It is well-known that the embedding of the Sobolev space of weakly differentiable functions into Hölder spaces holds if the integrability exponent is higher than the space dimension. In this paper, the embedding of the Sobolev functions into the Hölder spaces is expressed in terms of the minimal weak differentiability requirement independent of the integrability exponent. The proof is based on the generalization of the Newton-Leibniz formula to the n-dimensional rectangle and the inductive application of the Sobolev trace embedding results. The new method is applied to prove the embedding of the Sobolev spaces with dominating mixed smoothness into Hölder spaces. Counterexamples demonstrate that in terms of minimal weak regularity degree the Sobolev spaces with dominating mixed smoothness present the largest class of weakly differentiable functions with the upgrade of pointwise regularity to continuity. Remarkably, it also presents the largest class of weakly differentiable functions where the generalized Newton-Leibniz formula holds.
Citation: Ugur G. Abdulla. Generalized Newton-Leibniz formula and the embedding of the Sobolev functions with dominating mixed smoothness into Hölder spaces[J]. AIMS Mathematics, 2023, 8(9): 20700-20717. doi: 10.3934/math.20231055
[1] | Wen Teng, Jiulin Jin, Yu Zhang . Cohomology of nonabelian embedding tensors on Hom-Lie algebras. AIMS Mathematics, 2023, 8(9): 21176-21190. doi: 10.3934/math.20231079 |
[2] | Haikun Liu, Yongqiang Fu . Embedding theorems for variable exponent fractional Sobolev spaces and an application. AIMS Mathematics, 2021, 6(9): 9835-9858. doi: 10.3934/math.2021571 |
[3] | Boonyachat Meesuptong, Peerapongpat Singkibud, Pantiwa Srisilp, Kanit Mukdasai . New delay-range-dependent exponential stability criterion and $ H_\infty $ performance for neutral-type nonlinear system with mixed time-varying delays. AIMS Mathematics, 2023, 8(1): 691-712. doi: 10.3934/math.2023033 |
[4] | Zhaoyue Sui, Feng Zhou . Pointwise potential estimates for solutions to a class of nonlinear elliptic equations with measure data. AIMS Mathematics, 2025, 10(4): 8066-8094. doi: 10.3934/math.2025370 |
[5] | Xhevat Z. Krasniqi . Approximation of functions in a certain Banach space by some generalized singular integrals. AIMS Mathematics, 2024, 9(2): 3386-3398. doi: 10.3934/math.2024166 |
[6] | Jinguo Zhang, Dengyun Yang, Yadong Wu . Existence results for a Kirchhoff-type equation involving fractional $ p(x) $-Laplacian. AIMS Mathematics, 2021, 6(8): 8390-8403. doi: 10.3934/math.2021486 |
[7] | Ali Akgül, Esra Karatas Akgül, Sahin Korhan . New reproducing kernel functions in the reproducing kernel Sobolev spaces. AIMS Mathematics, 2020, 5(1): 482-496. doi: 10.3934/math.2020032 |
[8] | Messaoud Bounkhel . $ V $-Moreau envelope of nonconvex functions on smooth Banach spaces. AIMS Mathematics, 2024, 9(10): 28589-28610. doi: 10.3934/math.20241387 |
[9] | Kelei Zhang . Orlicz estimates for parabolic Schrödinger operators with non-negative potentials on nilpotent Lie groups. AIMS Mathematics, 2023, 8(8): 18631-18648. doi: 10.3934/math.2023949 |
[10] | Wei Fan, Kangqun Zhang . Local well-posedness results for the nonlinear fractional diffusion equation involving a Erdélyi-Kober operator. AIMS Mathematics, 2024, 9(9): 25494-25512. doi: 10.3934/math.20241245 |
It is well-known that the embedding of the Sobolev space of weakly differentiable functions into Hölder spaces holds if the integrability exponent is higher than the space dimension. In this paper, the embedding of the Sobolev functions into the Hölder spaces is expressed in terms of the minimal weak differentiability requirement independent of the integrability exponent. The proof is based on the generalization of the Newton-Leibniz formula to the n-dimensional rectangle and the inductive application of the Sobolev trace embedding results. The new method is applied to prove the embedding of the Sobolev spaces with dominating mixed smoothness into Hölder spaces. Counterexamples demonstrate that in terms of minimal weak regularity degree the Sobolev spaces with dominating mixed smoothness present the largest class of weakly differentiable functions with the upgrade of pointwise regularity to continuity. Remarkably, it also presents the largest class of weakly differentiable functions where the generalized Newton-Leibniz formula holds.
Let W1p(Rn),1≤p≤∞ be a Sobolev space of weakly differentiable functions u∈Lp(Rn) with first order weak derivatives in Lp(Rn),i=1,...,n. Originally discovered in the celebrated paper [1], the concept of Sobolev spaces became a trailblazing idea in many fields of mathematics. The goal of this paper is to analyze embedding of W1p(Rn) into Hölder spaces C0,μ(Rn),0≤μ≤1 [2]. A standard notation will be employed for the embedding of Banach spaces:
● B1↪B2 means bounded embedding of B1 into B2, i.e., B1⊂B2, and
‖u‖B2≤C‖u‖B1, ∀u∈B1, for some constant C. |
● B1⋐B2 denotes compact embedding of B1 into B2, meaning that B1↪B2, and every bounded subset of B1 is precompact in B2.
If n=1, the equivalency class of elements of W1p(R) always contain an absolutely continuous element, which is Hölder continuous with exponent 1−p−1, if p>1, i.e., there is a bounded embedding
W1p(R)↪C0,1−1p(R), if p>1; W11(R)↪C0(R). | (1.1) |
The embedding (1.1) easily follows from the Newton-Leibniz formula
u(x′)−u(x)=∫x′xdu(y)dxdy | (1.2) |
via the application of the Hölder inequality and compactness argument. The embedding (1.1) fails to be true if n≥2 and p≤n. However, there is a bounded embedding [3]
W1p(Rn)↪C0,1−np(Rn) if p>n. | (1.3) |
Hence, stretching the integrability exponent p beyond space dimension n implies the Hölder continuity. In particular, elements of the Hilbert space H1(Rn)=W12(Rn), are not continuous in general, if n≥2. The main goal of this paper is to express the continuity of elements of W1p(Rn) in terms of weak differentiability requirements.
