We link the QUMOND theory with the Helmholtz–Weyl decomposition and introduce a new formula for the gradient of the Mondian potential using singular integral operators. This approach allows us to demonstrate that, under very general assumptions on the mass distribution, the Mondian potential is well-defined, once weakly differentiable, with its gradient given through the Helmholtz–Weyl decomposition. Furthermore, we establish that the gradient of the Mondian potential is an Lp vector field. These findings lay the foundation for a rigorous mathematical analysis of various issues within the realm of QUMOND. Further, we prove that the once weakly differentiable Mondian potential solves a second-order partial differential equation in distribution sense. Thus, the question arises whether the potential has second-order derivatives. We affirmatively answer this question in the situation of spherical symmetry, although our investigation reveals that the regularity of the second derivatives is weaker than anticipated. We doubt that a similarly general regularity result can be proven without symmetry assumptions. In conclusion, we explore the implications of our results for numerous problems within the domain of QUMOND.

About 40 years ago Milgrom1 proposed MOND (MOdified Newtonian Dynamics), a non-linear modification of Newton’s law of gravity motivated by profound challenges in astrophysics. The basic MOND paradigm introduces a critical acceleration a0 = 1.2 × 10−10 m s−2, stating that the real gravitational acceleration greal of an object and its acceleration gN expected from Newtonian gravity are related as follows:
Thus, in the regime of large accelerations, MOND predicts behavior consistent with Newtonian gravity. However, at extremely low accelerations MOND predicts that greal is proportional to the square root of the acceleration gN expected from Newtonian physics. Through its single modification, MOND provides explanations for numerous astrophysical phenomena, many of which were predicted a priori by MOND.2 While MOND very effectively describes dynamics on the scales of galaxies, it faces more serious problems on scales slightly larger than the solar system. Recent debates have emerged regarding whether the data from GAIA3,4 on wide binary stars supports MOND5,6 or contradicts it.7 

The present paper focuses on mathematical questions, analysing the equations that are used to describe Mondian physics. We study in detail whether these equations are well posed, introduce a new formula for the Mondian gravitational field using the Helmholtz–Weyl decomposition, and analyze the regularity of the Mondian potential and its derivatives.

When considering to replace Newton’s law of gravity by Mondian gravity, one is tempted to replace the Newtonian field UρN, which corresponds to some density ρ on R3, by
(1)
In this formula we simply added a collinear correction term to the Newtonian field. In view of the basic MOND paradigm, one demands that λ: [0, ) → [0, ) is such that
Then we get
But with this simple implementation one runs into a problem. The field (1) is in general not the gradient of some potential, thus leading to a loss of classical conservation laws of physics like conservation of momentum [Ref. 8, Sec. 6]. A more refined approach is required. Milgrom9 proposed a theory, which is called QUMOND (QUasi linear formulation of MOND), where the Mondian potential UρM is defined as the solution of the partial differential equation (PDE)
(2)
Is the above PDE well posed? Milgrom provided an explicit formula for its solution UρM, namely
(3)
But is this UρM well defined? And if yes, which regularity properties does it have? These questions we answer in the present paper. To do so we develope a new mathematical foundation for the QUMOND theory using the Helmholtz–Weyl decomposition. Simply put, the Helmholtz–Weyl decomposition states that every well-behaved vector field v in R3 that vanishes at infinity can be uniquely decomposed into an irrotational vector field plus a solenoidal vector field. While the solenoidal field has a vector potential, the irrotational field has a scalar potential U, and U satisfies the PDE
(4)
Comparing the PDEs (2) and (4), we see that we can identify v with the vector field (1) and the potential U with UρM. Thus, the vector field UρM should be the irrotational part of the vector field (1).

In this paper we use the Helmholtz–Weyl decomposition in the form proven by Galdi10 for Lp vector fields11 and introduce a new explicit formula for the irrotational part of a vector field on R3 using singular integral operators. These operators are used to derive a new, explicit expression for the Mondian gravitational field UρM too. This new formulation is very useful to analyze the PDE (2) and the regularity properties of UρM. It enables us to prove the following theorem:

Theorem I.1.

For every density ρ on R3 that has finite mass and is an Lp function for some p > 1, the corresponding Mondian potential UρM – defined as in (3) – is well defined and once weakly differentiable with UρM being the irrotational part of the vector field (1) in the sense of the Helmholtz–Weyl decomposition. UρM=UρN+Uρλ can be decomposed into an Lq vector field UρN plus an Lr vector field Uρλ with q > 3/2 and r > 3. The potential UρM solves the PDE (2) in distribution sense

Further in this paper, we analyze second derivatives of UρM. Under the additional assumptions that ρ is bounded and spherically symmetric, we prove that UρM is twice weakly differentiable and D2UρM is an Lr function. Using handwaving arguments, one would expect that this should hold for 1 < r < 6. But this is wrong. It is only possible to prove that D2UρMLr(R3) for 1 < r < 2 and this result is really optimal. Through counterexamples, we show that it is impossible to achieve such a regularity result for r > 2. This is a surprising fact and it is due to the square root appearing above in the basic MOND paradigm. We discuss why achieving similarly general regularity results for the second derivatives of UρM without assuming spherical symmetry seems doubtful.

The regularity results for UρM, UρM and D2UρM presented in this paper are essential for addressing further important questions with mathematical rigor. For example they enable us to examine whether initial value problems using Mondian gravity are well-posed, whether corresponding solutions conserve energy, or what the stability properties of stationary solutions are.

