This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Real order total variation with applications to the loss functions in learning schemes

Pan Liu Medical Big Data Research Center,
Chinese PLA General Hospital,
Beijing 100853, China
[email protected]
Xin Yang Lu Department of Mathematical Sciences,
Lakehead University, 955 Oliver Road, Thunder Bay, ON, Canada
[email protected]
Kunlun He (corresponding author)
Medical Big Data Research Center,
Chinese PLA General Hospital,
Beijing 100853, China
[email protected]
Abstract.

Loss function are an essential part in modern data-driven approach, such as bi-level training scheme and machine learnings. In this paper we propose a loss function consisting of a rr-order (an)-isotropic total variation semi-norms TVrTV^{r}, r+r\in{\mathbb{R}}^{+}, defined via the Riemann-Liouville (R-L) fractional derivative. We focus on studying key theoretical properties, such as the lower semi-continuity and compactness with respect to both the function and the order of derivative rr, of such loss functions.

Key words and phrases:
total variation, fractional derivative, calculus of variations
2020 Mathematics Subject Classification:
26B30, 94A08, 47J20

1. Introduction

Medical image processing is a crucial step in routine clinical diagnoses, which includes tasks such as image denoising, image segmentation and reconstruction. Methods for such tasks can be roughly classified into two classes, based on the approach: one is the class of structure-driven approaches such as variational method proposed in [13]; the other is the class of data-driven approaches such as bi-level learning scheme and deep learning method [10, 15, 16, 6, 7, 17] which becomes more popular in recent image processing industry.

We take image denoising task as example: let uηu_{\eta} be a corrupted image, one of the classical structure-driven approach, the ROF model (see [13]) shown that the task of computing a denoised, or reconstructed, image from a noisy image uηu_{\eta} is equivalent to solving the problem

uα:=argmin{uuηL2(Q)2+αTV(u),uBV(Q)},u_{\alpha}:=\mathop{\rm arg\,min}\left\{\left\|u-u_{\eta}\right\|_{L^{2}(Q)}^{2}+\alpha TV(u),\,\,u\in BV(Q)\right\}, (1.1)

where Q:=(0,1)×(0,1)Q:=(0,1)\times(0,1) denotes a squared image and BV(Q):={uL1(Q):TV(u)<+}BV(Q):=\left\{u\in L^{1}(Q):\,\,TV(u)<+\infty\right\}, where

TV(u):=sup{Qu divφ𝑑x:φCc(Q)}TV(u):=\sup\left\{\int_{Q}u\text{ div}\varphi\,dx:\varphi\in C_{c}^{\infty}(Q)\right\} (1.2)

is the total variation semi-norm defined in [4].

Then, the minimizer uαu_{\alpha} in (1.1) is the denoised image for a given α>0\alpha>0. The advantage for such structure-driven method is it does not depends on image itself, robust so that for any given corrupted image uηu_{\eta}, we can certainly obtain a reasonable solution uαu_{\alpha}, however, the disadvantage is, without any additional supervision, choosing α\alpha too large would often result in the loss of relevant image details, and choosing α\alpha too small would produce an image with too much noise.

The data-driven approach, on the other hand, could incorporate the given prior knowledge of the problem in terms of a training set. In image denoising, a “training set”, or a “dataset” 𝒟\mathcal{D}, is a finite family of pair of images

𝒟={(uc,uη),{finitely many of pairs}},\mathcal{D}=\left\{(u_{c},u_{\eta}),\,\,\left\{\text{finitely many of pairs}\right\}\right\}, (1.3)

containing a corrupted one uηu_{\eta}, and the corresponding clean data ucu_{c}. A typical data-driven approach can be defined as follows:

ω~argmin{uc𝒟(uωuc):ωW},\tilde{\omega}\in\mathop{\rm arg\,min}\left\{\sum_{u_{c}\in\mathcal{D}}\mathcal{L}\left(u_{\omega}-u_{c}\right):\,\,\omega\in W\right\}, (1.4)

for

uω=𝒩ω(uη).u_{\omega}=\mathcal{N}_{\omega}(u_{\eta}). (1.5)

The functional 𝒩ω\mathcal{N}_{\omega} in (1.5) can be choosing from a widely range of options. For instance, 𝒩ω\mathcal{N}_{\omega} could be the functional from the equation (1.1) where we set ω=α+=W\omega=\alpha\in{\mathbb{R}}^{+}=W. Other possible choices for 𝒩ω\mathcal{N}_{\omega} can be, for example, a convolutional neural network such as AlexNet [10], VGG [15], GoogleNet [16], ResNet [6], DenseNet [7], and Transformers [17], in which the ω\omega is a set of matrix with real numbers. The purpose of this article is, however, not to study neural network, and hence we will not delve into the regularity issues of such functional 𝒩\mathcal{N} and leave the interested readers to the reference mentioned above.

The main focus of this article is the loss function \mathcal{L}, another critical part in data-driven approach which measures the difference between the clean image ucu_{c} and the reconstructed image uωu_{\omega}. The choice of loss function is a delicate and crucial one, as it decides how (1.4) to optimizes ω\omega. Popular choice of loss functions are L2L^{2} or L1L^{1} norm (also called MSE and MAE, respectively). However, such loss functions usually resulting in images that have high peak signal-to-noise ratios, but often lacking high-frequency details, and are therefore error-prone with respect to fine-scale, possibly clinically relevant details. Moreover, for medical images, especially the computed tomography angiography (CTA) scans for vessel reconstruction, the images are 3d dimensional and rather large in size. Therefore, the region of interested (ROI) constitutes a very small portion of the whole area, and hence, a suitable loss function must be able to accurately focus on both local area information, as well as global area information.

In [12], a loss function composed with real order derivative was first introduced and successfully applied in GAN [5], a generative adversarial network, and received improved fine-scale details in reconstruction images. The loss function is concisely defined as the rr-order total variation TVrTV^{r} (see Definition 3.4), with r(0,1)r\in(0,1) fixed. However, due to the limited literature at that time, no theoretical statements, such as compactness and semi continuity results, could be provided.

In this paper, we shall provide a full analysis of such real order derivative loss functions, and study our new functional TVrTV^{r}, i.e. the real order total variation semi-norm (of order rr).

Definition 1.1 (Definition 3.4).

Let r{\lfloor r\rfloor} denote the integer part of r+r\in{\mathbb{R}}^{+}. We define the rr-order total variation TVr(u)TV^{r}(u) on uL1(Q)u\in L^{1}(Q) as follows.

  1. 1.

    For r=s(0,1)r=s\in(0,1) (i.e. r=0{\lfloor r\rfloor}=0), we define

    TVs(u):=sup{Qudivsφdx:φCc(Q;N) and |φ|1},TV^{s}(u):=\sup\left\{\int_{Q}u\,{\operatorname{div}}^{s}\varphi\,dx:\,\,\varphi\in C_{c}^{\infty}(Q;{{{\mathbb{R}}}^{N}})\text{ and }\left\lvert\varphi\right\rvert\leq 1\right\}, (1.6)

    where divs{\operatorname{div}}^{s} is the “fractional divergence” operator, defined in (3.9) below.

  2. 2.

    For r=r+sr={\lfloor r\rfloor}+s where r1{\lfloor r\rfloor}\geq 1, we define

    TVr(u):=sup{Qudivs[divrφ]𝑑x:φCc(Q;𝕄N×(Nk)) and |φ|1}.TV^{r}(u):=\sup\left\{\int_{Q}u\,{\operatorname{div}}^{s}[{\operatorname{div}}^{\lfloor r\rfloor}\varphi]\,dx:\,\,\varphi\in C_{c}^{\infty}(Q;\mathbb{M}^{N\times(N^{k})})\text{ and }\left\lvert\varphi\right\rvert\leq 1\right\}. (1.7)

The TVrTV^{r} can be thus considered as an extension of the classic total variation semi-norm, which corresponds to the case r=1r=1, as well as the L1L^{1} norm, which corresponds to the case r=0r=0.

Similarly, we introduce the functional space BVrBV^{r}, defined as the space of L1L^{1} functions with bounded TVrTV^{r} semi-norm:

Definition 1.2.

Let r+r\in{\mathbb{R}}^{+} be given.

  1. 1.

    We define the (standard) BVr(Q)BV^{r}(Q) space by

    BVr(Q):={uL1(Q):TVr(u)<+}.BV^{r}(Q):=\left\{u\in L^{1}(Q):\,\,TV^{r}(u)<+\infty\right\}. (1.8)
  2. 2.

    Given a sequence {un}L1(Q)\left\{u_{n}\right\}\subseteq L^{1}(Q) such that TVr(un)<+TV^{r}(u_{n})<+\infty for each nn\in{\mathbb{N}}, we say it is strictly converging to uL1(Q)u\in L^{1}(Q) with respect to the TVrTV^{r} semi-norm, and write uns-suu_{n}\mathrel{\mathop{\to}\limits^{\text{s-s}}}u, if

    limnuunL1(Q)+|TVr(un)TVr(u)|=0.\lim_{n\rightarrow\infty}\left\|u-u_{n}\right\|_{L^{1}(Q)}+\left\lvert TV^{r}(u_{n})-TV^{r}(u)\right\rvert=0. (1.9)

    That is, uns-suu_{n}\mathrel{\mathop{\to}\limits^{\text{s-s}}}u if unuu_{n}\to u strongly in L1(Q)L^{1}(Q) and TVr(un)TVr(u)TV^{r}(u_{n})\to TV^{r}(u).

Thus, the space BVrBV^{r} can be therefore considered as an extension of the space BVBV. The aim of this article is twofold. First, in Sections 2 and 3, we study the basic properties of functions with bounded real order total variation, such as the lower semi-continuity (l.s.c.) with respect to the function, and the compactness of the embedding BVr(Q)L1(Q)BV^{r}(Q)\hookrightarrow L^{1}(Q). The main result is (in which the space SVsSV^{s} is defined in Definition 3.6):

Theorem 1.3 (see Theorem 4.3).

Given a non-integer order r+r\in{\mathbb{R}}^{+}\setminus\mathbb{N}, let s:=rr(0,1)s:=r-\lfloor r\rfloor\in(0,1) be the decimal part of rr. Consider a sequence {un}SVs(Q)BV(Q)\left\{u_{n}\right\}\subset SV^{s}(Q)\cap BV(Q) satisfying

sup{unL(Q)+unBVs(Q):n}<+.\sup\left\{\left\|u_{n}\right\|_{L^{\infty}(\partial Q)}+\left\|u_{n}\right\|_{BV^{s}(Q)}:\,\,n\in{\mathbb{N}}\right\}<+\infty. (1.10)

Then, there exists uBVs(Q)u\in BV^{s}(Q) such that, up to a sub-sequence (which we do not relabel), unuu_{n}\to u strongly in L1(Q)L^{1}(Q).

Moreover, the recent research on neural architecture search (NAS) [19, 11] further speeds up the process of designing more powerful structures, and in certain cases, it involves to perform optimize regarding a set of loss functions, i.e., use model to determine the most suitable loss function from a set of functions. Hence, we will, in Section 4, investigate the functional properties of {TVrn}\left\{TV^{r_{n}}\right\} with respect to a sequence of orders {rn}+\left\{r_{n}\right\}\subset{\mathbb{R}}^{+}. The main theorem is:

Theorem 1.4 (see Theorem 4.12).

Given sequences {rn}+\left\{r_{n}\right\}\subset{\mathbb{R}}^{+} and {un}L1(Q)\left\{u_{n}\right\}\subset L^{1}(Q) such that rnr+{0}r_{n}\to r\in{\mathbb{R}}^{+}\cup\left\{0\right\}, assume there exists p(1,+]p\in(1,+\infty] such that

sup{unLp(Q)+TVrn(un):n}<+.\sup\left\{\left\|u_{n}\right\|_{L^{p}(Q)}+TV^{r_{n}}(u_{n}):\,\,n\in{\mathbb{N}}\right\}<+\infty. (1.11)

Then, the following statements hold.

  1. 1.

    There exists uBVr(Q)u\in BV^{r}(Q), such that, up to a sub-sequence, unuu_{n}\rightharpoonup u weakly in Lp(Q)L^{p}(Q) and

    lim infnTVrn(un)TVr(u).\liminf_{n\to\infty}TV^{r_{n}}(u_{n})\geq TV^{r}(u). (1.12)
  2. 2.

    Assuming in addition that unSVrn(Q)BV(Q){u_{n}}\in SV^{r_{n}}(Q)\cap BV(Q), unL(Q)\left\|u_{n}\right\|_{L^{\infty}(\partial Q)} is uniformly bounded, and rnr>0r_{n}\to r>0. Then we have

    unu strongly in L1(Q).u_{n}\to u\text{ strongly in }L^{1}(Q). (1.13)

Theorems 1.3 and 1.4 provide the theoretical foundations for TVrTV^{r}. These results provide a rigorous justification that not only TVrTV^{r} can be a loss function for any fixed order rr, but also can be optimized regarding to its order in NAS, as more sophisticated approach in modern architecture design.

The paper is organized as follows: in Section 2 we collect some notations and preliminary results on the fractional order derivative. In Section 3 we analyze the main properties of the fractional rr-order total variation, with fixed r+r\in{\mathbb{R}}^{+}\setminus{\mathbb{N}}. The compact embedding, lower semi-continuity with respect to the order rr, and the relation between the fractional order total variation and its integer order counterpart will be the subjects of Section 4.

2. Preliminary results on fractional order derivatives

Through this article, r+r\in{\mathbb{R}}^{+} denotes a (positive) constant, whose integer and fractional parts are denoted by r{\lfloor r\rfloor} and s[0,1)s\in[0,1), respectively.

We recall the definitions of fractional order derivative in dimension one.

Definition 2.1 (the fractional order derivative on unit interval).

Let I:=(0,1)I:=(0,1) and xIx\in I be given.

  1. 1.

    The left (resp. right)-sided Riemann-Liouville derivatives of order r=r+s+r={\lfloor r\rfloor}+s\in{\mathbb{R}}^{+} (see [14]) are defined by

    dLrw(x)\displaystyle d^{r}_{L}w(x) =1Γ(1s)(ddx)r+10xw(t)(xt)s𝑑t,Γ(s):=0etts1𝑑t,\displaystyle=\frac{1}{\Gamma(1-s)}\left(\frac{d}{dx}\right)^{{\lfloor r\rfloor}+1}\int_{0}^{x}\frac{w(t)}{(x-t)^{s}}dt,\qquad\Gamma(s):=\int_{0}^{\infty}e^{-t}t^{s-1}dt, (2.1)
    dRrw(x)\displaystyle d^{r}_{R}w(x) =(1)r+1Γ(1s)(ddx)r+1x1w(t)(tx)s𝑑t.\displaystyle=\frac{(-1)^{{\lfloor r\rfloor}+1}}{\Gamma(1-s)}\left(\frac{d}{dx}\right)^{{\lfloor r\rfloor}+1}\int_{x}^{1}\frac{w(t)}{(t-x)^{s}}dt. (2.2)
  2. 2.

    The left (resp. right)-Riemann-Liouville fractional order integrals, of order r+r\in{\mathbb{R}}^{+}, are defined by

    (𝕀Lrw)(x):=1Γ(r)0xw(t)(xt)1r𝑑t and (𝕀Rrw)(x):=1Γ(r)x1w(t)(xt)1r𝑑t,(\mathbb{I}_{L}^{r}w)(x):=\frac{1}{\Gamma(r)}\int_{0}^{x}\frac{w(t)}{(x-t)^{1-r}}dt\quad\text{ and }\quad({\mathbb{I}_{R}^{r}}w)(x):=\frac{1}{\Gamma(r)}\int_{x}^{1}\frac{w(t)}{(x-t)^{1-r}}dt, (2.3)

    respectively.

  3. 3.

    The left (resp. right)-sided Caputo derivative, of order rr, are defined by

    dL,crw(x)\displaystyle d^{r}_{L,c}w(x) :=1Γ(1s)0x(dr+1)w(t)(xt)s𝑑t,\displaystyle:=\frac{1}{\Gamma(1-s)}\int_{0}^{x}\frac{(d^{{\lfloor r\rfloor}+1})w(t)}{(x-t)^{s}}dt,
    dR,crw(x)\displaystyle d^{r}_{R,c}w(x) :=(1)r+1Γ(1s)x1(dr+1)w(t)(tx)s𝑑t.\displaystyle:=\frac{(-1)^{{\lfloor r\rfloor}+1}}{\Gamma(1-s)}\int_{x}^{1}\frac{(d^{{\lfloor r\rfloor}+1})w(t)}{(t-x)^{s}}dt.

We collect some immediate results regarding Definition 2.1 from the literature.

Remark 2.2.

Let wC(I¯)w\in C^{\infty}(\bar{I}) and ϕCc(I)\phi\in C_{c}^{\infty}(I) be given.

  1. 1.

    The following integration by parts formulas hold (see, e.g.,[1]):

    IwdR,crϕ𝑑x\displaystyle\int_{I}w\,d_{R,c}^{r}\phi\,dx =(1)r+1I(dLrw)ϕ𝑑x,\displaystyle=(-1)^{{\lfloor r\rfloor}+1}\int_{I}(d_{L}^{r}w)\,\phi\,dx,
    IwdL,crϕ𝑑x\displaystyle\int_{I}w\,d_{L,c}^{r}\phi\,dx =(1)r+1I(dRrw)ϕ𝑑x.\displaystyle=(-1)^{{\lfloor r\rfloor}+1}\int_{I}(d_{R}^{r}w)\,{\phi}\,dx.
  2. 2.