Problem 1.1. What are the minimal weak differentiability requirements on elements of W1p(Rn) (1≤p≤n) to be continuous? In terms of weak differentiability requirements, what is the largest subspace of W1p(Rn) embedded into Hölder spaces for all p≥1?
The paper reveals that the anticipated subspace is the Sobolev-Nikol'skii space
S1p(Rn)={u∈W1p(Rn)| ∂ku∂xi1⋯∂xik∈Lp(Rn),i1<⋯<ik,k=¯2,n}, |
equipped with the norm
‖u‖S1p(Rn):={(‖u‖pLp(Rn)+n∑k=1n∑i1,...,ik=1i1<...<ik‖∂ku∂xi1⋯∂xik‖pLp(Rn))1p,if 1≤p<∞,‖u‖L∞(Rn)+n∑k=1n∑i1,...,ik=1i1<...<ik‖∂ku∂xi1⋯∂xik‖L∞(Rn),if p=∞. |
The space S1p(Rn) is a special case of Sobolev spaces with dominating mixed smoothness. The class was introduced by Nikol'skii in [4,5]. There is a vast literature on the analysis of these spaces. We refer to [6,7,8,9] and the references therein.
The main result of this paper is twofold. First, we introduce and prove a generalization of the celebrated Newton-Leibniz formula to n-dimensional rectangles (or n-rectangles). Then by using the new formula as a tool, we present a surprisingly simple and elegant proof of the embedding of the Sobolev spaces with dominating mixed smoothness into Hölder spaces. The proof resembles the proof of the embedding (1.1) in the one-dimensional case by using generalized Newton-Leibniz formula, Hölder inequality, and iterative application of the Sobolev trace embedding results. In particular, we prove that the generalized Newton-Leibniz formula is preserved in space S1p(Rn). Counterexamples support the claim that in terms of weak differentiability requirements, S1p is the largest class of Lebesgue's integrable and weakly differentiable functions in Rn which preserve generalized Newton-Leibniz formula, and upgrades the pointwise regularity to Hölder continuity.
● C0(Rn) is a Banach space of continuous and bounded functions with the norm
‖u‖C0(Rn):=supx∈Rn|u(x)|=‖u‖L∞(Rn). |
● For k∈N, Ck(Rn) is a Banach space of k times continuously differentiable functions, with all derivatives of order k bounded, and with the norm
‖u‖Ck(Rn):=k∑j=0supx∈Rn|Dju(x)|=k∑j=0‖Dju‖L∞(Rn), |
where Dju is a tensor of rank j, dimension n, and
|Dju|=(n∑i1,...,ij=1|∂ju(x)∂xi1⋯∂xij|2)12. |
● SC1(Rn) is a Banach space with the norm
‖u‖SC1(Rn):=∑α∈Zn+,αi≤1‖ ∂|α|u(x)∂xα11⋯∂xαnn‖C0(Rn). |
The following standard notation will be used for Hölder spaces:
● For 0≤γ≤1, Hölder space C0,γ(Rn) is the Banach space of elements u∈C0(Rn) with finite norm
‖u‖C0,γ(Rn):=‖u‖C0(Rn)+[u]C0,γ(Rn), |
where
[v]C0,γ(Rn):=supx,x′∈Rnx≠x′|v(x)−v(x′)||x−x′|γ. |
The space C0,0(Rn) is equivalent to C0(Rn).
● For k∈N, 0≤γ≤1, Hölder space Ck,γ(Rn) is a subspace of Ck(Rn) with finite norm
‖u‖Ck,γ(Rn):=k∑j=0‖Dju‖C0,γ(Rn) |
Throughout the paper we use standard notations for Lp(Q),1≤p≤∞ spaces; the following standard notations are used for Sobolev spaces [2]:
● For k∈N, 1≤p≤∞, Sobolev space Wkp(Rn) is the Banach space of measurable functions on Rn with finite norm
‖u‖Wkp(Rn):=k∑j=0‖Dju‖Lp(Rn). |
● For s=(s1,...,sn)∈Zn+,1≤p≤∞, anisotropic Sobolev space Wsp(Rn) is the Banach space of measurable functions on Rn with finite norm
‖u‖Wsp(Rn):=‖u‖Lp(Rn)+n∑i=1si∑k=1‖∂ku∂xki‖Lp(Rn). |
Note that the size of the vector s coincides with the dimension of the space. In particular, for 1≤k≤n, and fixed j∈{1,...,k}, we consider Sobolev spaces Wsp(Rk) of the weakly xj-differentiable functions on Rk, where s=(si)ki=1∈Zk+ and si=δij is a Kronecker symbol.
● For k=(k1,...,kn)∈Zn+,1≤p≤∞, Sobolev space Skp(Rn) with dominating mixed derivatives is a Banach space of measurable functions on Rn with the finite norm
‖u‖Skp(Rn):=∑α∈Zn+,αi≤ki‖ ∂|α|u(x)∂xα11⋯∂xαnn‖Lp(Rn). |
If k1=⋯=kn=k∈N, we shall write Skp(Rn)=Skp(Rn).
● Let Q⊂Rn be a bounded domain. For k=(k1,...,kn)∈Zn+,1≤p≤∞, Sobolev space Skp(Q) with dominating mixed derivatives is defined as
Skp(Q)={f∈D′(Q): ∃g∈Skp(Rn) with g|Q=f} |
and with
‖f‖Skp(Q):=inf‖g‖Skp(Rn), |
where the infimum is taken over all g∈Skp(Rn) such that its restriction g|Q to Q coincides with f in the space of distributions D′(Q). If k1=⋯=kn=k∈N, we shall write Skp(Q)=Skp(Q).