The outline of this paper is as follows. In Sec. II, we analyze Newtonian potentials, a prerequisite for analysing Mondian potentials later on, and we introduce the singular integral operators that are important for the rest of the paper. In Sec. III, we study the Helmholtz–Weyl theory from Ref. 10 and provide a new expression for the irrotational part of a vector field using the singular integral operators defined previously. In Sec. IV, we bring together the QUMOND theory from Ref. 9 and our new knowledge about the Helmholtz–Weyl decomposition to prove Theorem I.1. In Sec. V, we analyze the (non-)existence of second derivatives of the Mondian potential. In Sec. VI, we discuss how the results of this paper can be applied to many problems in QUMOND.

In this paper Newtonian potentials will play an important role in two different ways. On the one hand when we have a certain mass distribution with density ρ then UρN is the Newtonian gravitational potential that belongs to the density ρ. On the other hand in the QUMOND theory we must understand how to decompose a vector field v in its irrotational and its solenoidal part. Here the Newtonian potentials of the three components vi of the vector field play an important role. This we treat in Sec. III. Since we need Newtonian potentials of both densities ρ and components vi of vector fields, we use the “neutral” letter f for the source term of the Newtonian potential in the present section.

Given a measurable function f:R3R, the corresponding Newtonian gravitational potential UfN is given by
(5)
provided the convolution integral exists. Since the concrete value of the gravitational constant G does not affect our analysis we set it to unity. Next we want to introduce some useful singular integral operators. For ϵ > 0 and i, j = 1, 2, 3, we define
provided that the convolution integral on the right hand side exists. Since
the limit of Tijϵf for ϵ → 0 plays an important role in understanding the second derivatives of the Newtonian potential UfN. Further, it plays an important role for the Helmholtz–Weyl decomposition as we will see below and hence for the QUMOND theory. In the following two propositions we study this limit.

Proposition II.1.
For every ϵ > 0 and fCc1(R3)
and the limit
exists in L(R3). In particular

Proposition II.2.
Let 1 < p < . There is a Cp > 0 such that for every ϵ > 0 and fLp(R3)
and the limit
exists in Lp(R3) with
Thus, Tij:Lp(R3)Lp(R3) is a bounded, linear operator.

Proof of Proposition II.1 and II.2.
The statements follow quite directly from the literature. To apply the results from the literature, we have to verify that
satisfies the following four assumptions:
  1. Ωij must be homogeneous of degree 0, i.e., Ωij (δx) = Ωij(x) for all δ > 0, x ≠ 0. This is obviously true.

  2. Ωij must satisfy the cancellation property
    If ij, this is obviously true. If i = j this is also true, since
  3. Ωij must be bounded on {|x| = 1}. This is obviously true since Ωij is continuous on R3\{0}.

  4. Ωij must satisfy the following smoothness property: For
    must hold
    This is true since for x,xR3 with |x| = |x′| = 1 and |xx′| < δ we have
    Now Proposition II.2 follows directly from Ref. 12, Chap. II, Theorem 3 and Proposition II.1 follows from Ref. 13, Satz 2.2. In the formulation of the theorem in Ref. 13, Dietz does not mention the continuity of the Tijϵf, but studying her proof carefully one sees that she has proven the Hölder continuity of Tijϵf under the assumption that suppfB1 ≔ {|x| < 1}. This holds obviously also for every fCc1(R3) after a suitable scaling. If however one is interested solely in the continuity of Tijϵf, like we here in this paper, one could also simply apply the transformation yxy in the definition of Tijϵf and use standard results to deduce that Tijϵf is continuous.

Next we formulate regularity results for the Newtonian potential. Note that we have set the gravitational constant G to unity.

Lemma II.3.

Let fCc1+n(R3), nN0. Then the following holds

  • The Newtonian potential UfNC2+n(R3). Its first derivative is given by
    which, using integration by parts, can be written as
    The second derivative of UfN is given by
    where i, j = 1, 2, 3.
  • For every R > 0 there is a C > 0 such that
    and
    provided suppfBR where BR denotes the open ball about zero with radius R > 0.
  • UfN is the unique solution of
    in C2(R3)

Proof.
It is proven in Ref. 14, Lemma P1 that UfNC2(R3) if fCc1(R3) and that the formulae for the first derivatives hold. If fCc1+n(R3) with n ≥ 1, it follows directly from
that UfNC2+n(R3). To prove (a) it remains to verify the formula for the second derivatives. For every xR3 we have
Dominated convergences and integration by parts then yield
observe that the normal on {|y| = ϵ} is pointing inward and that there is no boundary term at infinity due to the compact support of f. Tijϵf converges uniformly to Tijf after Proposition II.1. If ij then
Hence the boundary term vanishes. If i = j then
As above the first term vanishes, however, the second one evaluates to 4πf(x)/3. In total we get
Let us turn to (b). Since suppfBR and f is bounded, one sees directly that
That
is proven in Ref. 14, Lemma P1.
It remains to show (c). It is stated in Ref. 14, Lemma P1 that UfN is the unique solution of
(6)
however the proof is omitted. So let us briefly summarize the proof of this well known fact. Since
we have
Thus
The asymptotic behavior of UfN(x) for |x| → follows from the compact support of f. That UfN is the unique solution of the PDE (6) follows from the strong maximum principle (Ref. 15, Theorem 2.2).□

Lemma II.4.