    The R-L and Caputo derivatives are equivalent for compactly supported functions ([14, Theorem 2.2]), i.e., for every xIx\in I, it holds

    dLrϕ(x)=dL,crϕ(x) and dRrϕ(x)=dR,crϕ(x).d^{r}_{L}\phi(x)=d^{r}_{L,c}\phi(x)\text{ and }d^{r}_{R}\phi(x)=d^{r}_{R,c}\phi(x). (2.4)

    Thus, the integration by parts formulas above can be rewritten as

    IwdRrϕ𝑑x\displaystyle\int_{I}w\,d_{R}^{r}\phi\,dx =(1)r+1I(dLrw)ϕ𝑑x,\displaystyle=(-1)^{{\lfloor r\rfloor}+1}\int_{I}(d_{L}^{r}w)\,\phi\,dx,
    IwdLrϕ𝑑x\displaystyle\int_{I}w\,d_{L}^{r}\phi\,dx =(1)r+1I(dRrw)ϕ𝑑x.\displaystyle=(-1)^{{\lfloor r\rfloor}+1}\int_{I}(d_{R}^{r}w)\,{\phi}\,dx.
  3. 3.

    For given functions w1w_{1}, w2C(I¯)w_{2}\in C^{\infty}(\bar{I}), and parameters aa, bb\in{\mathbb{R}}, the following linearity property holds:

    dr(aw1(x)+bw2(x))=adrw1(x)+bdrw2(x).d^{r}(aw_{1}(x)+bw_{2}(x))=ad^{r}w_{1}(x)+bd^{r}w_{2}(x). (2.5)
Remark 2.3.

In the following, we shall only work with left-sided fractional order derivative/integrations, as the arguments for the right-sided analogues are identical. Thus, for brevity, we drop the underlying LL and RR, and write drd^{r} and 𝕀r\mathbb{I}^{r}, instead of dLrd^{r}_{L} and 𝕀Lr\mathbb{I}^{r}_{L}, unless otherwise specified.

We next recall the definition of representable functions.

Definition 2.4 (Representable functions).

We denote by 𝕀r(L1(I))\mathbb{I}^{r}(L^{1}(I)), r>0r>0, the space of functions represented by the rr-order derivative of a summable function. That is,

𝕀r(L1(I)):={fL1(I):f=𝕀rw,wL1(I)}.\mathbb{I}^{r}(L^{1}(I)):=\left\{f\in L^{1}(I):\,\,f=\mathbb{I}^{r}w,\,\,w\in L^{1}(I)\right\}. (2.6)

Next we recall several theorems, from [14], on representable functions in dimension one.

Theorem 2.5.

For convenience, we unify our notation by writing 𝕀r=dr\mathbb{I}^{r}=d^{-r} for r<0r<0.

  1. 1.

    [14, Theorem 2.3] In order for w(x)𝕀r(L1(I))w(x)\in\mathbb{I}^{r}(L^{1}(I)), r>0r>0 to hold, it is necessary and sufficient to have

    𝕀r+1r[w](x)Wr+1,1(I),r=r+s\mathbb{I}^{{\lfloor r\rfloor}+1-r}[w](x)\in W^{{\lfloor r\rfloor}+1,1}(I),\qquad r={\lfloor r\rfloor}+s (2.7)

    and

    (dl𝕀r+1r[w])(0)=0,l=0,1,,r(d^{l}\mathbb{I}^{{\lfloor r\rfloor}+1-r}[w])(0)=0,\qquad l=0,1,\cdots,{\lfloor r\rfloor} (2.8)
  2. 2.

    [14, Theorem 2.5] The relation

    𝕀r1𝕀r2w=𝕀r1+r2w\mathbb{I}^{r_{1}}\mathbb{I}^{r_{2}}w=\mathbb{I}^{{r_{1}}+r_{2}}w (2.9)

    is valid if one of the following conditions holds:

    1. 1.

      r2>0r_{2}>0, r1+r2>0{r_{1}}+r_{2}>0, provided that wL1(I)w\in L^{1}(I),

    2. 2.

      r2<0r_{2}<0, r1>0{r_{1}}>0, provided that w𝕀r2(L1(I))w\in\mathbb{I}^{-r_{2}}(L^{1}(I)),

    3. 3.

      r1<0{r_{1}}<0, r1+r2<0{r_{1}}+r_{2}<0, provided that w𝕀r1r2(L1(I))w\in\mathbb{I}^{-{r_{1}}-r_{2}}(L^{1}(I)).

  3. 3.

    [14, Theorem 2.6] Let r+r\in{\mathbb{R}}^{+} be given.

    1. 1.

      The fractional order integration operator 𝕀r\mathbb{I}^{r} forms a semigroup in Lp(I)L^{p}(I), p1p\geq 1, which is continuous in the uniform topology for all r>0r>0, and strongly continuous for all r0r\geq 0.

    2. 2.

      ([14, (2.72)]) It holds

      𝕀rwLp(a,b)(ba)r1rΓ(r)wLp(a,b),p1.\left\|\mathbb{I}^{r}w\right\|_{L^{p}(a,b)}\leq(b-a)^{r}\frac{1}{r\Gamma(r)}\left\|w\right\|_{L^{p}(a,b)},\qquad{p\geq 1}. (2.10)

Next, we introduce the notations for (partial) fractional order derivative in higher dimension.

Definition 2.6 (fractional order partial derivative).

Given x=(x1,,xN)Q=(0,1)NNx=(x_{1},\ldots,x_{N})\in Q=(0,1)^{N}\subset{{{\mathbb{R}}}^{N}} and uC(Q)u\in C^{\infty}(Q), we define the rr-order partial derivative:

1ru(x):=drdtru(t,x2,x3,,xN),\displaystyle\partial^{r}_{1}u(x):=\frac{d^{r}}{{dt^{r}}}u(t,x_{2},x_{3},\ldots,x_{N}),

and similarly for iru(x)\partial_{i}^{r}u(x), i=2,,Ni=2,\ldots,N.

We next recall the integration by parts formula for fractional order derivatives from [3]: for uC(Q)u\in C^{\infty}(Q), vCc(Q)v\in C_{c}^{\infty}(Q), it holds

Qusvdx=Qsuvdx.\int_{Q}u\,\partial^{s}v\,dx=-\int_{Q}\partial^{s}u\,v\,dx. (2.11)

Then, by Theorem 2.5, we have the following multi-index integration by parts formula:

Qurvdx=(1)r+1Qruvdx.\int_{Q}u\,\partial^{r}v\,dx=(-1)^{{\lfloor r\rfloor}+1}\int_{Q}\partial^{r}u\,v\,dx. (2.12)

We conclude this section by recalling the following technical lemma.

Lemma 2.7 ([14]).

Let kk\in{\mathbb{N}} be given. The following assertions hold.

  1. 1.

    We have

    dLsxk=Γ(k+1)Γ(ks+1)xks,dRs(1x)k=Γ(k+1)Γ(ks+1)(1x)ks,for all xI,s(0,1).d^{s}_{L}x^{k}=\frac{\Gamma(k+1)}{\Gamma\left(k-s+1\right)}x^{k-s},\quad{d^{s}_{R}(1-x)^{k}=\frac{\Gamma(k+1)}{\Gamma\left(k-s+1\right)}(1-x)^{k-s},\qquad\text{for all }x\in I,s\in(0,1)}. (2.13)
  2. 2.

    If k=0k=0, i.e., w(x):=xk=1w(x):=x^{k}=1 is the function identically equal to 1, then dLrw(x)=dRrw(x)=0d^{r}_{L}w(x)=d^{r}_{R}w(x)=0 if and only if rr\in{\mathbb{N}}.

  3. 3.

    For all s(0,1)s\in(0,1), we have dLsxs1=dRs(1x)s1=0d^{s}_{L}x^{s-1}=d^{s}_{R}(1-x)^{s-1}=0.

3. The space of functions with bounded fractional-order total variation

3.1. Total variation with different underlying Euclidean norm

We start by recalling the definition of Euclidean p\ell^{p}-norm on N{{{\mathbb{R}}}^{N}}. Let p[1,+)p\in[1,+\infty) and x=(x1,x2,,xN)Nx=(x_{1},x_{2},\ldots,x_{N})\in{{{\mathbb{R}}}^{N}} be given, we define

|x|p:=(|x1|p+|x2|p+|xN|p)1/p.\left\lvert x\right\rvert_{\ell^{p}}:=(|x_{1}|^{p}+|x_{2}|^{p}+\cdots|x_{N}|^{p})^{1/p}. (3.1)

Note that ||p\left\lvert\cdot\right\rvert_{\ell^{p}} are equivalent norms on N{{{\mathbb{R}}}^{N}}, in the sense that, for all 1q<p1\leq q<p\leq\infty,

|x|p|x|qN1/q1/p|x|p.\left\lvert x\right\rvert_{\ell^{p}}\leq\left\lvert x\right\rvert_{\ell^{q}}\leq N^{1/q-1/p}\left\lvert x\right\rvert_{\ell^{p}}. (3.2)
Definition 3.1.

Let uL1(Q)u\in L^{1}(Q) be given. We recall the following definition: given kk\in{\mathbb{N}}, the kk-th order total variation, with underlying Euclidean p\ell^{p}-norm, is defined as

TVpk(u):=sup{Qudivkφdx:φCc(Q;𝕄N×Nk1) and |φ|p1}.TV_{\ell^{p}}^{k}(u):=\sup\left\{\int_{Q}u\,{\operatorname{div}}^{k}\varphi\,dx:\,\,\varphi\in C_{c}^{\infty}(Q;\mathbb{M}^{N\times N^{k-1}})\text{ and }\left\lvert\varphi\right\rvert_{\ell^{p}}^{\ast}\leq 1\right\}. (3.3)

Here ||p\left\lvert\cdot\right\rvert_{\ell^{p}}^{\ast} denotes the dual norm associated with ||p\left\lvert\cdot\right\rvert_{\ell^{p}},

in the sense that

|φ|p=supisupx|φi(x)|p,1p+1p=1,\left\lvert\varphi\right\rvert_{\ell^{p}}^{*}=\sup_{i}\sup_{x}|\varphi_{i}(x)|_{\ell^{p^{*}}},\qquad\frac{1}{p}+\frac{1}{p^{*}}=1,

where φi\varphi_{i}, i=1,,Nki=1,\cdots,N^{k}, are the individual components of φ\varphi.

Of particular interest are the first and second order total variation, corresponding to the cases k=1,2k=1,2:

TVp(u)\displaystyle TV_{\ell^{p}}(u) :=sup{Qudivφdx:φCc(Q;N) and |φ|p1},\displaystyle:=\sup\left\{\int_{Q}u\,{\operatorname{div}}\varphi\,dx:\,\,\varphi\in C_{c}^{\infty}(Q;{{{\mathbb{R}}}^{N}})\text{ and }\left\lvert\varphi\right\rvert_{\ell^{p}}^{\ast}\leq 1\right\},
TVp2(u)\displaystyle TV_{\ell^{p}}^{2}(u) :=sup{Qudiv2φdx:φCc(Q;𝕄N×N) and |φ|p1}.\displaystyle:=\sup\left\{\int_{Q}u\,{\operatorname{div}}^{2}\varphi\,dx:\,\,\varphi\in C_{c}^{\infty}(Q;\mathbb{M}^{N\times N})\text{ and }\left\lvert\varphi\right\rvert_{\ell^{p}}^{\ast}\leq 1\right\}.

As an example, when N=2N=2, we have φ=[φ1,φ2;φ3,φ4]\varphi=[\varphi_{1},\varphi_{2};\varphi_{3},\varphi_{4}], and

div2φ=div(div(φ1,φ2),div(φ3,φ4))\displaystyle{\operatorname{div}}^{2}\varphi={\operatorname{div}}({\operatorname{div}}(\varphi_{1},\varphi_{2}),{\operatorname{div}}(\varphi_{3},\varphi_{4})) =div(1φ1+2φ2,1φ3+2φ4)\displaystyle={\operatorname{div}}(\partial_{1}\varphi_{1}+\partial_{2}\varphi_{2},\partial_{1}\varphi_{3}+\partial_{2}\varphi_{4})
=11φ1+12φ2+21φ3+22φ4.\displaystyle=\partial_{1}\partial_{1}\varphi_{1}+\partial_{1}\partial_{2}\varphi_{2}+\partial_{2}\partial_{1}\varphi_{3}+\partial_{2}\partial_{2}\varphi_{4}.

We remark that it is possible to recover the classical isotropic total variation (TVTV) and an-isotropic total variation (ATVATV) by letting p=2p=2 and p=1p=1, respectively. Moreover, in dimension one, all TVpTV_{\ell^{p}} are the same.

More in general, the following equivalence result holds.

Lemma 3.2.

[Equivalence of the TVprTV_{\ell^{p}}^{r} norms] Given 1q<p1\leq q<p\leq\infty, we have

N1/p1/qTVqr(u)TVpr(u)TVqr(u),N^{1/p-1/q}TV_{\ell^{q}}^{r}(u)\leq TV_{\ell^{p}}^{r}(u)\leq TV_{\ell^{q}}^{r}(u), (3.4)

for all r+r\in{\mathbb{R}}^{+} and uBVr(Q)u\in BV^{r}(Q).

Proof.

In the following, we will use pp^{*} to denote the dual exponent of pp, i.e. 1p+1p=1\frac{1}{p}+\frac{1}{p^{*}}=1. Let 1q<p+1\leq q<p\leq+\infty be given. From Remark 3.7, we have

N1/p1/q|φ(x)|p|φ(x)|q|φ(x)|p,N^{1/p-1/q}\left\lvert\varphi(x)\right\rvert_{\ell^{p^{\ast}}}\leq\left\lvert\varphi(x)\right\rvert_{\ell^{q^{\ast}}}\leq\left\lvert\varphi(x)\right\rvert_{\ell^{p^{\ast}}}, (3.5)

for all φCc(Q;N)\varphi\in C_{c}^{\infty}(Q;{{{\mathbb{R}}}^{N}}), and xQx\in Q. That is, for any φCc(Q;N)\varphi\in C_{c}^{\infty}(Q;{{{\mathbb{R}}}^{N}}) such that |φ(x)|p1\left\lvert\varphi(x)\right\rvert_{\ell^{p^{\ast}}}\leq 1, we have |φ(x)|q1\left\lvert\varphi(x)\right\rvert_{\ell^{q^{\ast}}}\leq 1. Thus

{φCc(Q;𝕄N×(Nr)):|φ|q1}{φCc(Q;𝕄N×(Nr)):|φ|p1},\left\{\varphi\in C_{c}^{\infty}(Q;\mathbb{M}^{N\times(N^{{\lfloor r\rfloor}})}):\left\lvert\varphi\right\rvert_{\ell^{q}}^{\ast}\leq 1\right\}\supseteq\left\{\varphi\in C_{c}^{\infty}(Q;\mathbb{M}^{N\times(N^{{\lfloor r\rfloor}})}):\left\lvert\varphi\right\rvert_{\ell^{p}}^{\ast}\leq 1\right\},

which then gives TVqr(u)TVpr(u)TV^{r}_{\ell^{q}}(u)\geq TV^{r}_{\ell^{p}}(u). The proof for N1/p1/qTVqr(u)TVpr(u)N^{1/p-1/q}TV_{\ell^{q}}^{r}(u)\leq TV_{\ell^{p}}^{r}(u) is completely analogous.

We recall the usual trace operator for function with bounded total variation.

Theorem 3.3 ([4, Theorem 2, Page 181]).

Let uBV(Q)u\in BV(Q) be given. Then, for N1{\mathcal{H}^{N-1}}-a.e. x0Qx_{0}\in\partial Q,

limε0B(x0,ε)Q|uT[u](x)|𝑑x=0,\lim_{\varepsilon\rightarrow 0}\fint_{B(x_{0},\varepsilon)\cap Q}\left\lvert u-T[u](x)\right\rvert dx=0, (3.6)

where T[]T[\cdot] denotes the standard trace operator. That is,

T[u](x0)=limε0B(x0,ε)Qu(x)𝑑x.T[u](x_{0})=\lim_{\varepsilon\rightarrow 0}\fint_{B(x_{0},\varepsilon)\cap Q}u(x)\,dx. (3.7)
Definition 3.4.

We define the rr-order total variation TVpr(u)TV_{\ell^{p}}^{r}(u) on uL1(Q)u\in L^{1}(Q) as follows.

  1. 1.

    For r=s(0,1)r=s\in(0,1) (i.e. r=0{\lfloor r\rfloor}=0), we define

    TVps(u):=sup{Qudivsφdx:φCc(Q;N) and |φ|p1},TV_{\ell^{p}}^{s}(u):=\sup\left\{\int_{Q}u\,{\operatorname{div}}^{s}\varphi\,dx:\,\,\varphi\in C_{c}^{\infty}(Q;{{{\mathbb{R}}}^{N}})\text{ and }\left\lvert\varphi\right\rvert_{\ell^{p}}^{\ast}\leq 1\right\}, (3.8)

    where

    divsu:=[(11/N)s+1/N]i=1Ni,Rsφi;{\operatorname{div}}^{s}u:=[(1-1/N)s+1/N]\sum_{i=1}^{N}\partial^{s}_{i,R}\varphi_{i}; (3.9)
  2. 2.

    For r=r+sr={\lfloor r\rfloor}+s where r1{\lfloor r\rfloor}\geq 1, we define

    TVpr(u):=sup{Qudivs[divrφ]𝑑x:φCc(Q;𝕄N×(Nk)) and |φ|p1}.TV_{\ell^{p}}^{r}(u):=\sup\left\{\int_{Q}u\,{\operatorname{div}}^{s}[{\operatorname{div}}^{\lfloor r\rfloor}\varphi]\,dx:\,\,\varphi\in C_{c}^{\infty}(Q;\mathbb{M}^{N\times(N^{k})})\text{ and }\left\lvert\varphi\right\rvert_{\ell^{p}}^{\ast}\leq 1\right\}. (3.10)
Remark 3.5.

In (3.9) we applied the right-sided derivative on the test function φ\varphi. In this way, when uu is sufficient regular, e.g. uC(Q¯)u\in C^{\infty}(\bar{Q}), the integration by parts formula

Qudivsφdx=QLsuφdx\int_{Q}u\,{\operatorname{div}}^{s}\varphi\,dx=-\int_{Q}\nabla^{s}_{L}u\,\varphi\,dx (3.11)

holds. Similarly, if we choose to work primarily with the right-sided derivative, we shall use left-sided derivative on the test function φ\varphi in (3.9).

Definition 3.6.

Let r+r\in{\mathbb{R}}^{+} and p[1,+]p\in[1,+\infty] be given.

