• No results found

Analysis of finite element methods for vector Laplacians on surfaces

N/A
N/A
Protected

Academic year: 2022

Share "Analysis of finite element methods for vector Laplacians on surfaces"

Copied!
51
0
0

Loading.... (view fulltext now)

Full text

(1)

This is the published version of a paper published in IMA Journal of Numerical Analysis.

Citation for the original published paper (version of record):

Hansbo, P., Larson, M G., Larsson, K. (2020)

Analysis of finite element methods for vector Laplacians on surfaces IMA Journal of Numerical Analysis, 40(3): 1652-1701

https://doi.org/10.1093/imanum/drz018

Access to the published version may require subscription.

N.B. When citing this work, cite the original published paper.

Permanent link to this version:

http://urn.kb.se/resolve?urn=urn:nbn:se:umu:diva-174232

(2)

doi:10.1093/imanum/drz018

Advance Access publication on 25 April 2019

Analysis of finite element methods for vector Laplacians on surfaces

Peter Hansbo

Department of Mechanical Engineering, Jönköping University, SE-55111 Jönköping, Sweden peter.hansbo@ju.se

and

Mats G. Larson and Karl Larsson

Department of Mathematics and Mathematical Statistics, Umeå University, SE-90187 Umeå, Sweden mats.larson@umu.se Corresponding author: karl.larsson@umu.se

[Received on 21 October 2016; revised on 8 November 2018]

We develop a finite element method for the vector Laplacian based on the covariant derivative of tangential vector fields on surfaces embedded inR3. Closely related operators arise in models of flow on surfaces as well as elastic membranes and shells. The method is based on standard continuous parametric Lagrange elements that describe aR3vector field on the surface, and the tangent condition is weakly enforced using a penalization term. We derive error estimates that take into account the approximation of both the geometry of the surface and the solution to the partial differential equation. In particular, we note that to achieve optimal order error estimates, in both energy and L2norms, the normal approximation used in the penalization term must be of the same order as the approximation of the solution. This can be fulfilled either by using an improved normal in the penalization term, or by increasing the order of the geometry approximation. We also present numerical results using higher-order finite elements that verify our theoretical findings.

Keywords: vector Laplacian on surfaces; higher-order finite element method; a priori error estimates.

1. Introduction

In this contribution we develop a finite element method for the vector Laplacian on a surface. While there are several natural Laplacians acting on vector fields on surfaces, we in this work consider the rough Laplacian, a second-order elliptic operator based on covariant derivatives. In contrast, another natural Laplacian is the Hodge Laplacian that is based on exterior calculus, seeHolst & Stern (2012), and which differs from the rough Laplacian by a zeroth-order term depending only on the curvature of the surface.

The method is based on continuous parametric Lagrange elements with geometry and solution approximations, which are piecewise polynomial of orders kgand ku, respectively. Instead of defining an approximation space for tangent vector fields on the surface Γ , we seek solutions that are full vector fields Γ → R3and weakly enforce the tangential condition using a suitable penalty term, similar to our work on the Darcy problem, seeHansbo & Larson (2017). Note, however, that the Darcy problem does not involve any gradients of the velocity vector and is therefore easier to deal with. This approach leads to a convenient implementation without the need for special finite element spaces.

We prove a priori error estimates in the energy and L2norm, and we find that in order to obtain optimal order convergence in both norms it is necessary to use a discrete normal in the penalty term of order ku+ 1. For isoparametric finite elements this translates into a geometry approximation of

© The Author(s) 2019. Published by Oxford University Press.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.

0/), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited.

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(3)

the normal in the penalty term that is one degree higher than of the normal to the discrete surface Γh. Somewhat curiously, there is no loss of order in L2due to the fact that the covariant derivative is obtained by projecting the componentwise directional derivative onto the tangent plane, and that the approximation order of the projection is only hkg. To prove this, however, requires the use of non- standard techniques, which we developed inLarsson & Larson (2017).

