You are on page 1of 47

March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

Mathematical Models and Methods in Applied Sciences



c World Scientific Publishing Company
Sandia National Labs SAND 2010-8353J

A NONLOCAL VECTOR CALCULUS,


NONLOCAL VOLUME-CONSTRAINED PROBLEMS,
AND NONLOCAL BALANCE LAWS

QIANG DU
Department of Mathematics, Pennsylvania State University
University Park, PA 16802
qdu@math.psu.edu

MAX GUNZBURGER
Department of Scientific Computing, Florida State University
Tallahassee FL 32305-4120
gunzburg@fsu.edu

R. B. LEHOUCQ
Sandia National Laboratories
P.O. Box 5800, MS 1320, Albuquerque, NM 87185-1320
rblehou@sandia.gov

KUN ZHOU
Department of Mathematics, Pennsylvania State University
University Park, PA 16802
zhou@math.psu.edu

Received (Day Month Year)


Revised (Day Month Year)
Communicated by (xxxxxxxxxx)

A vector calculus for nonlocal operators is developed, including the definition of nonlocal
divergence, gradient, and curl operators and the derivation of the corresponding adjoint
operators. Nonlocal analogs of several theorems and identities of the vector calculus for
differential operators are also presented. Relationships between the nonlocal operators
and their differential counterparts are established, first in a distributional sense and
then in a weak sense by considering weighted integrals of the nonlocal adjoint operators.
The operators of the nonlocal calculus are used to define volume-constrained problems
that are analogous to elliptic boundary-value problems for differential operators; this
is demonstrated via some examples. Another application discussed is posing abstract
nonlocal balance laws and deriving the corresponding nonlocal field equations; this is
demonstrated for heat conduction and the peridynamics model for continuum mechanics.

Keywords: nonlocal operators, vector calculus, volume-constrained problems, balance


laws, peridynamics, nonlocal diffusion

AMS Subject Classification: 26B12, 26B15, 26B20, 45A05, 45P05, 45E99, 46F12

1
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

2 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

1. Introduction
Our principal goal is to develop a vector calculus for nonlocal operators that mimics
Sandia National Labs SAND 2010-8353J

the classical vector calculus for differential operators. We also show how the nonlocal
vector calculus can be used to define nonlocal balance laws and volume-constrained
problems that mimic boundary-value problems for differential operators.
Nonlocal analogs of the divergence, gradient, and curl operators are defined
and the corresponding nonlocal adjoint operators are deduced. Nonlocal analogs
of the Gauss theorem and Green’s identities of the vector calculus for differential
operators are also derived. Relationships between the nonlocal operators and their
differential counterparts are established. The nonlocal vector calculus can be used
to define nonlocal volume-constrained problems that are analogous to boundary-
value problems for partial differential operators. In addition, the nonlocal vector
calculus has an important application to balance lawsa that are nonlocal in the
sense that subregions not in direct contact may have a non-vanishing interaction.
This is accomplished by defining a nonlocal flux in terms of interactions between
regions having positive measure, possibly not sharing a common boundary. As a
result, the nonlocal vector calculus provides an alternative to standard approaches
for circumventing the technicalities associated with the lack of sufficient regularity
in local balance laws.
Preliminary attempts at a nonlocal calculus are found in Refs. 11 and 12 which
include applications to image processing and steady-state diffusion, respectively.
In particular, in Ref. 26, which is cited in Ref. 11, a discrete nonlocal divergence
and gradient are introduced within the context of machine learning; see also Refs.
4, 14, and 16 where a discrete calculus is also discussed. However, the discussion
in those papers is limited to scalar problems. In contrast, this paper extends the
ideas in Refs. 11 and 12 to vector and tensor fields and beyond the consideration of
image processing and steady-state diffusion. For example, the ideas presented here
enable an abstract formulation of the balance laws of momentum and energy in the
peridynamic continuum modelb for solid mechanics that parallels the classical vector
calculus formulation of the balance laws of elasticity. In fact, the nonlocal vector
calculus presented in this paper is sufficiently general that we envisage application
to balance laws beyond those of elasticity, e.g., to the laws of fluid mechanics and
electromagnetics.
The paper is organized as follows. The remainder of this section is devoted to es-
tablishing notation. In Section 2, the notions of local and nonlocal fluxes into or out
of a region are compared and contrasted. The next two sections contain our prin-

aA balance law postulates that the rate of change of an extensive quantity over any domain is
given by the rate at which that quantity is produced in the domain minus the flux out of the
domain.
b The peridynamics continuum model was introduced in Refs. 21 and 23; Ref. 24 reviews the peri-

dynamic balance laws of momentum and energy and provides many citations for the peridynamic
model and its applications. See Section 7.3 for a brief discussion.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 3

cipal contributions. In Section 3, nonlocal divergence, gradient, and curl operators


are introduced as are the corresponding adjoint operators, several vector identities,
Sandia National Labs SAND 2010-8353J

and other results about the operators. The nonlocal vector calculus is developed in
Section 4; in particular, nonlocal integral theorems and nonlocal Green’s identities
are derived. In Section 5, connections between the nonlocal operators and distribu-
tional and weak representations of the associated classical differential operators are
made. Sections 6 and 7 deal with applications of the nonlocal vector calculus. In
Section 6, examples are given of nonlocal volume-constrained problems formulated
in terms of the nonlocal operators. Then, in Section 7, a brief review of the con-
ventional notion of a balance law is provided after which abstract nonlocal balance
laws are discussed. The notion of nonlocal fluxes discussed in Section 2 is used in
developing nonlocal balance laws and the vector calculus developed in Section 4
plays a crucial role in transforming balance laws into field equations. Also, in Sec-
tion 7, a brief discussion is given of the application of our nonlocal vector calculus
to the peridynamic theory for continuum mechanics.
Throughout, wherever it is illuminating, we associate the definitions and results
of the nonlocal vector calculus with the analogous definitions and results of the
classical differential vector calculus.

1.1. Notation
We have need of two types of functions and two types of nonlocal operators. Point
functions refer to functions defined at points whereas two-point functions refer to
functions defined for pairs of points. Point operators map two-point functions to
point functions whereas two-point operators map point functions to two-point func-
tions so that the nomenclature for operators refer to their ranges. Point and two-
point operators are both nonlocal. Point operators involve integrals of two-point
functions whereas two-point operators explicitly involve point functions evaluated
at two different points.
We now make more precise the definitions given above. Let m, k, and n denote
positive integers. Points in Rn are denoted by the vectors x, y, or z and the natural
e ⊆ Rn , functions from Ω
Cartesian basis is denoted by e1 , . . . , en . For any domain Ω e
m×k m
into R or R or R are referred to as point functions or point mappings and are
denoted by Roman letters, upper-case bold for tensors, lower-case bold for vectors,
and plain face for scalars, respectively, e.g., U(x), u(x), and u(x), respectively.
Functions from Ω×e Ωe into Rm×k or Rm or R are referred to as two-point functions or
two-point mappings and are denoted by Greek letters, upper-case bold for tensors,
lower-case bold for vectors, and plain face for scalars, respectively, e.g., Ψ(x, y),
ψ(x, y), and ψ(x, y), respectively. Symmetric and antisymmetric scalar two-point
functions ψ(x, y) satisfy ψ(x, y) = ψ(y, x) and ψ(x, y) = −ψ(y, x), respectively,
and similarly for vector and tensor two-point functions.
For the sake of notational simplicity, in much of the rest of the paper we adopt
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

4 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

the following notation:

α := α(x, y) α0 := α(y, x) ψ := ψ(x, y) ψ 0 := ψ(y, x)


Sandia National Labs SAND 2010-8353J

u := u(x) u0 := u(y) u := u(x) u0 := u(y)

and similarly for other functions.


The dot (or inner) product of two vectors u, v ∈ Rm is denoted by u · v ∈ R;
the dyad (or outer) product is denoted by u ⊗ w ∈ Rm×k whenever w ∈ Rk ; given
a second-order tensor (matrix) U ∈ Rk×m , the tensor-vector (or matrix-vector)
product is denoted by U · v and is given by the vector whose components are the
dot products of the corresponding rows of U with v.c In R3 , the cross product of
two vectors u and v is denoted by u×v ∈ R3 . The Frobenius product of two second-
order tensors A ∈ Rm×k and B ∈ Rm×k , denoted by A : B, is given by the sum of
P
the element-wise product of the two tensors,i.e., A : B = i=1,...,m, j=1,...,k Aij Bij .
The trace of B ∈ Rm×m , denoted by tr B , is given by the sum of the diagonal
elements of B.
Inner products in L2 (Ω)e and L2 (Ω e × Ω)
e are defined in the usual manner. For
example, for vector functions, we have
 Z
 (u, v)Ω

 e = u · v dx for u(x), v(x) ∈ Ωe
Ωe
Z Z
 (µ, ν) e e = µ · ν dydx for µ(x, y), ν(x, y) ∈ Ω

 e
Ω×Ω

e Ω
e

with analogous expressions involving the Frobenius product and the ordinary prod-
uct for tensor and scalar functions, respectively.
Many of the results in the paper require that domains under consideration are
Lebesgue measurable and functions are Lebesgue integrable with their pointwise
properties interpreted in the Lebesgue measure sense, i.e., as holding almost every-
where.

2. Local and nonlocal fluxes and action-reaction principles


A key concept in the development of a vector calculus is the notion of a flux which
accounts for the interaction of points in a domain with points outside the domain.
As a result, the notion of a flux is also fundamental to the understanding of balance
laws in mechanics, heat transfer, and many other settings; see Section 7. In the
classical setting of local interactions, that interaction occurs at the boundary of the
domain whereas, in the nonlocal case, the interaction occurs over volumes external
to the domain. In order to compare and contrast the notion of a nonlocal flux with
the classical local flux, we begin by briefly reviewing the latter notion.

c In
matrix notation, the inner, outer, matrix-vector products are given by x·y = xT y, x⊗y = xyT ,
and U · v = Uv.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 5

2.1. Local fluxes


Let Ω1 ⊂ Rn and Ω2 ⊂ Rn denote two disjoint open regions having Lipschitz
Sandia National Labs SAND 2010-8353J

boundaries. If Ω1 and Ω2 have a nonempty common boundary ∂Ω12 = Ω1 ∩ Ω2 ,


then, for a vector-valued function q(x), the expression
Z
q · n1 dA (2.1)
∂Ω12

represents the classical local flux out of Ω1 into Ω2 , where n1 denotes the unit
normal on ∂Ω12 pointing outward from Ω1 and dA denotes a surface measure in
Rn ; q · n1 is referred to as the flux density along ∂Ω12 in the direction of the normal
vector n1 to the surface at that point. Because the local flux density q·n1 at a point
is defined in terms of any orientable surface passing through that point, the flux
conveys a notion of direction out of and into regions and is a proxy for the interaction
between Ω1 and Ω2 , i.e., the flux is an oriented interaction between two domains.
An important observation is that the local flux from Ω1 into Ω2 occurs across their
common boundary and that if the two disjoint regions have no common boundary,
then the flux from one to the other is zero. The classical flux (2.1) is then deemed
to be local because there is no interaction between Ω1 and Ω2 when separated by a
finite distance. The classical flux satisfies the action-reaction principled
Z Z
q · n1 dA + q · n2 dA = 0, (2.2)
∂Ω12 ∂Ω21

where, of course, ∂Ω12 = ∂Ω21 and n2 = −n R 1 denotes the unit normal on ∂Ω12
pointing outward from Ω2 . In words, the flux ∂Ω12 q · n1 dA from
R Ω1 into Ω2 across
their common boundary ∂Ω12 is equal and opposite to the flux ∂Ω21 q · n2 dA from
Ω2 into Ω1 across that same surface. The vector q is often expressed in terms of an
intensive variable through a constitutive relation.e

2.2. Nonlocal fluxes


For any point x ∈ Rn , for integrable ψ(x, y), we identify
Z
ψ(x, y) dy e ⊆ Rn
∀Ω (2.3)

e

as the nonlocal flux density f at x into a measurable volume Ω. e Because of nonlo-


cality, x does not necessarily have to belong to the closure of Ω.e The nonlocal flux
density (2.3) has units of an intensive quantity per unit volume, in contrast to the
local flux density q · n that has units of an intensive quantity per unit area. In (2.3),

d An example is in mechanics for which Newton’s third law, i.e., the force exerted upon an object
is equal and opposite to the force exerted by the object, is an action-reaction archetype.
e For example, if q · n denotes the heat flux density, then q is related to the temperature via the
1
Fourier heat law; see Section 7.2.2.
f In Section 3.2, we show how ψ can be expressed as an inner product of two vectors so that (2.3)

mimics the local flux density q · n.


March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

6 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

ψ(x, y) may be viewed as a flux density per unit volume; no such notion exists in
the local case because the flux density is not defined in terms of an integral as is
Sandia National Labs SAND 2010-8353J

the nonlocal flux density (2.3).


The following result is fundamental to the development of the nonlocal calculus.
Here and throughout, we assume that domains such as Ω1 , Ω2 , and Ω e are measur-
able.

Theorem 2.1. The following are equivalent:


(i) the antisymmetry of a two-point function
ψ(x, y) = −ψ(y, x) for almost all x, y ∈ Rn (2.4a)
(ii) the alternating or no self-interaction principle
Z Z
ψ(x, y) dydx = 0 e ⊆ Rn
∀Ω (2.4b)

e Ω
e

(iii) the nonlocal action-reaction principle


Z Z Z Z
ψ(x, y) dy dx + ψ(x, y) dy dx = 0 ∀ Ω1 , Ω2 ⊆ Rn (2.4c)
Ω1 Ω2 Ω2 Ω1

(iv) additivity for disjoint regions


Z Z
ψ(x, y) dydx
Ω1 ∪Ω2 Rn \(Ω1 ∪Ω2 )
Z Z Z Z
= ψ(x, y) dydx + ψ(x, y) dydx (2.4d)
Ω1 Rn \Ω1 Ω2 Rn \Ω2

∀ Ω1 , Ω2 ⊆ Rn such that Ω1 ∩ Ω2 = ∅.
Similar results hold for antisymmetric vector and tensor two-point functions.

Proof. A change of variables under integration followed by change in the order of


integration shows that
Z Z Z Z
ψ(x, y) dy dx = ψ(y, x) dy dx ∀Ωe ∈ Rn

e Ω
e Ω
e Ω
e

so that
Z Z Z Z
e ∈ Rn .

