefore . But doesn’t solve the PDE.