Related work. Finite elements for partial differential equations on surfaces is now a rapidly developing field that originates from the seminal work of Dziuk (1988), where surface finite elements for the Laplace–Beltrami operator was first developed. Most of the research is, however, focused on problems with scalar unknowns, see the recent review articleDziuk & Elliott (2013)and the references therein, which simplifies the differential calculus since the covariant derivative of a vector field, or more generally a tensor field, is not needed. Models of flow on surfaces as well as membranes and shells, however, involve vector unknowns, see for instance Hansbo & Larson (2014) (linear) and Hansbo et al. (2015)(nonlinear), for membrane models formulated using the same approach as used in this paper. Furthermore, we employ higher-order elements similar to the approaches presented in Nédélec (1976),Demlow (2009),Larsson & Larson (2017), andHansbo & Larson (2017). Concurrent to the present work, similar formulations for vector Laplace operators on surfaces, also using tangential differential calculus, were studied inJankuhn et al. (2018)motivated by their use in methods posed in an embedding space, and later such a method (TraceFEM) for a vector Laplacian problem was presented inGroß et al. (2018). As in the present work the formulation inGroß et al. (2018)assumes a full vector field on the surface, but instead of using a penalty term to enforce the field to be tangential a Lagrange multiplier approach is used. In addition our analysis includes the geometry approximation.

Paper outline. The remainder of this paper is organized as follows: in Section2 we introduce the vector Laplacian and results concerning the continuous problem; in Section3we introduce the finite element method; in Section4we recall some basic results regarding lifting and extension of functions between the discrete and continuous surfaces, present a nonstandard geometry approximation estimate and introduce the interpolant; in Section5we derive a sequence of necessary lemmas leading up to the a priori error estimate; and finally in Section6we present numerical examples confirming our theoretical findings.

2. Vector Laplacians on a surface

In this section we present the tools we need to work with vector Laplacians on surfaces in the setting of tangential differential calculus, which allows us to employ the Cartesian coordinates of the embedding R3space. In Section2.1we first define the surface and its assumptions; in Sections2.2–2.4we introduce the notations needed to describe tensor fields on the surface and derivatives, and covariant derivatives of such fields; in Section2.5we present the suitable Sobolev spaces. As these first five sections involve numerous definitions we, for clarity and compactness, present them in the form of bullet lists. In Section2.6we establish some lemmas fundamental to the analysis on surfaces, in particular a Poincaré inequality. Finally, in Section2.7we introduce our model variational problem, which involves certain vector Laplacians on a surface.

2.1 The surface

• Let Γ be a smooth compact surface embedded in R3without boundary and let ρ be the signed distance function, negative on the inside and positive on the outside. The exterior unit normal to the surface Γ is given by n= ∇ρ.

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(4)

• Let p : R3→ Γ be the closest point mapping onto Γ . Then there is a δ0 >0 such that p maps each point in Uδ

0(Γ )to precisely one point on Γ , where Uδ(Γ ) = {x ∈ R3:|ρ(x)| < δ} is the open tubular neighbourhood of Γ of thickness δ > 0.

• As ρ is a signed distance function within Uδ0(Γ ), the unit normal to Γ naturally extend to Uδ0(Γ ) through its original definition n(x)= ∇ρ.

• For each function u : Γ → Rm, m= 1, 2, . . . , we define the componentwise extension ueto the neighbourhood Uδ0(Γ )by the pull back ue= u ◦ p.

• The curvature tensor (or second fundamental form) is defined on Uδ0(Γ )by

κ = ∇ ⊗ ∇ρ (2.1)

and may be expressed in the form

κ(x)=

2 i=1

κie

1+ ρ(x)κieaei ⊗ aei, (2.2)

where κiare the principal curvatures with corresponding orthonormal principal curvature vectors ai, seeGilbarg & Trudinger (2001, Lemma 14.17, p. 355).

2.2 Tensors

• Let V, W be finite-dimensional vector spaces with bases {ei}mi=1, respectively{fi}ni=1. The tensor product V⊗ W is the vector space spanned by all pairs (ei, fj)of basis vectors, denoted by ei⊗ fj, and there is a bilinear product⊗ : V × W → V ⊗ W defined by

v⊗ w =

 m



i=1

viei



⎝n

j=1

wjfj

⎠ =m

i=1

n j=1

viwj(ei⊗ fj). (2.3)

The dimension of V⊗ W is dim(V ⊗ W) = dim(V)dim(W).