2 ψ(x, y) dy dx = ψ(x, y) + ψ(y, x) dy dx ∀Ω

e Ω
e Ω
e Ω
e

Thus, obviously, (i) implies (ii). This relation also shows that (ii) implies (i) because
if (ii) holds, we have
Z Z
e ∈ Rn

ψ(x, y) + ψ(y, x) dy dx = 0 ∀Ω

e Ω
e

which implies that ψ(x, y)R +Rψ(y, x) = 0 almost everywhere.


Setting F(Ω1 , Ω2 ) = Ω1 Ω2 ψ(x, y) dydx, we easily see that (2.4a) and (2.4b)
are satisfied. Then, the equivalence of (ii) and (iii) follows from Proposition (7.1)
and the equivalence of (ii) and (iv) follows from Proposition (7.3).
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 7

With (2.3) denoting a nonlocal flux density, for any two open regions Ω1 ⊆ Rn
and Ω2 ⊆ Rn , we identify
Sandia National Labs SAND 2010-8353J

Z Z
ψ(x, y) dydx (2.5)
Ω1 Ω2

as a scalar interaction or nonlocal flux from Ω1 into Ω2 . As in the local case, the
nonlocal flux is associated with the notion of an orientation so that nonlocal fluxes
are again oriented interactions between two domains.
Because we require no self interactions, i.e., the flux from a region into itself
vanishes so that (2.4b) holds whenever Ω1 = Ω2 , by Theorem 2.1 we equivalently
require
R that ψ : (Ω1 ∪ Ω2 ) × (Ω1 ∪ Ω2 ) → R be an antisymmetric function. By (2.3),
Ω2
ψ(x, y) dy is the nonlocal flux density at a point x ∈ Ω1 into the Rregion Ω2 . As
is the case for the local flux density q · n1 , the nonlocal flux density Ω2 ψ(x, y) dy
is related to an intensive variable through a constitutive relation; see Section 7.2.3.
Based on the above discussion, we see that (2.4c) is the nonlocal analogue of
(2.2). In words, (2.4c) states that the flux (or interaction) from Ω1 into Ω2 is equal
and opposite to the flux (or interaction) from Ω2 into Ω1 . The flux is nonlocal
because, by (2.5), the interaction may be nonzero even when the closures of Ω1 and
Ω2 have an empty intersection. This is in stark contrast to classical local interactions
for which we have seen that the interaction between Ω1 and Ω2 vanishes if their
closures have empty intersection, i.e., if they have no common boundary.

3. Nonlocal operators
The nonlocal vector calculus developed in Section 4 involves nonlocal operators
that mimic the classical local differential divergence, gradient, and curl operators.
An important distinction between local and nonlocal operators is that the adjoint
operators for the former involve the same operators, i.e., the adjoint of ∇· is −∇,
of ∇ is −∇·, and of ∇× is ∇×, whereas the adjoint of nonlocal operators involve
differently defined nonlocal gradient, divergence, and curl operators. See Section 3.3
for a further discussion.

3.1. Motivating the definitions of nonlocal operators


To provide motivation for the definitions given in Definition 3.1 for the basic oper-
ators of the nonlocal vector calculus, we recall that the Gauss theorem
Z Z
∇ · q dx = q · n dA ∀Ωe ⊂ Rn (3.1)

e ∂Ω
e

shows that the integral of the divergence of a vector q over any region Ω e ⊂ Rn is
equal to the flux out of that region. Of course, the latter is the integral over ∂ Ω
e of
the flux density q·n. We adopt the same word definition for the nonlocal divergence,
i.e., with D denoting the nonlocal divergence operator, we have that the integral of
the nonlocal divergence of a vector ν over any region Ωe ⊂ Rn is equal to the flux out
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

8 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

e can interact with all other points y ∈ Rn


of that region. In general, a point x ∈ Ω
so that, by (2.3), the nonlocal flux density out of a point x ∈ Ω
e is given by
Sandia National Labs SAND 2010-8353J

Z
ψ(x, y) dy.
Rn

Then, according to our definition of a the nonlocal divergence operator, we have


from (2.5) that
Z Z Z
D(ν)(x) dx = ψ(x, y) dydx ∀Ωe ⊂ Rn . (3.2)

e Ω
e Rn

In (3.1), the flux density q · n is obviously related to the vector q, i.e., to the
vector operand of the local differential divergence operator. In the nonlocal case, to
define the action of the nonlocal divergence operator on the vector ν, we have to
analogously determine how the operand ν of the nonlocal divergence operator D
is related to ψ(x, y), where, because of non-locality, ν(x, y) is a vector two-point
function. The information we have in hand to determine this relation is:
– because we assume no self interactions,
(3.3a)
(2.4a) requires that the scalar function ψ(x, y) be antisymmetric
– because of our definition of the nonlocal divergence operator,
(3.3b)
(3.2) relates D to ψ
– because we emulate the fact that ∇· is a linear operator,
(3.3c)
ψ(x, y) and therefore also D are linear in ν(x, y).
This information is sufficient to apply the Schwartz Kernel Theorem (see, e.g., Refs.
8, 10, 20) to determine the relation between ψ and ν that leads to a definition of
the nonlocal divergence operator. Before justifying this claim, we need to introduce
some notation.
Let
U = [C ∞ (Rn × Rn )]k and V = C ∞ (Ω)
e

so that U consists of vector two-point functions whereas V consists of scalar point


functions. The dual spaces are given by
U 0 = [D0 (Rn × Rn )]k and V 0 = D0 (Ω),
e

respectively, where D0 denotes the space of distributions. The corresponding duality


pairings are given as follows:
Z Z Z
hν, µiU,U 0 = ν(z, y) · µ(z, y) dzdy and hv, uiV,V 0 = v(x)u(x) dx
Rn Rn Ω
e
0 0
for ν ∈ U , µ ∈ U and v ∈ V , u ∈ V , respectively. We also use the product space
U × V , its dual (U × V )0 , and, for κ ∈ U × V and ρ ∈ (U × V )0 , the corresponding
duality pairing
Z Z Z
hκ, ρiU ×V,(U ×V )0 = κ(x, y, z) · ρ(x, y, z) dzdydx.

e Rn Rn
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 9

Specialized to our needs, the Schwartz Kernel Theorem is given as follows.

Theorem 3.1. [Schwartz Kernel Theorem] Let D : U → V 0 denote a con-


Sandia National Labs SAND 2010-8353J

tinuous linear map. Then, there exists a unique kernel ρ(x, y, z) ∈ (U × V )0 such
that

hv, DνiV,V 0 = hvν, ρi(U ×V ),(U ×V )0 ∀ v ∈ V, ν ∈ U

or, using the definitions of the duality pairings,


Z Z Z Z
v(x)(Dν)(x) dx = ρ(x, y, z) · ν(z, y)v(x) dzdydx

e Ω
e Rn Rn (3.4)
∀ v ∈ V, ν ∈ U. 

Corollary 3.1. Assume that (3.3) holds. The nonlocal flux density per unit volume
ψ(x, y) is uniquely expressed in terms of the vector ν(x, y) by
Z
1
ρ(x, y, z) · ν(z, y) − ρ(y, x, z) · ν(z, x) dz ∀ x, y ∈ Rn (3.5)

ψ(x, y) =
2 Rn
and the nonlocal divergence operator D is uniquely expressed in terms of the vector
ν(x, y) by
Z Z
1 
(Dν)(x) = ρ(x, y, z) · ν(z, y) − ρ(y, x, z) · ν(z, x) dzdy
2 Rn Rn (3.6)
n
∀x ∈ R .

e ⊂ Rn is arbitrary, we conclude from (3.4) that


Proof. Because Ω
Z Z
(Dν)(x) = ρ(x, y, z) · ν(z, y) dzdy ∀ x ∈ Rn . (3.7)
Rn Rn

Comparing (3.2) and (3.7), we see that


Z
ψ(x, y) = ρ(x, y, z) · ν(z, y) dz ∀ x, y ∈ Rn . (3.8)
Rn

The antisymmetry of ψ(x, y) with respect to x and y requires that

ρ(y, x, z) · ν(z, x) = −ρ(x, y, z) · ν(z, y) ∀ x, y, z ∈ Rn (3.9)

so that (3.5) and (3.6) follow from (3.8) and (3.7), respectively.

The general definition for the nonlocal divergence is thus given by (3.6). We have
found that we can simplify that definition and still develop a nonlocal vector calculus
that meets our needs; however, it is possible that in some applications, retaining
the more general definition (3.6) could prove useful. Our simplifying assumption is
to let

ρ(x, y, z) = 2α(x, y)δ(x − z).


March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

10 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

Substituting into (3.6), we obtain


Z Z
1 
(Dν)(x) = ρ(x, y, z) · ν(z, y) − ρ(y, x, z) · ν(z, x) dzdy
Sandia National Labs SAND 2010-8353J

2 n n
Z RZ R

= α(x, y) · ν(z, y)δ(x − z) − α(y, x) · ν(z, x)δ(y − z) dzdy
n n
ZR R

= α(x, y) · ν(x, y) − α(y, x) · ν(y, x) dy
Rn
At this point there are no restrictions on the two-point function α(x, y) as far as its
symmetry or antisymmetry with respect to its two arguments. However, in Section
5, we provide reasons for why α(x, y) should be antisymmetric so that we assume
this is true from the outset, leading to the definition
Z

(Dν)(x) = ν(x, y) + ν(y, x) · α(x, y) dy (3.10)
Rn
with α(x, y) antisymmetric.
We now have that ψ = (ν + ν 0 ) · α; note that the integrability of ψ does not
imply that α is necessarily integrable. For example, the integral appearing in (3.10)
could be interpreted as the L2 (Rn ) duality pairing of α with ν which would allow
for α to be nonintegrable or to be a distribution.

3.2. Nonlocal point divergence, gradient, and curl operators


Based on the discussion of Section 3.1, the nonlocal point divergence, gradient, and
curl operators mapping two-point functions to point functions are defined in terms
of their action on two-point functions as follows. These operators along with their
adjoints are the building blocks of our nonlocal calculus.

Definition 3.1. [Nonlocal operators] Given the vector two-point function


ν(x, y) : Rn × Rn → Rk and the antisymmetric vector two-point function
α(x, y) : Rn × Rn → Rk , the action of the nonlocal point divergence operator D
on ν is defined as
Z
ν + ν 0 · α dy
 
D ν (x) := for x ∈ Rn , (3.11a)
Rn

where D ν : R → R. Given the scalar two-point function η(x, y) : Rn × Rn → R
n

and the antisymmetric vector two-point function β(x, y) : Rn ×Rn → Rk , the action
of the nonlocal point gradient operator G on η is defined as
Z
η + η 0 β dy
 
G η (x) := for x ∈ Rn , (3.11b)
Rn

where G η : R → R . Given the vector two-point function µ(x, y) : Rn × Rn → R3
n k

and the antisymmetric vector two-point function γ(x, y) : Rn ×Rn → R3 , the action
of the nonlocal point curl operator C on µ is defined as
Z
γ × µ + µ0 dy
 
C µ (x) := for x ∈ Rn , (3.11c)
Rn
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 11


where C µ : Rn → R3 .

The nonlocal point operators D, G, and C map vectors to scalars, scalars to


Sandia National Labs SAND 2010-8353J

vectors, and vectors to vectors, respectively, as is the case for the divergence, gra-
dient, and curl differential operators. Relationships between the nonlocal point op-
erators and differential operators are made in Section 5 where we demonstrate
circumstances under which the nonlocal point operators are identified with the
corresponding differential operators in the sense of distributions and also as weak
representations.
Because the integrands in (3.11a)–(3.11c) are antisymmetric, (2.4b) immediately
implies that
Z Z Z
D(ν) dx = 0, G(η) dx = 0, and C(µ) dx = 0. (3.12)
Rn Rn Rn

These relations may be viewed as free-space nonlocal integral theorems.g

3.3. Nonlocal adjoint operators


The adjoint operators corresponding to nonlocal point operators are two-point op-
erators that are defined as follows.

Definition 3.2. Given a point operator Q that maps two-point functions F to


point functions defined over Rn , the adjoint operator Q∗ is a two-point operator
that maps point functions G to two-point functions defined over Rn × Rn that
satisfies

G, Q(F ) Rn − Q∗ (G), F Rn ×Rn = 0,


 
(3.13)

where, for Q = D, G, or C, we have that F and G denote pairs of vector-scalar,


scalar-vector, or vector-vector functions, respectively.

Definition 3.13 can be used to determine the nonlocal adjoint two-point operators
corresponding to the nonlocal point operators introduced in Definition 3.1.

Proposition 3.1. [Nonlocal adjoint operators] Given the scalar point function
u(x) : Rn → R, the adjoint of D is the two-point operator whose action on u is given
by

D∗ u (x, y) = −(u0 − u)α for x, y ∈ Rn ,



(3.14a)

where D∗ u : Rn × Rn → Rk . Given the vector point function v(x) : Rn → Rk , the




adjoint of G is the two-point operator whose action on v is given by

G ∗ v (x, y) = −(v0 − v) · β for x, y ∈ Rn ,



(3.14b)

g For
R
example, RRn D(ν) dx = 0 can be viewed as a nonlocal analog of the free space classical local
Gauss theorem Rn ∇ · u dx = 0.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

12 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

where G ∗ v : Rn × Rn → R. Given the vector point function w(x) : Rn → R3 , the




adjoint of C is the two-point operator whose action on w is given by


Sandia National Labs SAND 2010-8353J

C ∗ w (x, y) = γ × (w0 − w) for x, y ∈ Rn ,



(3.14c)
where C ∗ w : Rn × Rn → R3 .


Proof. Let ξ = uν. Then,


(ξ + ξ 0 ) · α = (uν + u0 ν 0 ) · α = u(ν + ν 0 ) · α + (u0 − u)ν 0 · α (3.15)
so that, from (3.11a) and the first equation in (3.12), we have
Z Z Z  
0= D(ξ) dx = u(ν + ν 0 ) · α + (u0 − u)ν 0 · α dydx
Rn n n
ZR RZ Z Z
0
= u (ν + ν ) · α dydx + (u0 − u)ν 0 · α dydx
Rn Rn Rn Rn
Z Z Z
= uD(ν) dx + (u0 − u)ν · α dydx,
Rn Rn Rn
(3.16)
where the last equality follows because, due to the antisymmetry of α, we have that
(u0 − u)(ν − ν 0 ) · α is an antisymmetric two-point function so that, by (2.4b),
Z Z
(u0 − u)(ν − ν 0 ) · α dydx = 0. (3.17)
Rn Rn
Associating F with ν, G with u, and Q with D, we see that (3.16) is exactly of the
form (3.13) with Q∗ = D∗ , where D∗ is given by (3.14a).
In a similar manner, (3.14b) can be derived from (3.11b) and the second equation
in (3.12) and (3.14c) can be derived from (3.11c) and the third equation in (3.12).