Let fL1Lp(R3) for a 1 < p < . Then UfNLloc1(R3) exists, is twice weakly differentiable and the formulae for UfN and xixjUfN from Lemma II.3 and the following estimates hold

  • If 1<p<32 and 3 < r < with 13+1p=1+1r then
  • If 1 < p < 3 and 32<s< with 23+1p=1+1s then
  • For every 1 < p <

Proof.

Reference 15, Theorem 9.9 implies that UfNLloc1(R3) is well defined, twice weakly differentiable and that the estimate for D2UfN holds. If UfN is defined by the formula from Lemma II.3, Ref. 12, Chap. V, Sec. 1.2, Theorem 1 implies that UfNLr(R3) and UfNLs(R3) with the desired estimates provided p < 3/2 and p < 3 respectively.

It remains to check that the weak derivatives UfN and D2UfN are indeed given by the formulae from Lemma II.3. Consider the first derivatives and take ϕCc(R3). The Hardy–Littlewood–Sobolev inequality (Ref. 16, Theorem 4.3) allows us to use Fubini:
Now Lemma II.3 implies
So the weak gradient of UfN is given by the formula for UfN from Lemma II.3.
Let 1 < p < . We study the second derivatives and take (fk)Cc1(R3) such that
Then Hölder, integration by parts and Lemma II.3 give
Thus the weak second derivatives of UfN are given by the same formula as in Lemma II.3.□

In the situation of spherical symmetry there is a second formula for the Newtonian field UfN, which often is quite useful.

Lemma II.5.
Let 1 < p < 3 and fL1Lp(R3), 0 be spherically symmetric. Then
for a.e. xR3 with r = |x| and
denoting the mass inside the ball with radius r.

Proof.

This lemma was first proven by Newton17 in a similar version. Below we give a proof of our modern version using Lp theory.

Assume that f would be continuous and compactly supported. Then MC1([0, )) with
Further
and
for r ≥ 0. For xR3 and r = |x|
is well defined. If r = |x| > 0, U is continuously differentiable with
Since
we have
Further
Since f is continuous,
and
for r → 0. Hence
for every i, j = 1, 2, 3. Thus
Further
and
Since by Lemma II.3 UfN is a solution of this PDE, too, and this solutions is unique
and
(7)
If now fL1Lp(R3), we take a sequence (fn)Cc(R3) of spherically symmetric densities such that
By Lemma II.4
(8)
where s > 3/2 with 1/p + 2/3 = 1 + 1/s. Set
Then for every r ≥ 0
Hence for all 0 < S < R
and
Together with (7) and (8) this implies that for a.e. xR3

Later on, we will make regular use of the following statement.

Lemma II.6.
If f,hL6/5(R3) then

Proof.

UfN,UhNL2(R3) and UfNL6(R3) according to Lemma II.4. Thus the first two integrals are well defined. By the Hardy–Littlewood–Sobolev inequality (Ref. 16, Theorem 4.3) also the third integral is well defined. If f,hCc(R3), integration by parts and ΔUhN=4πh give the above equalities of the integrals. Since Cc(R3)L6/5(R3) is dense and all three integrals above are continuous maps from L6/5×L6/5R, the above equalities hold for all f,hL6/5(R3).□

As stated in Theorem I.1, we want to prove that the gradient of the Mondian potential UρM is the irrotational part of the vector field UρN+λ(|UρN|)UρN. In this section we specify what we mean with the “irrotational part of a vector field.” To do so, we make use of the singular integral operators Tij introduced in the previous section about Newtonian potentials.

Definition III.1
(Irrotational part of a vector field). Let 1 < p < and vLp(R3) be a vector field. For i = 1, 2, 3 we define
We call the vector field HvLp(R3) the irrotational part of v. Observe that H is a bounded linear operator on Lp(R3) according to Proposition II.2.

We will see in Theorem III.4 that Hv is indeed the irrotational part of v in the sense of the Helmholtz–Weyl decomposition:

Notion.

We denote both the space of p-integrable functions on R3 and the space of vector fields on R3 having p-integrable components by Lp(R3). In expressions like “HvLp(R3)” or “UρNLq(R3),” it is evident from the context that we talk about vector fields. In ambiguous situations, we will always write “Let vLp(R3) be a vector field” if we talk about a vector field v, or “Let ρLp(R3)” if we talk about a function ρ.

Theorem III.2
(Helmholtz–Weyl decomposition). For every vector field vLp(R3), 1 < p < , exist uniquely determined vector fields v1Lirrp(R3) and v2Lsolp(R3) such that
where the two subspaces Lirrp(R3) and Lsolp(R3) of Lp(R3) are defined as follows:

Remark.

The space Wloc1,p(R3) denotes the Sobolev space of scalar functions U on R3 that are once weakly differentiable and that are locally integrable if taken to the power p, i.e., for each compact domain K the integral K|U|pdx is finite. Further, also the gradient of U must be locally integrable if taken to the power p, but observe that in the definition of Lirrp we additionally demanded that the gradient shall be an Lp vector field not only on every compact domain but on the entire space R3.

Proof.