• If V and W are inner product spaces, V ⊗ W is an inner product space with product

(a⊗ b, v ⊗ w)V⊗W = (a, v)V(b, w)W (2.4)

and the inner product norm is given by

v ⊗ wV⊗W= vVwW. (2.5)

• The dual space of V denoted by Vis the space of all linear functionals λ : V → R. The dual basisj}nj=1is defined by the identity λj(ei)= δij. When V is an inner product space, there is for each λ∈ Va unique vector ξλsuch that ξλ(v)= (v, ξλ)V,∀v ∈ V.

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(5)

• Tensors of type (k, l) are elements in the tensor product space Wk,l= V ⊗ · · · ⊗ V

k copies

⊗ V ⊗ · · · ⊗ V

l copies

. (2.6)

• If {ei}mi=1is an orthonormal basis in V, then{ei}mi=1is also the corresponding dual basis in V. If Q : V→ V is an orthogonal mapping, ei= Qeiis also an orthonormal basis in V, and hence also the corresponding dual basis in V.

• For v in V let [v] denote the array of coefficients in the expansion v =

iviei. If v=

iviei=



i vi ei = 

i viQei we find that vj = 

i vi(Qei, ej)V = 

i viQji, j = 1, . . . , m, and thus in matrix form [v] = Q[ v] or [ v] = Q−1[v]= QT[v]. The same transformation rules hold for the dual space Vand thus we do not have to distinguish between V and V, and we can restrict our attention to tensors of type k+ l of the form

Wk+l= V ⊗ · · · ⊗ V

k+l copies

. (2.7)

• Let v ∈ Vkand w∈ Vl, for n= 1, . . . , min(k,l) we define the n-contraction v ·nw∈ Vk+l−2nby

⊗ki=1vi

·n

lj=1wj

= Πin=1(vk−n+i, wi)V

ki=1−nvi

⊗

lj−n=1wj

. (2.8)

Special cases include n= l = 1 or n = k = 1 and k = l = 2 where we use the simplified notation

v· w ∈ Vk−1, v : w∈ R. (2.9)

2.3 Tensor fields Vector fields.

• Let {ei∈ R3}3i=1be a Cartesian basis, i.e., a fixed orthonormal basis, in the embedding spaceR3.

• For x ∈ Uδ0(Γ )let P(x)= I − n(x) ⊗ n(x) be the projection onto the tangential plane Tx(Γ )and Q= I − P the projection onto the normal line.

• The projected Cartesian basis {pi = Pei: Γ → Tx(Γ )}3i=1spans the tangential plane Tx(Γ ), but is not a basis for Tx(Γ )since the vectors in the set are linearly dependent. Note however that for b∈ Tx(Γ )we have the unique expansion b=3

i=1biei, which induces the canonical expansion b=3

i=1biPei =3

i=1bipi. Furthermore, inner products and norms are clearly independent of the choice of expansion in the projected basis.

• Define (a) the space of general smooth vector fields

T 1=

 a=

3 i=1

aiei: ai∈ C(Γ )



; (2.10)

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(6)

(b) the space of tangential smooth vector fields

Ttan1 =

 a=

3 i=1

aipi: ai∈ C(Γ )



. (2.11)

Tensor fields.

• Define (a) the vector space of smooth m tensor fields on Γ ,

Tm=

⎧⎨

X=

3 i1,...,im=1

Xi1,...,imei1⊗ · · · ⊗ eim, Xi1,...,im ∈ C,R)

⎫⎬

⎭; (2.12)

(b) the vector space of smooth tangential m tensor fields on Γ ,

Ttanm =

⎧⎨

X=

3 i1,...,im=1

Xi

1,...,impi

1⊗ · · · ⊗ pim, Xi

1,...,im ∈ C,R)

⎫⎬

⎭. (2.13)

• The projection P : Tm→ Ttanm is defined by

P

⎝ 3

i1,...,im=1

Xi1,...,imei1 ⊗ · · · ⊗ eim

⎠ = 3

i1,...,im=1

Xi1,...,impi1⊗ · · · ⊗ pim. (2.14)

2.4 Tangential calculus Tangential derivatives.

• The directional derivative of u ∈ T1, in the direction of a∈ T1, is defined by

au= (a · ∇)ue= (ue⊗ ∇) · a, (2.15)

where a· ∇ =3

i=1aiiand ue⊗ ∇ is the Jacobian of ue.