The operator −G ∗ , the negative of the adjoint of the nonlocal gradient operator
G, can be used to define a second form for a nonlocal divergence operator. Similarly,
the pairs G, −D∗ and C, C ∗ serve to define two forms for the nonlocal gradient and
curl operators, respectively. It is not surprising that there are two forms for each
operator because, in the nonlocal case, we have two different types of functions,
i.e., one-point and two-point functions. Naturally, one needs two sets of operators,
one operating on two-point functions, i.e., D, G, and C, and the other operating
on one-point functions, i.e., D∗ , G ∗ , and C ∗ .h Furthermore, the composition of op-
erators requires the two sets of operators. For example, if one wants to compose
a divergence and a gradient operator, the composition D(Gη) (where η denotes
a two-point function) involves the application of D to the one-point function Gη,
whereas in general, one expects D to operate on two-point functions. On the other
hand, the composition D(−D∗ u) (where u denotes a one-point function) involves
the application of D to the two-point function −D∗ u.

h Of course, in the local case, one only deals with one-point functions so that there is no need to
define two different sets of operators.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 13

3.4. Further observations and results about nonlocal operators


In this subsection, we collect several observations and results that can be deduced
Sandia National Labs SAND 2010-8353J

from the definitions and results of Sections 3.2 and 3.3.

3.4.1. A nonlocal divergence operator for tensor functions and a nonlocal


gradient operator for vector functions
A nonlocal divergence operator for tensor functions is defined by applying (3.11a)
to each row of the tensor.
Definition 3.3. Given the tensor two-point function Ψ : Rn × Rn → Rm×k and
the antisymmetric vector two-point function α : Rn × Rn → Rk , the action of the
nonlocal point divergence operator D for tensors on Ψ is defined as
Z
Ψ + Ψ0 · α dy for x ∈ Rn ,

D(Ψ)(x) := (3.18a)
Rn
n m
where D(Ψ)(x) : R → R .
Definition 3.2 implies that, given the vector point function v(x) : Rn → Rm , the
action of the adjoint of D on v is given by
D∗ (v)(x, y) = −(v0 − v) ⊗ α for x, y ∈ Rn , (3.18b)
∗ n n m×k ∗
where D (v) : R × R → R , i.e., D maps a vector point function to a second-
order tensor two-point function. One can also extend Definition (3.11b) of the non-
local gradient operator of a two-point scalar function to define the nonlocal gradient
operator Gv acting on a two-point vector function ψ(x, y) as the point tensor func-
tion
Z
G(ψ)(x) := (ψ 0 + ψ) ⊗ β dy.
Rn
The corresponding nonlocal adjoint operator G ∗ acting on a point tensor function
U(x) is then given by the two-point vector function
G ∗ (U)(x, y) = −(U0 − U) · β.

3.4.2. Nonlocal vector identities


Compositions of the point operators defined in Sections 3.2 and 3.4.1 with the
corresponding adjoint two-point operators derived in Sections 3.3 and 3.4.1 lead to
the following nonlocal vector identities. In this proposition, we set α = β = γ and,
for the first, second, and fourth results, i.e., those involving nonlocal curl operators,
we set m = k = 3; we also set m = k for the third result.i

i Thefour identities in (3.19) are analogous to the vector identities associated with the differential
divergence, gradient and curl operator:
∇ · (∇ × u) = 0, ∇ × (∇u) = 0, ∇ · u = tr(∇u),
− ∇ · (∇u) + ∇(∇ · u) = ∇ × (∇ × u),
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

14 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

Proposition 3.2. The nonlocal divergence, gradient, and curl operators and the
corresponding adjoint operators satisfy
Sandia National Labs SAND 2010-8353J

D C ∗ (u) = 0

for u : Rn → R3 (3.19a)
C D∗ (u) = 0

for u : Rn → R (3.19b)
G ∗ (u) = tr D∗ (u)

for u : Rn → Rk (3.19c)
D D∗ (u) − G G ∗ (u) = C C ∗ (u)
  
for u : Rn → R3 . (3.19d)

Proof. We prove (3.19d); the proofs of (3.19a)–(3.19c) are immediate after direct
substitution of the operators involved.
Let u : Rn → R3 . Then, by (3.11c), (3.14c), (3.18a), and (3.18b) and recalling
that α is an antisymmetric function, we have
Z  
D D∗ (u) − G G ∗ (u) = − (u0 − u) ⊗ α + (u − u0 ) ⊗ α0 · α dy
 
Rn
Z  
+ (u0 − u) · α + (u − u0 ) · α0 α dy
n
Z R
 
= −2 (u0 − u)(α · α) − (u0 − u) · α α dy
n
ZR 

= −2 α × (u0 − u) × α dy,
Rn

where, for the last equality, we have used the vector identity a × (b × c) = b(c · a) −
c(a · b). A simple computation shows that the last expression is equal to C C ∗ (u)
so that (3.19d) is proved.

Functions of the form C ∗ (u) do not entirely comprise the null space of the oper-
ator D. In fact,
 it is obvious that for any antisymmetric two-point function ν(x, y),
we have D ν = 0. However, functions of the form C ∗ (u) are the only symmetric
two-point functions belonging to the null space of D. Analogous statements can
be made for the null space of the operator C and two-point functions of  the form
∗ ∗
D (u). 
Note that, because of the nonlocality of the operators, G C µ 6= 0 and
C ∗ G η 6= 0.
Another set of results for the nonlocal operators that mimic obvious properties
of the corresponding differential operators are given in the following proposition

respectively. Because (∇·)∗ = −∇, ∇∗ = −∇·, and (∇×)∗ = (∇×), these identities can be written
in the form
∇ · (∇×)∗ u) = 0, ∇ × (∇·)∗ u) = 0, ∇∗ u = tr (∇·)∗ u ,
  

∇ · (∇·)∗ u − ∇ (∇)∗ u = ∇ × (∇×)∗ u


  

that more directly correspond to (3.19a)–(3.19d), respectively. In particular, the identities (3.19a),
(3.19b), and (3.19d) suggest that −D∗ , −G ∗ , and C ∗ can also be viewed as nonlocal analogs of the
differential gradient, divergence, and curl operators, respectively, that, when operating on point
functions, result in two-point functions.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 15

whose proof is a straightforward consequence of the results given in Proposition


3.1.
Sandia National Labs SAND 2010-8353J

Proposition 3.3. Let b and b denote a constant scalar and vector, respectively.
Then, the adjoints of the nonlocal divergence, gradient, and curl operators satisfy

D∗ (b) = 0, G ∗ (b) = 0, and C ∗ (b) = 0.

These results do not hold for the point divergence, gradient, and curl operators,
i.e., D(b), G(b), and C(b) do not necessarily vanish for constants b and constant
vectors b.

3.4.3. Nonlocal curl operators in two and higher dimensions


The nonlocal point and two-point curl operators defined in (3.11c) and (3.14c),
respectively, are defined in three dimensions. These definitions can be generalized to
arbitrary dimensions by replacing the vector cross product with the wedge product
so that, e.g., instead of (3.14c) we would have C ∗ (w) : Rn × Rn → Rr given by

C ∗ (w)(x, y) = γ ∧ (w0 − w),

where γ(x, y) : Rn × Rn → Rr with r a positive integer. This suggests a possible


exterior calculus-based formalism. However, such a formalism is beyond the scope
of this paper so that only the special cases of r = 3 and r = 2 are considered and
the vector cross product is retained.
We next define nonlocal point and two-point curl operators in two space dimen-
sions, i.e., for R2 ; in fact, we have two types of each kind of curl operator.j First,
we assume, without loss of generality, that µ · e3 = 0, w · e3 = 0, and γ · e3 = 0 in
n n 2
 Then, for a vector two-point function µ(x, y) : R × R → R ,
(3.11c) and (3.14c).
we can view C µ as the nonlocal scalar point function defined by
Z
(µ2 + µ02 )γ1 − (µ1 + µ01 )γ2 dy
 
C µ (x) = for x ∈ Rn
Rn

and, for a vector point function w : Rn → R2 , we can view C ∗ (w) as the nonlocal
scalar two-point function defined by

C ∗ (w)(x, y) = (w20 − w2 )γ1 − (w10 − w1 )γ2 for x, y ∈ Rn .

Next, assume, again without loss of generality, that µ = µe3 , w = we3 , and γ·e3 = 0
in (3.11c) and (3.14c). Then, for a scalar two-point function µ(x, y) : Rn × Rn → R,

j Thisis analogous to the two types of differential curl operators in two dimensions, one operating
on vectors, the other on scalars, respectively given by
∂u1 ∂u2 ∂u ∂u
curl u = − and curl u = − e1 + e2 .
∂x2 ∂x1 ∂x2 ∂x1
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

16 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

we can view C(µ) as the nonlocal vector point function defined by


Z
C(µ)(x) = (µ + µ0 )(γ2 e1 − γ1 e2 ) dy for x ∈ Rn
Sandia National Labs SAND 2010-8353J

Rn

and, for a scalar point function w(x) : Rn → R, we can view C ∗ (w) as the nonlocal
vector two-point function defined by

C ∗ (w)(x, y) = (w0 − w)(γ2 e1 − γ1 e2 ) for x, y ∈ Rn .

4. A nonlocal vector calculus


We develop a nonlocal vector calculus that mimics the classical vector calculus for
differential operators. In the classical calculus, interactions are local which is why,
e.g., in the Gauss theorem, the contribution coming from interactions with points
outside a domain Ω appears in the form of a flux through the boundary ∂Ω. If
interactions are nonlocal, then points outside of Ω interact with points in Ω so that
that interaction cannot be accounted for merely by an integral along the boundary
∂Ω. In fact, one must account for interactions with points in the complement domain
Rn \ Ω.
To treat the most general case, we introduce the notion of an interaction domain.

Definition 4.1. Given an open subset Ω ⊂ Rn , the interaction domain correspond-


ing to Ω is given by

ΩI := {y ∈ Rn \ Ω such that α(x, y) 6= 0 for some x ∈ Ω} (4.1)

so that ΩI consists of those points outside of Ω that interact with points in Ω.


For simplicity, we assume that the corresponding interactions domains for β
and γ are also given by (4.1). No assumption is made about the geometric relation
between Ω and ΩI so that, e.g., the four configurations of Figure 1 are possible.
We have that the situation ΩI = Rn \ Ω is allowable as is Ω = Rn in which case, of
course, ΩI = ∅.

ΩI
ΩI
Ω
Ω

ΩI
Ω Ω
ΩI

Fig. 1. Four of the possible configurations for Ω (the shaded regions) and ΩI (the white regions).
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 17

From (4.1), we see that points in Rn \ (Ω ∪ ΩI ) do not interact with points in


Ω, i.e.,
Sandia National Labs SAND 2010-8353J

α(x, y) = 0 for x ∈ Ω and y ∈ Rn \ (Ω ∪ ΩI ) (4.2)


and, by antisymmetry,
α(x, y) = 0 for y ∈ Ω and x ∈ Rn \ (Ω ∪ ΩI ).
We do not assume that points in ΩI interact only with points in Ω ∪ ΩI , i.e., (4.2)
does not hold for x ∈ ΩI , so that, in general, points in ΩI may interact with points
in Rn \ (Ω ∪ ΩI ) as well,.

4.1. Nonlocal interaction operators


We want to define operators that can be used to account for the interaction of
points in ΩI with points in Ω ∪ ΩI . To this end, we obtain, successively using the
antisymmetry of α and (2.4b), the linearity of the integral operator, (4.2), and
(3.11a),
Z Z
0= (ν + ν 0 ) · α dydx
Ω∪ΩI Ω∪ΩI
Z Z Z Z
0
= (ν + ν ) · α dydx + (ν + ν 0 ) · α dydx
Ω Ω∪ΩI ΩI Ω∪ΩI
Z Z Z Z (4.3)
= (ν + ν 0 ) · α dydx + (ν + ν 0 ) · α dydx
Ω Rn ΩI Ω∪ΩI
Z Z Z
= D(ν)dx + (ν + ν 0 ) · α dydx.
Ω ΩI Ω∪ΩI

We could not express the last term in terms of D because points in ΩI can interact
with points in Rn \ (Ω ∪ ΩI ). RBy (2.5), we identify the last term in (4.3) as the
0
flux from ΩI into Ω ∪ ΩI and Ω∪Ω R I (ν + ν ) · α dy as the flux density at x ∈ ΩI
into Ω ∪ ΩI . Thus, we could use Ω∪ΩI (ν +R ν 0 ) · α dy to define the operator we
are seeking. In (4.5a), we instead choose − Ω∪ΩI (ν + ν 0 ) · α dy for this purpose
because, using the antisymmetry of α, (2.1), and (2.4c),
Z Z Z Z
− (ν + ν 0 ) · α dydx = − (ν + ν 0 ) · α dydx
ΩI Ω∪ΩI Ω Ω
Z ZI (4.4)
= (ν + ν 0 ) · α dydx
Ω ΩI

i.e., it is convenient to have the integral over ΩI of the nonlocal operator we seek
to be the flux out of Ω into ΩI .
Similar discussions can be made for the nonlocal gradient and curl operators.
Thus, we define the nonlocal interaction operators as follows.

Definition 4.2. [Nonlocal interaction operators] Given a domain Ω ⊂ Rn ,


let the interaction domain ΩI be defined by (4.1). Then, corresponding to the
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

18 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou


point divergence operator D ν : Rn → R defined in (3.11a), we define the point
interaction operator N (ν) : ΩI → R through its action on ν by
Sandia National Labs SAND 2010-8353J

Z
(ν + ν 0 ) · α dy

N ν (x) := − for x ∈ ΩI . (4.5a)
Ω∪ΩI

Corresponding to the point gradient operator G η : Rn → Rk defined in (3.11b),
we define the point interaction operator S(η) : ΩI → Rk through its action on η by
Z
(η + η 0 )β dy

S η (x) := − for x ∈ ΩI . (4.5b)
Ω∪ΩI

Corresponding to the point curl operator C µ : Rn → R3 defined in (3.11c), we
define the point interaction operator T (µ) : ΩI → R3 through its action on µ by
Z
γ × (µ + µ0 ) dy

T µ (x) := − for x ∈ ΩI . (4.5c)
Ω∪ΩI

4.2. Nonlocal integral theorems


Definitions 3.1 and 4.2 lead to integral theorems of the nonlocal vector calculus that
mimic the basic integral theorems of the vector calculus for differential operators.