Let x,x′∈Rn with xi<x′i,i=¯1,n are fixed and P be n-rectangle
P={η∈Rn:xi≤ηi≤x′i,i=¯1,n} | (3.1) |
with vertex x (or x′) called a bottom (or top) corner of P. For any subset {i1,...,ik}⊂{1,...,n},k=¯1,n, let
Pi1…ik=P∩{η∈Rn:ηl=xl,l≠ij,j=¯1,k} |
be a k-rectangle with bottom corner x. Note that Pi1…ik is invariant with respect to permutation of multi-index i1⋯ik, and it coincides with P if k=n.
The following is the generalization of the celebrated Newton-Leibniz formula:
Theorem 3.1. Any function u∈SC1(Rn) satisfies the following generalized Newton-Leibniz formula:
u(x′)−u(x)=n∑k=1n∑i1,...,ik=1i1<...<ik ∫Pi1…ik∂ku(η)∂xi1⋯∂xikdηi1⋯dηik. | (3.2) |
If n=1, (3.2) coincides with the Newton-Leibniz formula (1.2). Note that for ∀k there are (nk) integrals in (3.2) along all k-rectangles Pi1…ik with bottom corner x. Therefore, altogether there are
n∑k=1(nk)=n∑k=0(nk)−1=(1+1)n−1=2n−1, |
integrals in (3.2) along all sub-rectangles of P with bottom corner at x.
Having generalized the Newton-Leibniz formula in the class of smooth functions, we can now formulate the major problem of classical analysis generated by the Newton-Leibniz formula:
Problem 3.1. What is the largest class of Lebesgue integrable and weakly differentiable functions in Rn which preserve generalized Newton-Leibniz formula?
In Theorem 3.2 we prove that the formula (3.2) remains valid in spaces S1p(Rn),1≤p≤∞, where the right-hand side is understood as trace integrals of respective weak derivatives. Counterexamples constructed in Section 3.3 support the claim that in terms of weak differentiability requirements S1p(Rn) is the largest class of Lebesgue's integrable and weakly differentiable functions in Rn which preserve generalized Newton-Leibniz formula. The formula (3.2) is a key to prove the Hölder continuity of elements of S1p(Rn).
Remark 3.1. Some variant of the formula (3.2) is proved in [10] in the class of tensor product space, i.e., linear cover of the space of separable (or product form) functions equipped with special norm consisting of some algebraic combination of one-dimensional W1,p norms selected in a way to guarantee the Lp-boundedness of mixed derivatives of product functions.
Theorem 3.2. The following bounded embedding holds
S1p(Rn)↪C0,1−1p(Rn); 1≤p≤∞. | (3.3) |
The equivalency class of every element of S1p(Rn) possesses a representative in C0,1−1p(Rn), which satisfies the generalized Newton-Leibniz formula (3.2), where P⊂Rn is an n-rectangle with bottom and top corner at x and x′ respectively. In particular, ∀k=1,...,n−1 and 1≤i1<⋯<ik≤n
∂ku∂xi1⋯∂xik∈Lp(Pi1…ik), | (3.4) |
in the sense of traces.
Corollary 3.1. For k∈N the following bounded embedding holds
Skp(Rn)↪Ck−1,1−1p(Rn); 1≤p≤∞. | (3.5) |
The following sharp embedding result holds for the anisotropic Sobolev spaces with dominating mixed smoothness:
Corollary 3.2. Let k=(k1,...,kn)∈Nn,1≤p≤∞, and u∈Skp(Rn). Then ∀m=1,...,n and ∀1≤i1<i2<⋯<im≤n
∂ki1+⋯+kim−mu∂xki1−1i1⋯∂xkim−1im∈C0,1−1p(Rn). | (3.6) |
Corollary 3.3. Let Q⊂Rn be a bounded domain. For k∈N the following bounded and compact embeddings hold
Skp(Q)↪Ck−1,1−1p(¯Q), if 1≤p≤∞; | (3.7) |
Skp(Q)⋐Ck−1,μ(¯Q),0<μ<1−1p, if 1<p≤∞. | (3.8) |
The goal of this section is to provide counterexamples to support the claim that S1p(Rn) is a subspace of W1p(Rn),1≤p≤n with minimal increase of the weak differentiability requirements in order that every equivalency class has an element which is
● Hölder continuous;
● satisfy generalized Newton-Leibniz formula (3.2);
● satisfy the trace regularity (3.4);
Example 3.1. Consider a function
u(x)=loglog(1+1|x|)∈W1n(B(0,1)) | (3.9) |
where B(0,1)⊂Rn is a unit ball with center 0. It is discontinuous at 0. Direct calculation demonstrates that for arbitrary k∈{1,...,n−1} we have
∂ku∂xi1⋯∂xik∈Lp(B(0,1)), 1≤p≤nk, | (3.10) |
for all 1≤i1≤⋯≤ik≤n. However, we have
∂nu∂x1⋯∂xn∉L1(B(0,1)). | (3.11) |
Since for all k∈{1,...,n−1} all k-th order weak derivatives are in L1(B(0,1)), by using standard extension theorem [2] function u can be extended to Rn by possessing the regularity
∂ku∂xi1⋯∂xik∈Lp(Rn), 1≤p≤nk, | (3.12) |
for all 1≤i1≤⋯≤ik≤n. From (3.11) it follows that the n-th order mixed derivative of the extended function is not integrable, i.e.,
∂nu∂x1⋯∂xn∉L1(Rn). | (3.13) |
It can also be verified that if P is an n-rectangle with bottom corner at 0, then ∀k=1,...,n−1 and 1≤i1<⋯<ik≤n, we have
∂ku∂xi1⋯∂xik∉L1(Pi1…ik). | (3.14) |
Clearly, Newton-Leibniz formula (3.2) is not satisfied at the origin. Hence, the extended function presents the desired counterexample when the n-th order mixed derivative is removed from the definition of the space S1p(Rn).