  1. 1.

    Given a sequence {un}L1(Q)\left\{u_{n}\right\}\subseteq L^{1}(Q) such that TVr(un)<+TV^{r}(u_{n})<+\infty for each nn\in{\mathbb{N}}, we say it is strictly converging to uL1(Q)u\in L^{1}(Q) with respect to the TVprTV^{r}_{\ell^{p}} semi-norm, and write uns-s (p)uu_{n}\mathrel{\mathop{\to}\limits^{\text{s-s }(p)}}u, if

    limnuunL1(Q)+|TVpr(un)TVpr(u)|=0.\lim_{n\rightarrow\infty}\left\|u-u_{n}\right\|_{L^{1}(Q)}+\left\lvert TV^{r}_{\ell^{p}}(u_{n})-TV_{\ell^{p}}^{r}(u)\right\rvert=0. (3.12)

    That is, uns-s (p)uu_{n}\mathrel{\mathop{\to}\limits^{\text{s-s }(p)}}u if unuu_{n}\to u strongly in L1(Q)L^{1}(Q) and TVpr(un)TVpr(u)TV^{r}_{\ell^{p}}(u_{n})\to TV_{\ell^{p}}^{r}(u).

  2. 2.

    We define the space SVr(Q)SV^{r}(Q) by

    SVr(Q):=p[1,+]C(Q)¯ s-s (p).SV^{r}(Q):=\bigcap_{p\in[1,+\infty]}\overline{C^{\infty}(Q)}^{\text{ s-s }(p)}. (3.13)
  3. 3.

    We define the (standard) BVr(Q)BV^{r}(Q) space by

    BVr(Q):=p[1,+]{uL1(Q):TVpr(u)<+}.BV^{r}(Q):=\bigcap_{p\in[1,+\infty]}\left\{u\in L^{1}(Q):\,\,TV^{r}_{\ell^{p}}(u)<+\infty\right\}. (3.14)
Remark 3.7 (Equivalence between TVprTV_{\ell^{p}}^{r}).

Definition 3.6 has several consequences.

  1. 1.

    By (3.2) we have, for any 1q<p+1\leq q<p\leq+\infty,

    N1/p1/qTVpr(u)TVqr(u)TVpr(u).N^{1/p-1/q}TV_{\ell^{p}}^{r}(u)\leq TV_{\ell^{q}}^{r}(u)\leq TV_{\ell^{p}}^{r}(u). (3.15)

    That is, the set {uL1(Q):TVpr(u)<+}\left\{u\in L^{1}(Q):\,\,TV^{r}_{\ell^{p}}(u)<+\infty\right\} is actually independent of pp. As a consequence, the functional space BVr(Q)BV^{r}(Q), defined in (3.14), satisfies

    BVr(Q)={uL1(Q):TV2r(u)<+},BV^{r}(Q)=\left\{u\in L^{1}(Q):\,\,TV^{r}_{\ell^{2}}(u)<+\infty\right\}, (3.16)

    without any dependence on the underlying p\ell^{p}-norm.

  2. 2.

    The space SVr(Q)SV^{r}(Q), defined in (3.13), enjoys the “smooth approximation” property: for each uSVr(Q)u\in SV^{r}(Q), there exists a sequence {un}C(Q)BVr(Q)\left\{u_{n}\right\}\subseteq C^{\infty}(Q)\cap BV^{r}(Q) such that (3.12) holds. In the case of integer order, i.e. r=kr=k\in{\mathbb{N}}, we do have SVk(Q)=BVk(Q)SV^{k}(Q)=BV^{k}(Q) (see for instance [4]). However, due to the singularities at the boundary arising from the definition of fractional derivatives, we are unable to prove a smooth approximation result in this case. In particular, the construction from [4] would not work, unless additional conditions are imposed.

We conclude this subsection with some definitions and properties of fractional order Sobolev semi-norms.

Definition 3.8.

Given q[1,+]q\in[1,+\infty], r>0r>0, we define the rr-order fractional Sobolev space

Wr,q(Q):={uLq(Q):\displaystyle W^{r,{q}}(Q):=\bigg{\{}u\in L^{q}(Q):\,\, there exists gL1(Q;𝕄N×Nr) such that QudivRrφdx=Qgφ𝑑x\displaystyle\text{there exists }g\in L^{1}(Q{;{\mathbb{M}}^{N\times N^{\lfloor r\rfloor}}})\text{ such that }\int_{Q}u\,{\operatorname{div}}_{R}^{r}\varphi\,dx=\int_{Q}g{\cdot}\varphi\,dx
for all test functions φCc(Q;𝕄N×Nr)}.\displaystyle\text{ for all test functions }\varphi\in C_{c}^{\infty}(Q{;{\mathbb{M}}^{N\times N^{\lfloor r\rfloor}}})\bigg{\}}.

Here gLq(Q;𝕄N×Nr)g\in L^{q}(Q{;{\mathbb{M}}^{N\times N^{\lfloor r\rfloor}}}) is the weak rr-order fractional derivative of uu. We equip it with norm

uWpr,q(Q):=uLq(Q)+Q|g|p𝑑x.\left\|u\right\|_{W^{r,{q}}_{\ell^{p}}(Q)}:=\left\|u\right\|_{L^{q}(Q)}+\int_{Q}\left\lvert g\right\rvert_{\ell^{p}}dx. (3.17)
Definition 3.9.

Let r=r+sr={\lfloor r\rfloor}+s. By ACr,1(I)AC^{r,1}(I) we denote the set of all functions w:Iw:I\to{\mathbb{R}} admitting a representation of the form

w(t)=i=0rciΓ(s+i)ts1+i+𝕀rϕ(t),tIa.e.,w(t)=\sum_{i=0}^{\lfloor r\rfloor}\frac{c_{i}}{\Gamma(s+i)}t^{s-1+i}+\mathbb{I}^{r}\phi(t),\,\,t\in I\,\,a.e., (3.18)

where c0,,ckc_{0},\ldots,c_{k}\in{\mathbb{R}} and ϕL1(I)\phi\in L^{1}(I).

We recall the following results form [8].

Theorem 3.10.

Let r=r+sr={\lfloor r\rfloor}+s be given.

  1. 1.

    [8, Theorem 7] A function wL1(I)w\in L^{1}(I) admits the rr-order derivative if and only if wACr,1(I)w\in AC^{r,1}(I). In this case, ww has the representation (3.18), and

    di+sw(0)=ci,i=0,,k2, and drw(t)=ϕ(t),tIa.e..\displaystyle d^{i+s}w(0)=c_{i},\,\,i=0,\cdots,k-2,\text{ and }d^{r}w(t)=\phi(t),\,\,t\in I\,\,a.e.. (3.19)
  2. 2.

    [8, Theorem 19] We have

    Wr,1(I)=ACr,1(I)L1(I).W^{r,1}(I)=AC^{r,1}(I)\cap L^{1}(I). (3.20)

We remark that, from (3.19) and (3.20), for a given wWr,1(I)w\in W^{r,1}(I), the corresponding ϕ=drwL1(I)\phi{=d^{r}w}\in L^{1}(I) satisfies 𝕀rϕL1(I)<+\left\|\mathbb{I}^{r}\phi\right\|_{L^{1}(I)}<+\infty, and hence w𝕀r(L1(I)){w}\in\mathbb{I}^{r}(L^{1}(I)) in view of Definition 2.4.

Remark 3.11.

Let p[1,+]p\in[1,+\infty] be given, and assume that uC(Q)BVs(Q)u\in C^{\infty}(Q)\cap BV^{s}(Q). Then, by [18, Proposition 3.5], we have

TVps(u)=Q|su|p𝑑x=Q|(1su,,Nsu)|p𝑑x.TV^{s}_{\ell^{p}}(u)=\int_{Q}\left\lvert\nabla^{s}u\right\rvert_{\ell^{p}}dx=\int_{Q}\left\lvert(\partial_{1}^{s}u,\cdots,\partial^{s}_{N}u)\right\rvert_{\ell^{p}}dx. (3.21)


Equation (3.21) allows us to use the fact that the an-isotropic total variation TV1TV_{\ell^{1}} can be computed axis by axis. For example, for N=2N=2, by (3.21) we have

TV1s(u)=Q|(1su,2su)|1𝑑x=Q|1su|𝑑x+Q|2su|𝑑x.TV^{s}_{\ell^{1}}(u)=\int_{Q}\left\lvert(\partial_{1}^{s}u,\partial_{2}^{s}u)\right\rvert_{\ell^{1}}dx=\int_{Q}\left\lvert\partial^{s}_{1}u\right\rvert dx+\int_{Q}\left\lvert\partial^{s}_{2}u\right\rvert dx. (3.22)

That is, we are able to separate the integration in the two variables. This will be crucial in allowing us to study properties of real order total variation in the multi-dimensional setting, by using results from the (often easier) one-dimensional case.

3.2. Basic properties of TVrTV^{r} with fixed order of derivative

For brevity, we will only write TVrTV^{r} without explicit reference to the underlying Euclidean p\ell^{p}-norm. However, in several arguments it will be advantageous to use the TV1rTV_{\ell^{1}}^{r} semi-norm (see Remark 3.11). In such instances, we will thus write TV1rTV_{\ell^{1}}^{r} to clarify that we are relying on the underlying norm being the Euclidean 1\ell^{1}-norm. We first prove a lower semi-continuity result with fixed order r+r\in{\mathbb{R}}^{+}.

Theorem 3.12.

Given ub(Q)u\in{\mathcal{M}_{b}}(Q), i.e. the space of finite Radon measures on QQ, and sequence {un}BVr(Q)\left\{u_{n}\right\}\subseteq BV^{r}(Q) satisfying one of the following conditions:

  1. 1.

    uL1(Q)u\in L^{1}(Q) and {un}\left\{u_{n}\right\} is locally uniformly integrable and unuu_{n}\to u a.e.,

  2. 2.

    unuu_{n}\mathrel{\mathop{\rightharpoonup}\limits^{*}}u in b(Q){\mathcal{M}_{b}}(Q)

  3. 3.

    unuu_{n}\rightharpoonup u in Lp(Q)L^{p}(Q), for some p>1p>1,

then we have

lim infnTVr(un)TVr(u).\liminf_{n\to\infty}TV^{r}(u_{n})\geq TV^{r}(u). (3.23)

Note that (3) is stronger than (2), but we stated it explicitly since it is a special case widely used in this paper. To prove Theorem 3.12, a preliminary result is required.

Lemma 3.13.

Let φCc(Q)\varphi\in C_{c}^{\infty}(Q) be given. Then the following statements hold.

  1. 1.

    For any fixed TT\in{\mathbb{N}}, we have

    sup{drφL(Q):r(0,T)}<+.\sup\left\{\left\|d^{r}\varphi\right\|_{L^{\infty}(Q)}:\,\,{r\in(0,T)}\right\}<+\infty. (3.24)
  2. 2.

    For a.e. xQx\in Q, we have

    divrφ(x)divrφ(x) as rr+{\operatorname{div}}^{r}\varphi(x)\to{\operatorname{div}}^{\lfloor r\rfloor}\varphi(x)\text{ as }r\to{\lfloor r\rfloor}^{+} (3.25)

    and

    divrφ(x)divrφ(x) as rr.{\operatorname{div}}^{r}\varphi(x)\to{\operatorname{div}}^{\lceil r\rceil}\varphi(x)\text{ as }r\to\lceil r\rceil^{-}. (3.26)
Proof.

We first consider the one-dimensional case, i.e. Q=I=(0,1)Q=I=(0,1). In view of (2.4), for any r=r+sr={\lfloor r\rfloor}+s, we have

|dLrφ(x)|\displaystyle\left\lvert d^{r}_{L}\varphi(x)\right\rvert =|dL,cr(x)|1Γ(1s)0x|(dr+1)φ(t)|(xt)s𝑑t\displaystyle=\left\lvert d^{r}_{L,c}(x)\right\rvert\leq\frac{1}{\Gamma(1-s)}\int_{0}^{x}\frac{\left\lvert(d^{{\lfloor r\rfloor}+1}){\varphi}(t)\right\rvert}{(x-t)^{s}}dt
φWr+1,(I)Γ(1s)0x1(xt)s𝑑tφWr+1,(I)(s1)Γ(1s)[(xt)1s|0x]φWr+1,(I)(1s)Γ(1s),\displaystyle\leq\frac{\left\|{\varphi}\right\|_{W^{{\lfloor r\rfloor}+1,\infty}(I)}}{\Gamma(1-s)}\int_{0}^{x}\frac{1}{(x-t)^{s}}dt\leq\frac{\left\|{\varphi}\right\|_{W^{{\lfloor r\rfloor}+1,\infty}(I)}}{(s-1)\Gamma(1-s)}\left[(x-t)^{1-s}\bigg{|}^{x}_{0}\right]\leq\frac{\left\|{\varphi}\right\|_{W^{{\lfloor r\rfloor}+1,\infty}(I)}}{(1-s)\Gamma(1-s)},

which implies

|dLrφ(x)|dr+1φL(I)1(1s)Γ(1s).\left\lvert d^{r}_{L}{\varphi}(x)\right\rvert\leq\left\|d^{{\lfloor r\rfloor}+1}{\varphi}\right\|_{L^{\infty}(I)}\cdot\frac{1}{(1-s)\Gamma(1-s)}. (3.27)

Note also that, by Euler’s reflection formula,

sups(0,1)1Γ(1s)(1s)=sups(0,1)1Γ(2s)1,\sup_{s\in(0,1)}\frac{1}{\Gamma(1-s)(1-s)}=\sup_{s\in(0,1)}\frac{1}{\Gamma(2-s)}\leq 1, (3.28)

hence,

dLrφL(I)dr+1φL(I)<+,\left\|d_{L}^{r}{\varphi}\right\|_{L^{\infty}(I)}\leq\left\|d^{{\lfloor r\rfloor}+1}{\varphi}\right\|_{L^{\infty}(I)}<+\infty, (3.29)

and

sup{dLrφL(I):r(0,T)}l=0Tdl+1φL(I)<+.\sup\left\{\left\|d^{r}_{L}\varphi\right\|_{L^{\infty}(I)}:\,\,{r\in(0,T)}\right\}\leq\sum_{l=0}^{T}\left\|d^{l+1}\varphi\right\|_{L^{\infty}(I)}<+\infty. (3.30)

The same arguments give the desired results for the right and central sided RLR-L derivatives.

We next prove Statement 2, for the case 0<r<10<r<1. The case r1r\geq 1 is proven using similar arguments. In view of (2.4), we have dsφ=dcsφd^{s}\varphi=d^{s}_{c}\varphi, with dcsd^{s}_{c} denoting the Caputo fractional derivative. Recall the Laplace transform gives

{dcsφ}(y)=ys{φ}(y)ys1φ(0)=ys{φ}(y).\mathcal{L}\left\{d^{s}_{c}\varphi\right\}(y)=y^{s}\mathcal{L}\left\{\varphi\right\}(y)-y^{s-1}\varphi(0)=y^{s}\mathcal{L}\left\{\varphi\right\}(y). (3.31)

Therefore, we have

lims1{dcsφ}(y)=y{φ}(y) and lims0{dcsφ}(y)={φ}(y),\lim_{s\nearrow 1}\mathcal{L}\left\{d^{s}_{c}\varphi\right\}(y)=y\,\mathcal{L}\left\{\varphi\right\}(y)\qquad\text{ and }\qquad\lim_{s\searrow 0}\mathcal{L}\left\{d^{s}_{c}\varphi\right\}(y)=\mathcal{L}\left\{\varphi\right\}(y), (3.32)

and hence we conclude

lims1dcs(φ)(x)=dφ(x) and lims0dcs(φ)(x)=φ(x),\lim_{s\nearrow 1}d^{s}_{c}(\varphi)(x)=d\varphi(x)\qquad\text{ and }\qquad\lim_{s\searrow 0}d^{s}_{c}(\varphi)(x)=\varphi(x), (3.33)

as desired.

The multi-dimensional case can be directly inferred from the one-dimensional case. We discuss the two-dimensional case (i.e. N=2N=2) as an example. To show the first statement, i.e.

sup{drφL(Q):r(0,T)}<+,for any φCc(Q),\sup\left\{\|d^{r}\varphi\|_{L^{\infty}(Q)}:r\in(0,T)\right\}<+\infty,\qquad\text{for any }\varphi\in C_{c}^{\infty}(Q),

it suffices to note that

|1rφ(x1,x2)|\displaystyle|\partial_{1}^{r}\varphi(x_{1},x_{2})| 1Γ(1s)0x1|1r+1φ(t,x2)|(x1t)s𝑑t(r=r+s)\displaystyle\leq\frac{1}{\Gamma(1-s)}\int_{0}^{x_{1}}\frac{|\partial_{1}^{{\lfloor r\rfloor}+1}\varphi(t,x_{2})|}{(x_{1}-t)^{s}}dt\qquad(r={\lfloor r\rfloor}+s)
φL(Q)1Γ(1s)0x11(x1t)s𝑑tφL(Q)1(1s)Γ(1s).\displaystyle\leq\|\nabla\varphi\|_{L^{\infty}(Q)}\frac{1}{\Gamma(1-s)}\int_{0}^{x_{1}}\frac{1}{(x_{1}-t)^{s}}dt\leq\|\nabla\varphi\|_{L^{\infty}(Q)}\frac{1}{(1-s)\Gamma(1-s)}.

To show the second statement, recall that

divsφ(x1,x2)=1sφ(x1,x2)+2sφ(x1,x2),{\operatorname{div}}^{s}\varphi(x_{1},x_{2})=\partial^{s}_{1}\varphi(x_{1},x_{2})+\partial_{2}^{s}\varphi(x_{1},x_{2}),

thus we have 1sφ(x1,x2)=dsφ(,x2)\partial^{s}_{1}\varphi(x_{1},x_{2})=d^{s}\varphi(\cdot,x_{2}) for fixed x2(0,1)x_{2}\in(0,1). Then, by using the conclusion in the 1D case, we infer that, for s0+s\to 0^{+},

1sφ(x1,x2)φ(x1,x2),for a.e. x1 and fixed x2,\partial_{1}^{s}\varphi(x_{1},x_{2})\to\varphi(x_{1},x_{2}),\qquad\text{for a.e. $x_{1}$ and fixed $x_{2}$},

and similarly,

2sφ(x1,x2)φ(x1,x2),for a.e. x2 and fixed x1.\partial_{2}^{s}\varphi(x_{1},x_{2})\to\varphi(x_{1},x_{2}),\qquad\text{for a.e. $x_{2}$ and fixed $x_{1}$}.