In Ref. 10, Theorem III.1.2 it is proven that the Helmholtz–Weyl decomposition in the sense of Ref. 10, Eq. III.1.5 holds. This form of the theorem makes use of a different definition of the space Lsolp. However, in Ref. 10, Theorem III.2.3 it is proven that our definition here coincides with the definition used in Ref. 10, Theorem III.1.2.□

Let us study how the Helmholtz–Weyl decomposition looks like for smooth vector fields with compact support.

Lemma III.3.
Let vCc2(R3) be a vector field. For every 1 < p <
is the uniquely determined, irrotational part of the vector field v according to the Helmholtz–Weyl decomposition. Further

Proof.
From Definition III.1 follows directly that HvLp(R3) for every 1 < p < . Since by Lemma II.3
we have also
In particular HvLirrp(R3). Since Hv is a gradient, its rotation is zero. For the divergence we have with Lemma II.3
Further we get
for every 1 < p < with div(v2) = 0 classically. Hence v2Lsolp(R3).□
We are particularly interested in the Helmholtz–Weyl decomposition of vector fields vLp(R3). If p < 3, we can show that xjUvjN exists and, using the same formula as in Lemma III.3, it is easy to deduce that in this situation, too, the Helmholtz–Weyl decomposition of v is given by Hv + (vHv). If p ≥ 3 however (and this is the case of special interest in Mondian physics), the integral
does not necessarily converge. Nevertheless, also in this situation the Helmholtz–Weyl decomposition of v is given by Hv + (vHv) but we can no longer make use of the formula from Lemma III.3. The key ingredients to prove this explicit form of the Helmholtz–Weyl decomposition for vLp(R3) is that Lirrp(R3) and Lsolp(R3) are closed subsets of Lp(R3). For Lsolp(R3) this is proven in Ref. 10. For Lirrp(R3), Ref. 10 leaves this as an exercise to the reader (Exercise III.1.2). This exercise can be solved using Poincaré’s inequality. Using that Lsolp(R3) and Lirrp(R3) are closed subsets, leads to the following theorem:

Theorem III.4
(Explicit Helmholtz–Weyl decomposition). Let 1 < p < and vLp(R3) be a vector field. Then the uniquely determined Helmholtz–Weyl decomposition of v is given by
with

Proof.
We can approximate the vector field vLp(R3) with a sequence of vector fields (vk)Cc2(R3). By Lemma III.3 the Helmholtz–Weyl decomposition of vk is given by
with
By Ref. 10, Exercise III.1.2, Lirrp(R3) is a closed subset of Lp(R3). By Ref. 10, Theorem III.2.3, Lsolp(R3) is the closure of the set
with respect to the Lp-norm on R3. Hence, also Lsolp(R3) is a closed subset of Lp(R3). Since vkv in Lp(R3) for k and H is a bounded, linear operator on Lp vector fields, also HvkHv in Lp(R3). Thus,
and
is the uniquely determined Helmholtz–Weyl decomposition of v.□

Thus the vector field Hv as defined in Definition III.1 is indeed the irrotational part of the vector field v in the sense of the Helmholtz–Weyl decomposition. Before we close this section we prove two useful lemmas. First: For spherically symmetric vector fields the Helmholtz–Weyl decomposition is trivial.

Lemma III.5.
Let 1 < p < . Then for every spherically symmetric vector field vLp(R3)

Proof.
Let vLp(R3) be a spherically symmetric vector field. There exists a sequence (vk)Cc(R3) of spherically symmetric vector fields with
Since the vk are spherically symmetric
Hence, by standard results for vector calculus, there exist potentials (Uk)C(R3) such that for every kN
in particular vkLirrp(R3) and the uniqueness of the Helmholtz–Weyl decomposition implies
Since H:Lp(R3)Lp(R3) is continuous

And last in this section we prove the useful fact that the operator H is symmetric.

Lemma III.6.
Let 1 < p, q < with 1p+1q=1, and let vLp(R3), wLq(R3) be vector fields. Then

Proof.
Assume that vL1Lp(R3) and wL1Lq(R3). Since v,wL1(R3) we can apply Fubini and get that for every ϵ > 0 and i, j = 1, 2, 3
Hence by Hölder
Since L1Lp(R3)Lp(R3) and L1Lq(R3)Lq(R3) are dense, and H is continuous, it follows that for every vLp(R3) and wLq(R3)

In Ref. 9, Milgrom introduced the QUMOND theory where the Mondian potential UρM belonging to some density ρ is given as the solution of the PDE (2). Milgrom gave an explicit formula for the solution of this PDE - our Eq. (3). Here in this paper we take another way to approach the Mondian potential UρM. This new approach enables us to place the entire QUMOND theory on a more robust, mathematical foundation. We define
The theory of the previous section guarantees that the field UρM is indeed the gradient of some potential UρM. Further, it guarantees that UρM is an Lp vector field for some p > 1. To bring this new definition of UρM together with the theory from Ref. 9, we have to take a closer look on the potential UρM. How do we get an explicit formula for it? In Lemma III.3 we have seen that for a vector field
Hv is the gradient of
(9)
But if ρL1Lp(R3) for a 1 < p < 3, then
for a 3/2 < q < . If now λ(σ)a0/σ then
with 2q > 3. But then xjUvjN is not well defined; for this it would be necessary that v is an Lp vector field for some p < 3. However we can still recover a slight variation of Lemma III.3. If we compare Eq. (9) with Eq. (3), which is the definition of the Mondian potential in Ref. 9, we see that both are quite similar. We prove that Eq. (3) really yields a well defined potential and that its gradient is given by UρM as defined above.