• Define the tangential gradient operator ∇Γ =n

j=1ejp

j and the total derivative of a vector field

u⊗ ∇Γ =

3 j=1

(∂p

ju)⊗ ej=

3 i,j=1

(∂p

jui)ei⊗ ej= (ue⊗ ∇)P. (2.16)

We note that

au= (u ⊗ ∇Γ)· a ∀a ∈ Ttan1. (2.17)

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(7)

• More generally, for X ∈ Tmwe define in the same way the directional derivative

aX=

3 i1,...,im=1

(∂aXi1,...,im)ei

1⊗ · · · ⊗ eim (2.18)

and the total derivative X⊗ ∇Γ ∈ Tm+1,

X⊗ ∇Γ =

3 j=1

3 i1,...,im=1

(∂p

jXi1,...,im)ei

1⊗ · · · ⊗ eim⊗ ej (2.19)

and we note that

aX= (X ⊗ ∇Γ)· a ∀a ∈ Ttanm (2.20) since a= aiei = aipiand ai= ei· a.

• Higher-order derivatives of X ∈ Tmare obtained by repeated application of (2.19), (Γ)kX= X ⊗ ∇Γk = X ⊗ ∇ Γ ⊗ · · · ⊗ ∇ Γ

k gradients

, (2.21)

which gives (∇Γ)kX∈ Tm+kof the form

(Γ)kX=

3 j1,...,jk=1

3 i1,...,im=1

(∂p

jk. . . ∂p

j1Xi1,...,im)(ei

1⊗ · · · ⊗ eim)⊗ (ej1⊗ · · · ⊗ ejk). (2.22)

Covariant derivatives.

• For u ∈ Ttan1 we define the covariant derivative of u in the direction a by

Dau= P∂au. (2.23)

• Writing u =3

i=1uipiwe have using the product rule

Dau= P∂au= P∂a

3 i=1

uipi=

3 i=1

(∂aui)pi+ uiP(∂api). (2.24)

We note that the covariant derivative includes a lower-order term multiplied by a projected directional derivative of a tangent basis vector pi. Writing a = 3

j=1ajpj we have ∂api =

3

j=1ajp

jpiand, using the identity pi= ei− nin, we find that

p

jpi= ∂pj(ei− nin)= −pj· κin+ niκ· pj= −κijn− niκj, (2.25)

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(8)

where κ = n ⊗ ∇ = ∇2ρ is the tangential curvature tensor, see (2.1), with elements κij and columns (and rows) κj. Thus, P(∂p

jpi)= −niκjand expanding the right-hand side in the Cartesian basis, we obtain

P(∂p

jpi)=

3 k=1

γij,kpk, (2.26)

where the coefficients γij,k = −niκjk correspond to the Christoffel symbols of the Levi–Civita connection.

• Furthermore, note that in the case of the canonical expansion u = 3

i=1uipi = 3

i=1uiei we have the simplified identity

Dau= P∂au= P

 3



i=1

(∂aui)ei



=

3 i=1

(∂aui)pi. (2.27)

Using the fact3

i=1uini= 0, we also note that the second term in the right-hand side of (2.24) is indeed zero since

3 i=1

ui

3 j=1

ajP(∂pjpi)= −

3 i=1

ui

3 j=1

ajniκj= −

 3



i=1

uini

 ⎛

⎝3

j=1

ajκj

⎠ = 0. (2.28)

• Define the total covariant derivative DΓu∈ Ttan2,

DΓu=

3 i=1

(Dp

iu)⊗ pi= P(u ⊗ ∇Γ)= P(ue⊗ ∇)P (2.29) and we note that

Dau= (DΓu)· a ∀a ∈ Ttan1. (2.30) In contrast to u⊗ ∇Γ, DΓu is a tangential tensor.

• The symmetric part of DΓu is defined by

Γ(u)=1 2

DΓu+ (DΓu)T

, (2.31)

which is the tangential strain tensor used in modelling of solids and fluids, seeHansbo & Larson (2014).