The proof is thus complete. ∎

Proof of Theorem 3.12.

Let r=r+sr={\lfloor r\rfloor}+s be given, and fix an arbitrary φCc(Q;𝕄N×Nr)\varphi\in C_{c}^{\infty}(Q;\mathbb{M}^{N\times N^{{{\lfloor r\rfloor}}}}), |φ|21|\varphi|_{\ell^{2}}^{*}\leq 1. In view of (3.8) (or (3.10)) we have

TVr(un)Qun[divsdivrφ]𝑑x.TV^{r}(u_{n})\geq\int_{Q}u_{n}\,[{\operatorname{div}}^{s}{\operatorname{div}}^{\lfloor r\rfloor}\varphi]\,dx. (3.34)

By Lemma 3.13, from φCc(Q;𝕄N×Nr)\varphi\in C_{c}^{\infty}(Q;\mathbb{M}^{N\times N^{{{\lfloor r\rfloor}}}}), we get divrφ{\operatorname{div}}^{\lfloor r\rfloor}\varphi is also smooth and compactly supported, which in turn gives divsdivrφL(Q){\operatorname{div}}^{s}{\operatorname{div}}^{\lfloor r\rfloor}\varphi\in L^{\infty}(Q).

Assume condition (1) holds. Since {un}\left\{u_{n}\right\} is locally uniformly integrable and unuu_{n}\to u a.e., by the dominated convergence theorem

limnQun[divsdivrφ]𝑑x=Qu[divsdivrφ]𝑑x.\displaystyle\lim_{n\rightarrow\infty}\,\int_{Q}u_{n}\,[{\operatorname{div}}^{s}{\operatorname{div}}^{\lfloor r\rfloor}\varphi]\,dx=\int_{Q}u\,[{\operatorname{div}}^{s}{\operatorname{div}}^{\lfloor r\rfloor}\varphi]\,dx. (3.35)

If we assume either condition (2), or the stronger (3), then note that since φCc(Q;𝕄N×Nr)\varphi\in C_{c}^{\infty}(Q;\mathbb{M}^{N\times N^{\lfloor r\rfloor}}), we have divsdivrφC(Q¯){\operatorname{div}}^{s}{\operatorname{div}}^{\lfloor r\rfloor}\varphi\in C(\bar{Q}). Hence, since {un}L1(Q)\left\{u_{n}\right\}\subset L^{1}(Q), in view of the weak\text{weak}^{\ast} convergence in b(Q){\mathcal{M}_{b}}(Q) (see [2, Page 116]), we have again (3.35).

Thus, in all cases, we have

lim infnTV1r(un)lim infnQun[divsdivrφ]𝑑x=Qu[divsdivrφ]𝑑x.{\liminf_{n\to\infty}}\,TV_{\ell^{1}}^{r}(u_{n})\geq{\liminf_{n\to\infty}}\,\int_{Q}u_{n}\,[{\operatorname{div}}^{s}{\operatorname{div}}^{\lfloor r\rfloor}\varphi]\,dx=\int_{Q}u\,[{\operatorname{div}}^{s}{\operatorname{div}}^{\lfloor r\rfloor}\varphi]\,dx. (3.36)

Taking the supremum over all φCc(Q;𝕄N×Nr)\varphi\in C_{c}^{\infty}(Q;\mathbb{M}^{N\times N^{\lfloor r\rfloor}}) with |φ|21|\varphi|_{\ell^{2}}^{*}\leq 1, we conclude (3.23), as desired. ∎

Now we show that it is possible to approximate functions in SVr(Q)SV^{r}(Q) with smooth functions.

Proposition 3.14 (strict approximation with smooth functions).

Let r+r\in{\mathbb{R}}^{+} and uSVr(Q)u\in SV^{r}(Q) be given. Then there exists a sequence {un}C(Q¯)BVr(Q)\left\{u_{n}\right\}\subset C^{\infty}(\bar{Q})\cap BV^{r}(Q) such that

unu strongly in L1(Q) and limnTVr(un)=TVr(u).u_{n}\to u\text{ strongly in }L^{1}(Q)\text{ and }\lim_{n\rightarrow\infty}{TV^{r}}(u_{n})={TV^{r}}(u). (3.37)

To this aim, two preliminary results are required. First, we recall the Riesz representation theorem.

Theorem 3.15 ([4, Theorem 1, Section 1.8]).

Let L:L: Cc(N,M)C_{c}({{{\mathbb{R}}}^{N}},{\mathbb{R}}^{M})\to{\mathbb{R}} be a linear functional satisfying

sup{L(φ):φCc(N;M),|φ|21,support(φ)K}<+\sup\left\{L(\varphi):\,\,\varphi\in C_{c}({{{\mathbb{R}}}^{N}};{\mathbb{R}}^{M}),\,\,{|\varphi|_{\ell^{2}}^{*}\leq 1},\,\,{\operatorname{support}}(\varphi)\subset K\right\}<+\infty (3.38)

for some compact set KNK\subset{{{\mathbb{R}}}^{N}}. Then there exists a Radon measure μ\mu on N{{{\mathbb{R}}}^{N}}, and a μ\mu-measurable function σ\sigma: NM{{{\mathbb{R}}}^{N}}\to{\mathbb{R}}^{M} such that

  1. 1.

    |σ(x)|=1\left\lvert\sigma(x)\right\rvert=1 for μ\mu-a.e. xx, and

  2. 2.

    L(φ)=Nφσ𝑑μL(\varphi)=\int_{{{\mathbb{R}}}^{N}}\varphi\cdot\sigma d\mu.

Next, we show the following crucial result:

Lemma 3.16.

Given a function uBVr(Q)u\in BV^{r}(Q), there exists a Radon measure μ\mu on QQ and a μ\mu-measurable function σ:QN\sigma:Q\to{{{\mathbb{R}}}^{N}} such that

  1. 1.

    |σ(x)|=1\left\lvert\sigma(x)\right\rvert=1 μ\mu-a.e., and

  2. 2.

    Qudivrφdx=Qφσ𝑑μ\int_{Q}u\,{\operatorname{div}}^{r}\varphi\,dx=-\int_{Q}\varphi\cdot\sigma\,d\mu for all φCc(Q;N)\varphi\in C_{c}^{\infty}(Q;{{{\mathbb{R}}}^{N}}), |φ|21|\varphi|_{\ell^{2}}^{*}\leq 1.

Proof.

We first define the linear functional

L:Cc(Q;𝕄N×Nr),L(φ):=Qudivrφdx.L:\,\,C_{c}^{\infty}(Q;\mathbb{M}^{N\times N^{\lfloor r\rfloor}})\to{\mathbb{R}},\qquad L(\varphi):=-\int_{Q}u\,{\operatorname{div}}^{r}\varphi\,dx. (3.39)

Since TVr(u)<+TV^{r}(u)<+\infty, we have, in view of (3.10),

sup{1φL(Q)Qudivrφdx:φCc(Q;𝕄N×Nr)}=TVr(u)<+.\sup\left\{\frac{1}{\left\|\varphi\right\|_{L^{\infty}(Q)}}\int_{Q}u\,{\operatorname{div}}^{r}\varphi\,dx:\varphi\in C_{c}^{\infty}(Q;\mathbb{M}^{N\times N^{\lfloor r\rfloor}})\right\}=TV^{r}(u)<+\infty. (3.40)

Thus,

|L(φ)|TVr(u)φL(Q).\left\lvert L(\varphi)\right\rvert\leq TV^{r}(u)\left\|\varphi\right\|_{L^{\infty}(Q)}. (3.41)

Now we extend by continuity the definition of LL to the entire space Cc(Q;N)C_{c}(Q;{{{\mathbb{R}}}^{N}}): for an arbitrary φCc(Q;N)\varphi\in C_{c}(Q;{{{\mathbb{R}}}^{N}}), we consider the mollifications φε:=φηε\varphi_{\varepsilon}:=\varphi\ast\eta_{\varepsilon} (for some uninfluential mollifier ηe\eta_{e}) and, by [4, Theorem 1, item (ii), Section 4.2],

φεφ uniformly on Q.\varphi_{\varepsilon}\to\varphi\text{ uniformly on }Q. (3.42)

Therefore, by defining

L¯(φ):=limε0L(φε) for φCc(Q;N),\bar{L}(\varphi):=\lim_{\varepsilon\rightarrow 0}L(\varphi_{\varepsilon})\text{ for }\varphi\in C_{c}(Q;{{{\mathbb{R}}}^{N}}), (3.43)

in view of (3.41) and (3.42), we conclude that

sup{L¯(φ): for φCc(Q;𝕄N×Nk) and |φ|21}<+.\sup\left\{\bar{L}(\varphi):\text{ for }\varphi\in C_{c}(Q;\mathbb{M}^{N\times N^{k}})\text{ and }{|\varphi|_{\ell^{2}}^{*}\leq 1}\right\}<+\infty. (3.44)

Thus, by Theorem 3.15, the proof is complete. ∎

Proof.

( of Proposition 3.14) Let uSVr(Q)u\in SV^{r}(Q) be given. In view of Definition 3.6, there exists a sequence {un}C(Q)\left\{u_{n}\right\}\subset C^{\infty}(Q) such that

uns-su.u_{n}\mathrel{\mathop{\to}\limits^{\text{s-s}}}u. (3.45)

Next, let x0:=(1/2,,1/2)x_{0}:=(1/2,\cdots,1/2) be the center of QQ. For sufficiently small ε>0\varepsilon>0, we define

unε(x):=un(xx01+ε+x0) for xQ.u_{n}^{\varepsilon}(x):=u_{n}\Big{(}\frac{x-x_{0}}{1+\varepsilon}+x_{0}\Big{)}\text{ for }x\in Q. (3.46)

Note that by construction, unεu_{n}^{\varepsilon} will be a scaled version of the restriction of unu_{n} to Qε:=Q/(1+ε)Q_{\varepsilon}:={Q}/\left(1+\varepsilon\right). Since unC(Qε¯)u_{n}\in C^{\infty}(\overline{Q_{\varepsilon}}), we have unεC(Q¯)u_{n}^{\varepsilon}\in C^{\infty}(\bar{Q}) too.

We next show that TVr(unε)TVr(un)TV^{r}(u^{\varepsilon}_{n})\to TV^{r}(u_{n}). Let ε>0\varepsilon>0 be fixed, then by Lemma 3.16,

TVr(un)sup{Qεundivrφdx:φCc(Qε),|φ|21}.TV^{r}(u_{n})\geq\sup\left\{\int_{Q_{\varepsilon}}u_{n}{\operatorname{div}}^{r}\varphi\,dx:\,\,\varphi\in C_{c}^{\infty}(Q_{\varepsilon}),{|\varphi|_{\ell^{2}}^{*}\leq 1}\right\}. (3.47)

On the other hand, for any φCc(Q)\varphi\in C_{c}^{\infty}(Q), we have

Qunεdivrφdx\displaystyle\int_{Q}u_{n}^{\varepsilon}{\operatorname{div}}^{r}\varphi\,dx =Qun((xx0)/(1+ε)+x0)divrφ(x)𝑑x\displaystyle=\int_{Q}u_{n}\left((x-x_{0})/(1+\varepsilon)+x_{0}\right){\operatorname{div}}^{r}\varphi(x)\,dx
(1+ε)r+1Qεun(y)divrφ((yx0)(1+ε)+x0)𝑑y\displaystyle\leq(1+\varepsilon)^{{\lfloor r\rfloor}+1}\int_{Q_{\varepsilon}}u_{n}(y)\,{\operatorname{div}}^{r}\varphi((y-x_{0})(1+\varepsilon)+x_{0})dy
(3.47)(1+ε)r+1TVr(un).\displaystyle\overset{\eqref{smaller_set_tv}}{\leq}(1+\varepsilon)^{{\lfloor r\rfloor}+1}TV^{r}(u_{n}).

Therefore,

TVr(unε)(1+ε)r+1TVr(un).TV^{r}(u_{n}^{\varepsilon})\leq(1+\varepsilon)^{{\lfloor r\rfloor}+1}TV^{r}(u_{n}). (3.48)

On the other hand, in view of (3.46), we have unεunu_{n}^{\varepsilon}\to u_{n} strongly in L1(Q)L^{1}(Q). By Theorem 3.12, we obtain

lim infε0TVr(unε)TVr(un).\liminf_{\varepsilon\to 0}TV^{r}(u_{n}^{\varepsilon})\geq TV^{r}(u_{n}). (3.49)

Combined with (3.48), this implies unεs-sunu_{n}^{\varepsilon}\mathrel{\mathop{\to}\limits^{\text{s-s}}}u_{n}. Combined with (3.45), we infer the existence of a sub-sequence {uεn}C(Q)\left\{u_{\varepsilon_{n}}\right\}\subset C^{\infty}(Q) such that uεns-suu_{\varepsilon_{n}}\mathrel{\mathop{\to}\limits^{\text{s-s}}}u, hence (3.37) is proven. ∎

Remark 3.17.

We again emphasize that, in view of Remark 3.11 and [18, Proposition 3.5], for any uC(Q)BVr(Q)u\in C^{\infty}(Q)\cap BV^{r}(Q), it holds

TV1l+s(u)=Q|l+su|𝑑x=|αs|=l+10101|αsu(x1,,xN)|𝑑x1𝑑xN.TV_{\ell^{1}}^{l+s}(u)=\int_{Q}\left\lvert\nabla^{l+s}u\right\rvert dx=\sum_{\left\lvert\alpha_{s}\right\rvert=l+1}\int_{0}^{1}\cdots\int_{0}^{1}\left\lvert\partial^{\alpha_{s}}u(x_{1},\ldots,x_{N})\right\rvert dx_{1}\cdots dx_{N}. (3.50)

This allows us to compute TV1l+s(u)TV_{\ell^{1}}^{l+s}(u) by computing TVl+s(u)TV^{l^{\prime}+s}(u), l=0,1,,ll^{\prime}=0,1,\cdots,l, along each coordinate axis.

4. Analytic properties of the functional space SVr(Q)SV^{r}(Q)

Remark 4.1.

As the goal of this article is to construct models for imaging applications, we assume the function uL1(Q)u\in L^{1}(Q) analyzed here represents an image. That is, we could restrict ourself to consider only functions in the space

IM(Q):={uBV(Q):T[u]L(Q)1}.{\operatorname{IM}}(Q):=\left\{u\in BV(Q):\,\,\left\|T[u]\right\|_{L^{\infty}(\partial Q)}\leq 1\right\}. (4.1)

Here T:QQT:Q\longrightarrow{\partial}Q denotes the trace operator. We remark that most of the conclusions in this article are independent of (4.1). However, certain results can be improved with (4.1). We will always make it explicit when we do assume (4.1). We also note that assumption (4.1) is crucial in the numerical realization of fractional order derivatives. We refer the interested reader to [3, Section 4] and the references therein.

4.1. Compact embedding for BVsBV^{s} with fixed s(0,1)s\in(0,1)

We start by recalling the following theorem from [2].

Theorem 4.2 ([2, Theorem 4.26]).

Let \mathcal{F} be a bounded set in Lp(N)L^{p}({{{\mathbb{R}}}^{N}}) with 1p<+1\leq p<+\infty. Assume that

lim|h|0τhffLp(N)=0,τhf():=f(+h)\lim_{\left\lvert h\right\rvert\to 0}\left\|\tau_{h}f-f\right\|_{L^{p}({{{\mathbb{R}}}^{N}})}=0,\qquad{\tau_{h}f(\cdot):=f(\cdot+h)} (4.2)

uniformly in ff\in\mathcal{F}. Then, the closure of Ω\mathcal{F}\lfloor\Omega in Lp(Ω)L^{p}(\Omega) is compact for any measurable set ΩN\Omega\subset{{{\mathbb{R}}}^{N}} with finite measure.

The main result of Section 4.1 reads as follows.

Theorem 4.3 (compact embedding BVs(Q)L1(Q)BV^{s}(Q)\hookrightarrow L^{1}(Q)).

Let s(0,1)s\in(0,1) be given. Assume {un}ss-BVs(Q)\left\{u_{n}\right\}\subset\text{ss-}BV^{s}(Q) satisfy

sup{unL(Q)+unBVs(Q):n}<+.\sup\left\{\left\|u_{n}\right\|_{L^{\infty}(\partial Q)}+\left\|u_{n}\right\|_{BV^{s}(Q)}:\,\,n\in{\mathbb{N}}\right\}<+\infty. (4.3)

Then, there exists uBVs(Q)u\in BV^{s}(Q) such that, upon sub-sequence, unuu_{n}\to u strongly in L1(Q)L^{1}(Q).

We prove Theorem 4.3 in several steps. We first recall a revisited left sided RLR-L ss-order derivative from [9], which we denote by d^Lsw(x)\hat{d}_{L}^{s}w(x), as follows:

d^Lsw(x):=1Γ(1s)ddx0xw(t)w(0)(xt)s𝑑t.\hat{d}^{s}_{L}w(x):=\frac{1}{\Gamma(1-s)}\frac{d}{dx}\int_{0}^{x}\frac{w(t)-w(0)}{\left(x-t\right)^{s}}dt. (4.4)

That is, the singularity of dLsd^{s}_{L} at the boundary t=0t=0 is removed. Moreover, we remind that in dimension one, the semi-norms TVTV and TV1TV_{\ell^{1}} are equivalent.

Proposition 4.4.