For better readability we decompse UρM and write UρM=UρN+Uρλ where the gradient of UρN is UρN and the gradient of Uρλ shall be H(λ(|UρN|)UρN). In the next lemma we analyze Uρλ.

Lemma IV.1.
Assume that λ: (0, ) → (0, ) is measurable and that there is Λ > 0 such that λ(σ)Λ/σ, for every σ > 0 (thus the function λ remains in its physically motivated regime). Let ρL1Lp(R3) for a 1 < p < 3 and let 3/2 < q < with 2/3 + 1/p = 1 + 1/q. Set
Then
and

Remark.

The formula for Uρλ above is the same one as in Eq. (3) and it is almost the same one as (9). The only difference is the term “+y/|y|3,” which guarantees that the integral is finite.

Proof.
Let R > 0. First we prove that for
it holds
(10)
Let p, q be as stated above and let r be the dual exponent of 2q. Since 3 < 2q < ,
Then
Next observe that for all yR3\{0}, i, j = 1, 2, 3
Thus for x,yR3 with |x| ≤ R, |y| > 2R holds
Since for all 0 ≤ s ≤ 1
we estimate further
Hence
Thus Ref. 10 holds. Fubini then implies that
Hence the potential Uρλ is well defined. It remains to prove that
We write shortly
Let ϕCc(R3). Then
Thanks to Ref. 10 we can apply Fubini and, since
we have
Lemma II.3 and Proposition II.2 imply
As in the Proof of Lemma III.6 we have
Hence
Thus
In particular, the Helmholtz–Weyl decomposition Theorem III.4 implies
and

Taking Lemmas IV.1 and II.4 together gives

Corollary IV.2.
Let λ be as in Lemma IV.1 and ρL1Lp(R3) for some p > 1. Then
is well defined and once weakly differentiable. The gradient UρM can be decomposed into an Lq plus an Lr vector field, where UρN is an Lq vector field for some q > 3/2 and Uρλ is an Lr vector field for some r > 3.

Proof.

When ρL1Lp(R3) it follows from the interpolation formula that ρLp′ for every 1 < p′ < p, in particular for p′ < 3. Thus Lemmas II.4 and IV.1 imply that UρN and Uρλ are both well defined and that UρNLq(R3) for some q > 3/2 and UρλLr(R3) for some r > 3.□

To summarize, we have now proven that the potential UρM as given by Eq. (3) is really well defined. It is once weakly differentiable and its gradient UρM is the irrotational part of the vector field UρN+λ(|UρN|)UρN in the sense of the Helmholtz–Weyl decomposition. Further, when we write UρM=UρN+Uρλ, we have that UρN is an Lp vector field for some p > 3/2 and that Uρλ is an Lp vector field for some p > 3. Lastly, from the definition of the operator H (Definition III.1) we have a new explicit formula for Uρλ using singular integral operators. And this new formula is not only useful to analyze the regularity of Uρλ as done above but it is also very useful to verify that UρM is really a solution of the PDE (2) from the introduction. This we do in the last lemma of this section.

Lemma IV.3.
Let λ be as in Lemma IV.1 and ρL1Lp(R3) for some p > 1, then UρM solves the PDE
in distribution sense, i.e., for every ϕCc(R3)

Proof.
We have
Since UρN is already a gradient, we get H(UρN)=UρN. Thus
The operator H is symmetric (Lemma III.6) and hence
H(∇ϕ) = ∇ϕ since ∇ϕ is already a gradient. Thus it follows that
This means that UρM solves the PDE
in distribution sense.□

Taking all statements of the lemmas proven in this section together implies that Theorem I.1 from the introduction holds.

In the previous section we have shown that in QUMOND the potential UρM, which corresponds to some density ρ on R3, is well defined, once weakly differentiable and solves the second order PDE (2) in distribution sense. Does UρM also have second order derivatives?

Consider a density ρL1L(R3) with finite support. Such a density is a reasonable model for the distribution of mass in systems like globular clusters or galaxies. If we have such a density the interpolation formula and Lemma II.4 tell us that the second derivatives of the corresponding Newtonian potential are Lp-functions:
for every p ∈ (1, ). Therefore
for every α ∈ (0, 1) by Morrey’s inequality; the space C0,α(R3) denotes the space of bounded and Hölder continuous functions with Hölder exponent α on R3. Under quite general assumptions (see Lemma V.2) the function
is Hölder continuous with Hölder exponent 12. Thus
for every β(0,12). Assuming for simplicity that we are in spherical symmetry, we have by Lemma III.5
Taking a second look on Morrey’s inequality one could now expect that
for all 1 < p < 6. But this expectation proves deceptive. Why? Let us remain in the situation of spherical symmetry and for ρ = ρ(x) spherically symmetric study the divergence of
In view of Lemma II.5
where for convenience we assumed λ(σ)=1/σ, σ > 0. So
ρ is just fine, the second term can be controlled as expected above, but the third one will cause problems. We prove the following

Proposition V.1.
Let R > 0, 1 < p, q < and ρC1Lp(R3), 0, spherically symmetric. Then
if 1 < q < 6 and p > 3q/(6 − q) with C = C(p, q, R) > 0, and
if 1 < q < 2 and p > q/(2 − q) + q with C = C(p, q, R) > 0. With M(r) we denote the function