Example 3.2. Consider a function
u(x)=|x|−nn∏k=1xk∈L∞(Rn) | (3.15) |
It is discontinuous at 0. Along the each hyperplane {xk=0} it is zero, but for ∀C>0
limxk=Ct,t↓0u=1. | (3.16) |
Direct calculation demonstrates that for arbitrary k∈{1,...,n−1} we have
∂ku∂xi1⋯∂xik∈Lp(B(0,1)), 1≤p<nk, | (3.17) |
for all 1≤i1≤⋯≤ik≤n. However, we have
∂nu∂x1⋯∂xn∉L1(B(0,1)). | (3.18) |
Since for all k∈{1,...,n−1} all k-th order weak derivatives are in L1(B(0,1)), by using standard extension theorem [2] function u can be extended to Rn by possessing the regularity
∂ku∂xi1⋯∂xik∈Lp(Rn), 1≤p<nk, | (3.19) |
for all 1≤i1≤⋯≤ik≤n. From (3.18) it follows that the n-th order mixed derivative of the extended function is not integrable, i.e.
∂nu∂x1⋯∂xn∉L1(Rn). | (3.20) |
It can also be verified that if P is an n-rectangle with bottom corner at 0, then ∀k=1,...,n−1 and 1≤i1<⋯<ik≤n, we have
∂ku∂xi1⋯∂xik≡0∈L1(Pi1…ik). | (3.21) |
In particular, this implies that the Newton-Leibniz formula is satisfied in any k-rectangle Pi1…ik with k≤n−1 if we assign u(0)=0. However, Newton-Leibniz formula is not satisfied in an n-rectangle P due to the fact that
∂nu∂x1⋯∂xn∉L1(P). | (3.22) |
Example 3.3. Consider a function
u(x)=k∏s=1sign xs∈L∞(Rn), | (3.23) |
with k∈{1,...,n−1}. It is discontinuous along all the coordinates axis x1,...,xk. For arbitrary s∈{1,...,k} the weak derivatives
∂su∂x1⋯∂xs | (3.24) |
do not exist. All the other mixed derivatives which involve differentiation with respect to any other variables xl with l∈{k+1,...,n} exists and equal to zero. Clearly, the Newton-Leibniz formula is not satisfied in any n-rectangle whose interior intersects any of the coordinate axis x1,...,xk.
Proof of Theorem 3.1. Assuming that u∈SC1(Rn), we prove (3.2) by induction in terms of the space dimension n. If n=1, it coincides with the Newton-Leibniz formula. Assume that (3.2) is true, and demonstrate that it is true if n is replaced with n+1. Let x,x′∈Rn+1 with xi<x′i,i=¯1,n+1, are fixed. We have
u(x′)−u(x)=(u(x′)−u(˜x,x′n+1))+(u(˜x,x′n+1)−u(x)), | (4.1) |
where ˜x=(x1,...,xn). Applying (3.2) to the first term and the Newton-Leibniz formula to the second term in (4.1), we derive
u(x′)−u(x)=n∑k=1n∑i1,...,ik=1i1<...<ik ∫Pi1…ik∂ku(˜η,x′n+1)∂xi1⋯∂xikdηi1⋯dηik+x′n+1∫xn+1∂u(˜x,η)∂xn+1dη. | (4.2) |
Applying the Newton-Leibniz formula to all but the last integrand, we have
u(x′)−u(x)=n∑k=1n∑i1,...,ik=1i1<...<ik ∫Pi1…ikx′n+1∫xn+1∂k+1u(η)∂xi1⋯∂xik∂xn+1dηi1⋯dηikdηn+1+x′n+1∫xn+1∂u(˜x,η)∂xn+1dη+n∑k=1n∑i1,...,ik=1i1<...<ik ∫Pi1…ik∂ku(η)∂xi1⋯∂xikdηi1⋯dηik, | (4.3) |
which imply that
u(x′)−u(x)=n+1∑k=1n+1∑i1,...,ik=1i1<...<ik ∫Pi1…ik∂ku(η)∂xi1⋯∂xikdηi1⋯dηik, | (4.4) |
where we use the same notation for the (n+1)-rectangle P, as well as its corresponding sub-rectangles in Rn+1. Indeed, divide all 2n+1−1 sub-rectangles of P with the bottom corner at x into two groups depending on whether or not the edge pn+1 joining vertices x and (˜x,x′n+1) is contained in it. The first two terms on the right-hand side of (4.3) consist of all 2n terms of (4.4) with integrals along sub-rectangles containing the edge pn+1, and the last term on the right-hand side of (4.3) is identical with the remaining 2n−1 integrals in (4.4) along sub-rectangles which do not contain the edge pn+1. This completes the proof by induction.
Proof of Theorem 3.2. First, we prove the Theorem assuming that 1≤p<∞. The proof will be pursued in four steps.