• The covariant derivative DaX ∈ Ttanm of a tangential tensor X ∈ Ttanm in the direction a ∈ Ttan1 is defined by

DaX= P(∂aX) (2.32)

=

3 i1,...,im=1

(∂aXi

1···im)(pi

1⊗ · · · ⊗ pim)+ Xi1···imP

a(pi

1⊗ · · · ⊗ pim)

, (2.33)

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(9)

where the projectionP of a tensor field is defined in (2.14). We use the product rule ∂a(Y⊗ Z) = (∂aY)⊗ Z + Y ⊗ (∂aZ), X ∈ Tm, Y ∈ Tn, to compute the second term. The total covariant derivative DΓX∈ Ttanm+1, is defined by

DΓX=

3 j=1

(Dp

jX)⊗ pj (2.34)

and note that, since pj· a = Pej· a = ej· Pa = ej· a = ajwe have

DaX= (DΓX)· a (2.35)

for all tangential vector fields a.

• Iterating this definition we can represent covariant derivatives of order m as (DΓ)mu= D Γ · · · DΓ

m covariant derivatives

u. (2.36)

2.5 Function spaces

For ω ⊂ Γ let (·, ·)ω and · L2(ω)denote the usual L2inner product and norm on ω and let · L(ω)

denote the usual Lnorm on ω. We define the following Sobolev spaces:

• Hs(ω), with ω ⊂ Γ , denotes the standard Sobolev spaces of scalar- or vector-valued functions with componentwise derivatives and norm

v2Hs(ω)=

s j=0

(∇Γ)jv2L2(ω). (2.37)

• Htans (ω), with ω ⊂ Γ , denotes the Sobolev space of tangential vector fields with covariant derivatives and norm

v2Htans (ω)=

s j=0

(DΓ)jv2L2(ω). (2.38)

We employ the standard notation L2(ω)= H0(ω)andvL2(ω)= vω. 2.6 Basic lemmas

We here prove three fundamental lemmas. In Lemma2.1 we show that the kernel of the covariant derivative of a tangential vector field is empty, a fact then used in the proof of Lemma2.2, which is a Poincaré inequality. Finally, in Lemma2.3 we show that Sobolev norms based on tangential, respectively, covariant derivatives are equivalent.

Lemma 2.1 If v∈ H1tan(Γ )satisfies DΓv= 0 then v = 0.

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(10)

Proof. Step 1. Claim: if v ∈ Ttan1 is a smooth tangential vector field that is covariantly constant, DΓv= 0, there is a point x ∈ Γ such that v(x) = 0.

To verify this claim we introduce the Riemannian curvature tensor, seedo Carmo(1992, Def. 2.1, p. 89), which is the mapping R :Ttan1 × Ttan1 × Ttan1 → Ttan1 defined by

R(a, b, v)= DaDbv− DaDbv− D[a,b]v, (2.39) where Dav= P∂av is the covariant derivative in the direction of the tangential vector field a and [a, b]

is the tangent vector field given by the Lie bracket

[a, b]= ∂ba− ∂ab, (2.40)

where we recall that ∂ba= (a ⊗ ∇Γ)· b, see (2.17). To see that the Lie bracket is indeed a tangential vector field we note that, since n·a = 0, we have 0 = ∂b(n·a) = (∂bn)·a+n·(∂ba)= b·κ ·a+n·(∂ba) and thus n· (∂ba)= −b · κ · a, from which it follows that n · [a, b] = 0.