Let wBVs(I)C(I)w\in BV^{s}(I)\cap C^{\infty}(I) be given. Let TVs(w,I)TV^{s}(w,I) denote the ss-order total variation of ww in I=(0,1)I=(0,1) and

w~(x):={w(x) for xI,0 if xI.\tilde{w}(x):=\begin{cases}w(x)&\text{ for }x\in I,\\ 0&\text{ if }x\in{\mathbb{R}}\setminus I.\end{cases} (4.5)

Then we have

τhw~w~L1()hs(TVs(w)+wL(I))+O(|h|),\left\|\tau_{h}\tilde{w}-\tilde{w}\right\|_{L^{1}({\mathbb{R}})}\leq h^{s}\left(TV^{s}(w)+\left\|w\right\|_{L^{\infty}(\partial I)}\right){+O(|h|)}, (4.6)

provided that wL(I)<+\left\|w\right\|_{L^{\infty}(\partial I)}<+\infty.

Proof.

We start by recalling the fractional mean value formula for 0<s<10<s<1 from [9, Corollary 4.3]: for hh\in{\mathbb{R}} small,

w(x+h)=w(x)+hs1Γ(1+s)d^Lsw(x+θh),w(x+h)=w(x)+h^{s}\frac{1}{\Gamma(1+s)}\hat{d}_{L}^{s}w(x+\theta h), (4.7)

where the revised left-sided R-L ss-order derivative d^Ls\hat{d}^{s}_{L} is defined in (4.4) and θ\theta\in{\mathbb{R}}, say θ(h)\theta(h), depends upon hh, and satisfies

limh0θs(h)=Γ(1+s)2Γ(1+2s).\lim_{h\to 0}\theta^{s}(h)=\frac{\Gamma(1+s)^{2}}{\Gamma(1+2s)}. (4.8)

We note that

d^Lsw(x)\displaystyle\hat{d}^{s}_{L}w(x) =1Γ(1s)ddx0xw(t)w(0)(xt)s𝑑t\displaystyle=\frac{1}{\Gamma(1-s)}\frac{d}{dx}\int_{0}^{x}\frac{w(t)-w(0)}{\left(x-t\right)^{s}}dt
=1Γ(1s)ddx0xw(t)(xt)s𝑑t1Γ(1s)ddx0xw(0)(xt)s𝑑t\displaystyle=\frac{1}{\Gamma(1-s)}\frac{d}{dx}\int_{0}^{x}\frac{w(t)}{\left(x-t\right)^{s}}dt-\frac{1}{\Gamma(1-s)}\frac{d}{dx}\int_{0}^{x}\frac{w(0)}{\left(x-t\right)^{s}}dt
=dLsw(x)w(0)1Γ(1s)1xs.\displaystyle=d^{s}_{L}w(x)-w(0)\frac{1}{\Gamma(1-s)}\frac{1}{x^{s}}. (4.9)

We also, by the definition of w~\tilde{w}, summarize the following 4 cases (w.l.o.g we only consider |h|<0.1\left\lvert h\right\rvert<0.1).

  1. Case 1.

    Both xx and (x+h)I(x+h)\in{\mathbb{R}}\setminus I. In this case we have

    |w~(x+h)w~(x)|=0;\left\lvert\tilde{w}(x+h)-\tilde{w}(x)\right\rvert=0; (4.10)
  2. Case 2.

    xIx\in I, h<0h<0 and x+h<0x+h<0. In this case we have 0<x<h0<x<-h and

    |w~(x+h)w~(x)||w~(x+h)w(0)|+|w(0)w(x)|=|w(0)|+|w(0)w(x)|;\left\lvert\tilde{w}(x+h)-\tilde{w}(x)\right\rvert\leq\left\lvert\tilde{w}(x+h)-w(0)\right\rvert+\left\lvert w(0)-w(x)\right\rvert=\left\lvert w(0)\right\rvert+\left\lvert w(0)-w(x)\right\rvert; (4.11)
  3. Case 3.

    xIx\in I, h>0h>0 and x+h>1x+h>1. In this case we have 0<1x<h0<1-x<h and

    |w~(x+h)w~(x)||w~(x+h)w(1)|+|w(1)w(x)|=|w(1)|+|w(1)w(x)|;\left\lvert\tilde{w}(x+h)-\tilde{w}(x)\right\rvert\leq\left\lvert\tilde{w}(x+h)-w(1)\right\rvert+\left\lvert w(1)-w(x)\right\rvert=\left\lvert w(1)\right\rvert+\left\lvert w(1)-w(x)\right\rvert; (4.12)
  4. Case 4.

    Both xx and x+hIx+h\in I. In this case we have

    |w~(x+h)w~(x)|=|w(x+h)w(x)|.\left\lvert\tilde{w}(x+h)-\tilde{w}(x)\right\rvert=\left\lvert w(x+h)-w(x)\right\rvert. (4.13)

We now claim (4.6). In any of above cases, we could deduce that

τ~hww~L1()|IhIh|wL()+τhwwL1(Ih),\left\|\tilde{\tau}_{h}w-\tilde{w}\right\|_{L^{1}({\mathbb{R}})}\leq\left\lvert I^{h}\setminus I_{h}\right\rvert\left\|w\right\|_{L^{\infty}({\mathbb{R}})}+\left\|\tau_{h}w-w\right\|_{L^{1}(I_{h})}, (4.14)

where Ih=(|h|,1|h|)I_{h}=(\left\lvert h\right\rvert,1-\left\lvert h\right\rvert) and Ih=(|h|,1+|h|)I^{h}=(-\left\lvert h\right\rvert,1+\left\lvert h\right\rvert). Clearly, the first term |IhIh|wL()\left\lvert I^{h}\setminus I_{h}\right\rvert\left\|w\right\|_{L^{\infty}({\mathbb{R}})} is of order O(|h|)O(|h|). Next, for xx and x+hIx+h\in I, we observe that

|w~(x+h)w~(x)|\displaystyle\left\lvert\tilde{w}(x+h)-\tilde{w}(x)\right\rvert hs1Γ(1+s)|d^Lsw(x+θh)|\displaystyle\leq h^{s}\frac{1}{\Gamma(1+s)}\left\lvert\hat{d}_{L}^{s}w(x+\theta h)\right\rvert
hs1Γ(1+s)[|dLsw(x)|+w(0)1Γ(1s)1xs],\displaystyle\leq h^{s}\frac{1}{\Gamma(1+s)}\left[\left\lvert d^{s}_{L}w(x)\right\rvert+w(0)\frac{1}{\Gamma(1-s)}\frac{1}{x^{s}}\right],

where we used (4.7) and (4.9).

That is, we have

w~(x+h)w~(x)L1(Ih)hs1Γ(1+s)[TVs(w,I)+wL(I)[Γ(1s)(1s)]1],\left\|\tilde{w}(x+h)-\tilde{w}(x)\right\|_{L^{1}(I_{h})}\leq h^{s}\frac{1}{\Gamma(1+s)}\left[TV^{s}(w,I)+\left\|w\right\|_{L^{\infty}(\partial I)}[\Gamma(1-s)(1-s)]^{-1}\right], (4.15)

and combining with (4.14) we obtain

τ~hww~L1()hs[[Γ(1s)(1s)]1+[Γ(1+s)]1][TVs(w)+wL(I)]+O(|h|),\left\|\tilde{\tau}_{h}w-\tilde{w}\right\|_{L^{1}({\mathbb{R}})}\leq h^{s}\left[[\Gamma(1-s)(1-s)]^{-1}+[\Gamma(1+s)]^{-1}\right]\left[TV^{s}(w)+\left\|w\right\|_{L^{\infty}(\partial I)}\right]{+O(|h|)}, (4.16)

hence (4.6). ∎

Proof.

(of Theorem 4.3) The main idea of the proof is to combine an approximation and a slicing argument. We only write the proof for the case N=2N=2, as the general case N3N\geq 3 can be obtained similarly. Let {un}BVs(Q)\left\{u_{n}\right\}\subset BV^{s}(Q) be given, such that the assumption (4.3) holds. Define

u~n(x):={un(x) if xQ,0 if x2Q.\tilde{u}_{n}(x):=\begin{cases}u_{n}(x)&\text{ if }x\in Q,\\ 0&\text{ if }x\in{\mathbb{R}}^{2}\setminus Q.\end{cases} (4.17)

Then, by Proposition 4.4, we have, for any given x2x_{2}\in{\mathbb{R}} and small ε\varepsilon,

τεu~n(x1,x2)u~n(x1,x2)L1()εsCs[TV1(un(x1,x2),I)+unL(Q)]+O(|ε|),\left\|\tau_{\varepsilon}\tilde{u}_{n}({x_{1}},x_{2})-\tilde{u}_{n}({x_{1}},x_{2})\right\|_{L^{1}({\mathbb{R}})}\leq\varepsilon^{s}C_{s}[TV_{\ell^{1}}(u_{n}({x_{1}},x_{2}),I)+\left\|u_{n}\right\|_{L^{\infty}(\partial Q)}]{+O(|\varepsilon|)}, (4.18)

where Cs:=[[Γ(1s)(1s)]1+[Γ(1+s)]1]C_{s}:=\left[[\Gamma(1-s)(1-s)]^{-1}+[\Gamma(1+s)]^{-1}\right]. Thus, we have, for vector h2h\in{\mathbb{R}}^{2} with small norm and parallel to the x1x_{1} axis,

τhu~nu~nL1(2)hsCs[TV1s(un)+unL(Q)]+O(|h|).\left\|\tau_{h}\tilde{u}_{n}-\tilde{u}_{n}\right\|_{L^{1}({\mathbb{R}}^{2})}\leq h^{s}C_{s}\left[TV_{\ell^{1}}^{s}(u_{n})+\left\|u_{n}\right\|_{L^{\infty}(\partial Q)}\right]{+O(|h|)}. (4.19)

Similarly, the above estimate holds for h2h\in{\mathbb{R}}^{2} with small norm and parallel to the x2x_{2} axis. For a general direction vector h2h\in{\mathbb{R}}^{2}, we write h=(h1,h2)h=(h_{1},h_{2}), and

τhu~nu~nL1(2)\displaystyle\left\|\tau_{h}\tilde{u}_{n}-\tilde{u}_{n}\right\|_{L^{1}({\mathbb{R}}^{2})} τh1u~nu~nL1(2)+τh2u~nu~nL1(2)\displaystyle\leq\left\|\tau_{h_{1}}\tilde{u}_{n}-\tilde{u}_{n}\right\|_{L^{1}({\mathbb{R}}^{2})}+\left\|\tau_{h_{2}}\tilde{u}_{n}-\tilde{u}_{n}\right\|_{L^{1}({\mathbb{R}}^{2})}
2hsCsC[unBVs(Q)+unL(Q)]+O(|h|).\displaystyle\leq 2h^{s}C_{s}C\left[\left\|u_{n}\right\|_{BV^{s}(Q)}+\left\|u_{n}\right\|_{L^{\infty}(\partial Q)}\right]{+O(|h|)}.

Thus, in view of Theorem 4.2, there exists uL1(Q)u\in L^{1}(Q) such that u~nu\tilde{u}_{n}\to u strongly in L1(Q)L^{1}(Q). That is, unuu_{n}\to u strongly in L1(Q)L^{1}(Q), and the proof is complete. ∎

Corollary 4.5.

Given a sequence {un}BVs(Q)\left\{u_{n}\right\}\subset BV^{s}(Q) such that the assumptions of Theorem 4.3 hold, then, up to a sub-sequence, there exists uBVs(Q)u\in BV^{s}(Q) such that

unu strongly in L1(Q) and lim infnTVs(un)TVs(u).u_{n}\to u\text{ strongly in }L^{1}(Q)\quad\text{ and }\quad{\liminf_{n\to\infty}}\,TV^{s}(u_{n})\geq TV^{s}(u). (4.20)
Proof.

The proof is done by combining Theorems 4.3 and 3.12. ∎

4.2. Lower semi-continuity with respect to sequences of orders

We first perform our analysis for the an-isotropic total variation TV1rTV_{\ell^{1}}^{r}, and then use the equivalence condition (Remark 3.7) to recover the case of isotropic total variation TVrTV^{r}.

4.2.1. Interpolation properties of an-isotropic total variation

We start by studying an monotonicity result of TV1sTV_{\ell^{1}}^{s} with respect to the order s(0,1)s\in(0,1), for function uIM(Q)u\in IM(Q) (recall space IM(Q)IM(Q) from (4.1)).

Proposition 4.6 (monotonicity of an-isotropic TV1sTV_{\ell^{1}}^{s}).

Let 0<s<t<10<s<t<1 be given and uSV1t(Q)IM(Q)u\in SV_{\ell^{1}}^{t}(Q)\cap\,IM(Q). Then we have uBVs(Q)IM(Q)u\in BV^{s}(Q)\cap IM(Q) and

TV1s(u)TV1t(u).TV_{\ell^{1}}^{s}(u)\leq TV_{\ell^{1}}^{t}(u). (4.21)

In particular, at s=0s=0, we have

uL1(Q)TV1t(u).\left\|u\right\|_{L^{1}(Q)}\leq TV_{\ell^{1}}^{t}(u). (4.22)
Proof.

For a moment we assume that 0<s<t<10<s<t<1. We deal in one dimension first and we assume also that wBVt(I)C(I)w\in BV^{t}(I)\cap C^{\infty}(I). We claim w𝕀t(L1(I))w\in\mathbb{I}^{t}(L^{1}(I)). Indeed, we have

d(𝕀1sw(x))=ddx0xw(t)(xt)s𝑑t=dsw(x).d(\mathbb{I}^{1-s}w(x))=\frac{d}{dx}\int_{0}^{x}\frac{w(t)}{(x-t)^{s}}dt=d^{s}w(x). (4.23)

Thus, by assumption we have

𝕀1swW1,1(I)=TVs(w)<+,\|\mathbb{I}^{1-s}w\|_{W^{1,1}(I)}=TV^{s}(w)<+\infty, (4.24)

and hence (2.7) holds.

We next claim (2.8). Indeed, we only need to consider the case l=0l=0 and we observe that

|𝕀1sw(ε)|=|0εw(t)(εt)s𝑑t|wL(I)0ε1(εt)s𝑑twL(I)ε1s0\left\lvert\mathbb{I}^{1-s}w(\varepsilon)\right\rvert=\left\lvert\int_{0}^{\varepsilon}\frac{w(t)}{(\varepsilon-t)^{s}}dt\right\rvert\leq\left\|w\right\|_{L^{\infty}(\partial I)}\int_{0}^{\varepsilon}\frac{1}{(\varepsilon-t)^{s}}dt\leq\left\|w\right\|_{L^{\infty}(\partial I)}\varepsilon^{1-s}\to 0 (4.25)

which implies that

d0𝕀1sw(0)=0.d^{0}\mathbb{I}^{1-s}w(0)=0. (4.26)

Thus, (4.24) and (4.26) allows us to use Theorem 2.5, Assertion 1 to infer the existence of fL1(I)f\in L^{1}(I) such that w=𝕀t[f]w=\mathbb{I}^{t}[f]. Moreover, in view of Theorem 2.5, Assertion 2, we have

w=𝕀t[f]=𝕀s𝕀ts[f].w=\mathbb{I}^{t}[f]=\mathbb{I}^{s}\mathbb{I}^{t-s}[f]. (4.27)

We observe that

dswL1(I)=𝕀ts[f]L1(I)|𝕀ts|fL1(I),\left\|d^{s}w\right\|_{L^{1}(I)}=\left\|\mathbb{I}^{t-s}[f]\right\|_{L^{1}(I)}\\ \leq\left\lvert\mathbb{I}^{t-s}\right\rvert\left\|f\right\|_{L^{1}(I)}, (4.28)

where by |𝕀ts|\left\lvert\mathbb{I}^{t-s}\right\rvert we denote the operator norm. Thus, in view of (2.10), and setting δ=ts>0\delta=t-s>0, we have

|𝕀δ|supφL1>0𝕀δφL1(a,b)φL1(a,b)(ba)δ1δΓ(δ)=(ba)δ1Γ(δ+1)1,\left\lvert\mathbb{I}^{\delta}\right\rvert\leq\sup_{\left\|\varphi\right\|_{L^{1}}>0}\frac{\left\|\mathbb{I}^{\delta}\varphi\right\|_{L^{1}(a,b)}}{\left\|\varphi\right\|_{L^{1}(a,b)}}\leq(b-a)^{\delta}\frac{1}{\delta\Gamma(\delta)}=(b-a)^{\delta}\frac{1}{\Gamma(\delta+1)}\leq 1, (4.29)

whenever ba1b-a\leq 1. This, together with (4.28), gives

dswL1(I)dtwL1(I),\left\|d^{s}w\right\|_{L^{1}(I)}\leq\left\|d^{t}w\right\|_{L^{1}(I)}, (4.30)

as desired.