Proof.
Let 1 < q < 6 and 1 < p < with p > 3q/(6 − q). For r ≥ 0
Thus
Since
we have
Now we turn to the second estimate. Let 1 < q < 2, p > q/(2 − q) + q and r0 ≥ 0 be such that M(r0) = 0 and M(r) > 0 for all r > r0. Since ρC1(R3), M(r) ∈ C1 ([0, )) with
Hence (M(r))C1((r0,)) with
Assume that R > r0, then
with
Obviously α > 0, and further α < q since
Now we apply Hölder’s inequality and get
note that 0 < α < q < p. Since
we have
here we have used that pq/(2p − 2α) < 1 since
Thus
Since
and
we finally have

Now we can analyze derivatives of λ(|UρN|)UρN. However, before we do so, we need to strengthen the assumptions on λ that we made in Lemma IV.1. This we do in the next lemma, where we take a look on the derivative λ′. The assumptions of the next lemma imply that λ has the same regularity as in Lemma IV.1 and that additionally the function R3uλ(|u|)u is Hölder continuous.

Lemma V.2.
Assume that λC1((0, )), λ(σ) → 0 as σ and there is Λ > 0 such that −Λ/(2σ3/2) ≤ λ′(σ) ≤ 0, for σ > 0. Then
for every σ > 0 (as in Lemma IV.1) and there is a C > 0 such that for all u,vR3
with λ(|u|)u = 0 if u = 0.

We postpone the proof of this Lemma to the  Appendix and return our attention to the analysis of the second derivatives of UρM. Using Proposition V.1 we can control Lq-norms of the derivatives of the Mondian part
of the field UρM provided 1 < q < 2 and ρ ≥ 0 is spherically symmetric.

Lemma V.3.
Let 1 < q < 2, p>q2q+q, R > 0 and ρL1Lp(R3), 0, spherically symmetric. Assume that λ is as in Lemma V.2, then
with
where C = C (p, q, R) > 0.

Proof.
Since we are in spherical symmetry, Lemma II.5 gives
for better readability we suppress the x-argument on the left side. Using the abbreviation
we have
Thanks to Lemma V.2
(11)
for a C > 0 where
From this lemma follows further
(12)
Using the bounds for λ and λ′ we get
(13)
for a C > 0. Thanks to Eq. (12), for every R > 0 holds
Next we approximate ρ by smooth densities ρn and study the (weak) derivatives of λ(|UρnN|)UρnN. Let (ρn)Cc1(R3) be a sequence of spherically symmetric densities such that
As above λ(|UρnN|)UρnNLlocq(R3). Denote by
the mass of ρn inside the ball with radius r. Then MnC1(R3) with
Let rn ≥ 0 be such that Mn (rn) = 0 and Mn(r) > 0 for all r > rn. Then
with
(14)
if |x| < rn, and
(15)
if |x| > rn and i, j = 1, 2, 3. Denote by
the functions that are pointwise almost everywhere on R3 defined by Eqs. (14) and (15); the only region where the xi[] are not defined by Eqs. (14) and (15) is {|x| = rn} where they might have a discontinuity. Using Eqs. (12) and (13) we get for |x| > rn
Since p > q/(2 − q) + q and
we can apply Proposition V.1 and get for every R > 0
(16)
Now we prove that the functions given by Eqs. (14) and (15) are indeed the weak derivatives of λ(|UρnN|)UρnN. For every ϕCc(R3)
If rn = 0, we use
and get
If rn > 0, we use
and get, too, that the boundary term in the above integration by parts vanishes. Hence, the Eqs. (14) and (15) pointwise almost everywhere defined functions are indeed the weak derivatives of
It remains to prove that
and that the estimate Eq. (16) holds with ρn replaced by ρ. Using Eq. (11) and Hölder we have for R > 0
Thus
Since
independent of nN, Eq. (16) implies that there is a subsequence [again denoted by (ρn)] such that
V is the weak derivative of
with respect to xi and hence
Since the Lq-norm is weakly lower semi-continuous, Eq. (16) implies
Thus, in spherical symmetry it follows from Lemma V.3 (and Lemma II.4) that the Mondian potential UρM is always twice weakly differentiable. But, as we have argued in the introduction to this section, one might expect from a naïve argumentation that
for 1 < q < 6 if
However, in Lemma V.3 we were only able to prove an estimate of the type
if 1 < q < 2. In the following Lemma we show that this estimate is indeed optimal; there is no such estimate if q > 2. The subsequent Lemma will further show that it is unlikely that any such estimate can be proven if we drop the assumption of spherical symmetry.

Lemma V.4.
Let λ(σ)=1/σ, σ > 0. Then there is a sequence of spherically symmetric densities (ρn)L1L(R3) such that for all nNρn0, supp ρnB2 andρn ≤ 1, but
if 2 < q < 6.

Remark V.5.
The idea behind the Proof of Lemma V.4 is the following: In ΔUρM appears the term
where we have introduced the notion
Reference 18, Satz 7.7 implies that for a.e. yR3
So we could expect that
Assuming for the moment that ρ(0) > 0 and that ‖ρ < this would guarantee that M(r)/rq is bounded for all 1 < q < 6. Together with Proposition V.1 this would give us a bound for D2UρMq for all 1 < q < 6. However, the pointwise representation of an Lp-function ρ is tricky:
Let us take an open set Ωn ⊂ [0, 2] such that for all ϵ > 0
and set
Then there is no well defined value of ρ(0) and we get
with a constant C > 0 independent of n. Thus
and when we send n this is unbounded in Lq for 2 < q < 6.