Theorem 4.1. [Nonlocal integral theorems] Assuming the notations and def-
initions found in Definitions 3.1 and 4.2, we havek
Z Z
D(ν) dx = N (ν) dx (4.6a)
Ω ΩI
Z Z
G(η) dx = S(η) dx (4.6b)
Ω Ω
Z Z I
C(µ) dx = T (µ) dx. (4.6c)
Ω ΩI

Proof.
R The proof is contained in (4.3) because, by (4.5a), the last term is equal to
− ΩI N (ν) dx. In a similar manner, (4.6b) and (4.6c) can be derived. Alternately,
they can be derived from (4.6a); one simply chooses ν = ηb and ν = b × µ,
respectively, in that equation, where b is a constant vector; one also has to associate
β and γ with α.
R 
From (4.4), we have that ΩI N ν dx is the nonlocal flux from Ω into ΩI . Thus,
in words, (4.6a) states that the integral of the nonlocal divergence of ν over Ω is

k The nonlocal integral theorems (4.6) are analogous to the classical differential integral theorems
given by
Z Z Z Z Z Z
∇ · v dx = v · n dx, ∇v dx = vn dx, and ∇ × v dx = n × v dx
Ω ∂Ω Ω ∂Ω Ω ∂Ω

for functions v and v defined on Rn , with n = 3 for the third one, for which the integrals are well
defined.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 19

equal to the total flux out of Ω into ΩI .l Similar interpretations of (4.6b) and (4.6c)
hold.
Sandia National Labs SAND 2010-8353J

The nonlocal integral theorems (4.6a)–(4.6c) are action-reaction principles. For


example, from (4.6a) and the assumptions made, we have that

Z Z
 
0= D ν dx − N ν dx
Ω ΩI
Z Z Z Z
0
= (ν + ν ) · α dydx + (ν + ν 0 ) · α dydx (4.7)
Ω Rn ΩI Ω∪ΩI
Z Z Z Z
0
= (ν + ν ) · α dydx + (ν + ν 0 ) · α dydx
Ω ΩI ΩI Ω

which is exactly of the form (2.4c) with Ω1 = Ω, Ω2 = ΩI , and ψ = (ν + ν 0 ) · α;


(4.7) simply states that the flux out of Ω into ΩI is equal and opposite to the flux
from ΩI into Ω.
The nonlocal integral theorems (4.6a)–(4.6c) have the simple consequences listed
in the following corollary that can be viewed as providing nonlocal integration by
parts formulas.

Corollary 4.1. [Nonlocal integration by parts formulas] Adopt the hypothe-


ses of Theorem 4.1 and let the nonlocal adjoint operators be given as in Propo-
sition 3.1. Then, given the point functions u(x) : Rn → R, v(x) : Rn → Rk , and
w(x) : Rn → R3 , we havem

Z Z Z Z
D∗ u · ν dydx =
 
uD ν dx − uN (ν) dx (4.8a)
Ω Ω∪ΩI Ω∪ΩI ΩI
Z Z Z Z
G ∗ v η dydx =
 
v · G η dx − v · S(η) dx (4.8b)
Ω Ω∪ΩI Ω∪ΩI ΩI
Z Z Z Z
C ∗ w · µ dydx =
 
w · C µ dx − w · T (µ) dx. (4.8c)
Ω Ω∪ΩI Ω∪ΩI ΩI

l This
R
R observation is analogous to the observation for the classical Gauss theorem Ω ∇ · v =
∂Ω v · n dA that, by (2.1), the integral of the local divergence of v over Ω is equal to the total
flux out of Ω.
m If ν and α are scalar-valued functions and for free space, a version of the integration by parts

formula (4.8a) appears in Lemma 2.1 of Ref. 13.


The classical analog of (4.8a) is given by, for a scalar function u(x) and vector function v(x),
Z Z Z
u∇ · v dx + v · ∇u dx = uv · n dA
Ω Ω ∂Ω

and similarly for (4.8b) and (4.8c).


March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

20 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

Proof. Let ξ = uν; then, from (3.11a) and (3.15), we have


Z  
D(ξ) = u(ν + ν 0 ) · α + (u0 − u)ν 0 · α dy
Sandia National Labs SAND 2010-8353J

Rn
Z Z
=u (ν + ν 0 ) · α dy + (u0 − u)ν 0 · α dy
Rn Rn
Z (4.9)
= uD(ν) + (u0 − u)ν 0 · α dy
Rn
Z
= uD(ν) + (u0 − u)ν 0 · α dy ∀ x ∈ Ω,
Ω∪ΩI

where the last equality follows from (4.1). Also, from (3.15) and (4.5a), we have
Z  
N (ξ) = − u(ν + ν 0 ) · α + (u0 − u)ν 0 · α dy
Ω∪ΩI
Z (4.10)
= uN (ν) − (u0 − u)ν 0 · α dy ∀ x ∈ ΩI .
Ω∪ΩI

Then,
Z Z

0= D(ξ) dx − N ξ dx
Ω ΩI
Z Z

= uD(ν) dx − uN ν dx
Ω ΩI
Z Z Z Z
+ (u − u)ν 0 · α dydx +
0
(u0 − u)ν 0 · α dydx
Ω Ω∪ΩI ΩI Ω∪ΩI
Z Z Z Z
(u0 − u)ν 0 · α dydx,

= uD(ν) dx − uN ν dx +
Ω ΩI Ω∪ΩI Ω∪ΩI

where the first equality follows from (4.6a), the second from (4.9) and (4.10), and
the third from the linearity of the integration operator. Then, (4.8a)
R follows
R from the
antisymmetry of α, the fact that, similar to (3.17), we have that Ω∪ΩI Ω∪ΩI (u0 −
u)(ν − ν 0 ) · α dydx = 0, and (3.14a).
In a similar manner, (4.8b) and (4.8c) can be derived from (4.6a) along with
(3.11b) and (3.11c), respectively. Alternately, (4.8b) and (4.8c) easily follow by
setting, for an arbitrary constant vector b, ν = ηb and ν = b × µ, respectively, in
(4.8a) and also associating β and γ with α.

4.3. Nonlocal Green’s identities


Nonlocal (generalized) Green’s identities are simple consequences of Corollary 4.1.
For the following two corollaries, we carry over the notations, definitions, and results
obtained above.

Corollary 4.2. [Nonlocal (generalized) Green’s first identities] Given the


scalar point functions u(x) : Rn → R and v(x) : Rn → R and the two-point second-
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 21

order tensor function Θ(x, y) : Rn × Rn → Rk×k , thenn


Z Z Z
uD Θ · D∗ (v) dx − D∗ (u) · Θ · D∗ (v) dydx
 
Sandia National Labs SAND 2010-8353J

Ω Ω∪ΩI Ω∪ΩI
Z (4.11a)
uN Θ · D∗ (v) dx.

=
ΩI

Given the vector point functions u(x) : Rn → Rk and v(x) : Rn → Rk and the
two-point scalar function θ(x, y) : Rn × Rn → R, then
Z Z Z

θG ∗ (v)G ∗ (u) dydx

v · G θG (u) dx −
Ω Ω∪ΩI Ω∪ΩI
Z (4.11b)
v · S θG ∗ (u) dx.

=
ΩI

Given the vector point functions u(x) : Rn → R3 and w(x) : Rn → R3 and the
two-point second-order tensor function Θ(x, y) : Rn × Rn → R3×3 , then
Z Z Z
w · C Θ · C ∗ (u) dx − C ∗ (w) · C Θ · C ∗ (u) dydx
 
Ω Ω∪ΩI Ω∪ΩI
Z (4.11c)
w · T Θ · C ∗ (u) dx.

=
ΩI

Proof. The results (4.11a)–(4.11c) follow by setting ν = Θ · D∗ (u) in (4.8a), η =


θG ∗ (u) in (4.8b), and µ = Θ · C ∗ (u) in (4.8c), respectively.

Corollary 4.3. [Nonlocal (generalized) Green’s second identities] We as-


sume the same notation as in Corollary 4.2 and we also assume that the tensor Θ
appearing in (4.11a) and (4.11c) is symmetric. Then,
Z Z

vD Θ · D∗ (u) dx
 
uD Θ · D (v) dx −
Ω Ω
Z Z (4.12a)
uN Θ · D∗ (v) dx − vN Θ · D∗ (u) dx
 
=
ΩI ΩI

n Wehave that (4.11a) and (4.12a) are the nonlocal analogs of the local classical (generalized) first
Green’s identity
Z Z Z
u∇ · (C · ∇v) dx + ∇u · (C · ∇v) dx = un · (C · ∇v) dA
Ω Ω ∂Ω

and of the local classical (generalized) second Green’s identity


Z Z Z Z
u∇ · (C · ∇v) dx − v∇ · (C · ∇u) dx = un · (C · ∇u) dA − vn · (C · ∇u) dA,
Ω Ω ∂Ω ∂Ω

respectively, and similarly for (4.11b), (4.12b), (4.11c) and (4.12c). Here, C(x) denotes a second-
order tensor function with C(x) symmetric for the second identity. These are “generalizations” of
the classical local Green’s identities for which C is the identity tensor.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

22 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

Z Z
v · G θG ∗ (u) dx − u · G θG ∗ (v) dx
 
Ω Ω
(4.12b)
Sandia National Labs SAND 2010-8353J

Z Z

u · S θG ∗ (v) dx
 
= v · S θG (u) dx −
ΩI ΩI
Z Z
w · C Θ : C ∗ (u) dx − u · C Θ : C ∗ (w) dx
 
Ω Ω
Z Z (4.12c)
w · T Θ : C ∗ (u) dx − u · T Θ : C ∗ (v) dx.
 
=
ΩI ΩI

Proof. The result (4.12a) is obtained by reversing the roles of u and v in (4.11a)
and then subtracting the result from (4.11a). The results (4.12b) and (4.12c) are
obtained from (4.11b) and (4.11c), respectively, in a similar manner.

4.4. Special cases of the vector calculus


We consider some special cases of the vector calculus developed in Sections 4.1–4.3.

4.4.1. The free space vector calculus


As was mentioned previously, the vector calculus allows for the choice Ω = Rn in
which case we also simply set ΩI = ∅ in all definitions and results. The resulting
nonlocal integral theorems have already been stated; see (3.12). Corresponding to
the nonlocal divergence operator D, we have the nonlocal generalized Green’s first
identity
Z Z Z

D∗ (u) · Θ · D∗ (v)ν dydx = 0

uD Θ · D (v) dx −
Rn Rn Rn
and the generalized nonlocal Green’s second identity
Z Z
uD Θ · D∗ (v) dx − vD Θ · D∗ (u) dx = 0
 
Rn Rn

corresponding to (4.11a) and (4.12a), respectively, and similarly for the other oper-
ators.

4.4.2. The vector calculus for interactions of infinite extent


In case points in Ω interact with all points in Rn , we have that ΩI = Rn \ Ω
and all definitions and results remain unchanged. In this case, N (ν) = −D(ν) on
ΩI = Rn \ Ω.

4.4.3. The vector calculus for localized kernels


An important special case is that of localized kernels for which we have that
α(x, y) = 0 if |y − x| ≥ ε (4.13)
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 23

and similarly for β and γ; here, ε > 0 denotes a cut-off or interaction or horizono
parameter which is not necessarily small and which defines the extent of interactions.
Sandia National Labs SAND 2010-8353J

We then have that


ΩI = {y ∈ Rn \ Ω : |y − x| < ε for x ∈ Ω}.
We also define the domain
Ω− := {x ∈ Ω : |y − x| < ε for y ∈ Rn \ Ω} = {x ∈ Ω : |y − x| < ε for y ∈ ΩI },
where the last equality follows from (4.13). See Figure 2 for an illustration.

ΩI
ε
Ω
ε Ω-

Fig. 2. For a localized kernel, the domain Ω and the interaction regions ΩI and Ω− whose thick-
nesses are given by the horizon ε.

From (4.1), (4.5a), and (4.13), we have that


Z Z Z Z Z
N (ν) dx = − (ν + ν 0 ) · α dydx = − (ν + ν 0 ) · α dydx
ΩI Ω Ω∪ΩI ΩI Ω
Z ZI Z Z
0
= (ν + ν ) · α dydx = (ν + ν 0 ) · α dydx.
Ω ΩI Ω− ΩI

The first term is the nonlocal flux from Ω into ΩI ; according to the last term, only
points in Ω− ⊂ Ω contribute to that flux. Similar observations can be made for the
operators S and T .
Of course, all the definitions and results of Sections 3.2–4.3 hold for the case of
localized kernels.

5. Relations between nonlocal and differential operators


In this section, we choose particular kernel functions α, β, and γ that lead to iden-
tifications of the nonlocal operators with their differential counterparts. In Section
5.1, the identification is made in a distributional sense whereas, in Section 5.2,
we demonstrate circumstances under which the nonlocal point operators are weak

o Theterminology “horizon” is introduced in the context of cutt-off parameters in peridynamic


models for continuum mechanics; see, e.g., Refs. 21 and 24.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

24 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

representations of the divergence, gradient, and curl differential operators. For the
sake of brevity, we mostly consider the nonlocal point divergence operator D and its
adjoint operator D∗ . However, analogous results also hold for the nonlocal gradient
Sandia National Labs SAND 2010-8353J

and curl operators G and C, respectively, and their adjoints. We then consider, in
Section 5.3, a connection between the nonlocal and local Gauss theorems made with
the help of two results given in Ref. 17.
Throughout this section, we set k = n in Proposition 3.1, Theorem 3.1, and all
subsequent results.

5.1. Identification of nonlocal operators with differential operators


in a distributional sense
We begin with the identification of the nonlocal point divergence operator D with
the divergence differential operator ∇·. This identification is subject to the under-
standing that the former operates on two-point functions whereas the latter operates
on point functions. Where there might be ambiguity in applying differential opera-
tors to two-point function, we identify explicitly the variable of differentiation, e.g.,
∇y · ν(x, y) denotes the divergence of ν(x, y) with respect to y.
We consider special cases of localized kernels (see Section (4.4.3)) for which
ε → 0. As such, the results and “proofs” provided in this subsection are purely
formal excepting Proposition 5.1; a limiting process involving sequences of kernels
similar to that presented in the proof of Proposition 5.1 would make all other proofs
rigorous.
We first relate, the nonlocal divergence with the local differential divergence.
Let ρε = ρε (x) be a standard Gaussian with variance ε.