We next deal with the multi-dimensional case. We shall only write in details for N=2N=2, as the case N3N\geq 3 is similar. Assume that uBVt(Q)C(Q)u\in BV^{t}(Q)\cap C^{\infty}(Q), and in view of Remark 3.17, we have

TV1t(u)=Q|su|𝑑x=0101|1tu(x)|𝑑x1𝑑x2+0101|2tu(x)|𝑑x1𝑑x2.TV_{\ell^{1}}^{t}(u)=\int_{Q}\left\lvert\nabla^{s}u\right\rvert dx=\int_{0}^{1}\int_{0}^{1}\left\lvert\partial^{t}_{1}u(x)\right\rvert dx_{1}dx_{2}+\int_{0}^{1}\int_{0}^{1}\left\lvert\partial^{t}_{2}u(x)\right\rvert dx_{1}dx_{2}. (4.31)

Since uC(Q)BVt(Q)u\in C^{\infty}(Q)\cap BV^{t}(Q), we have, for each x2(0,1)x_{2}\in(0,1), the slice wx2(x1):=u(x1,x2)w_{x_{2}}(x_{1}):=u(x_{1},x_{2}) is well defined and belongs to BVt(I)BV^{t}(I). Thus

01|1su(x1,x2)|𝑑x1=TV1s(wx2(x1))TV1t(wx2(x1))=01|1tu(x1,x2)|𝑑x1.\int_{0}^{1}\left\lvert\partial^{s}_{1}u(x_{1},x_{2})\right\rvert dx_{1}=TV_{\ell^{1}}^{s}(w_{x_{2}}(x_{1}))\leq TV_{\ell^{1}}^{t}(w_{x_{2}}(x_{1}))=\int_{0}^{1}\left\lvert\partial^{t}_{1}u(x_{1},x_{2})\right\rvert dx_{1}. (4.32)

Integrating over x2Ix_{2}\in I, we have

0101|1su(x)|𝑑x1𝑑x20101|1tu(x)|𝑑x1𝑑x2.\int_{0}^{1}\int_{0}^{1}\left\lvert\partial^{s}_{1}u(x)\right\rvert dx_{1}dx_{2}\leq\int_{0}^{1}\int_{0}^{1}\left\lvert\partial^{t}_{1}u(x)\right\rvert dx_{1}dx_{2}. (4.33)

The same conclusion, but with 2s{\partial}_{2}^{s} (resp. 2t{\partial}_{2}^{t}) instead of 1s{\partial}_{1}^{s} (resp. 1t{\partial}_{1}^{t}), can be proven using the exact same arguments. This, together with (4.33) and (4.31), gives

TV1s(u)TV1t(u) for uC(Q)BVt(Q).TV_{\ell^{1}}^{s}(u)\leq TV_{\ell^{1}}^{t}(u)\text{ for $u\in C^{\infty}(Q)\cap BV^{t}(Q)$}. (4.34)

Finally, assume uSV1t(Q)u\in SV_{\ell^{1}}^{t}(Q) only. From Definition 3.6 we could obtain a sequence {un}BVt(Q)C(Q)\left\{u_{n}\right\}\subset BV^{t}(Q)\cap C^{\infty}(Q) such that unuu_{n}\to u strongly in L1(Q)L^{1}(Q) and

TV1t(un)TV1t(u).TV_{\ell^{1}}^{t}(u_{n})\to TV_{\ell^{1}}^{t}(u). (4.35)

Then, in view of (4.34), for each unu_{n} we have

TV1s(un)TV1t(un).TV_{\ell^{1}}^{s}(u_{n})\leq TV_{\ell^{1}}^{t}(u_{n}). (4.36)

Also, since unuu_{n}\to u strongly in L1(Q)L^{1}(Q), in view of Theorem 3.12, and together with (4.35), we conclude that

TV1s(u)lim infnTV1s(un)lim supnTV1t(un)=TV1t(u).TV_{\ell^{1}}^{s}(u)\leq{\liminf_{n\to\infty}}\,TV_{\ell^{1}}^{s}(u_{n})\leq\limsup_{n\to\infty}TV_{\ell^{1}}^{t}(u_{n})=TV_{\ell^{1}}^{t}(u). (4.37)

In the end, take any φCc(Q;2)\varphi\in C_{c}^{\infty}(Q;{\mathbb{R}}^{2}), we observe, in view of Lemma 3.13 and (3.9), that

lim infs0TV1s(u)lim infs0Qudivsφdx=12Qu(φ1+φ2)𝑑x,\liminf_{s\searrow 0}TV_{\ell^{1}}^{s}(u)\geq\liminf_{s\searrow 0}\int_{Q}u\,{\operatorname{div}}^{s}\varphi\,dx=\frac{1}{2}\int_{Q}u\left(\varphi_{1}+\varphi_{2}\right)\,dx, (4.38)

and hence, we conclude, by the arbitrariness of φ\varphi, that

TV1s(u)TV10(u)=uL1(Q),TV_{\ell^{1}}^{s}(u)\geq TV_{\ell^{1}}^{0}(u)={\|u\|_{L^{1}(Q)}}, (4.39)

which concludes the case that s=0s=0. The proof is thus complete. ∎

We next investigate the lower semi-continuity and compactness with respect to the order.

Proposition 4.7.

Given sequences {sn}(0,1)\left\{s_{n}\right\}\subset(0,1), sns[0,1]s_{n}\to s\in[0,1], and {un}L1(Q)\left\{u_{n}\right\}\subset L^{1}(Q) such that there exists p(1,+]p\in(1,+\infty] satisfying

sup{unLp(Q)+TVsn(un):n}<+,\sup\left\{\left\|u_{n}\right\|_{L^{p}(Q)}+TV^{s_{n}}(u_{n}):\,\,n\in{\mathbb{N}}\right\}<+\infty, (4.40)

then the following statements hold.

  1. 1.

    There exists uBVs(Q)u\in BV^{s}(Q) and, up to a sub-sequence, unuu_{n}\rightharpoonup u weakly in Lp(Q)L^{p}(Q), and

    lim infnTVsn(un)TVs(u).\liminf_{n\to\infty}TV^{s_{n}}(u_{n})\geq TV^{s}(u). (4.41)
  2. 2.

    Assuming in addition that unSVsn(Q)u_{n}\in SV^{s_{n}}(Q) for each nn\in{\mathbb{N}}, sns(0,1]s_{n}\to s\in(0,1], and {un}IM(Q)\left\{u_{n}\right\}\subset IM(Q), we have

    unu strongly in L1(Q).u_{n}\to u\text{ strongly in }L^{1}(Q). (4.42)
Proof.

Let {un}L1(Q)\left\{u_{n}\right\}\subset L^{1}(Q) be given, such that (4.40) holds. Since unLp(Q)\left\|u_{n}\right\|_{L^{p}(Q)} (with p>1p>1) is uniformly bounded, there exists uLp(Q)u\in L^{p}(Q) such that, upon sub-sequence,

unu weakly in Lp(Q).u_{n}\rightharpoonup u\text{ weakly in }L^{p}(Q). (4.43)

We now prove Statement 1. In view of Lemma 3.13 we get that for any φCc(Q)\varphi\in C_{c}^{\infty}(Q) and s[0,1]s\in[0,1], it holds divsφLp(Q){\operatorname{div}}^{s}\varphi\in L^{p^{\prime}}(Q) (where 1/p+1/p=11/p+1/{p^{\prime}}=1) and hence, by the weak Lp(Q)L^{p}(Q)-convergence in (4.43), we have that

limnQundivsφdx=Qudivsφdx.\lim_{n\rightarrow\infty}\int_{Q}u_{n}\,{\operatorname{div}}^{s}\varphi\,dx=\int_{Q}u\,{\operatorname{div}}^{s}\varphi\,dx. (4.44)

Statement 2 of Lemma 3.13 gives divsnφdivsφ{\operatorname{div}}^{s_{n}}\varphi\to{\operatorname{div}}^{s}\varphi a.e., while Statement 1 of Lemma 3.13 gives

supndivsnφL(Q)<+.\sup_{n}\|{\operatorname{div}}^{s_{n}}\varphi\|_{L^{\infty}(Q)}<+\infty.

Thus, by dominated convergence theorem, divsnφdivsφ{\operatorname{div}}^{s_{n}}\varphi\to{\operatorname{div}}^{s}\varphi strongly in Lp(Q)L^{p^{\prime}}(Q). By Hölder inequality,

lim supnQ|un||divsnφdivsφ|𝑑xsupnunLp(Q)lim supndivsnφdivsφLp(Q)=0.\displaystyle\limsup_{n\to\infty}\int_{Q}\left\lvert u_{n}\right\rvert\left\lvert{\operatorname{div}}^{s_{n}}\varphi-{\operatorname{div}}^{s}\varphi\right\rvert dx\leq\sup_{n\in{\mathbb{N}}}\left\|u_{n}\right\|_{L^{p}(Q)}\limsup_{n\to\infty}\left\|{\operatorname{div}}^{s_{n}}\varphi-{\operatorname{div}}^{s}\varphi\right\|_{L^{p^{\prime}}(Q)}=0.

Therefore, we have

limn\displaystyle\lim_{n\rightarrow\infty} Qundivsnφdx\displaystyle\int_{Q}u_{n}\,{\operatorname{div}}^{s_{n}}\varphi\,dx
lim infnQundivsφdx+lim infnQun[divsnφdivsφ]𝑑x\displaystyle\geq{\liminf_{n\to\infty}}\int_{Q}u_{n}\,{\operatorname{div}}^{s}\varphi\,dx+{\liminf_{n\to\infty}}\int_{Q}u_{n}\left[{\operatorname{div}}^{s_{n}}\varphi-{\operatorname{div}}^{s}\varphi\right]dx
=Qudivsφdx,\displaystyle=\int_{Q}u\,{\operatorname{div}}^{s}\varphi\,dx, (4.45)

and hence

lim infnTVsn(un)lim infnQundivsnφdx=limnQundivsnφdx=Qudivsφdx,{\liminf_{n\to\infty}}\,TV^{s_{n}}(u_{n})\geq{\liminf_{n\to\infty}}\,\int_{Q}u_{n}\,{\operatorname{div}}^{s_{n}}\varphi\,dx=\lim_{n\rightarrow\infty}\int_{Q}u_{n}\,{\operatorname{div}}^{s_{n}}\varphi\,dx=\int_{Q}u\,{\operatorname{div}}^{s}\varphi\,dx, (4.46)

and (4.41) follows by the arbitrariness of φCc(Q;N)\varphi\in C_{c}^{\infty}(Q;{{{\mathbb{R}}}^{N}}).

We next prove Statement 2. By Remark 3.7 and (4.40) we have that

sup{unLp(Q)+TV1sn(un):n}<+,\sup\left\{\left\|u_{n}\right\|_{L^{p}(Q)}+TV_{\ell^{1}}^{s_{n}}(u_{n}):\,\,n\in{\mathbb{N}}\right\}<+\infty, (4.47)

Since sns(0,1]s_{n}\to s\in(0,1], in view of Proposition 4.6, there exist a (sufficiently large) n0{n_{0}}\in{\mathbb{N}} and sn0<1s_{n_{0}}<1, such that

TV1sn0/2(un)TV1sn(un)+1.for all nn0.TV_{\ell^{1}}^{s_{n_{0}}/2}(u_{n})\leq TV_{\ell^{1}}^{s_{n}}(u_{n})+1.\qquad\text{for all }n\geq{n_{0}}. (4.48)

Combined with the assumption that {un}IM(Q)\left\{u_{n}\right\}\subset IM(Q) allows us to use Theorem 4.3 to infer Statement 2. ∎

The next theorem provides a connection between the lower order TVsTV^{s} to the higher order TVrTV^{r}, where we recall again that r=r+sr={\lfloor r\rfloor}+s.

Theorem 4.8.

Let r=r+sr={\lfloor r\rfloor}+s be given. There exists a constant Cr>0C_{r}>0 such that

TVs(u)Cr[uL1(Q)+TVr(u)],TV^{s}(u)\leq C_{r}\left[\left\|u\right\|_{L^{1}(Q)}+TV^{r}(u)\right], (4.49)

for all uSVr(Q)u\in SV^{r}(Q).

We will first show an one-dimensional version.

Lemma 4.9.

Let I=(0,1)I=(0,1) and r=r+sr={\lfloor r\rfloor}+s be given. There exists a constant Cr>0C_{r}>0 such that

l=0r1TVs+l(w)Cr[wL1(I)+TVr(w)],\sum_{l=0}^{{\lfloor r\rfloor}-1}TV^{s+l}(w)\leq C_{r}\left[\left\|w\right\|_{L^{1}(I)}+TV^{r}(w)\right], (4.50)

for all wSVr(I)w\in SV^{r}(I).

Proof.

We deal with the case r=1{\lfloor r\rfloor}=1 first. That is, we have r=1+sr=1+s.

Assume no such CrC_{r} exists, i.e., there exists a sequence {wn}SV1+s\left\{w_{n}\right\}\subset SV^{1+s} such that, for each nn\in{\mathbb{N}},

TVs(wn)=1 and wnL1(I)+TV1+s(wn)<1/n.TV^{s}(w_{n})=1\text{ and }\left\|w_{n}\right\|_{L^{1}(I)}+TV^{1+s}(w_{n})<1/n. (4.51)

In view of Theorem 3.14, we may as well assume that {wn}C(I)BV1+s(I)\left\{w_{n}\right\}\subset C^{\infty}(I)\cap BV^{1+s}(I) and we may write

dswnL1(I)1/2 and wnL1(I)+d1+swnL1(I)<2/n.\left\|d^{s}w_{n}\right\|_{L^{1}(I)}\geq 1/2\quad\text{ and }\quad\left\|w_{n}\right\|_{L^{1}(I)}+\left\|d^{1+s}w_{n}\right\|_{L^{1}(I)}<2/n. (4.52)

Thus, we have that {wn}W1+s(I)\left\{w_{n}\right\}\subset W^{1+s}(I), and

wn0 and d1+swn0 strongly in L1(I).w_{n}\to 0\quad\text{ and }\quad d^{1+s}w_{n}\to 0\quad\text{ strongly in }L^{1}(I). (4.53)

Recall from Theorem 3.10, we may write that, for each wnw_{n},

wn(t)=c0,nΓ(s)ts1+c1,nΓ(s+1)ts+𝕀1+sϕn(t),tIa.e.,w_{n}(t)=\frac{c_{0,n}}{\Gamma(s)}t^{s-1}+\frac{c_{1,n}}{\Gamma(s+1)}t^{s}+\mathbb{I}^{1+s}\phi_{n}(t),\,\,t\in I\,\,a.e., (4.54)

and

d1+swn(t)=ϕn(t).d^{1+s}w_{n}(t)=\phi_{n}(t). (4.55)

Thus, in view of (4.53) we have

ϕn0 strongly in L1(I),\phi_{n}\to 0\text{ strongly in }L^{1}(I), (4.56)

and together with (2.10), we have, for any 0ss0\leq s^{\prime}\leq s,

𝕀1+sϕn(t)L1(I)1Γ(2+s)ϕnL1(I)0.\left\|\mathbb{I}^{1+s^{\prime}}\phi_{n}(t)\right\|_{L^{1}(I)}\leq\frac{1}{\Gamma(2+s^{\prime})}\left\|\phi_{n}\right\|_{L^{1}(I)}\to 0. (4.57)

We next claim that

c0,n0 and c1,n0.c_{0,n}\to 0\text{ and }c_{1,n}\to 0. (4.58)

By the mean value theorem we have

(𝕀1sw)(t0)=I(𝕀1sw)(l)𝑑l.(\mathbb{I}^{1-s}w)(t_{0})=\int_{I}(\mathbb{I}^{1-s}w)(l)dl. (4.59)

Since wW1+s(I)w\in W^{1+s}(I), we have that (𝕀1sw)(\mathbb{I}^{1-s}w) is absolutely continuous, hence

(𝕀1sw)(t)=(𝕀1sw)(t0)+t0td1(𝕀1sw)(l)𝑑l=(𝕀1sw)(t0)+t0t(dsw)(l)𝑑l.(\mathbb{I}^{1-s}w)(t)=(\mathbb{I}^{1-s}w)(t_{0})+\int_{t_{0}}^{t}d^{1}(\mathbb{I}^{1-s}w)(l)dl=(\mathbb{I}^{1-s}w)(t_{0})+\int_{t_{0}}^{t}(d^{s}w)(l)dl. (4.60)

Thus, for any t[0,1]t\in[0,1],

|(𝕀1sw)(t)|𝕀1swL1(I)+dsuL11Γ(2s)wL1(I)+dswL1(I),\left\lvert(\mathbb{I}^{1-s}w)(t)\right\rvert\leq\left\|\mathbb{I}^{1-s}w\right\|_{L^{1}(I)}+\left\|d^{s}u\right\|_{L^{1}}\leq\frac{1}{\Gamma(2-s)}\left\|w\right\|_{L^{1}(I)}+\left\|d^{s}w\right\|_{L^{1}(I)}, (4.61)

where at the last inequality we used (2.10). This, and together with (3.19), gives that

|c0,n|=|(𝕀1swn)(0)|1Γ(2s)wnL1(I)+dswnL1(I).\left\lvert c_{0,n}\right\rvert=\left\lvert(\mathbb{I}^{1-s}w_{n})(0)\right\rvert\leq\frac{1}{\Gamma(2-s)}\left\|w_{n}\right\|_{L^{1}(I)}+\left\|d^{s}w_{n}\right\|_{L^{1}(I)}. (4.62)

Similarly, we may show that

|c1,n|=|(d(𝕀1s)wn)(0)|dswnL1(I)+d1+swnL1(I).\left\lvert c_{1,n}\right\rvert=\left\lvert(d(\mathbb{I}^{1-s})w_{n})(0)\right\rvert\leq\left\|d^{s}w_{n}\right\|_{L^{1}(I)}+\left\|d^{1+s}w_{n}\right\|_{L^{1}(I)}. (4.63)

Thus, in view of (4.52), we have

sup{|c0,n|+|c1,n|:n}<+.\sup\left\{\left\lvert c_{0,n}\right\rvert+\left\lvert c_{1,n}\right\rvert:\,\,n\in{\mathbb{N}}\right\}<+\infty. (4.64)

We also notice that

c0,nΓ(s)ts1+c1,nΓ(s+1)tsL1(I)\displaystyle\left\|\frac{c_{0,n}}{\Gamma(s)}t^{s-1}+\frac{c_{1,n}}{\Gamma(s+1)}t^{s}\right\|_{L^{1}(I)} =c0,nΓ(s)ts1+c1,nΓ(s+1)ts+𝕀1+sϕn(t)𝕀1+sϕn(t)L1(I)\displaystyle=\left\|\frac{c_{0,n}}{\Gamma(s)}t^{s-1}+\frac{c_{1,n}}{\Gamma(s+1)}t^{s}+\mathbb{I}^{1+s}\phi_{n}(t)-\mathbb{I}^{1+s}\phi_{n}(t)\right\|_{L^{1}(I)}
=wn𝕀1+sϕn(t)L1(I)\displaystyle=\left\|w_{n}-\mathbb{I}^{1+s}\phi_{n}(t)\right\|_{L^{1}(I)}
wnL1(I)+𝕀1+sϕn(t)L1(I)0,\displaystyle\leq\left\|w_{n}\right\|_{L^{1}(I)}+\left\|\mathbb{I}^{1+s}\phi_{n}(t)\right\|_{L^{1}(I)}\to 0, (4.65)

which, combined with (4.64), implies

c0,n0 and c1,n0,c_{0,n}\to 0\text{ and }c_{1,n}\to 0,

and hence (4.58). Then, in view of Lemma 2.7, we have

dsts1=0 and dsts=Γ(s+1).d^{s}t^{s-1}=0\text{ and }d^{s}t^{s}=\Gamma(s+1). (4.66)

Next, in view of Definition 2.4 and Theorem 2.5, Assertion 1, we have ϕ𝕀s(L1(I))\phi\in\mathbb{I}^{s}(L^{1}(I)). Combined with Statement 2 of Theorem 2.5, Assertion 2, we obtain

ds𝕀1+sϕn(t)=𝕀1ϕn(t).d^{s}\mathbb{I}^{1+s}\phi_{n}(t)=\mathbb{I}^{1}\phi_{n}(t). (4.67)

Thus, we have

dswn=c1,n+𝕀1ϕn(t),d^{s}w_{n}=c_{1,n}+\mathbb{I}^{1}\phi_{n}(t), (4.68)

and together with (4.57) and (4.58), we conclude that

lim supn0dswnL1(I)|c1,n|+𝕀1ϕnL1(I)0,\limsup_{n\to 0}\left\|d^{s}w_{n}\right\|_{L^{1}(I)}\leq\left\lvert c_{1,n}\right\rvert+\left\|\mathbb{I}^{1}\phi_{n}\right\|_{L^{1}(I)}\to 0, (4.69)

which contradicts (4.52).