Step 1. Prove that for u∈S1p(Rn), each of the 2n−1 integrals on the right-hand side of (3.2) is finite, and in particular, (3.4) is satisfied. Existence of the integral with k=n on the right hand side of (3.2) follows from the definition of S1p(Rn) and Hölder inequality. We prove the existence of the remaining 2n−2 trace integrals in (3.2) by mathematical induction and Sobolev trace embedding result. First, we demonstrate that the claim is true if k=n−1. Then we show that the claim is true for any k<n−1, provided it is true for k+1. Indeed, if k=n−1, for each of the n integrals we select a unique integer j satisfying
j∈{1,...,n}∩{i1,...,ik}c, | (4.5) |
and define a multi-index s=(s1,...,sn)∈Zn+, where si=δij is a Kronecker symbol. We have
∂n−1u∂xi1⋯∂xin−1∈Wsp(P). | (4.6) |
Note that (n−1)-rectangle Pi1⋯in−1 is a boundary of P on the hyperplane ηj=xj. Existence of the trace
∂n−1u∂xi1⋯∂xin−1∈Lp(Pi1⋯in−1) | (4.7) |
is a consequence of the Sobolev trace embedding result:
Wsp(P)↪Lp(Pi1⋯in−1). | (4.8) |
For completeness, we present a proof of (4.8). Consider a function
ζ(η)=1−ηj−xjx′j−xj, | (4.9) |
which satisfy
0≤ζ≤1, |∂ζ∂ηj|≤1x′j−xj, η∈P. | (4.10) |
Assuming that u∈SC1(P), we have
∫Pi1…in−1|∂n−1u(η)∂xi1⋯∂xin−1|pdηi1⋯dηin−1=∫Pi1…in−1ζ|∂n−1u(η)∂xi1⋯∂xin−1|pdηi1⋯dηin−1=−∫Pi1…in−1x′j∫xj∂∂xj(ζ|∂n−1u(η)∂xi1⋯∂xin−1|p)dηjdηi1⋯dηin−1=−∫Pi1…in−1x′j∫xj[∂ζ∂xj|∂n−1u(η)∂xi1⋯∂xin−1|p+ζp|∂n−1u(η)∂xi1⋯∂xin−1|p−1×sign(∂n−1u(η)∂xi1⋯∂xin−1)∂nu(η)∂xi1⋯∂xik∂xj]dηjdηi1⋯dηin−1 | (4.11) |
If p>1, by using Young's inequality and (4.10), from (4.11) it follows
‖∂n−1u(η)∂xi1⋯∂xin−1‖Lp(Pi1⋯in−1)≤C‖∂n−1u(η)∂xi1⋯∂xin−1‖Wsp(P), | (4.12) |
where C=max(p−1+|x′j−xj|−1;1). If p=1, (4.12) follows directly from (4.10) and (4.11). In general, we can approximate u∈S1p(Rn) with the sequence uϵ=u∗ϕϵ∈C∞loc(Rn), where ϕϵ is a standard rescaled mollifier, and derive (4.12) for uϵ. Since uϵ converges to u in the norm given on the right-hand side of (4.12), it is so in the norm of the left-hand side as well, and passing to the limit as ϵ→0, (4.12), (4.8) and (4.7) follow. Hence, n relations of (3.4) with k=n−1 are established. Next, we prove that the claim is true for k if it is so for k+1. For any of the (nk) integrals in (3.2) along the k-dimensional prism Pi1⋯ik we select any integer j satisfying (4.5), and define a multiindex s=(s1,...,sk+1)∈Zk+1+, where si=δij is a Kronecker symbol. Noting that Pi1⋯ik+1 is invariant with respect to permutations of the multi-index i1⋯ik+1, and due to the induction assumption we have
∂ku∂xi1⋯∂xik∈Wsp(Pi1⋯ikj). | (4.13) |
k-rectangle Pi1⋯ik is a boundary of (k+1)-rectangle Pi1⋯ikj on the hyperplane xj=const. Sobolev trace embedding result implies:
Wsp(Pi1⋯ikj)↪Lp(Pi1⋯ik), | (4.14) |
The proof of (4.14) is identical to the proof of (4.8). Hence, (3.4) is proved for all k-dimensional integrals.
Step 2. In this step we prove that any u∈S1p(Rn)∩SC1loc(Rn),p>1 satisfies the estimate
|u(x)−u(x′)|≤[((1+p)1p+|x−x′|p−1p)n−(1+p)np]‖u‖S1p(Rn), | (4.15) |
for all x,x′∈Rn. Similarly, any u∈S11(Rn)∩SC1loc(Rn) satisfy the estimate
|u(x)−u(x′)|≤(3n−2n)‖u‖S11(Rn), | (4.16) |
for all x,x′∈Rn. Note that the estimate (4.16) is a formal limit of the estimate (4.15) as p→1.
To prove (4.15) (or (4.16)) without loss of generality we can assume that xi<x′i,i=¯1,n. Indeed, if xi≠x′i,i=¯1,n, then we can transform the space via finitely many translations
˜y:Rn→Rn, ˜yi={yi,if xi<x′i,−yi,if xi>x′i, | (4.17) |
and note that the space S1p(Rn) is invariant under this transformation. Then we can apply (4.15) (or (4.16)) to the ϵ-mollification of the transformed function ˜u(˜x)=u(˜x), and passing to limit as ϵ→0 deduce (4.15) (or (4.16)) for ˜u. Applying inverse transformation of (4.17) implies (4.15) (or (4.16)) for u. If, on the other side xi=x′i for some i, we can replace x′i with x′i+δ, prove (4.15) (or (4.16)) and pass to limit as δ→0.
The proof of (4.15) and (4.16) under the assumption that xi<x′i,i=¯1,n is based on the generalized Newton-Leibniz formula (3.2). The following is the proof of the estimate (4.15). Let P be a n-rectangle (3.1), and
P1:={η∈Rn:xi≤ηi≤x′i+1, i=¯1,n}. |
By using Hölder inequality the integral on the right-hand side of (3.2) with k=n is estimated as follows