Proposition 5.1. Let ν ∈ [C0∞ (Rn × Rn )]n , i.e., the components of ν belongs
to the space of compactly supported infinitely differentiable functions. Select the
(antisymmetric) distribution

α(x, y) = −∇y ρε (y − x), (5.1)

where ∇y denotes the differential gradient with respect to y. Then,

∀ x ∈ Rn ,

lim D ν (x) = ∇ · ν(x, x) (5.2)
ε→0

where ∇· denotes the differential divergence operator.

Proof. First, by the chain rule, we have


 
∇ · ν(x, x) = ∇x · ν(x, y) |y=x + ∇y · ν(x, y) |y=x
 
= ∇y · ν(y, x) |y=x + ∇y · ν(x, y) |y=x

= ∇y · ν(y, x) + ν(x, y) |y=x .
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 25

Then, given the convergence of ρε to the Dirac delta measure δ in the sense of
distributions, we obtain
Sandia National Labs SAND 2010-8353J


∇ · ν(x, x) = ∇y · ν(y, x) + ν(x, y) |y=x
Z

= lim ∇y · ν(y, x) + ν(x, y) ρε (y − x) dy
ε→0 Rn
Z

= − lim ν(y, x) + ν(x, y) ∇y ρε (y − x) dy
ε→0 Rn

which gives (5.4) by the definition given in (3.11a).

For any x ∈ Rn , given that in the sense of distributions ∇y ρε (y − x) converges


to ∇y δ(y − x) with δ the Dirac delta measure, we may let

α(x, y) = −∇y δ(y − x), (5.3)

and rewrite (5.2) formally as

∀ x ∈ Rn .

D ν (x) = ∇ · ν(x, x) (5.4)

We may further extend (5.4) in the sense of distributions to a more general function
space for ν, again via a passage to limit or density argument.
Along the same spirit, the following proposition relates D∗ formally to −∇ =
(∇·)∗ .

Proposition 5.2. Let u ∈ C0∞ (Rn ) and select α as in (5.3). Then, for any ν ∈
[C ∞ (Rn × Rn )]n ,
Z
D∗ (u) · ν dy = −ν(x, x) · ∇u(x) ∀ x ∈ Rn . (5.5)
Rn

Proof. Starting with (3.14a) and (5.3), we have that


Z Z


D (u) · ν dy = u(y) − u(x) ν(x, y) · ∇y δ(y − x) dy
R n R n
Z  
=− δ(y − x)∇y · ν(x, y) u(y) − u(x) dy
n
ZR
=− δ(y − x)ν(x, y) · ∇y u(y) dy
Rn
= −ν(x, x) · ∇u(x) = ν(x, x) · (∇·)∗ u(x)

Choosing ν to be the natural Cartesian unit vectors in Rn , we have, from (5.5),


that
Z
D∗ (u) dy = −∇u(x) = (∇·)∗ u(x) ∀ x ∈ Rn , (5.6)
Rn
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

26 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

i.e., the average of the action of D∗ on a point function is related, in a distributional


sense, to the action of −∇ on that function.p
We next relate, in a distributional sense, the interaction operator N with the
Sandia National Labs SAND 2010-8353J

normal component along the boundary.

Proposition 5.3. Assume the hypotheses of Proposition 5.2. Then, for all u(x) ∈
C0∞ (Rn ),
Z Z
uN (ν) dx = u(x)ν(x, x) · n dA, (5.7)
ΩI ∂Ω

where n denotes the unit outward-pointing normal vector along ∂Ω. Thus, N (ν) is
a delta measure concentrated on ∂Ω with coefficient (weight) ν(x, x) · n.

Proof. With α chosen as in (5.3) and using the definition (3.14) for the operator
D∗ , we have that
Z Z Z Z
D∗ (u) · ν dydx =

− u(y) − u(x) ν(x, y) · α(x, y) dydx
Ω∪ΩI Ω∪ΩI Ω∪ΩI Ω∪ΩI
Z Z

=− u(y) − u(x) ν(x, y) · ∇y δ(y − x) dydx
Ω∪ΩI Ω∪ΩI
Z Z

= ∇y u(y) − u(x) · ν(x, y)δ(y − x) dydx
Ω∪ΩI Ω∪ΩI
Z Z

+ u(y) − u(x) δ(y − x)∇y · ν(x, y) dydx
Ω∪ΩI Ω∪ΩI
Z Z

− u(y) − u(x) δ(y − x)ν(x, y) · n dydx
Ω∪ΩI ∂(Ω∪ΩI )
Z Z
= ∇y u(y) · ν(x, y)δ(y − x) dydx
Ω∪ΩI Ω∪ΩI
Z
= ∇u(x) · ν(x, x) dx,
Ω∪ΩI

where the fourth equality follows because u(y) − u(x) = 0 whenever y = x and
because ∇y u(x) = 0. Also, with α chosen as in (5.3) and using (5.4), we have that
Z Z Z Z
uD(ν) dx = u(x)∇·ν(x, x) dx = u(x)ν(x, x)·n dA− ∇u(x)·ν(x, x) dx
Ω Ω ∂Ω Ω

p We see that α(x, y) as given by (5.3) is antisymmetric. Had we chosen α to be a symmetric

distribution or a distribution having an non-vanishing symmetric part, we would not have obtained
the results (5.4) and (5.5), and (5.6). This provides at least partial justification for the assumption,
made at the end of Section 3.1, that α(x, y) should be antisymmetric.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 27

Substituting the last two results into (4.8a) results in


Z Z Z Z
uN (ν) dx = uD(ν) dx − D∗ (u) · ν dydx
Sandia National Labs SAND 2010-8353J

ΩI Ω Ω∪ΩI Ω∪ΩI
Z
= u(x)ν(x, x) · n dA (5.8)
∂Ω
Z Z
+ ∇u(x) · ν(x, x) dx − ∇u(x) · ν(x, x) dx.
Ω∪ΩI Ω

Because α is a singular measure, the measure of ΩI vanishes, i.e., it becomes a


lower-dimensional domain, in fact, just the boundary ∂Ω of Ω, so that the last two
terms in (5.8) cancel each other out and we obtain (5.7).

Now consider the composition of the nonlocal divergence operator and its ad-
joint, i.e., D D∗ , which, according to the next proposition, can be identified, in the
sense of distributions, with −∆ = ∇ · (−∇), where ∆ denotes the Laplace operator.

Proposition 5.4. Let u ∈ C0∞ (Rn ) and select |α(x, y)|2 = α(x, y) · α(x, y)=
1
2 ∆y δ(y − x). Then,

D D∗ u(x) = −∆u(x) ∀ x ∈ Rn .

(5.9)

Proof. From (3.11a) and (3.14a) we have


Z
D D∗ u (x) = −2 u(y) − u(x) |α(x)|2 dy
 
n
Z R

=− u(y) − u(x) ∆y δ(y − x) dy
n
ZR

=− δ(y − x)∆y u(y) − u(x) dy
n
ZR
=− δ(y − x)∆y u(y) dy
Rn
= −∆u(x) = −∇ · ∇u(x) = ∇ · (∇·)∗ u(x),
where the differential Green’s second identity is used for the third equality.

We again note that the relations between nonlocal and local differential operators
given in (5.4)–(5.7) and (5.9) are for specially selected α.

5.2. Relations between weighted nonlocal operators and weak


representations of differential operators
The classical differential calculus involves only operators mapping point functions
to point functions; on the other hand, the nonlocal operators defined earlier map
two-point functions to point functions or, in case of the nonlocal adjoint operators,
point functions to two-point functions. To further demonstrate that the nonlocal
operators correspond to nonlocal versions of the classical divergence, gradient, and
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

28 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

curl differential operators, we use the nonlocal operators D, G, and C to define, in


Section 5.2.1, corresponding nonlocal weighted operators that map point functions
Sandia National Labs SAND 2010-8353J

to point functions. We also show that the adjoint operators corresponding to the
weighted operators are weighted integrals of the nonlocal adjoint operators D∗ , G ∗ ,
and C ∗ . Then, in Section 5.2.2, the weighted operators are rigorously shown to be
nonlocal versions of the corresponding differential operators.

5.2.1. Nonlocal weighted operators


The nonlocal point operators defined in (3.11) induce new operators, referred to
as weighted operators. Throughout this section, we asssume that α = β = γ.
Definitions 5.1 and 5.2 and Proposition 5.5 hold for general k and m although,
when we apply them in Section 5.2.2, we set k = m = n.

Definition 5.1. Let ω(x, y) : Rn × Rn → R denote a non-negative scalar-valued


two-point function. Let the operators D, G, and C be defined as in (3.11). Then,
given the point function u(x) : Rn → Rk , the weighted nonlocal divergence operator
Dω (u) : Rn → R is defined by its action on u by

Dω (u)(x) := D ω(x, y)u(x) for x ∈ Rn . (5.10a)
Given the point function v(x) : Rn → R, the weighted nonlocal gradient operator
Gω (v) : Rn → Rk is defined by its action on u by

Gω (v)(x) := G ω(x, y)v(x) for x ∈ Rn . (5.10b)
Given the point function w(x) : Rn → R3 , the weighted nonlocal curl operator
Cω (u) : Rn → R3 is defined by its action on w by

Cω (w)(x) := C ω(x, y)w(x) for x ∈ Rn . (5.10c)
The following result shows that the adjoint operators corresponding to the
weighted nonlocal operators are determined as weighted integrals of the correspond-
ing nonlocal two-point adjoint operators (3.14).

Proposition 5.5. Let ω(x, y) : Rn × Rn → R denote a non-negative scalar two-


point function and let D∗ , G ∗ , and C ∗ denote the nonlocal adjoint operators given
in (3.14). The action on u(x) of the adjoint operator Dω∗ (u)(x) : Rn → Rk corre-
sponding to the weighted nonlocal divergence operator Dω is given by
Z
Dω∗ (u)(x) = D∗ (u)(x, y) ω(x, y) dy for x ∈ Rn (5.11a)
Rn

for scalar point functions u(x) : Rn → R. The action on v(x) of the adjoint operator
Gω∗ (v)(x) : Rn → R corresponding to the weighted nonlocal gradient operator Gω is
given by
Z
Gω∗ (v)(x) = G ∗ (v)(x, y) ω(x, y) dy for x ∈ Rn (5.11b)
Rn
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 29

for vector point functions v(x) : Rn → Rk . The action on w(x) of the adjoint
operator Cω∗ (w)(x) : Rn → R3 corresponding to the weighted nonlocal curl operator
Sandia National Labs SAND 2010-8353J

Cω is given by
Z

Cω (w)(x) = C ∗ (w)(x, y) ω(x, y) dy for x ∈ Rn (5.11c)
Rn

for vector point functions w(x) : Rn → R3 .

Proof. We have that


Z
 
Dω (u), u Rn = u(x)D ω(x, y)u(x) dx
n
ZR Z
D∗ (u)(x, y) · ω(x, y)u(x) dydx

=
n n
ZR RZ 

= ω(x, y)D (u)(x, y) dy · u(x) dx,
Rn Rn

where (5.10a) is used for the first equality and (3.13) and (3.14a) for the second.
But, by definition, the adjoint operator Dω∗ (·) corresponding to the operator Dω (·)
satisfies

u, Dω (u) Rn = Dω∗ (u), u Rn .


 

Comparing the last two results yields (5.10a). The conclusions (5.10b) and (5.10c)
are derived in a similar fashion.

The definition given in (5.10a) and the result (5.11a) can be extended to tensors
and vectors, respectively.

Definition 5.2. Let ω(x, y) : Rn × Rn → R denote a non-negative scalar two-


point function and let D be given as in (3.18b). Given the tensor point function
U(x) : Rn → R`×k , the weighted nonlocal divergence operator Dω (U) : Rn → R` for
tensors is defined by
 
Dω U (x) := D ω(x, y)U(x) (x) for x ∈ Rn . (5.12)

Corollary 5.1. The adjoint operator Dω∗ (u)(x) : Rn → R`×k corresponding to the
operator Dω is given by
Z

D∗ (u)(x, y) ω(x, y) dy for x ∈ Rn

Dω u (x) = (5.13)
Rn

for vector point functions u : Rn → R` .

Similar results hold for the operator Gv .


March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

30 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

5.2.2. Relationships between weighted operators and differential operators


We have seen that the nonlocal operators satisfy many properties that mimic their
Sandia National Labs SAND 2010-8353J

differential counterparts. In particular, we established, in Section 5.1, that the nonlo-


cal point operators can be identified with the corresponding differential operators in
a distributional sense. Here, we establish rigorous connections between the weighted
operators and their differential counterparts for the case of special localized kernels
(see Section 4.13), i.e., we introduce the horizon parameter ε > 0 and analyze what
occurs as ε → 0, that is, in the local limit. To this end, we define
Bε (x) := {y ∈ Rn : |y − x| < ε}
and let φ denote a positive radial function satisfying the normalization condition
Z
|y − x|2 φ(|y − x|) dy = n, (5.14)
Bε (x)

where n denotes the space dimension. We then chooseq

 y−x
 α(x, y) = for x 6= y
|y − x|



( (5.15)
|y − x|φ(|y − x|) y ∈ Bε (x)
 ω(x, y) =



0 otherwise.
We have that α is an antisymmetric functionr whereas ω is a symmetric  function.
We then have, for a scalar function u, that the components of Gω u and Dω∗ u
given by (5.10b) and (5.11a), respectively, are given by, for j = 1, . . . , n,
Z

dj u(x) := u(x + z) + u(x) zj φ(|z|) dz (5.16a)
Bε (0)
Z
d∗j u(x) := −

u(x + z) − u(x) zj φ(|z|) dz, (5.16b)
Bε (0)

where zj denotes the j-th component of z.


The following result ensues.

Lemma 5.1. Assume that ω and α are defined as in (5.15). Let u : Rn → R and
let dj u(x) and d∗j u(x) denote the j-th components of Gω (u) and Dω∗ (u), respectively.
Then,
dj u = −d∗j u. (5.17)

q Synergistic with the results of Section 5.1, if we instead choose the singular distribution
ω(|z|)αj (z) = zj φ(|z|) = −∂j δ(z) for j = 1, 2, 3,
where ∂j u denotes the weak derivative of u with respect to xj , we obtain
dj = ∂j and d∗j = −∂j .

r Again, the antisymmetry of α(x, y) is necessary for the results of this section, giving further justi-
fication for the assumption, made at the end of Section 3.1, that α(x, y) should be antisymmetric.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 31

Proof. From (5.16) we have


Z
Sandia National Labs SAND 2010-8353J

dj u(x) + d∗j u(x) = −2u(x) zj φ(|z|)dz = 0


Bε (0)

from which the result directly follows.