All derivatives in (2.39) cancel so that R(a, b, v) is a tangential vector field that does not depend on any derivatives of v. In the case of an embedded codimension one surface inR3, we have the identity

R(a, b, v)= (b · κ · v)κ · a − (a · κ · v)κ · b, (2.41) where κ is the curvature tensor of Γ , and we note in particular that there are no derivatives of v. To verify (2.41) we first recall the directional and covariant derivatives introduced in Section2.4, i.e.,

av= (v ⊗ ∇Γ)· a, Dav= P∂av (2.42) for a tangential vector field a. We then have

DaDbv= P∂aP∂bv (2.43)

= P∂a(∂bv− (n · ∂bv)n) (2.44)

= P∂abv− P(∂a(n· ∂bv)n+ (n · ∂bv)∂an) (2.45)

= P∂abv− (n · ∂bv)κ· a (2.46)

= P∂abv+ (b · κ · v)κ · a. (2.47)

Here we used that identities

Pn= 0, = κ, an= κ · a, n· ∂bv= −b · κ · v, (2.48) where the last formula follows from the fact that v· n = 0, which leads to

0= ∂b(n· v) = n · (∂bv)+ ∂bn· v = n · (∂bv)+ b · κ · v. (2.49)

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(11)

We thus obtain

R(a, b, v)= DaDbv− DbDav− D[a,b]v (2.50)

= P(∂abv− ∂bav− ∂[a,b]v)



(2.51)

+ (a · κ · v)κ · b − (b · κ · v)κ·

= (a · κ · v)κ · b − (b · κ · v)κ · a. (2.52)

Here we used the identity

abvi= ∂bavi+ ∂abvi (2.53) for each component viin v, to conclude that

abv− ∂bav= ∂abv− ∂bav= ∂[a,b]v (2.54) and thus = 0. This concludes the verification of (2.41).

Next let{ti}2i=1be a smooth orthonormal basis to Tx(Γ )in the vicinity of a point x ∈ Γ , i.e., all tangential vector fields can be written as a linear combination v =2

i=1vitiwith coordinate functions vi. We then have the identity

R(t2, t1, t1)· t2= R(t1, t2, t2)· t1= K, (2.55) where K = κ1κ2is the Gauss curvature and it also holds

R(t2, t1, t2)· t2= R(t1, t2, t1)· t1= 0 (2.56) and we get the corresponding identities if we interchange t1and t2. In verification of (2.55), we directly obtain

R(t2, t1, t1)· t2= (t1· κ · t1)(t2· κ · t2)− (t2· κ · t1)(t2· κ · t2)= det(κ) = κ1κ2, (2.57) where det(κ) is the determinant of the 2× 2 tangential part of κ, and we used the fact that the matrix T = [t1, t2] is orthogonal and det(TTκT)= det κ. For (2.56), we get

R(t1, t2, t1)· t1= (t2· κ · t1)(t1· κ · t1)− (t1· κ · t1)(t2· κ · t1)= 0 (2.58) and we note that both verifications hold also if we switch t1and t2.

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(12)

If DΓv = 0 we have Dav = 0 for all tangential vector fields a, and thus we can conclude that R(a, b, v)= 0 for all tangential vector fields a, b. Expanding v in the orthonormal frame we also have the identity

0= R(a, b, v) · w = R

 a, b,

2 i=1

viti



· w =

2 i=1

viR(a, b, ti)· w. (2.59)

Setting a= t1, b= t2and w= t1, we get

0= v2K (2.60)

and setting a= t2, b= t1and w= t2, we get

0= v1K. (2.61)

We can therefore conclude that in a point with nonzero Gauss curvature a covariantly constant vector field must be zero. For any closed compact smooth surface embedded inR3there is at least one point x∈ Γ where K = 0, seeThorpe(1994, Theorem 4, p. 88), and thus v(x) = 0, which concludes the verification of the claim in Step 1.

Step 2. Claim: if v∈ Ttan1 is a smooth tangential vector field, which is covariantly constant, DΓv= 0, and there exists a point x∈ Γ such that v(x) = 0, then v(y) = 0 for all y ∈ Γ .

We will use so-called parallel transport of vectors along curves to verify this claim. First, using the fact that a closed compact manifold is geodesically complete in the sense that each point y ∈ Γ is connected to x by a geodesic, i.e., a length minimizing curve γ : I t → γ (t) ∈ Γ , where I = [a, b] is an interval inR and γ (a) = x, γ (b) = y. Consider now the transport problem: find w ∈ {Tx(Γ ): x∈ γ } such that