|∫P∂nu(η)∂x1⋯∂xndη|≤|P|p−1p‖∂nu∂x1⋯∂xn‖Lp(P), | (4.18) |
where |P| denotes volume of the n-rectangle P. For k=1,...,n−1, estimation of any of the k-dimensional integrals on the right-hand side of (3.2) will be pursued in n−k steps. Consider typical k-dimensional integral in (3.2) along the k-rectangle Pi1⋯ik. The idea is based on successive application of the trace embedding result (4.14) n−k times. First, we select any integer j from (4.5) and assign it to the multi-index component ik+1. By using Hölder inequality we have
|∫Pi1⋯ik∂ku(η)∂xi1⋯∂xikdηi1⋯dηik|≤|Pi1⋯ik|p−1p‖∂ku∂xi1⋯∂xik‖Lp(Pi1⋯ik). | (4.19) |
Consider a function
ζ(η)=1−ηik+1−xik+1x′ik+1−xik+1+1, | (4.20) |
which satisfy
0≤ζ≤1, |∂ζ∂ηik+1|≤1. | (4.21) |
We have
∫Pi1…ik|∂ku(η)∂xi1⋯∂xik|pdηi1⋯dηik=∫Pi1…ikζ|∂ku(η)∂xi1⋯∂xik|pdηi1⋯dηik=−∫Pi1…ikx′ik+1+1∫xik+1∂∂xik+1(ζ|∂ku(η)∂xi1⋯∂xik|p)dηik+1dηi1⋯dηik=−∫Pi1…ikx′ik+1+1∫xik+1[∂ζ∂xik+1|∂ku(η)∂xi1⋯∂xik|p+ζp|∂ku(η)∂xi1⋯∂xik|p−1×sign(∂ku(η)∂xi1⋯∂xik)∂k+1u(η)∂xi1⋯∂xik∂xik+1]dηik+1dηi1⋯dηik | (4.22) |
By using Young's inequality and (4.21), from (4.22) it follows
∫Pi1⋯ik|∂ku(η)∂xi1⋯∂xik|pdηi1⋯dηik≤∫Pi1⋯ikx′ik+1+1∫xik+1[|∂k+1u(η)∂xi1⋯∂xik∂xik+1|p+p |∂ku(η)∂xi1⋯∂xik|p]dηik+1dηi1⋯dηik. | (4.23) |
From (4.19), (4.23) it follows that
|∫Pi1⋯ik∂ku(η)∂xi1⋯∂xikdηi1⋯dηik|≤|Pi1⋯ik|p−1p×(‖∂k+1u∂xi1⋯∂xik+1‖pLp(P1i1⋯ik)+p‖∂ku∂xi1⋯∂xik‖pLp(P1i1⋯ik))1p, | (4.24) |
where P1i1⋯ik=Pi1⋯ik×(xik+1,x′ik+1+1) is a (k+1)-rectangle. This completes one out of n−k steps for the estimation of the k-dimensional integral in (3.2) along the k-rectangle Pi1⋯ik. In the next step, we select any integer j from (4.5) with k replaced with k+1 and assign it to multi-index component ik+2. Then for each of the k+1-dimensional integrals on the right-hand side of (4.24) we derive the estimation similar to (4.23), where Pi1⋯ik is replaced with (k+1)-rectangle P1i1⋯ik, and integration interval (xik+1,x′ik+1+1) is replaced accordingly with (xik+2,x′ik+2+1). Application of these estimations to the right-hand side of (4.24) would complete the second out of n−k steps. By repeating the procedure after m=1,...,n−k steps we derive the following estimate:
|∫Pi1⋯ik∂ku(η)∂xi1⋯∂xikdηi1⋯dηik|≤|Pi1⋯ik|p−1p×[m∑j=0(mj)pj‖∂k+m−ju∂xi1⋯∂xik+m−j‖pLp(Pmi1⋯ik)]1p, | (4.25) |
where
Pmi1⋯ik=Pi1⋯ik×(xik+1,x′ik+1+1)×⋯×(xik+m,x′ik+m+1), |
be a (k+m)-rectangle. Let us prove the estimation (4.25) by induction. If m=1, the estimation (4.25) coincides with (4.24). Prove that (4.25) is true for m+1 if it is so for any m<n−k. Each of the k+m-dimensional integrals on the right-hand side of (4.25) satisfy the following estimate
∫Pmi1⋯ik|∂k+m−ju(η)∂xi1⋯∂xik+m−j|pdηi1⋯dηik+m≤∫Pmi1⋯ikx′ik+1+m+1∫xik+1+m[|∂k+1+m−ju(η)∂xi1⋯∂xik+m−j∂xik+1+m−j|p+p |∂k+m−ju(η)∂xi1⋯∂xik+m−j|p]dηik+1+mdηi1⋯dηik+m. | (4.26) |
Using (4.26), we have
m∑j=0(mj)pj‖∂k+m−ju∂xi1⋯∂xik+m−j‖pLp(Pmi1⋯ik)≤m∑j=0(mj)pj[‖∂k+1+m−ju∂xi1⋯∂xik+1+m−j‖pLp(Pm+1i1⋯ik)+p‖∂k+m−ju∂xi1⋯∂xik+m−j‖pLp(Pm+1i1⋯ik)]=m∑j=0(mj)pj‖∂k+1+m−ju∂xi1⋯∂xik+1+m−j‖pLp(Pm+1i1⋯ik)+m+1∑j=1(mj−1)pj‖∂k+1+m−ju∂xi1⋯∂xik+1+m−j‖pLp(Pm+1i1⋯ik). | (4.27) |
Since
(mj)+(mj−1)=(m+1j), j=1,...,m, |
from (4.27) it follows
m∑j=0(mj)pj‖∂k+m−ju∂xi1⋯∂xik+m−j‖pLp(Pmi1⋯ik)≤m+1∑j=0(m+1j)pj‖∂k+m+1−ju∂xi1⋯∂xik+1+m−j‖pLp(Pm+1i1⋯ik), | (4.28) |
which completes the proof of (4.25) by mathematical induction. By choosing m=n−k in (4.25), we derive an upper bound of the right-hand side by replacing the integration domain with Rn:
|∫Pi1⋯ik∂ku(η)∂xi1⋯∂xikdηi1⋯dηik|≤|Pi1⋯ik|p−1p[n−k∑j=0(n−kj)pj]1p‖u‖S1p(Rn)≤|x−x′|k(p−1)p(1+p)n−kp‖u‖S1p(Rn). | (4.29) |
Note that the estimation (4.29) holds for k = n as well in view of (4.18). By using (4.29) from the generalized Newton-Leibniz formula (3.2) it follows the estimate
|u(x′)−u(x)|≤n∑k=1n∑i1,...,ik=1i1<...<ik |x−x′|k(p−1)p(1+p)n−kp‖u‖S1p(Rn)=n∑k=1 (nk)|x−x′|k(p−1)p(1+p)n−kp‖u‖S1p(Rn)=[n∑k=0 (nk)|x−x′|k(p−1)p(1+p)n−kp−(1+p)np]‖u‖S1p(Rn)=[((1+p)1p+|x−x′|p−1p)n−(1+p)np]‖u‖S1p(Rn), | (4.30) |
which proves the desired estimate (4.15). The proof of the estimate (4.16) is almost identical to the proof of (4.15).
Step 3. In this step we prove
● the uniform C0,1−1p(Rn)-estimate for any u∈S1p(Rn)∩SC1loc(Rn),1<p<∞;
● the uniform C0(Rn)-estimate for any u∈S11(Rn)∩SC1loc(Rn).