Based on this lemma and the definition of the weighted operators, we have the
following results.

Corollary 5.2. Under the same conditions as in Lemma 5.1, we have

Dω = −Gω∗ , Gω = −Dω∗ , and Cω = Cω∗ . (5.18)

Thus, under the hypotheses of this corollary, the identities in (5.18) mimic their
counterparts in the vector calculus for differential operators. Identities such as those
in Proposition 3.2 do not generally hold for the weighted operators due to their
nonlocal properties.s
The following proposition demonstrates that the components of the weighted
gradient operator Gω or, by (5.18), Dω∗ , converge, as ε → 0, to the corresponding
spatial derivatives.

Proposition 5.6. Let ω be defined as in (5.15) with φ satisfying (5.14). Then, for
j = 1, . . . , n, the weighted operators dj and d∗j defined by (5.16) are bounded linear
operators from H 1 (Rn ) to L2 (Rn ). Moreover, if u ∈ H 1 (Rn ), then as ε → 0

kdj u − ∂j ukL2 (Rn ) → 0 (5.19a)


kd∗j u + ∂j ukL2 (Rn ) → 0, (5.19b)

where ∂j u denotes the weak partial derivative of u with respect to xj . Moreover, if


Z
|z|1+s φ(|z|) dz < ∞ for some 0 ≤ s ≤ 1, (5.19c)
Bε (0)

then, for j = 1, . . . , n, the weighted operators dj and d∗j are bounded linear operators
from H t (Rn ) to H t−s (Rn ) for any t ≥ 0.

Proof. The conclusion (5.19a) is equivalent to kdd j u − ∂j ukL2 (Rn ) → 0, where dj u


d d
and ∂j u denote the Fourier transforms of dj u and ∂j u, respectively.
d

s Theidentities in Lemma 5.1 and Corollary 5.2 also hold for more general choices of α and ω,
namely, those such that α is antisymmetric and ω is a radial function with support in Bε (x).
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

32 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

Under the condition (5.19c) for s ∈ [0, 1], if u ∈ H t (Rn ) for t ≥ 0, then
Z
Sandia National Labs SAND 2010-8353J

dd
ju = (eiy·ξ + 1) u
b(ξ) yj φ(|y|) dy
Bε (0)
Z
=i sin(y · ξ) u
b(ξ) yj φ(|y|) dy
Bε (0)
Z  X
=i sin(yj ξj ) cos( yk ξk )
Bε (0) k6=j
X 
+ cos(yj ξj ) sin( yk ξk ) ub(ξ) yj φ(|y|) dy
k6=j
Z X
= ib
u(ξ) sin(yj ξj ) cos( yi ξk ) yj φ(|y|) dy,
Bε (0) k6=j

where the second and fourth equalities hold because of the symmetry of the domain
of integration and the antisymmetry of the integrand with respect to the integration
P
variable, respectively. Because for any 0 ≤ s ≤ 1, | sin(yj ξj ) cos( k6=j yi ξk )| ≤
|yj ξj |s , we have
Z
s
|dd
j u| ≤ |ξ ub(ξ)| |yj |1+s φ(|y|) dy
Bε (0)

t−s
so that ddju ∈ H (Rn ). In particular, for t = s = 1, we obtain that, under the
condition (5.14), the weighted operators dj , j = 1, . . . , n, defined in (5.16a) are
bounded linear operators from H 1 (Rn ) to L2 (Rn ).
If u ∈ H 1 (Rn ), we have that

|dd
j u − ∂j u| ≤ |dj u| + |∂j u|
d d d
Z X

≤ sin(yj ξj ) cos( b(ξ) yj φ(|y|) dy + |ξj u
yk ξ k ) u b|
Bε (0) k6=j
Z
≤ |ξj u
b| yj2 φ(|y|) dy + |ξj u
b|
Bε (0)

b| ∈ L2 (Rn ),
≤ 2|ξj u

where the third inequality holds because | sin(x)| ≤ |x| and | cos(x)| ≤ 1. By Taylor’s
theorem, we have
Z X
dj u = i
d (yj ξj + yj3 ξj3 cos(θ1 )/6)(1 + ( yk ξk )2 cos(θ2 )/2) u
b yj φ(|y|) dy
Bε (0) k6=j
Z
= iξj u
b yj2 φ(|y|) dy + Iε (ξ) = iξj u
b + Iε (ξ),
Bε (0)
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 33

where, for some θ1 and θ2 ,


Z
i
Iε (ξ) := y 4 cos(θ1 )φ(|y|) dy ξj3 u
Sandia National Labs SAND 2010-8353J

6 Bε (0) j
b
Z
i X
+ yj2 ( yi ξi )2 cos(θ2 )φ(|y|) dy ξj u
b
2 Bε (0)
k6=j
Z
i X
+ yj4 ( yk ξk )2 cos(θ1 ) cos(θ2 )φ(|y|) dy ξj3 u
b.
12 Bε (0)
k6=j

Because ub is bounded a.e., we see that, for any ξ,


X X
|Iε (ξ)| ≤ ε2 |ξj3 u
b| + ε2 |( ξk )2 ξj u
b| + ε4 |( ξk )2 ξj3 u
b| → 0 as ε → 0.
k6=j k6=j

Hence, we have
n
j u → ∂j u a.e. for ξ ∈ R
dd as ε → 0.
d

Then, by the dominated convergence theorem, we obtain


kdj u − ∂j ukL2 (Rn ) → 0 as ε → 0.
Lemma 5.1 implies the same results hold for the operator d∗j .

The above lemma implies that if ω satisfies (5.19c) for s = 0, then the weighted
operators are bounded operators from L2 (Rn ) to L2 (Rn ). More generally, for φ
satisfying (5.19c) with positive s, the operators dj and d∗j actually map a subspace
of L2 (Rn ), for instance the fractional Sobolev space H s (Rn ), to L2 (Rn ), or even
map L2 (Rn ) to H −s (Rn ). We refer to Ref. 7 for related work.
A direct consequence of Lemma 5.6 is the following result.

Corollary 5.3. Under the condition of Lemma 5.6, the weighted operators Dω , Gω ,
and Cω and their adjoint operators Dω∗ , Gω∗ , and Cω∗ are bounded linear operators from
H t (Rn ) to H t−s (Rn ) for 0 ≤ s ≤ 1, where n = 3 for the weighted curl operators.
Moreover, if u ∈ H 1 (Rn ) and u ∈ [H 1 (Rn )]n , then
Dω (u) → ∇ · u Dω∗ (u) → −∇u

Gω (u) → ∇u Gω∗ (u) → −∇ · u (5.20)

Cω (u) → ∇ × u Cω∗ (u) → ∇ × u,


where the convergence as ε → 0 is with respect to L2 (Rn ).

Proof. We prove the convergence of the operator Dω (u) to the divergence differ-
ential operators. First, note that
n
X n
X
Dω (u) = di ui and ∇·u= ∂i ui
i=1 i=1
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

34 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

so that
n
Sandia National Labs SAND 2010-8353J

X
kDω (u) − ∇ · ukL2 (Rn ) ≤ kdi ui − ∂i ui kL2 (Rn ) → 0 as ε → 0.
i=1

The remaining results can be proved in a similar fashion.

For the purpose of discussing nonlocal equations, we also need to consider com-
binations of the weighted operators such as Dω (C1 · Dω∗ (u)), where C1 (x) is a
“constitutive” tensor point function that, for example, describes a point property
of a material. Note that the two-point property is involved in the definition of the
weighted operators. In the next corollary, we illustrate that whenever the horizon
ε goes to zero, the combinations of the weighted operators converge to their local
counterparts.

Corollary 5.4. Let u ∈ H 1 (Rn ), u ∈ [H 1 (Rn )]n , C1 : Rn → Rn × Rn in L∞ (Rn ×


Rn ), and c2 : Rn → R in L∞ (Rn ). Then,

Dω C1 · Dω∗ (u) → −∇ · (C1 · ∇u)




Gω c2 Gω∗ (u) → −∇(c2 ∇ · u)



(5.21)
Cω C1 · Cω∗ (u) → ∇ × C1 · (∇ × u) ,
 

where the convergence as ε → 0 is with respect to H −1 (Rn ).

Proof. Using Proposition 5.3 and a similar method of proof as for Lemma 5.6, the
results are obtained.

We also have the following result.

Corollary 5.5. Let L denote a linear operator that commutes with the differ-
ential and nonlocal operators. Then, if Lu ∈ [H 1 (Rn )]n , Lu ∈ H 1 (Rn ), and
Lu ∈ [H 1 (R3 )]3 as need be, we have

Dω (Lu) → ∇ · Lu Dω∗ (Lu) → −∇Lu


Gω (Lu) → ∇Lu Gω∗ (Lu) → −∇ · Lu (5.22)
Cω (Lu) → ∇ × Lu Cω∗ (Lu) → ∇ × Lu,

where the convergence as ε → 0 is with respect to L2 (Rn ).

If L is selected as a differential operator with constant coefficients, or its formal


inverse, we can observe convergence of the weighted operators in either stronger or
weaker norms.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 35

5.3. A connection between the nonlocal and local Gauss theorems


Two interesting lemmas found in Ref 17t allow us to make another connection
Sandia National Labs SAND 2010-8353J

between nonlocal and local operators or, more precisely, between the nonlocal Gauss
theorem (4.6a) and the classical Gauss theorem.
Given a two-point function ν : Rn × Rn → Rk , we define the vector-valued point
function q(x) : Rn → Rk by
Z
q(x) := − (y − x)τ (x, y − x) dy, (5.23)
Rn
where, with p(x, y) = (ν + ν 0 ) · α and z = y − x, the function τ : Rn × Rn → R is
given by
Z 1

τ (x, z) = p x + λz, x − (1 − λ)z dλ.
0
Then, Lemmas I and II in Ref. 17 state that
Z
∇ · q(x) = (ν + ν 0 ) · α dy ∀ x ∈ Rn (5.24)
Rn
and
Z Z Z
q(x) · n dA = (ν + ν 0 ) · α dydx, (5.25)
∂Ω Ω Rn \Ω
respectively, where n denotes the outward pointing unit normal vector along ∂Ω.
From (3.11a) and (5.24) we then have that
∇ · q = D(ν) (5.26)
and from (4.5a) and (5.25), using the by now familiar sequence of steps, we have
that
Z Z Z Z Z
q · n dA = (ν + ν 0 ) · α dydx = − (ν + ν 0 ) · α dydx
∂Ω n
Ω R \Ω n
R \Ω Ω
Z Z Z Z
=− (ν + ν 0 ) · α dydx = − (ν + ν 0 ) · α dydx (5.27)
ΩI Ω ΩI Ω∪ΩI
Z
= N (ν) dx.
ΩI
Then, the nonlocal Gauss theorem (4.6a) for ν, (5.26), and (5.27) imply that
Z Z Z Z
0= D(ν) dx − N (ν) dx = ∇ · q dx − q · n dA,
Ω ΩI Ω ∂Ω
i.e., the classical Gauss’s theorem for the vector-valued function q. Thus, we have
shown that the nonlocal Gauss’s theorem (4.6a) for the nonlocal vector two-point
function ν(x, y) formally implies the classical Gauss theorem for the local vector
point function q(x) derived from ν through (5.23). Evidently, the Gauss theorem
can be given a meaning without the notions of the divergence operator, unit normal
vector, or surface.

t See Ref. 15 for an English translation of Ref. 17.


March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

36 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

6. Examples of nonlocal volume-constrained problems


The nonlocal point operators, the corresponding nonlocal adjoint operators, and
Sandia National Labs SAND 2010-8353J

the corresponding nonlocal interaction operators given in (3.11), (3.14), and (4.5),
respectively, can be used to define nonlocal problems that are analogous to classical
boundary-value problems for partial differential equations. Here, we merely state
problems involving scalar and vector “second-order” operators so that we are in the
setting of “elliptic” problems. Specifically, we define nonlocal problems that are are
analogous to the second-order differential boundary-value problems
 
 −∇ · C2 · ∇u = b
 in Ω
u = gd on ∂Ωd (6.1a)
 
C2 · ∇u · n = gn

on ∂Ωn

 
 −∇ · C4 : ∇u = b
 in Ω
u = gd on ∂Ωd (6.1b)
 
C4 : ∇u · n = gn

on ∂Ωn

  

 ∇ × C2 · ∇ × u − ∇ c0 ∇ · u = b in Ω
u = gd on ∂Ωd (6.1c)
 
n × C2 · ∇ × u × n + c0 (∇ · u)n = gn

on ∂Ωn ,
respectively, where ∂Ω = ∂Ωd ∪∂Ωn denotes the boundary of Ω with ∂Ωd ∩∂Ωn = ∅;
∂Ωd and ∂Ωn are the parts of the boundary ∂Ω on which Dirichlet and Neumann
boundary conditions are applied, respectively. In (6.1a)–(6.1c), C4 , C2 , and c0 de-
note fourth-order tensor, second-order tensor, and scalar point functions, respec-
tively.
Each of the problems (6.1a)–(6.1c) for an unknown function, e.g., u in (6.1a),
consists of a partial differential equation and boundary conditions, where the latter
may be viewed as constraints placed on possible solutions of the partial differential
equation. Thus, in the case of local operators, constraints are applied along the
boundary ∂Ω of the domain Ω on which the partial differential equation is applied.
A consequence of the nonlocality of operators is that constraints analogous to the
boundary conditions in (6.1a)–(6.1c) are applied instead over sets with positive
measure in Rn . Thus, in the case of nonlocal operators, constraints are applied on the
interaction domain ΩI that has positive volume and that corresponds to the domain
Ω on which the nonlocal operator equation is applied. As a result, we refer to nonlocal
problems with constraints imposed on volumes as volume-constrained problems in
contrast to local problems with constraints imposed on boundary surfaces that are
universally referred to as boundary-value problems.
We now describe the nonlocal problems analogous to (6.1a)–(6.1c). In the dif-
ferential equation setting we had to divide the boundary ∂Ω in the two parts ∂Ωd
and ∂Ωn over which we applied Dirichlet and Neumann conditions, respectively.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 37