D˙γw= 0 on γ , w(a)= v(x), (2.62)

where ˙γ = dt is the tangent vector to γ . We note that dwdt◦γ = ∂˙γw and thus D˙γw = Pdwdt◦γ. Setting w◦ γ (t) =2

i=1wi(t)ti◦ γ (t), we get

0= Pdw◦ γ

dt =

2 i=1

dwi(t)

dt ti◦ γ (t) + wi(t)(D˙γ(t)ti) (2.63)

and using the fact that{ti}2i=1is orthonormal, we obtain

dwi(t) dt +

2 j=1

wj(t)(D˙γ(t)tj)· ti, (2.64)

which is a standard system of linear ordinary differential equations with a unique solution since the coefficients are smooth. We say that w is the parallel transport of v along the curve γ . Now let w1and

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(13)

w2be solutions to (2.62) with initial data v1and v2, we then have d(w1· w2)

dt = dw1

dt · w2+ w1·dw2 dt =

 Pdw1

dt



· w2+ w1·

 Pdw2

dt



= 0, (2.65)

where we used the fact that w1and w2are tangent vectors to insert P. Thus, the scalar product of w1 and w2is constant along γ and in particular we havew(t)R3 = v(x)R3. We conclude that v(y)= 0, since v(x) = 0 and v(y) is obtained by parallel transport of v(x) along γ , since DΓv= 0 on Γ implies D˙γv= 0 on γ .

Step 3. Using the fact that smooth tangent vector fields are dense in Htan1 (Γ ) the desired result

follows. 

Lemma 2.2 (Poincaré inequality). For all v∈ H1tan(Γ )there is a constant such that

vΓ  DΓvΓ. (2.66)

Proof. Assume that (2.66) does not hold. Then there is a sequence{vk}k=1in Htan1 (Γ )such that

vkΓ  kDΓvkΓ. (2.67)

Setting wk= vk/vkΓ, we obtain

DΓwkΓ  k−1 (2.68)

and therefore{wk}k=1is bounded in Htan1 (Γ ). Using Rellich’s compactness theorem, seeTaylor(2011, Ch. 4, Prop. 4.4, p. 334), there is a subsequence{wkj}j=1and a tangential vector field w∈ L2(Γ )such that

wkj → w in L2(Γ ). (2.69)

ThenwΓ = 1 and DΓwΓ = 0, but this is a contradiction in view of Lemma2.1.  Lemma 2.3 (Sobolev norm equivalence). For all tangential vector fields v∈ Htanm (Γ ), and m= 1, 2, . . . there are constants such that

∇ΓmvΓ 

m k=0

DkΓvΓ (2.70)

DmΓvΓ 

m k=0

∇ΓkvΓ (2.71)

and as a consequence

vHm(Γ )∼ vHmtan(Γ ). (2.72) Proof. Let X∈ Ttann be a smoothly varying tangential tensor on Γ .

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(14)

Bound (2.70). Taking k derivatives on X, adding and subtracting a projection on the innermost derivative and using the triangle inequality, we obtain

∇ΓkXΓ = ∇Γk−1(P(∇ΓX)+ (I − P)(∇ΓX))Γ (2.73)

 ∇Γk−1(DΓX)Γ + ∇Γk−1((I− P)(∇ΓX))Γ, (2.74) where we in the first term use the definition of the covariant derivative DΓX = P(∇ΓX). Next we show that the second term is actually of lower order. Expressing X using the canonical expansion in the spanning set, see (2.13), we, by the product rule, have the total derivative

ΓX=

3 i1,...,in,j=1

p

j(Xi

1···in) pi

1⊗ · · · ⊗ pin ⊗ pj

∈Ttann+1

+Xi1···inp

j( pi

1⊗ · · · ⊗ pin)⊗ pj. (2.75)

As the first term in this sum is tangential we, after subtraction of a projection, get the expression

(I− P)(∇ΓX)=

3 i1,...,in,j=1

Xi1···in(I− P)

p

j( pi

1⊗ · · · ⊗ pin)⊗ pj

, (2.76)

which has no derivatives acting on the coordinates of X. Furthermore, we note that we have the identities

(I− P)∂pjpi= −κijn= −

3 k=1

κijnkek, pj=

3 l=1

pjlel, (2.77)

where the expansion coefficients are smooth since Γ is smooth. Thus, there are smooth functions αi