Proof of Lemma V.4.
For nN set
and define
Denote by
the mass of ρn inside the ball with radius r ≥ 0. Let n,jN, then
Let r ∈ [2j, 2j+1) for a j ≥ 0. Then
Thus
and
(17)
Let now 2 < q < 6, then
Since ρn, Mn ≥ 0 and Mn is monotonic increasing
Now we use the estimate Eq. (17) and get
For 2 < q ≤ 4 we have
and for 4 < q < 6
Hence
and this is divergent if q > 2.□
So it is not possible for any q > 2 to prove an estimate of the form
even if ρ is spherically symmetric (and non-negative). Will the situation get even worse if we leave spherical symmetry?
Let us look at the difficulties that one can encounter. D2UρM causes difficulties when UρN(x)=0 for an xR3 because then
if |y| is small and λ(σ)=1/σ for σ > 0. Consider now the following, non-spherically symmetric case: For every nN place a point mass at position
Then for every nN there is 0 < αn < 1 such that for
we have
UN denotes the Newtonian gravitational potential created by all the masses at the points xn. So for every nND2UN(yn) will cause difficulties.
The exact treatment of such a non-spherically symmetric case is difficult. Can we perhaps mimic the above difficulties in spherical symmetry? The answer is yes, if we do not demand that ρ has to be non-negative. Then the next lemma shows that it is no more possible for any 1 ≤ p, q to prove an estimate of the form

Lemma V.6.
Let λ(σ)=1/σ, σ > 0. Then there exists a ρL1L(R3), spherically symmetric, which takes positive and negative values, such that

Proof.
For nN set
and let mn be the center between an and an+1, i.e.,
Then a1 = 2 and
Set M(r) ≔ 0 if r ∈ [0, 2) or r ∈ [π2/3, ). If r ∈ [2, π2/3) set
Then M is continuous and
(18)
Set ρ(r) ≔ 0 if r ∈ [0, 2) or r ∈ [π2/3, ). If r ∈ [2, π2/3) set
Then ρL1L(R3). Further for r ≥ 0
and thus
In view of Eq. (18)
will have a zero for all x = (an, 0, 0), nN. Let us see how this troubles the second derivatives of the Mondian potential:
As in the introduction to this section we have
for r = |x| > 0. But
since
Hence
and

Since the density ρ constructed in Lemma V.6 mimics the difficulties that one can encounter in a situation without symmetry assumptions, we suspect that it is impossible to prove the existence of weak, integrable derivatives of UρM for general ρL1L(R3), 0. Thus the assumption of spherical symmetry in Lemma V.3 seems indeed to be necessary if one wants to prove that UρM is twice weakly differentiable.

We have conducted an extensive analysis of the QUMOND theory, focusing initially on the gradient UρM of the Mondian potential instead of directly studying the potential UρM. Our investigation reveals that this gradient is the irrotational part of the vector field (1) in the sense of the Helmholtz–Weyl decomposition. This assures that UρM is an Lp vector field and indeed the weak gradient of a potential. Our findings show that the corresponding potential is given by Eq. (3), which was introduced by Milgrom,9 and that it is well defined.

These results were attained through a careful examination of the operator H responsible for extracting the irrotational part of a vector field. We developed a new, explicit expression for this operator using singular integral operators. Using the operator H also significantly aided in demonstrating that the Mondian potential solves the PDE (2) in distribution sense. Thus by linking the QUMOND theory with the Helmholtz–Weyl decomposition, we established a robust mathematical foundation for QUMOND.

Furthermore, we investigated second-order derivatives of the Mondian potential UρM. Under the additional assumption of spherical symmetry, we proved that the Mondian potential is twice weakly differentiable. However, the regularity of the second derivatives was found to be weaker than anticipated. Additionally, we illustrated why proving a similarly general regularity result for the second derivatives without symmetry assumptions seems impossible.

Our findings can be applied to many problems in QUMOND. For instance, in an accompanying paper,19 we establish the stability of a large class of spherically symmetric models. The perturbations permitted are still confined to spherical symmetry and removing this restriction draws heavily upon the results presented in this paper, a discussion of which is provided in the accompanying work. Moreover, our results can be applied to analyze initial value problems. Recent work by Carina Keller in her master’s thesis demonstrates the existence of global weak solutions to the initial value problem for the collisionless Boltzmann equation. Her result is limited to spherically symmetric solutions. Generalizing it to solutions devoid of symmetry restrictions necessitates the use of the theory presented here and a further generalization of it: We have to use that the operator H also preserves Hölder continuity. This is work in progress.

Our research contributes to the investigation of solutions to the initial value problem for the collisionless Boltzmann equation in yet another way. Building upon the theory from DiPerna and Lions,20 we have established that weak Lagrangian solutions conserve energy. This unpublished result, not imposing any symmetry restrictions, heavily relies on the results proven in this paper. Further, the question of whether every Eulerian solution of the collisionless Boltzmann equation is also a Lagrangian one, and vice versa, is of considerable interest. DiPerna and Lions20 have shown that this equivalence holds if the Mondian potential has second-order weak derivatives. Thus, our findings confirm this equivalence for spherically symmetric solutions, but cast doubt on extending this conclusion to nonsymmetric scenarios.