Similarly, in the nonlocal case, we divide the interaction domain ΩI into two sub-
domains that we use to apply Dirichlet-like and Neumann-like volume constraints.
Let ΩI = ΩId ∪ ΩIn , where ΩId ∩ ΩIn = ∅ although either ΩId or ΩIn may be
Sandia National Labs SAND 2010-8353J

empty. Over ΩId we specify function values, e.g., for a given function gd , we set
u(x) = gd (x) for x ∈ ΩId . (6.2)
This is, of course, a straightforward generalization of the Dirichlet boundary condi-
tion u(x) = gd (x) for x ∈ ∂Ωd for the partial differential equation case; see, e.g.,
(6.1a).
The Neumann boundary condition in (6.1a) implies that −q · n = (C2 · ∇u) · n =
hn so that it specifies the negative of the flux density at a point x ∈ ∂Ω; see Section
(2.1). To follow the local case, in the nonlocal case, we similarly specify the negative
of the nonlocal flux density at a point x ∈ Ω. Thus, from the discussion in Section
4.1 about the operator N , we specify, for a given function gn (x) and tensor function
Θ2 (x, y), we set −NΩIn (ν)(x) = gn (x) for x ∈ ΩIn and ν = Θ2 · D∗ (u) so that
NΩIn − Θ2 · D∗ (u) = gn

for x ∈ ΩIn . (6.3)
Thus, to define volume-constrained problems analogous to the boundary-value
problems (6.1a)–(6.1c), we let Θ4 , Θ2 , and θ0 denote fourth-order tensor, second-
order tensor, and scalar two-point functions, respectively, where the tensors are
symmteric in the function and matrix senses and are positive definite. Using
(6.2) and (6.3), the nonlocal volume-constrained problems corresponding to the
boundary-value problems (6.1a)–(6.1c) are then given by
D Θ2 · D∗ (u) = b
 

 in Ω
u = gd in ΩId (6.4a)



NΩIn − Θ2 · D (u) = gn

in ΩIn

D Θ4 : D∗ (u) = b
 

 in Ω
u = gd in ΩId (6.4b)



NΩIn ,t − Θ4 : D (u) = gn

in ΩIn

C Θ2 · C ∗ (u) + G θ0 G ∗ (u) = b
  

 in Ω
u = gd in ΩId (6.4c)

∗ ∗
 
TΩIn − Θ2 · C (u) + SΩIn − θ0 G (u) = gn

in ΩIn ,
respectively.
The nonlocal Green’s first identities (4.2) are useful for defining variational for-
mulations of the nonlocal volume-constrained problems. For example, from (4.11a)
and (6.4a), we have that u(x) satisfies u = gd on ΩId , and for suitable test functions
v(x) that vanish on ΩId ,
Z Z Z Z
∗ ∗

D (u) · Θ2 · D (v) dydx = vb dx + vgn dx. (6.5)
Ω∪ΩI Ω∪ΩI Ω ΩI
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

38 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

The kernel function α dictates the choice of function spaces for the trial and test
functions u and v, respectively, so that the problem (6.5) is well posed.
Sandia National Labs SAND 2010-8353J

A special form of the nonlocal volume-constrained problem (6.4a) correspond-


ing to a scalar-valued solution is studied in Ref. 12 by appealing to a variational
formulation. Well-posedness results are provided in Ref. 12 for the case in which
the natural energy space (used to define the variational problem) is equivalent to
L2 (Ω), the space of square integrable functions. Although the notion of a nonlocal
vector calculus is not introduced, the recent book Ref. 2 describes, using semigroup
analysis, substantial recent work on nonlocal diffusion and its relationship to (6.1a).
The report Ref. 5 extends these previous results to the case when the natural energy
space is equivalent a fractional Sobolev space, a proper subspace of L2 (Ω). In this
case, the variational problem possess smoothing properties akin to those for elliptic
partial differential equations but with possibly reduced order of smoothing.

7. Local and nonlocal balance laws with applications


A balance law postulates that the rate of change of the amount of a quantity in any
domain is given by the rate at which that quantity is produced within the domain
minus the rate at which the quantity exits the domain; the latter is referred to as
the flux out of the domain. In the classical differential equation setting, the flux is
local, i.e., the quantity exits or enters the subdomain through its boundary; see Sec-
tion 2.1. Difficulties arise, however, in classical settings due to, e.g., discontinuities
and corner singularities, all of which are troublesome when defining an appropriate
notion for the “flux through the boundary of a subdomain.”
The nonlocal vector calculus we developed has an important application to non-
local balance laws for which subdomains not in direct contact may have a non-zero
interaction, i.e., there is a nonzero flux between the subdomains. This is accom-
plished by defining the flux in terms of interactions between open regions of positive
measure that are possibly a finite distance apart; recall Section 2.2. An important
feature of nonlocal balance laws is that the significant technical details associated
with determining normal and tangential traces along boundaries of suitable re-
gions is obviated when fluxes are induced through interactions between volumes.
Our nonlocal calculus, then, provides an alternative to standard approaches for cir-
cumventing the technicalities associated with lack of sufficient regularity in local
balance laws such as measure-theoretic generalizations of the Gauss-Green theorem
(see, e.g., Refs. 3 and 19) or the use of the fractional calculus (see, e.g., Refs. 1 and
25).

7.1. Abstract flux operators


In this section, we define an abstract flux operator which serves as a proxy for
interactions between two domains and then derive several useful properties that
such operators satisfy.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 39

Definition 7.1. For any two not necessarily disjoint open domains Ω1 ⊆ Rn and
Ω2 ⊆ Rn , a flux operator is any operator F(Ω1 , Ω2 ) satisfying
Sandia National Labs SAND 2010-8353J

F(Ω,
e ∅) = F(∅, Ω)
e =0 e ⊆ Rn ,
∀Ω (7.1a)

F(Ω1 ∪ Ω2 , Ω3 ) 


= F(Ω1 , Ω3 ) + F(Ω2 , Ω3 ) − F(Ω1 ∩ Ω2 , Ω3 ) 

∀ Ω1 , Ω2 , Ω3 ⊆ Rn , (7.1b)
F(Ω3 , Ω1 ∪ Ω2 ) 



= F(Ω3 , Ω1 ) + F(Ω3 , Ω2 ) − F(Ω3 , Ω1 ∩ Ω2 )

and the alternating or no self-interaction property


F(Ω,
e Ω)
e =0 e ⊆ Rn .
∀Ω (7.1c)
The following result shows that the defining relation (7.1c) can be replaced with
an antisymmetry property.

Proposition 7.1. Assume that the flux operator F(·, ·) satisfies (7.1a) and (7.1b).
Then, the antisymmetry property or action-reaction principle
F(Ω1 , Ω2 ) + F(Ω2 , Ω1 ) = 0 ∀ Ω1 , Ω2 ⊆ Rn (7.2)
is equivalent to the alternating property (7.1c).

Proof. The fact that (7.2) implies (7.1c) is obvious; simply set Ω1 = Ω2 = Ωe in
the former.
On the other hand, through repeated applications of (7.1a)–(7.1c), we obtain
0 = F(Ω1 ∪ Ω2 , Ω1 ∪ Ω2 )
= F(Ω1 , Ω1 ∪ Ω2 ) + F(Ω2 , Ω1 ∪ Ω2 ) − F(Ω1 ∩ Ω2 , Ω1 ∪ Ω2 )
= F(Ω1 , Ω1 ) + F(Ω1 , Ω2 ) − F(Ω1 , Ω1 ∩ Ω2 )
+ F(Ω2 , Ω1 ) + F(Ω2 , Ω2 ) − F(Ω2 , Ω1 ∩ Ω2 )
− F(Ω1 ∩ Ω2 , Ω1 ) − F(Ω1 ∩ Ω2 , Ω2 ) + F(Ω1 ∩ Ω2 , Ω1 ∩ Ω2 )
= F(Ω1 , Ω2 ) + F(Ω2 , Ω1 ) − F(Ω1 , Ω1 ∩ Ω2 ) − F(Ω1 ∩ Ω2 , Ω1 )
− F(Ω2 , Ω1 ∩ Ω2 ) − F(Ω1 ∩ Ω2 , Ω2 )
= F(Ω1 , Ω2 ) + F(Ω2 , Ω1 ) (7.3)
− F(Ω1 \ (Ω1 ∩ Ω2 ), Ω1 ∩ Ω2 ) − F(Ω1 ∩ Ω2 , Ω1 ∩ Ω2 )
− F(Ω1 ∩ Ω2 , Ω1 \ (Ω1 ∩ Ω2 )) − F(Ω1 ∩ Ω2 , Ω1 ∩ Ω2 )
− F(Ω2 \ (Ω1 ∩ Ω2 ), Ω1 ∩ Ω2 ) − F(Ω1 ∩ Ω2 , Ω1 ∩ Ω2 )
− F(Ω1 ∩ Ω2 , Ω2 \ (Ω1 ∩ Ω2 )) − F(Ω1 ∩ Ω2 , Ω1 ∩ Ω2 )
= F(Ω1 , Ω2 ) + F(Ω2 , Ω1 )
− F(Ω1 \ (Ω1 ∩ Ω2 ), Ω1 ∩ Ω2 ) − F(Ω1 ∩ Ω2 , Ω1 \ (Ω1 ∩ Ω2 ))
− F(Ω2 \ (Ω1 ∩ Ω2 ), Ω1 ∩ Ω2 ) − F(Ω1 ∩ Ω2 , Ω2 \ (Ω1 ∩ Ω2 )).
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

40 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

Now, assume that the regions Ω1 and Ω2 are disjoint, i.e., Ω1 ∩ Ω2 = ∅. Then, (7.2)
immediately follows from (7.3). Thus, we have proven (7.2) for disjoint regions.
Sandia National Labs SAND 2010-8353J

If Ω1 ∩ Ω2 6= ∅, we have that Ω1 \ (Ω1 ∩ Ω2 ) and Ω1 ∩ Ω2 are disjoint regions as


are Ω2 \ (Ω1 ∩ Ω2 ) and Ω1 ∩ Ω2 . Then, from (7.2) for disjoint regions, which we just
proved, and (7.3), we obtain (7.2) for the case Ω1 ∩ Ω2 6= ∅.

The following results hold only for disjoint regions.

Proposition 7.2. Assume that the flux operator F(·, ·) satisfies (7.1a) and (7.1b).
Then, F(·, ·; q) is bilinear for disjoint regions, i.e.,
)
F(Ω1 ∪ Ω2 , Ω3 ; q) = F(Ω1 , Ω3 ; q) + F(Ω2 , Ω3 ; q) ∀ Ω1 , Ω2 , Ω3 ⊆ Rn such
(7.4)
F(Ω3 , Ω1 ∪ Ω2 ; q) = F(Ω3 , Ω1 ; q) + F(Ω3 , Ω2 ; q) that Ω1 ∩ Ω2 = ∅.

Proof. The results immediately follow from (7.1a) and (7.1b).

Proposition 7.3. Assume that the flux operator F(·, ·) satisfies (7.1a) and (7.1b).
Then, F(·, ·) is additive for disjoint regions, i.e.,

F(Ω1 ∪ Ω2 , Rn \ (Ω1 ∪ Ω2 )) = F(Ω1 , Rn \ Ω1 ) + F(Ω2 , Rn \ Ω2 )


(7.5)
∀ Ω1 , Ω2 ⊆ Rn such that Ω1 ∩ Ω2 = ∅,

if and only if F(·, ·) is symmteric, i.e, if and only if F(·, ·) satisfies (7.2).

Proof. We have that if F(·, ·) satisfies (7.2), then

F(Ω1 ∪ Ω2 , Rn \ (Ω1 ∪ Ω2 )) = F(Ω1 , Rn \ (Ω1 ∪ Ω2 )) + F(Ω2 , Rn \ (Ω1 ∪ Ω2 ))


= F(Ω1 , Rn \ (Ω1 ∪ Ω2 )) + F(Ω2 , Rn \ (Ω1 ∪ Ω2 )) + F(Ω1 , Ω2 ) + F(Ω2 , Ω1 )
= F(Ω1 , Rn \ (Ω1 ∪ Ω2 ) ∪ Ω2 ) + F(Ω2 , Rn \ (Ω1 ∪ Ω2 ) ∪ Ω1 )
 

= F(Ω1 , Rn \ Ω1 ) + F(Ω2 , Rn \ Ω2 ),

i.e., F(·, ·) satisfies (7.5). Here, we have used (7.4) for the first and third equalities,
(7.2) for the second, and simple set theory for the fourth.
If, on the other hand, if F(·, ·) satisfies (7.5), then reversing the above steps

0 = F(Ω1 , Rn \ Ω1 ) + F(Ω2 , Rn \ Ω2 ) − F(Ω1 ∪ Ω2 , Rn \ (Ω1 ∪ Ω2 ))


= F(Ω1 , Rn \ (Ω1 ∪ Ω2 )) + F(Ω2 , Rn \ (Ω1 ∪ Ω2 ))
− F(Ω1 ∪ Ω2 , Rn \ (Ω1 ∪ Ω2 )) + F(Ω1 , Ω2 ) + F(Ω2 , Ω1 )
= F(Ω1 , Ω2 ) + F(Ω2 , Ω1 ),

where the first equality follows from (7.5), the second from simple set theory, and
the third from (7.4).
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 41

7.2. Local and nonlocal balance laws


In this section, we pose an abstract balance law, discuss local and nonlocal balance
Sandia National Labs SAND 2010-8353J

laws, and show how the classical vector calculus and the nonlocal vector calculus
are used to derive local and nonlocal field equations, respectively. The process is
illustrated in the context of heat conduction; another illustration is given in Section
7.3 where the peridynamics model of continuum mechanics is considered.