Now we consider the general case r{\lfloor r\rfloor}\in{\mathbb{N}}. In view of (3.18), we write

wn(t)=l=0rcl,nΓ(s+l)ts1+l+𝕀rϕn(t),tIa.e..w_{n}(t)=\sum_{l=0}^{\lfloor r\rfloor}\frac{c_{l,n}}{\Gamma(s+l)}t^{s-1+l}+\mathbb{I}^{r}\phi_{n}(t),\,\,t\in I\,\,a.e.. (4.70)

Similarly to (4.65), we may show that

|c0,n|1Γ(2s)wnL1(I)+dswnL1(I),\left\lvert c_{0,n}\right\rvert\leq\frac{1}{\Gamma(2-s)}\left\|w_{n}\right\|_{L^{1}(I)}+\left\|d^{s}w_{n}\right\|_{L^{1}(I)}, (4.71)

and

|cl,n|ds+l1wnL1(I)+ds+lwnL1(I),\left\lvert c_{l,n}\right\rvert\leq\left\|d^{s+l-1}w_{n}\right\|_{L^{1}(I)}+\left\|d^{s+l}w_{n}\right\|_{L^{1}(I)}, (4.72)

for l=1,,r1l=1,\ldots,{\lfloor r\rfloor}-1, and further deduce that

cl,n0 for i=0,,r.c_{l,n}\to 0\text{ for }i=0,\ldots,{\lfloor r\rfloor}. (4.73)

Note that, for each i=0,,r1i=0,\ldots,{\lfloor r\rfloor}-1, we have

  1. 1.

    dl+sts1+j=0d^{l+s}t^{s-1+j}=0, for j=0,,lj=0,\ldots,l;

  2. 2.

    dl+stl+s=Γ(s+1)d^{l+s}t^{l+s}=\Gamma(s+1);

  3. 3.

    dl+stl+s+j=Γ(l+s+j+1)Γ(j+1)tjd^{l+s}t^{l+s+j}=\frac{\Gamma(l+s+j+1)}{\Gamma(j+1)}t^{j}, for j=1,,r1lj=1,\ldots,{\lfloor r\rfloor}-1-l,

  4. 4.

    dl+s(𝕀r[ϕn](t))=𝕀rl[ϕn](t)d^{l+s}(\mathbb{I}^{r}[\phi_{n}](t))=\mathbb{I}^{{\lfloor r\rfloor}-l}[\phi_{n}](t), a.e., tIt\in I.

Therefore, we obtain that

lim supn0dl+swnL1(I)j=0r1l|cl+s+j,n|+𝕀rl[ϕn]L1(I)0,\limsup_{n\to 0}\left\|d^{l+s}w_{n}\right\|_{L^{1}(I)}\leq\sum_{j=0}^{{\lfloor r\rfloor}-1-l}\left\lvert c_{l+s+j,n}\right\rvert+\left\|\mathbb{I}^{{\lfloor r\rfloor}-l}[\phi_{n}]\right\|_{L^{1}(I)}\to 0, (4.74)

which is a contradiction, and hence we conclude the proof. ∎

Proof.

(of Theorem 4.8) We only deal with the case N=2N=2, as the case N3N\geq 3 is similar. We first show that

TV1s(u)Cr[uL1(Q)+TV1r(u)].TV_{\ell^{1}}^{s}(u)\leq C_{r}\left[\left\|u\right\|_{L^{1}(Q)}+TV_{\ell^{1}}^{r}(u)\right]. (4.75)

We assume for a moment that uBVr(Q)C(Q)u\in BV^{r}(Q)\cap C^{\infty}(Q). We start with the case r=1{\lfloor r\rfloor}=1. That is, r=1+sr=1+s. In view of Remark 3.17, we have

TV1s(u)=Q|1su|𝑑x+Q|2su|𝑑xTV_{\ell^{1}}^{s}(u)=\int_{Q}\left\lvert\partial_{1}^{s}u\right\rvert dx+\int_{Q}\left\lvert\partial_{2}^{s}u\right\rvert dx (4.76)

and

TV11+s(u)=Q|11+su|𝑑x+Q|21+su|𝑑x+Q|1s2u|𝑑x+Q|2s1u|𝑑x.TV_{\ell^{1}}^{1+s}(u)=\int_{Q}\left\lvert\partial_{1}^{1+s}u\right\rvert dx+\int_{Q}\left\lvert\partial_{2}^{1+s}u\right\rvert dx+\int_{Q}\left\lvert\partial_{1}^{s}\partial_{2}u\right\rvert dx+\int_{Q}\left\lvert\partial_{2}^{s}\partial_{1}u\right\rvert dx. (4.77)

Let w(t):=u(t,x2)w(t):=u(t,x_{2}), with a fixed x2Ix_{2}\in I. Then, by Lemma 4.9, we have

TV1s(w)Cr[wL1(I)+TV11+s(w)].TV_{\ell^{1}}^{s}({w})\leq C_{r}\left[\left\|{w}\right\|_{L^{1}(I)}+TV_{\ell^{1}}^{1+s}({w})\right]. (4.78)

That is,

01|1su(x1,x2)|𝑑x1Cr[01|u(x1,x2)|𝑑x1+01|11+su(x1,x2)|𝑑x1],\int_{0}^{1}\left\lvert\partial_{1}^{s}u(x_{1},x_{2})\right\rvert dx_{1}\leq C_{r}\left[\int_{0}^{1}\left\lvert u(x_{1},x_{2})\right\rvert dx_{1}+\int_{0}^{1}\left\lvert\partial_{1}^{1+s}u(x_{1},x_{2})\right\rvert dx_{1}\right], (4.79)

and hence

Q|1su|𝑑x\displaystyle\int_{Q}\left\lvert\partial_{1}^{s}u\right\rvert dx =0101|1su(x1,x2)|𝑑x1𝑑x2\displaystyle=\int_{0}^{1}\int_{0}^{1}\left\lvert\partial_{1}^{s}u(x_{1},x_{2})\right\rvert dx_{1}dx_{2}
Cr[0101|u(x1,x2)|𝑑x1𝑑x2+0101|11+su(x1,x2)|𝑑x1𝑑x2]\displaystyle\leq C_{r}\left[\int_{0}^{1}\int_{0}^{1}\left\lvert u(x_{1},x_{2})\right\rvert dx_{1}dx_{2}+\int_{0}^{1}\int_{0}^{1}\left\lvert\partial_{1}^{1+s}u(x_{1},x_{2})\right\rvert dx_{1}dx_{2}\right]
=Cr[uL1(Q)+Q|11+su|𝑑x].\displaystyle=C_{r}\left[\left\|u\right\|_{L^{1}(Q)}+\int_{Q}\left\lvert\partial_{1}^{1+s}u\right\rvert dx\right].

We may analogously prove

Q|2su|𝑑xCr[uL1(Q)+Q|21+su|𝑑x],\int_{Q}\left\lvert\partial_{2}^{s}u\right\rvert dx\leq C_{r}\left[\left\|u\right\|_{L^{1}(Q)}+\int_{Q}\left\lvert\partial_{2}^{1+s}u\right\rvert dx\right], (4.80)

and, in view of (4.76) and (4.77),

TV1s(u)Cr[uL1(Q)+TV11+s(u)], for each uC(Q)BVr(Q)TV_{\ell^{1}}^{s}(u)\leq C_{r}\left[\left\|u\right\|_{L^{1}(Q)}+TV_{\ell^{1}}^{1+s}(u)\right],\text{ for each $u\in C^{\infty}(Q)\cap BV^{r}(Q)$. } (4.81)

To conclude, we take an approximating sequence {un}C(Q)BV1+s(Q)\left\{u_{n}\right\}\subset C^{\infty}(Q)\cap BV^{1+s}(Q) such that unuu_{n}\to u strongly in L1(Q)L^{1}(Q) and TV11+s(un)TV11+s(u)TV_{\ell^{1}}^{1+s}(u_{n})\to TV_{\ell^{1}}^{1+s}(u). The former implies that

lim infnTV1s(un)TV1s(u),{\liminf_{n\to\infty}}TV_{\ell^{1}}^{s}(u_{n})\geq TV_{\ell^{1}}^{s}(u), (4.82)

and together with (4.81) we conclude that, for uSV1+s(Q)u\in SV^{1+s}(Q),

TV1s(u)lim infnTV1s(un)lim infn2Cr[unL1(Q)+TV11+s(un)]lim supn2Cr[unL1(Q)+TV11+s(un)]=2Cr[uL1(Q)+TV11+s(u)],TV_{\ell^{1}}^{s}(u)\leq{\liminf_{n\to\infty}}TV_{\ell^{1}}^{s}(u_{n})\leq\liminf_{n\to\infty}2C_{r}\left[\left\|u_{n}\right\|_{L^{1}(Q)}+TV_{\ell^{1}}^{1+s}(u_{n})\right]\\ \leq\limsup_{n\to\infty}2C_{r}\left[\left\|u_{n}\right\|_{L^{1}(Q)}+TV_{\ell^{1}}^{1+s}(u_{n})\right]=2C_{r}\left[\left\|u\right\|_{L^{1}(Q)}+TV_{\ell^{1}}^{1+s}(u)\right], (4.83)

as desired.

Now we assume r=2{\lfloor r\rfloor}=2. Similarly to the case r=1{\lfloor r\rfloor}=1, we observe that

Q|jsu|𝑑xCr[uL1(Q)+Q|j2+su|𝑑x],j=1,2.\int_{Q}\left\lvert\partial_{j}^{s}u\right\rvert dx\leq C_{r}\left[\left\|u\right\|_{L^{1}(Q)}+\int_{Q}\left\lvert\partial_{j}^{2+s}u\right\rvert dx\right],\qquad j=1,2. (4.84)

That is, we conclude that

TV1s(u)2Cr[uL1(Q)+Q|12+su|𝑑x+Q|22+su|𝑑x]\displaystyle TV_{\ell^{1}}^{s}(u)\leq 2C_{r}\left[\left\|u\right\|_{L^{1}(Q)}+\int_{Q}\left\lvert\partial_{1}^{2+s}u\right\rvert dx+\int_{Q}\left\lvert\partial_{2}^{2+s}u\right\rvert dx\right]
2Cr[uL1(Q)+Q|12+su|𝑑x+Q|22+su|𝑑x+Q|11+s2u|𝑑x+Q|21+s1u|𝑑x]\displaystyle\leq 2C_{r}\left[\left\|u\right\|_{L^{1}(Q)}+\int_{Q}\left\lvert\partial_{1}^{2+s}u\right\rvert dx+\int_{Q}\left\lvert\partial_{2}^{2+s}u\right\rvert dx+\int_{Q}\left\lvert\partial_{1}^{1+s}\partial_{2}u\right\rvert dx+\int_{Q}\left\lvert\partial_{2}^{1+s}\partial_{1}u\right\rvert dx\right]
=2Cr[uL1(Q)+TV1r(u)].\displaystyle=2C_{r}\left[\left\|u\right\|_{L^{1}(Q)}+TV_{\ell^{1}}^{r}(u)\right].

For the general case that r{\lfloor r\rfloor}\in{\mathbb{N}}, we shall always have, in view of Lemma 4.9, that

TV1s(u)2Cr[uL1(Q)+Q|1r+su|𝑑x+Q|2r+su|𝑑x],TV_{\ell^{1}}^{s}(u)\leq 2C_{r}\left[\left\|u\right\|_{L^{1}(Q)}+\int_{Q}\left\lvert\partial_{1}^{{\lfloor r\rfloor}+s}u\right\rvert dx+\int_{Q}\left\lvert\partial_{2}^{{\lfloor r\rfloor}+s}u\right\rvert dx\right], (4.85)

and we conclude (4.75) as the right hand side is bounded by uL1(Q)+TV1r(u)\left\|u\right\|_{L^{1}(Q)}+TV_{\ell^{1}}^{r}(u). Finally, we conclude our thesis by invoking again condition (3.15). ∎

We close sub-section 4.2.1 by proving that the constant CrC_{r} from Theorem 4.8 can be taken independently of rr.

Proposition 4.10.

Let r=r+sr=\lfloor r\rfloor+s, s(0,1)s\in(0,1), be given. Then there exists C>0C>0, independent of rr, such that

TVs(u)C[uL1(Q)+TVr(u)]TV^{s}(u)\leq C\left[\left\|u\right\|_{L^{1}(Q)}+TV^{{r}}(u)\right] (4.86)

for all uSVr(Q)u\in SV^{{r}}(Q).

Proof.

We only proof this proposition for case of dimension one, i.e., N=1N=1. The case in which N2N\geq 2 can be obtained from the one-dimensional result, and the arguments from Theorem 4.8.

Let wBVr(I)w\in BV^{{r}}(I) be given. In view of Theorem 4.8 we have, for each r{r}, a constant Cr>0{C_{r}}>0 (depending on rr) such that

l=0r1TVs+l(w)Cr[wL1(I)+TVr(w)]\sum_{l=0}^{{\lfloor r\rfloor}-1}TV^{s+l}(w)\leq C_{r}\left[\left\|w\right\|_{L^{1}(I)}+TV^{{r}}(w)\right] (4.87)

for all wL1(I)w\in L^{1}(I).

We shall only deal with the case r=1{\lfloor r\rfloor}=1, as the case r>1{\lfloor r\rfloor}>1 can be treated analogously. Suppose (4.87) fails, i.e. there exist sequences {wn}L1(I)\left\{w_{n}\right\}\subset L^{1}(I) and {sn}(0,1)\left\{s_{n}\right\}\subset(0,1) such that

TVsn(wn)=1 and wnL1(I)+TV1+sn(wn)<1/n.TV^{s_{n}}(w_{n})=1\qquad\text{ and }\qquad\left\|w_{n}\right\|_{L^{1}(I)}+TV^{1+s_{n}}(w_{n})<1/n. (4.88)

In view of (4.64), we have

|c0,n|+|c1,n|\displaystyle\left\lvert c_{0,n}\right\rvert+\left\lvert c_{1,n}\right\rvert [1Γ(2sn)+1][wnL1(I)+TVsn(wn)+TV1+sn(wn)]\displaystyle\leq\left[\frac{1}{\Gamma(2-s_{n})}+1\right]\left[\left\|w_{n}\right\|_{L^{1}(I)}+TV^{s_{n}}(w_{n})+TV^{1+s_{n}}(w_{n})\right]
1Γ(2sn)+1+1n2+1n.\displaystyle\leq\frac{1}{\Gamma(2-s_{n})}+1+\frac{1}{n}\leq 2+\frac{1}{n}. (4.89)

Hence, there exist c0c_{0} and c1c_{1} such that c0,nc0c_{0,n}\to c_{0} and c1,nc1c_{1,n}\to c_{1}. Then, we may reach the contradiction by using the same arguments from (4.58) to (4.69). ∎

4.2.2. Compact embedding and lower semi-continuity

We start again with a result on the an-isotropic total variation. Recall the image space IM(Q)IM(Q) from (4.1).

Proposition 4.11.

Given sequences {rn}+\left\{r_{n}\right\}\subset{\mathbb{R}}^{+} and {un}L1(Q)\left\{u_{n}\right\}\subset L^{1}(Q) such that rnr+r_{n}\to r\in{\mathbb{R}}^{+} and, for some p>1p>1,

sup{unLp(Q)+TV1rn(un):n}<+,\sup\left\{\left\|u_{n}\right\|_{L^{p}(Q)}+TV_{\ell^{1}}^{r_{n}}(u_{n}):\,\,n\in{\mathbb{N}}\right\}<+\infty, (4.90)

then, the following statement hold.

  1. 1.

    There exists uBVr(Q)u\in BV^{r}(Q), such that, up to a sub-sequence, unuu_{n}\rightharpoonup u in Lp(Q)L^{p}(Q) and

    lim infnTV1rn(un)TV1r(u).\liminf_{n\to\infty}TV_{\ell^{1}}^{r_{n}}(u_{n})\geq TV_{\ell^{1}}^{r}(u). (4.91)
  2. 2.

    Assuming in addition that unIM(Q)SVrn(Q){u_{n}}\subset IM(Q)\cap SV^{r_{n}}(Q) and rnr>0r_{n}\to r>0, then

    unu strongly in L1(Q).u_{n}\to u\text{ strongly in }L^{1}(Q). (4.92)
Proof.

Write rn=rn+snr_{n}=\lfloor r_{n}\rfloor+s_{n} for sn[0,1)s_{n}\in[0,1). By (4.90), there exists uLp(Q)u\in L^{p}(Q) such that, up to a sub-sequence,

unu weakly in Lp(Q).u_{n}\rightharpoonup u\text{ weakly in }L^{p}(Q). (4.93)

Then Statement 1 can be proved by using the same arguments from Proposition 4.7.