In summary, with QUMOND now placed on a robust mathematical foundation, it is possible to analyze many interesting yet unsolved questions with mathematical rigor.

Funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Projektnummer RE 885/4-1.

The authors have no conflicts to disclose.

Joachim Frenkler: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Writing – original draft (equal); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

We omitted the Proof of Lemma V.2 and we give it now in the Appendix.

Proof of Lemma V.2.
Let σ > 0, then
as desired. Further, the function λ(|u|)u is continuously differentiable on R3\{0}, and for uR3, u ≠ 0, holds
where E3 denotes the identity matrix of dimension 3. Using the bounds for λ and λ′, we have
Let now u,vR3 be such that for all t ∈ [0, 1]
is different from zero. Then
Set
then |b| = 1 and we have
Thus for a.e. u,vR3\{0}
By continuity this holds for all u, v ≠ 0 and due to the Hölder continuity this holds for all u,vR3.□

1.
M.
Milgrom
, “
A modification of the Newtonian dynamics as a possible alternative to the hidden mass hypothesis
,”
Astrophys. J.
270
,
365
370
(
1983
).
2.
S.
McGaugh
, “
Predictions and outcomes for the dynamics of rotating galaxies
,”
Galaxies
8
,
35
(
2020
); arXiv:2004.14402 [astro-ph.GA].
3.
Gaia Collaboration
, “
The Gaia mission
,”
Astron. Astrophys.
595
,
A1
(
2016
); arXiv:1609.04153 [astro-ph.IM].
4.
Gaia Collaboration
, “
Gaia Data Release 3: Summary of the content and survey properties
,”
Astron. Astrophys.
674
,
A1
(
2023
); arXiv:2208.00211 [astro-ph.GA].
5.
K.-H.
Chae
, “
Breakdown of the Newton-Einstein standard gravity at low acceleration in internal dynamics of wide binary stars
,”
Astrophys. J.
952
,
128
(
2023
); arXiv:2305.04613 [astro-ph.GA].
6.
X.
Hernandez
, “
Internal kinematics of GAIA DR3 wide binaries: Anomalous behaviour in the low acceleration regime
,”
Mon. Not. R. Astron. Soc.
525
,
1401
1415
(
2023
); arXiv:2304.07322 [astro-ph.GA].
7.
I.
Banik
,
C.
Pittordis
,
W.
Sutherland
,
B.
Famaey
,
R.
Ibata
,
S.
Mieske
, and
H.
Zhao
, “
Strong constraints on the gravitational law from Gaia DR3 wide binaries
,”
Mon. Not. R. Astron. Soc.
527
,
4573
4615
(
2023
); arXiv:2311.03436 [astro-ph.SR].
8.
B.
Famaey
and
S. S.
McGaugh
, “
Modified Newtonian dynamics (MOND): Observational phenomenology and relativistic extensions
,”
Living Rev. Relativ.
15
,
10
(
2012
); arXiv:1112.3960 [astro-ph.CO].
9.
M.
Milgrom
, “
Quasi-linear formulation of MOND
,”
Mon. Not. R. Astron. Soc.
403
,
886
895
(
2010
); arXiv:0911.5464 [astro-ph.CO].
10.
G. P.
Galdi
, in
An Introduction to the Mathematical Theory of the Navier-Stokes Equations
, 2nd ed.,
Springer Monographs in Mathematics
(
Springer
,
New York, NY
,
2011
).
11.

As usual, we say that some measurable function f is an Lp function, i.e., fLp(R3) for some 1 < p < ∞, if |f(x)|pdx < ∞. We say that a vector field.vLp(R3) if all three components are Lp functions. Further, we say that some measurable function f is locally p-integrable, i.e., fLlocp(R3), if fLp(K) for every KR3 compact.

12.
E. M.
Stein
,
Singular Integrals and Differentiability Properties of Functions
(
Princton University Press
,
Princton, NJ
,
1970
).
13.
S.
Dietz
, “
Flache Lösungen des Vlasov-Poisson-Systems
,” Ph.D. thesis,
University of Munich
,
2001
.
14.
G.
Rein
, “
Collisionless kinetic equations from astrophysics – The Vlasov–Poisson system
,” in
Handbook of Differential Equations: Evolutionary Equations
(
Elsevier, Amsterdam
,
2007
), pp.
383
476
.
15.
D.
Gilbarg
and
N. S.
Trudinger
,
Elliptic Partial Differential Equations of Second Order
, 2nd ed. (
Springer
,
Berlin
,
2001
).
16.
E.
Lieb
and
M.
Loss
, in
Analysis
, 2nd ed. (
American Mathematical Society
,
Providence, RI
,
2010
).
17.
I.
Newton
,
Philosophiae Naturalis Principia Mathematica
(
Jussu Societatis Regiae ac Typis Josephi Streater, London
,
1687
).
18.
W.
Rudin
,
Reelle und Komplexe Analysis
(
Oldenburg, München
,
1999
).
19.
J.
Frenkler
, “
Stability of spherical models in MOND
,”
Kinet. Relat. Models
(published onine
2024
).
20.
R.
DiPerna
and
P.
Lions
, “
Ordinary differential equations, transport theory and Sobolev spaces
,”
Invent. Math.
98
,
511
547
(
1989
).