1...in,k1...kn,lsuch that

(I− P)

⎝3

j=1

p

j( pi

1⊗ · · · ⊗ pin)⊗ pj

⎠ = αi1...in,k1...kn,lek1⊗ · · · ⊗ ekn⊗ el. (2.78)

Defining the smooth 2n+ 1 tensor A by A= αi1...in,k1...kn,l(ei

1⊗ · · · ⊗ ein)⊗ (ek1⊗ · · · ⊗ ekn)⊗ el (2.79) we have the identity

ei1⊗ · · · ⊗ ein·nA= αi1...in,j;k1...kn,lek1⊗ · · · ⊗ ekn⊗ el (2.80) and thus, using the canonical expansion of the tangential tensor X, we obtain the identity

X·nA= (I − P)(∇ΓX). (2.81)

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

(15)

Using the product rule, we get

∇Γk−1(I− P)(∇ΓX)Γ = ∇Γk−1(X·nA)Γ (2.82)



k−1



l=0

∇Γl XΓ∇Γk−1−lAL(Γ )

1

(2.83)



k−1



l=0

∇Γl XΓ. (2.84)

Combined with (2.73)–(2.74), this yields

∇ΓkXΓ  ∇Γk−1(DΓX)Γ +

k−1



l=0

∇Γl XΓ (2.85)

with a constant depending only on Γ . Inequality (2.70) now follows by induction. For k= 1 estimate (2.70) follows directly from (2.85). Assuming that (2.70) holds for k− 1, we have the estimate

∇ΓkXΓ  ∇Γk−1(DΓX)Γ +

k−1



l=0

∇Γl XΓ (2.86)



k−1



l=0

DlΓ(DΓX)Γ +

k−1



l=0

DlΓXΓ (2.87)



k l=0

DlΓXΓ (2.88)

and thus (2.70) holds for k as well.

Bound (2.71). By adding and subtractingΓX inside the gradients and applying the triangle inequality, we have

∇Γk−1DΓXΓ  ∇Γk−1(ΓX)Γ + ∇Γk−1((I− P)(∇ΓX))Γ (2.89)

 ∇ΓkXΓ + ∇Γk−1XΓ, (2.90)

where we use the same lower-order bound on the second term in (2.89) as above. The inequality readily follows by iterating this formula, starting with k = 1 and X = DmΓ−1v, and applying the Poincaré inequalityvΓ  v ⊗ ∇ΓΓ. This Poincaré inequality clearly holds as we, by Lemma2.2, have

v2Γ  DΓv2Γ  DΓv2Γ + (I − P)(v ⊗ ∇ Γ)2Γ

0

= v ⊗ ∇Γ2Γ (2.91)

by the orthogonality between tangential and nontangential tensors. 

Downloaded from https://academic.oup.com/imajna/article/40/3/1652/5435894 by Umea University Library user on 19 August 2020

References

Related documents

In case of collusion attack, multiple compromised nodes may share their individual pieces of information to regain access to the group key. The compromised nodes can all belong to

Om de fyra kvinnorna i vår undersökning som var arbetslösa istället haft sysselsättning av något slag skulle de i enlighet med Hiltes (1996) förklaring av kontrollteorin kanske

Då denna studie syftade till att redogöra för hur stora svenska företag, som är topprankade inom CSR, designar sina MCS- paket för att styra medarbetare i linje med

Dessa anses dock inte vara lika stora problem inom det svenska samhället eftersom det finns stor kunskap kring riskerna och att Sverige ligger långt fram inom dessa punkter,

Flera kvinnor i vår studie upplevde att deras omgivning inte var villiga att stötta dem i den känslomässiga processen och detta kan ha lett till att det uppkommit känslor av skam

156 Anledningen till detta är, enligt utredningen, att en lagstiftning vilken innebär skolplikt för gömda barn skulle vara mycket svår att upprätthålla: ”Det kan inte krävas

For example, Lusch and Nambi- san (2015) cautioned that the product–service distinction should not constrain a broader view of innovation. Similarly, Rubalcaba et al. 697) noted

The adherends are modelled using 4-node shell elements of type S4 and the adhesive layer is modelled with cohesive user elements.. In the simulations discussed in