We next prove Statement 2. We only study the case of 1rn21\leq r_{n}\leq 2, or equivalently r=1{\lfloor r\rfloor}=1, as the case in which r2{\lfloor r\rfloor}\geq 2 can be dealt analogously. Again by applying Proposition 4.10 we have

supnunBVsn(Q)\displaystyle\sup_{n}\left\|u_{n}\right\|_{BV^{s_{n}}(Q)} CsupnunBVrn(Q)supn{unLp(Q)+TV1rn(un)}=:M<+,\displaystyle\leq C\sup_{n}\left\|u_{n}\right\|_{BV^{r_{n}}(Q)}\leq\sup_{n}\{\left\|u_{n}\right\|_{L^{p}(Q)}+TV_{\ell^{1}}^{r_{n}}(u_{n})\}=:M<+\infty, (4.94)

where C>0C>0, obtained from Proposition 4.10, is a constant independent of rnr_{n}. Assume in addition that there exists ε>0\varepsilon>0 such that {sn}[ε,1]\left\{s_{n}\right\}\subset[\varepsilon,1]. Then, in view of Proposition 4.7, there exists uL1(Q)u\in L^{1}(Q) such that, up to a sub-sequence, that unuu_{n}\to u strongly in L1(Q)L^{1}(Q), which gives Statement 2.

We now deal with the situation that sn0s_{n}\searrow 0, that is, rn1r_{n}\searrow 1. In this case, although (4.94) still holds, Proposition 4.7 does not produce a sub-sequence strongly converging in L1(Q)L^{1}(Q). We proceed by using Theorem 3.14 to relax unu_{n}, for each nn\in{\mathbb{N}}, such that unBV1+sn(Q)C(Q)u_{n}\in BV^{1+s_{n}}(Q)\cap C^{\infty}(Q). Hence, we have, for arbitrary φCc(Q,2)\varphi\in C_{c}^{\infty}(Q,{\mathbb{R}}^{2}), that (recall (3.10))

Qsnundivφdx=Qun[divsndiv]φ𝑑x=Qundiv1+snφdxTV11+sn(un),-\int_{Q}\nabla^{s_{n}}u_{n}\,{\operatorname{div}}\varphi\,dx=\int_{Q}u_{n}[{\operatorname{div}}^{s_{n}}{\operatorname{div}}]\varphi\,dx=\int_{Q}u_{n}{\operatorname{div}}^{1+s_{n}}\varphi\,dx\leq TV_{\ell^{1}}^{1+s_{n}}(u_{n}), (4.95)

which implies

TV1(snun)TV11+sn(un)M<+.TV_{\ell^{1}}(\nabla^{s_{n}}u_{n})\leq TV_{\ell^{1}}^{1+s_{n}}(u_{n})\leq M<+\infty. (4.96)

Moreover, by Proposition 4.10,

snunL1(Q)=TV1sn(un)CunBV1+sn(Q)M<+.\left\|\nabla^{s_{n}}u_{n}\right\|_{L^{1}(Q)}=TV_{\ell^{1}}^{s_{n}}(u_{n})\leq C\left\|u_{n}\right\|_{BV^{1+s_{n}}(Q)}\leq M<+\infty. (4.97)

We claim

lim supnunsnunL1(Q)=0.\limsup_{n\to\infty}\left\|u_{n}-\nabla^{s_{n}}u_{n}\right\|_{L^{1}(Q)}=0. (4.98)

We start from the one-dimensional case, i.e. Q=I=(0,1)Q=I=(0,1), and use wnw_{n} to represent unu_{n}. By Theorem 3.10, for each nn\in{\mathbb{N}}, there exists ϕn(t)L1(I)\phi_{n}(t)\in L^{1}(I) such that

wn(t)=c0,nΓ(sn)ts1+c1,nΓ(sn+1)tsn+𝕀1+snϕn(t)w_{n}(t)=\frac{c_{0,n}}{\Gamma(s_{n})}t^{s-1}+\frac{c_{1,n}}{\Gamma(s_{n}+1)}t^{s_{n}}+\mathbb{I}^{1+s_{n}}\phi_{n}(t) (4.99)

for a.e. tt, and

dsnwn(t)=c1,n+𝕀1ϕn(t).d^{s_{n}}w_{n}(t)=c_{1,n}+\mathbb{I}^{1}\phi_{n}(t). (4.100)

Next, in view of (2.3), we have, for each nn\in{\mathbb{N}} fixed, that

𝕀1+sϕn(t)=1Γ(1+s)0tϕn(z)(tz)1(1+s)𝑑z=1Γ(1+s)0tϕn(z)(tz)s𝑑z.\mathbb{I}^{1+s}\phi_{n}(t)=\frac{1}{\Gamma(1+s)}\int_{0}^{t}\frac{\phi_{n}(z)}{\left(t-z\right)^{1-(1+s)}}dz=\frac{1}{\Gamma(1+s)}\int_{0}^{t}{\phi_{n}(z)}\left(t-z\right)^{s}dz. (4.101)

That is, we have

limt0|𝕀1+sϕn(t)|<+, for each n.\lim_{t\to 0}\left\lvert\mathbb{I}^{1+s}\phi_{n}(t)\right\rvert<+\infty\text{, for each }n\in{\mathbb{N}}. (4.102)

However, ts1+t^{s-1}\to+\infty as t0+t\to 0^{+}. Hence, c0,n=0c_{0,n}=0 must hold, as the opposite gives wn(t)+w_{n}(t)\to+\infty as t0t\to 0, which contradicts our assumption wnIM(Q)w_{n}\in IM(Q). Thus, we have

wn(t)=c1,nΓ(s+1)ts+𝕀1+sϕn(t),for a.e. t.w_{n}(t)=\frac{c_{1,n}}{\Gamma(s+1)}t^{s}+\mathbb{I}^{1+s}\phi_{n}(t),\qquad\text{for a.e. }t. (4.103)

Then, direct computation gives

wndsnwnL1(Q)=c1,nΓ(sn+1)tsnc1,n+𝕀1+snϕn𝕀1ϕnL1(Q)c1,nΓ(sn+1)tsnc1,nL1(Q)+𝕀1+snϕn𝕀1ϕnL1(Q)|c1,n||1Γ(sn+1)(sn+1)1|+|𝕀1+sn𝕀1|ϕnL1(Q).\displaystyle\begin{split}\left\|w_{n}-d^{s_{n}}w_{n}\right\|_{L^{1}(Q)}&=\left\|\frac{c_{1,n}}{\Gamma(s_{n}+1)}t^{s_{n}}-c_{1,n}+\mathbb{I}^{1+s_{n}}\phi_{n}-\mathbb{I}^{1}\phi_{n}\right\|_{L^{1}(Q)}\\ &\leq\left\|\frac{c_{1,n}}{\Gamma(s_{n}+1)}t^{s_{n}}-c_{1,n}\right\|_{L^{1}(Q)}+\left\|\mathbb{I}^{1+s_{n}}\phi_{n}-\mathbb{I}^{1}\phi_{n}\right\|_{L^{1}(Q)}\\ &\leq\left\lvert c_{1,n}\right\rvert\left\lvert\frac{1}{\Gamma(s_{n}+1)(s_{n}+1)}-1\right\rvert+\left\lvert\mathbb{I}^{1+s_{n}}-\mathbb{I}^{1}\right\rvert\left\|\phi_{n}\right\|_{L^{1}(Q)}.\end{split} (4.104)

Moreover, using the same argument from the proof of (4.64), we have

|c1,n|+ϕnL1(Q)2[dsnwnL1(Q)+ds+n+1wnL1(Q)].\left\lvert c_{1,n}\right\rvert+\left\|\phi_{n}\right\|_{L^{1}(Q)}\leq 2\left[\left\|d^{s_{n}}w_{n}\right\|_{L^{1}(Q)}+\left\|d^{s+n+1}w_{n}\right\|_{L^{1}(Q)}\right]. (4.105)

This, combined with (4.104), implies that

unsnunL1(Q)=01unx1dsnunx1L1(Q)dx1\displaystyle\|u_{n}-\nabla^{s_{n}}u_{n}\|_{L^{1}(Q)}=\int_{0}^{1}\left\|u_{n}\lfloor_{x_{1}}-d^{s_{n}}u_{n}\lfloor_{x_{1}}\right\|_{L^{1}(Q)}dx_{1}
2[|1Γ(sn+1)(sn+1)1|+|𝕀1+sn𝕀1|]01[dsnunx1L1(Q)+ds+n+1unx1L1(Q)]dx1\displaystyle\leq 2\left[\left\lvert\frac{1}{\Gamma(s_{n}+1)(s_{n}+1)}-1\right\rvert+\left\lvert\mathbb{I}^{1+s_{n}}-\mathbb{I}^{1}\right\rvert\right]\int_{0}^{1}\left[\left\|d^{s_{n}}u_{n}\lfloor_{x_{1}}\right\|_{{L^{1}(Q)}}+\left\|d^{s+n+1}u_{n}\lfloor_{x_{1}}\right\|_{{L^{1}(Q)}}\right]dx_{1}
2[|1Γ(sn+1)(sn+1)1|+|𝕀1+sn𝕀1|]unBVrn(Q).\displaystyle\leq 2\left[\left\lvert\frac{1}{\Gamma(s_{n}+1)(s_{n}+1)}-1\right\rvert+\left\lvert\mathbb{I}^{1+s_{n}}-\mathbb{I}^{1}\right\rvert\right]\left\|u_{n}\right\|_{BV^{r_{n}}(Q)}.

By Theorem 2.5, Assertion 3, we have |𝕀sn+1𝕀1|0\left\lvert\mathbb{I}^{s_{n}+1}-\mathbb{I}^{1}\right\rvert\to 0, as sn0s_{n}\to 0, and hence we conclude (4.98). Next, by (4.96) and (4.97), and the compact embedding in standard BV(Q)BV(Q) space, up to a sub-sequence, there exists u¯BV(Q)\bar{u}\in BV(Q), that

snunu¯ strongly in L1(Q).\nabla^{s_{n}}u_{n}\to\bar{u}\text{ strongly in }L^{1}(Q). (4.106)

Hence, by (4.98), we have

unu¯L1(Q)unsnunL1(Q)+snunu¯L1(Q)0,\left\|u_{n}-\bar{u}\right\|_{L^{1}(Q)}\leq\left\|u_{n}-\nabla^{s_{n}}u_{n}\right\|_{L^{1}(Q)}+\left\|\nabla^{s_{n}}u_{n}-\bar{u}\right\|_{L^{1}(Q)}\to 0, (4.107)

that is, we have unu¯u_{n}\to\bar{u} strongly in L1(Q)L^{1}(Q). Finally, in view of (4.93), we have u=u¯u=\bar{u}, and hence the thesis. ∎

We conclude this section by proving the second main result of this article.

Theorem 4.12 (lower semi-continuity and compact embedding in TVprTV^{r}_{\ell^{p}} semi-norms).

Given sequences {rn}+\left\{r_{n}\right\}\subset{\mathbb{R}}^{+}, {pn}[1,+]\left\{p_{n}\right\}\subset[1,+\infty], and {un}L1(Q)\left\{u_{n}\right\}\subset L^{1}(Q) such that rnr+{0}r_{n}\to r\in{\mathbb{R}}^{+}\cup\left\{0\right\} and pnp[1,+]p_{n}\to p\in[1,+\infty], and there exists q(1,+]q\in(1,+\infty] such that

sup{unLq(Q)+TVpnrn(un):n}<+,\sup\left\{\left\|u_{n}\right\|_{L^{q}(Q)}+TV_{\ell^{p_{n}}}^{r_{n}}(u_{n}):\,\,n\in{\mathbb{N}}\right\}<+\infty, (4.108)

then, the following statements hold.

  1. 1.

    There exists uBVr(Q)u\in BV^{r}(Q) such that, up to a sub-sequence, unuu_{n}\rightharpoonup u weakly in Lq(Q)L^{q}(Q) and

    lim infnTVpnrn(un)TVpr(u).\liminf_{n\to\infty}TV_{\ell^{p_{n}}}^{r_{n}}(u_{n})\geq TV_{\ell^{p}}^{r}(u). (4.109)
  2. 2.

    Assuming in addition that unIM(Q)SVrn(Q){u_{n}}\subset IM(Q)\cap SV^{r_{n}}(Q) and rnr>0r_{n}\to r>0, we have

    unu strongly in L1(Q).u_{n}\to u\text{ strongly in }L^{1}(Q). (4.110)
Proof.

By applying Remark 3.7, we deduce that (4.108) implies (4.90), and the thesis follows by combining (4.45) and Theorem 3.12. ∎

Acknowledgements

Xin Yang Lu acknowledges the support of NSERC Grant “Regularity of minimizers and pattern formation in geometric minimization problems”, and of the Startup funding, and Research Development Funding of Lakehead University. The authors are grateful to Todd J. Falkenholt, and the NSERC USRA grant supporting him, for many useful comments and suggestions.

References

  • [1] O. P. Agrawal, Fractional variational calculus in terms of riesz fractional derivatives, Journal of Physics A: Mathematical and Theoretical, 40 (2007), p. 6287.
  • [2] H. Brezis, Functional analysis, Sobolev spaces and partial differential equations, Universitext, Springer, New York, 2011.
  • [3] D. Chen, S. Sun, C. Zhang, Y. Chen, and D. Xue, Fractional-order TVTV-L2L^{2} model for image denoising, Central European Journal of Physics, 11 (2013), pp. 1414–1422.
  • [4] L. C. Evans and R. F. Gariepy, Measure theory and fine properties of functions, CRC press, 2015.
  • [5] I. J. Goodfellow, J. Pouget-Abadie, M. Mirza, B. Xu, D. Warde-Farley, S. Ozair, A. C. Courville, and Y. Bengio, Generative adversarial networks, CoRR, abs/1406.2661 (2014).
  • [6] K. He, X. Zhang, S. Ren, and J. Sun, Deep residual learning for image recognition, in 2016 IEEE Conference on Computer Vision and Pattern Recognition, CVPR 2016, Las Vegas, NV, USA, June 27-30, 2016, IEEE Computer Society, 2016, pp. 770–778.
  • [7] G. Huang, Z. Liu, L. van der Maaten, and K. Q. Weinberger, Densely connected convolutional networks, in 2017 IEEE Conference on Computer Vision and Pattern Recognition, CVPR 2017, Honolulu, HI, USA, July 21-26, 2017, IEEE Computer Society, 2017, pp. 2261–2269.
  • [8] D. Idczak and S. a. Walczak, Fractional Sobolev spaces via Riemann-Liouville derivatives, J. Funct. Spaces Appl., (2013), pp. Art. ID 128043, 15.
  • [9] G. Jumarie, Modified Riemann-Liouville derivative and fractional Taylor series of nondifferentiable functions further results, Comput. Math. Appl., 51 (2006), pp. 1367–1376.
  • [10] A. Krizhevsky, I. Sutskever, and G. E. Hinton, Imagenet classification with deep convolutional neural networks, in Advances in Neural Information Processing Systems 25: 26th Annual Conference on Neural Information Processing Systems 2012. Proceedings of a meeting held December 3-6, 2012, Lake Tahoe, Nevada, United States, P. L. Bartlett, F. C. N. Pereira, C. J. C. Burges, L. Bottou, and K. Q. Weinberger, eds., 2012, pp. 1106–1114.
  • [11] H. Liu, K. Simonyan, and Y. Yang, DARTS: differentiable architecture search, CoRR, abs/1806.09055 (2018).
  • [12] P. Liu, C. Li, and C.-B. Schönlieb, GANReDL: Medical image enhancement using a generative adversarial network with real-order derivative induced loss functions, in Medical Image Computing and Computer Assisted Intervention – MICCAI 2019, D. Shen, T. Liu, T. M. Peters, L. H. Staib, C. Essert, S. Zhou, P.-T. Yap, and A. Khan, eds., Cham, 2019, Springer International Publishing, pp. 110–117.
  • [13] L. I. Rudin, S. Osher, and E. Fatemi, Nonlinear total variation based noise removal algorithms, Phys. D, 60 (1992), pp. 259–268. Experimental mathematics: computational issues in nonlinear science (Los Alamos, NM, 1991).
  • [14] S. G. Samko, A. A. Kilbas, and O. I. Marichev, Fractional integrals and derivatives, Gordon and Breach Science Publishers, Yverdon, 1993. Theory and applications, Edited and with a foreword by S. M. Nikolskiĭ, Translated from the 1987 Russian original, Revised by the authors.
  • [15] K. Simonyan and A. Zisserman, Very deep convolutional networks for large-scale image recognition, in 3rd International Conference on Learning Representations, ICLR 2015, San Diego, CA, USA, May 7-9, 2015, Conference Track Proceedings, Y. Bengio and Y. LeCun, eds., 2015.
  • [16] C. Szegedy, W. Liu, Y. Jia, P. Sermanet, S. E. Reed, D. Anguelov, D. Erhan, V. Vanhoucke, and A. Rabinovich, Going deeper with convolutions, in IEEE Conference on Computer Vision and Pattern Recognition, CVPR 2015, Boston, MA, USA, June 7-12, 2015, IEEE Computer Society, 2015, pp. 1–9.
  • [17] A. Vaswani, N. Shazeer, N. Parmar, J. Uszkoreit, L. Jones, A. N. Gomez, L. Kaiser, and I. Polosukhin, Attention is all you need, in Advances in Neural Information Processing Systems 30: Annual Conference on Neural Information Processing Systems 2017, December 4-9, 2017, Long Beach, CA, USA, I. Guyon, U. von Luxburg, S. Bengio, H. M. Wallach, R. Fergus, S. V. N. Vishwanathan, and R. Garnett, eds., 2017, pp. 5998–6008.
  • [18] J. Zhang and K. Chen, A total fractional-order variation model for image restoration with nonhomogeneous boundary conditions and its numerical solution, SIAM J. Imaging Sci., 8 (2015), pp. 2487–2518.
  • [19] B. Zoph and Q. V. Le, Neural architecture search with reinforcement learning, in 5th International Conference on Learning Representations, ICLR 2017, Toulon, France, April 24-26, 2017, Conference Track Proceedings, OpenReview.net, 2017.