Assume p>1 and fix x,x′∈Rn such that |x−x′|≤1. From (4.15) it follows that
|u(x)−u(x′)|≤[((1+p)1p+|x−x′|p−1p)n−(1+p)np]‖u‖S1p(Rn)=n∑k=1 (nk)|x−x′|k(p−1)p(1+p)n−kp‖u‖S1p(Rn)=n∑k=1 (nk)|x−x′|(k−1)(p−1)p(1+p)n−kp‖u‖S1p(Rn)|x−x′|pp−1≤n∑k=1 (nk)(1+p)n−kp‖u‖S1p(Rn)|x−x′|pp−1=[n∑k=0 (nk)(1+p)n−kp−(1+p)np]‖u‖S1p(Rn)|x−x′|pp−1=[(1+(1+p)1p)n−(1+p)np]‖u‖S1p(Rn)|x−x′|pp−1. | (4.31) |
Hence, we have
sup|x−x′|≤1x≠x′|u(x)−u(x′)||x−x′|pp−1≤[(1+(1+p)1p)n−(1+p)np]‖u‖S1p(Rn). | (4.32) |
Now fix x∈Rn. By using (4.32) and Hölder inequality we deduce
|u(x)|≤∮|y−x|≤1|u(x)−u(y)|dy+∮|y−x|≤1|u(y)|dy≤[(1+(1+p)1p)n−(1+p)np]‖u‖S1p(Rn)+Γ−1pn‖u‖Lp(Rn)≤[(1+(1+p)1p)n−(1+p)np+Γ−1pn]‖u‖S1p(Rn). | (4.33) |
where Γn is a volume of the unit ball in Rn. Hence, we have
‖u‖C0(Rn)≤[(1+(1+p)1p)n−(1+p)np+Γ−1pn]‖u‖S1p(Rn). | (4.34) |
From (4.34) it follows that
sup|x−x′|≥1|u(x)−u(x′)||x−x′|pp−1≤2‖u‖C0(Rn)≤2[(1+(1+p)1p)n−(1+p)np+Γ−1pn]‖u‖S1p(Rn). | (4.35) |
From (4.32) and (4.35) we deduce the following Hölder seminorm estimate for u:
[u]C0,1−1p(Rn)≤2[(1+(1+p)1p)n−(1+p)np+Γ−1pn]‖u‖S1p(Rn). | (4.36) |
Finally, (4.34), (4.36) imply the following Hölder norm estimate for u∈S1p(Rn)∩SC1loc(Rn),1<p<∞:
‖u‖C0,1−1p(Rn)≤3[(1+(1+p)1p)n−(1+p)np+Γ−1pn]‖u‖S1p(Rn). | (4.37) |
If p=1 from the estimate (4.16) with the similar argument as in (4.33) we derive the following C0(Rn)-estimate for any u∈S11(Rn)∩SC1loc(Rn):
‖u‖C0(Rn)≤[3n−2n+Γ−1n]‖u‖S11(Rn). | (4.38) |
Step 4. We complete the proof of the embedding (3.3) by using estimates (4.37), (4.38) and smooth approximation of elements of S1p(Rn). Given u∈S1p(Rn),1≤p<∞, we select a sequence vm∈C∞0(Rn) such that
‖vm−u‖S1p(Rn)→0, as m→∞. | (4.39) |
For example, the sequence vm can be given explicitly as in [11] (Lemma 23):
vm(x)=u1m(x)η(xm), |
where u1m=u∗ϕ1m∈C∞loc(Rn)∩S1p(Rn) is the 1m-mollification of u, ϕ1m is a standard rescaled mollifier, η∈C∞0(Rn) be a compactly supported function which equals 1 near the origin. If p>1, then by applying the estimate (4.37) to vm, we have
‖vm‖C0,1−1p(Rn)≤3[(1+(1+p)1p)n−(1+p)np+Γ−1pn]‖vm‖S1p(Rn). | (4.40) |
Equivalently, we have
‖vm−vl‖C0,1−1p(Rn)≤3[(1+(1+p)1p)n−(1+p)np+Γ−1pn]‖vm−vl‖S1p(Rn), | (4.41) |
for all m,l≥1, whence there exists a function u∗∈C0,1−1p(Rn) such that
‖vm−u∗‖C0,1−1p(Rn)→0, as m→∞. | (4.42) |
From (4.39) it follows that u∗=u, a.e. on Rn, so that u∗ is in the equivalency class of u. Passing to limit as m→∞, from (4.40) it also follows that
‖u∗‖C0,1−1p(Rn)≤3[(1+(1+p)1p)n−(1+p)np+Γ−1pn]‖u‖S1p(Rn). | (4.43) |
which proves the bounded embedding (3.3). Step 1 of the proof implies that the traces of u∗ satisfy (3.4), and each of them is an Lp-limit of the corresponding sequence of traces of vm. Therefore, writing (3.2) for vm, and passing to limit as m→∞, it follows that u∗ satisfies the generalized Newton-Leibniz formula (3.2).
Proof in the case p=1 is identical by using an estimate (4.38). This completes the proof of the theorem in the case 1≤p<∞.
Assume that p=∞. In this case, the embedding (3.3) is not new, and it is contained in the well-known fact that [2]
W1∞(Rn)↪C0,1(Rn). |
Since S1∞(Rn) is a subspace of W1∞(Rn), its elements are bounded and Lipschitz continuous functions and the embedding (3.3) holds. The assertion that u satisfies (3.2) follows from the proof given for the case p<∞. It only remains to show that (3.4) holds with p=∞. Note that from the given proof it follows that (3.4) holds for any p<∞. In particular for the smoothing sequence uϵ=u∗ϕϵ∈S1∞(Rn)∩C∞loc(Rn) all the traces indicated on the left hand side of (3.4) are uniformly bounded in L∞(Pi1⋯ik), and converge to corresponding traces of u in Lp(Pi1⋯ik) with any 1<p<∞. Such limits are also limits in the sense of distributions. Since L∞(Pi1⋯ik) is a dual space of L1(Pi1⋯ik), distributional limit of the sequence bonded in L∞(Pi1⋯ik) remains in L∞(Pi1⋯ik). Therefore, (3.4) holds with p=∞. Theorem is proved.
Corollaries 3.1, 3.2 and 3.3 are the direct consequence of the Theorem 3.2 due to the fact that if u∈Skp(Rn), then all the weak partial derivatives of order k−1 are elements of S1p(Rn), and if u∈Skp(Rn) the indicated partial derivative on the left hand side of (3.6) is an element of S1p(Rn). The bounded embedding (3.7) is a direct consequence of (3.5) and the definition of the space S1p(Q). The compact embedding (3.8) follows from (3.7) and Arzela-Ascoli's theorem.
The following remarks are added following on the recommendation of the reviewer:
Remark 4.1. Reviewer of the paper writes: The techniques used to prove the Theorem 3.2 are nice alternative to Fourier-based methods. The results of Theorem 3.2 and Corollary 3.1 are new if p=1. If 1<p<∞, the results can be derived from the known embedding results of Besov and Triebel-Lizorkin spaces as follows: The theorem on page 104 in [7] tells us that SmpW(Rn)=Smp,2F(Rn) if m∈N and 1<p<∞. By Proposition 1.15 in [9] we have Smp,2(Rn)↪Smp,max(p,2)B(Rn)↪Smp,∞B(Rn). Next, by the theorem on p.131 in [7] we have Smp,∞B(Rn)↪Sm−1/p∞,∞B(Rn). The latter space Sm−1/p∞,∞B(Rn) is a dominating mixed smoothness type Hölder-Zygmund space and sometimes denoted by Sm−1,1−1/pC(Rn). Either way, relying on Theorem 3.1 from [12] we get the embedding Sm−1,1−1/pC(Rn)↪Bm−1/p∞,∞(Rn)=Cm−1,1−1/p(Rn), where the last equality is a classical result (see e.g., [13]). In summary, we have the embedding SmpW(Rn)↪Cm−1,1−1/p(Rn). Regarding the assertion on traces, the cases with p=1,∞ are novel in the literature. But the cases 1<p<∞ are contained in Lemma 4.1 of [14].
Remark 4.2. The task of embedding large subspaces of the Sobolev spaces Wmp into Lebesgue spaces Lq is touched on e.g., in [15] or [6].
The concept of Sobolev spaces became a trailblazing idea in many fields of mathematics. The goal of this paper is to gain insight into the embedding of the Sobolev spaces into Hölder spaces-a very powerful concept that reveals the connection between weak differentiability and integrability (or weak regularity) of the function with its pointwise regularity. It is well-known that the embedding of the Sobolev space of weakly differentiable functions into Hölder spaces holds if the integrability exponent is higher than the space dimension. Otherwise speaking, one can trade one degree of weak regularity with an integrability exponent higher than the space dimension to upgrade the pointwise regularity to Hölder continuity. In this paper, the embedding of the Sobolev functions into the Hölder spaces is expressed in terms of the minimal weak differentiability requirement independent of the integrability exponent. Precisely, the question asked is what is the minimal weak regularity degree of Sobolev functions which upgrades the pointwise regularity to Hölder continuity independent of the integrability exponent. The paper presents proof of the embedding of the Sobolev spaces with dominating mixed smoothness into Hölder spaces. The new method of proof is based on the generalization of the Newton-Leibniz formula to the n-dimensional rectangle and inductive application of the Sobolev trace embedding results. Counterexamples demonstrate that in terms of minimal weak regularity degree the Sobolev spaces with dominating mixed smoothness present the largest class of weakly differentiable functions which preserve generalized Newton-Leibniz formula, and upgrades the pointwise regularity to Hölder continuity.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The author states that there is no conflict of interest.
[1] | S. L. Sobolev, On a theorem of functional analysis, Mat. Sbornik, 4 (1938), 471–497. |
[2] | R. A. Adams, J. F. Fournier, Sobolev spaces, Elsevier, 2003. |
[3] | C. B. Morrey, Jr, On the solutions of quasi-linear elliptic partial differential equations, Trans. Amer. Math. Soc., 43, 1938,126–166. https://doi.org/10.2307/1989904 |
[4] | S. M. Nikol'skii, On boundary properties of differentiable functions of several variables, Russ. Acad. Sci., 146 (1962), 542–545. |
[5] | S. M. Nikol'sky, On stable boundary values of differentiable functions of several variables, Mat. Sb., 61 (1963), 224–252. |
[6] | O. V. Besov, V. P. Il'in, S. M. Nikol'skii, Integral representations of functions and imbedding theorems, Washington: Winston Sons, 1978. |
[7] | H. J. Schmeisser, H. Triebel, Topics in Fourier analysis and function spaces, Wiley, 1987. |
[8] | J. Vybiral, Function spaces with dominating mixed smoothness, Diss. Math., 436, 2006, 1–73. |
[9] | H. Triebel, Function spaces with dominating mixed smoothness, EMS Series of Lectures in Mathematics, Zürich: Eur. Math. Soc., 2019. |
[10] | F. J. Hickernell, I. H. Sloan, G. W. Wasilkowski, On tractability of weighted integration over bounded and unbounded regions in Rs, Math. Comput., 73 (2004), 1885–1901. |
[11] | T. Tao, Sobolev Spaces, 2009, Available from: https://terrytao.wordpress.com/2009/04/30/245c-notes-4-sobolev-spaces/ |
[12] | V. K. Nguyen, W. Sickel, Isotropic and dominating mixed besov spaces-a comparison, 2016, https://arXiv.org/abs/1601.04000 |
[13] | H. Triebel, Theory of Function Spaces II, Birkhäuser Basel, 1992. https://doi.org/10.1007/978-3-0346-0419-2 |
[14] |
V. K. Nguyen, W. Sickel, Pointwise multipliers for Sobolev and Besov spaces of dominating mixed smoothness, J. Math. Anal. Appl., 452 (2017), 62–90. https://doi.org/10.1016/j.jmaa.2017.02.046 doi: 10.1016/j.jmaa.2017.02.046
![]() |
[15] | R. A. Adams, Reduced Sobolev inequalities, Can. Math. Bull., 31 (1988), 159–167. https://doi.org/10.4153/CMB-1988-024-1 |