7.2.1. Abstract balance laws


We start with an abstract balance law for an open subset Ω ⊆ Rn given by
A(Ω;
e q) = P(Ω) e Rn \ Ω;
e − F(Ω, e q) ∀ Ω
e⊂Ω (7.6)
which postulates that A(Ω; e q) (the rate of change of the amount of the intensive
quantity q in any open subdomain Ω e ⊂ Ω) is equal to P(Ω)
e (the rate at which the
intensive quantity is produced in the subdomain by sources) minus F(Ω, e Rn \ Ω;
e q)
(the rate at which the intensive quantity exits the subdomain). In (7.6), q(x, t) may
be a scalar or vector time-dependent point function and we assume that P(Ω), e the
rate of production of q in any subdomain Ω, is given, e.g., through the specification
e
of an external source function. Here, we take the usual choices
Z Z
e q) = d
A(Ω; q(x, t) dx and P(Ω)e = b(x, t) dx (7.7)
dt Ω e Ω
e

for the rate of change and production terms, respectively, where b(x, t) denotes a
given source density function. If we assume the general case for which points in Ω
only interact with points in ΩI ⊆ Rn \ Ω, then (7.6) reduces to
A(Ω;
e q) = P(Ω)
e − F(Ω,
e ΩI ; q) ∀ Ω
e ⊂ Ω. (7.8)
The assumption (7.1c) postulates that there are no self-interactions. The result
(7.2) is an abstract action-reaction principle, i.e., the flux from Ω1 into Ω2 is equal
and opposite to the flux from Ω2 into Ω1 . The result (7.4) states that the total
interaction or flux from two disjoint regions Ω1 and Ω2 into a region Ω3 is simply
the sum of the individual fluxes from Ω1 into Ω3 and from Ω2 into Ω3 ; (7.4) also
states that the total flux from a region Ω3 into two disjoint regions Ω1 and Ω2 is the
sum of the individual fluxes from Ω3 into the two subdomains. Finally, (7.5) implies
the following proposition which shows that the abstract balance law is additive for
disjoint regions.

Proposition 7.4. Assume the flux operator F(·, ·) satisfies (7.1c)–(7.1b) and that
Ω1 and Ω2 are disjoint. Then, the balance law (7.6) and (7.7) is additive in the
following sense:
A(Ω1 ∪ Ω2 ; q) − P(Ω1 ∪ Ω2 ) + F(Ω1 ∪ Ω2 , Rn \ (Ω1 ∪ Ω2 ); q)
= A(Ω1 ; q) − P(Ω1 ) + F(Ω1 , Rn \ Ω1 ; q) (7.9)
n
+ A(Ω2 ; q) − P(Ω2 ) + F(Ω2 , R \ Ω2 ; q).
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

42 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

Proof. The additivity of A(·; q) and P(·) given by (7.7) is clear by the linearity of
the integral. Combining with (7.5) then yields (7.9).
Sandia National Labs SAND 2010-8353J

7.2.2. Local balance laws


For disjoint open regions Ω1 , Ω2 ⊂ Rn , we assume that the flux operator F(Ω1 , Ω2 ; q)
is given by (2.1) so that
Z
Floc (Ω1 , Ω2 ; q) = q · n dA for Ω1 ∩ Ω2 = ∅ with ∂Ω12 = Ω1 ∩ Ω2 (7.10)
∂Ω12

for some vector q, where ∂Ω12 denotes the common boundary of Ω1 and Ω2 and n
the unit normal along ∂Ω12 pointing out of Ω1 ; q is related to the intensive variable
q through a constitutive equation; see below for an example. We have that
if ∂Ω12 = ∅, then Floc (Ω1 , Ω2 ; q) = 0.
The flux operator given in (7.10) satisfies (7.4) and, by (2.2), is antisymmetric.
From (7.6), (7.7), (7.10), and the Gauss theorem, we have that
Z
(qt + ∇ · q − b) dx = 0 ∀Ωe⊂Ω

e

from which, because Ω


e is arbitrary in Ω, we obtain the local field equation

qt + ∇ · q = b ∀x ∈ Ω (7.11)
corresponding to the balance law (7.6) and the flux operator (7.10).
To obtain a field equation in terms of the intensive variable q, a constitutive
equation must be postulated relating the flux vector q to q. For example, consider
the case of heat conduction for which the variable q denotes the temperature. Then,
a constitutive equation relating the heat flux vector q to the temperature q is given
by the Fourier heat law q = −κ∇q, where κ denotes the thermal diffusivity. Then,
from (7.11), we have the field equation
qt = κ∆q + b (7.12)
for the temperature q, i.e., the heat equation.

7.2.3. Nonlocal balance laws


From (2.5), in the nonlocal case, we define the flux operatoru
Z Z
Fnonloc (Ω1 , Ω2 ; q) := (ν + ν 0 ) · α dydx ∀ Ω1 , Ω2 ⊂ Rn (7.13)
Ω1 Ω2

that gives the nonlocal flux from Ω1 into Ω2 ; here, the vector ν(x, y) has to be
related to q through a constitutive relation. The flux operator given in (7.13) satisfies

u Comparing Floc (Ω1 , Ω2 ; q) and Fnonloc (Ω1 , Ω2 ; q), we see that the role of the flux density q · n
for the local balance is now assumed by the nonlocal flux density N (ν).
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 43

(7.4), i.e., is bilinear for disjoint regions, and, because α(x, y) is antisymmetric, is
alternating and therefore satisfies the action-reaction principle.
Sandia National Labs SAND 2010-8353J

In stark contrast to that for the flux operator Floc (·, ·; q) given in (7.10), the flux
induced by Fnonloc (·, ·; q) dispenses with the need for determining a unit normal
to an orientable surface and instead considers volume interactions among regions.
Furthermore, the local flux operator Floc (·, ·; q) vanishes whenever Ω1 ∩ Ω2 = ∅
whereas, in general, the nonlocal flux operator Fnonloc (·, ·; q) does not.
We derive a nonlocal field equation analogous to the classical heat equation. Due
to (7.1c) with Ω1 = Ω e and Ω2 = Ωe I , we have from (7.13) that,
Z Z Z Z
Fnonloc (Ω,e Ωe I ; q) = (ν + ν 0 ) · α dydx = − (ν + ν 0 ) · α dydx
Ω ΩI
e e ΩI Ω
e e
Z Z Z (7.14)
0
=− (ν + ν ) · α dydx = N (ν) dx ∀ Ω e ⊂ Ω,

eI Ω∪
e ΩeI Ω
eI

where Ω e v Substituting
e I ⊆ ΩI denotes the interaction domain corresponding to Ω.
(7.14) along with (7.7) into (7.6), we obtain
Z Z Z
qt dx + N (ν) = b dx ∀Ωe ⊂ Ω.

e Ω
eI Ω
e

Then, using the nonlocal Gauss theorem (4.6a), we obtain


Z
(qt + D(ν) − b) dx = 0 ∀Ω
e ⊂Ω

e

from which it follows that, because Ω


e is arbitrary in Ω,

qt + D(ν) = b ∀ x ∈ Ω.

The interaction vector ν is related to the intensive quantity q through a constitutive


relation such as ν = κD∗ (q), where the adjoint operator D∗ is defined in (3.14a).
We then obtain the field equation for q given by

qt + κDD∗ (q) = b ∀x ∈ Ω

which may be viewed as a nonlocal heat equation analogous to the classical local heat
equation (7.12). Note that that the operator κDD∗ (·) is exactly that that appears
in the steady-state volume constrained problem (6.4a) with Θ2 = κI, where κ is
constant and I denotes the identity tensor.

7.3. The peridynamics nonlocal theory of continuum mechanics


We now demonstrate that the nonlocal calculus and the related nonlocal balance
laws developed in the earlier sections are synergistic with the peridynamic nonlocal
theory of continuum mechanics.

v See also page 85 in Ref. 24 and page 29 in Ref. 18.


March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

44 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

The abstract balance law (7.6) and the discussion that ensues generalize, in
straightforward manner, to vector valued intensive quantities. In particular, we
Sandia National Labs SAND 2010-8353J

obtain the field equation


∂m
+ D(Ψ) = b ∀ x ∈ Ω, (7.15)
∂t
where the operator Dt is defined in (3.18a). In particular, consider the case of
m denoting the momentum density for a continuum solid material so that m =

ρ(x) ∂t u(x, t), where ρ(·) denotes the material density and u(x, t) the displacement.
T
A constitutive equation relating Ψ(x, y) to u(x) is given by Ψ(x, y) = − 21 D∗ (u) ,
where the adjoint operator D∗ (·) is defined in (3.18b). We then have, from (7.15),
that
∂2u 1 T 
ρ 2 + D (D∗ u =b ∀ x ∈ Ω. (7.16)
∂t 2
Substituting the definitions of Dt and Dt∗ , we obtain, invoking (4.1),
Z

T 
D (D u = −2 (α ⊗ α) · (u0 − u) dx (7.17)
Ω∪ΩI

so that, from (7.16), we have that


∂2u
Z
ρ 2 = (α ⊗ α) · (u0 − u) dx + b ∀ x ∈ Ω. (7.18)
∂t Ω∪ΩI
T
A mechanical perspective indicates that D∗ u describes the deformation of u and
that the constitutive relation maps the deformation to the force density given by
the integral operator. Because the integrand of (7.17) is antisymmetric with respect
to the arguments x and y, the operator (7.17) induces an interaction, i.e., that of
forces between subdomains.w
Equation (7.18) is a generalization of the conservation of momentum equation in
the linearized bond-based peridynamic theory introduced in Ref. 21. In fact, if we
require that the null space of the integral operator in (7.16) contains rigid motions,
i.e., that
D (D∗ (Ax + c))T = 0


for all constant skew-symmetric tensors A and all constant vectors c,


then a sufficient condition is that α(x, y) = (y − x)ζ(|y − x|), where ζ : R+ → R.
Then, from (7.18), we have that
 
∂2u y−x ⊗ y−x
Z
ρ 2 = · (u0 − u) dx + b ∀ x ∈ Ω, (7.19)
∂t Ω∪ΩI σ(|y − x|)

w In the local case, (7.10) is an abstraction of Cauchy’s postulate in mechanics. There, the intensive
variable q is the vector momentum density and q is the stress tensor with q · n then being the
stress force density at a point of a surface. Then, Cauchy’s postulate states that the interaction
between two abutting regions occurs at the common interface between the two regions and is given
by the integral of the stress force along that interface. This is exactly a word description of (7.10).
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 45

where σ := ζ −2 , which is the linearized peridynamic bond-based model of Ref. 21.


In Ref. 7, analytical conditions on σ describing the amount of smoothing associated
Sandia National Labs SAND 2010-8353J

with the integral operator are given and the well posedness of the balance of linear
momentum is discussed. That paper also demonstrates that, for the case of localized
kernels, as ε → 0,
1
D (D∗ u)T → −µ∇ · (∇u) − 2µ∇(∇ · u)

2
which is the Navier operator of linear elasticity for Poisson ratio one-quarter; see
also Ref. 9. The subsequent paper Ref. 27 considers volume-constrained problems
on bounded domains in R and squares in R2 .
The state-based peridynamic theory (see Ref. 23) requires the consideration of
both volumetric and shear deformations. Similar to the above discussion for bond-
based peridynamics models, we may also formulate the state-based peridynamic
model in terms of the nonlocal operators. Let α and ω be given by (5.15). Then,
the linear state-based peridynamic integral operator (see Ref. 22) is given by
T 
−D η D∗ u − Dω (λtr Dω∗ u I ,
 
(7.20)
where η and λ are materials constants and D, D∗ , Dω , and Dω∗ are given by (3.18a),
(3.18b), (5.12), and (5.13), respectively. The scalar tr Dω∗ u  measures the volumet-
ric change, or dilatation, in the material so that tr Dω∗ u I is a diagonal tensor
representing volumetric stress. This allows us to readily apply the nonlocal calculus
to study the well-posedness of both free-space and volume-constrained linear peri-
dynamic state-based balance laws, and also suggests why, in the limit as ε → 0,
the operator given by (7.20) leads to the linear Navier operator of elasticity for
linear isotropic materials with general Poisson ratios, e.g., not relegated to a value
of one-quarter; see Ref. 6 for details.

Acknowledgments
Q. Du was supported in part by the U.S. Department of Energy grant DE-
SC0005346 and the U.S. National Science Foundation grant DMS-1016073.
M. Gunzburger was supported by the U.S. Department of Energy grant number
DE-SC0004970 and the U.S. National Science Foundation grant DMS-1013845. Also
supported in part while in residence at Yonsei University (Republic of Korea) as
part of the the WCU (World Class University) program under award number R31-
2008-000-10049-0 of the National Research Foundation of Korea.
Sandia is a multiprogram laboratory operated by Sandia Corporation, a Lock-
heed Martin Company, for the U.S. Department of Energy under contract DE-AC04-
94AL85000. Supported in part by the U.S. Department of Energy grant number
FWP-09-014290 through the Office of Advanced Scientific Computing Research,
DOE Office of Science.
K. Zhou was supported in part by the U.S. Department of Energy grant DE-
SC0005346 and the U.S. National Science Foundation grant DMS-1016073.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

46 Q. Du, M. Gunzburger, R. Lehoucq, K. Zhou

References
1. R. Almeida, A. Malinowska, and D. Torres, A fractional calculus of variations for
Sandia National Labs SAND 2010-8353J

multiple integrals with application to vibrating string, J. Math. Physics 51 (2010),


033503, doi:10.1063/1.3319559.
2. F. Andreu, J. Mazón, J. Rossi, and J. Toledo, Nonlocal diffusion problems, Mathe-
matical Surveys and Monographs, 165, American Mathematical Society, 2010.
3. G. Chen, M. Torres, and W. Ziemer, Gauss-Green theorem for weakly differentiable
vector fields, sets of finite perimeter, and balance laws, Comm. Pure Appl. Math. 62
(2009), 242–304, doi:10.1002/cpa.20262.
4. T. Coulhon and A. Grigoryan, Random walks on graphs with regular volume growth,
Geom. Funct. Anal. 8 (1998), 656–701, 10.1007/s000390050070.
5. Q. Du, M. Gunzburger, R. Lehoucq, and K. Zhou, Analysis and approximation of
nonlocal diffusion problems with volume constraints, Technical Report 2011-3168J,
Sandia National Laboratories, Albuquerque, (2011). To appear in the SIAM Review.
6. Q. Du, M. Gunzburger, R. Lehoucq, and K. Zhou, Application of a nonlocal vector
calculus to the analysis of linear peridynamic materials, Sandia National Laboratories,
Technical report SAND 2011-3870J.
7. Q. Du and K. Zhou, Mathematical analysis for the peridynamic nonlocal continuum
theory, Math. Model. Numer. Anal. 45 (2011), 217–234.
8. L. Ehrenpreis, On the theorem of kernels of Schwartz, Proc. Amer. Math. Soc. 7
(1956), 713–718.
9. E. Emmrich and O. Weckner, On the well-posedness of the linear peridynamic model
and its convergence towards the Navier equation of linear elasticity, Commun. Math.
Sci. 5 (2007), 851–864.
10. H. Gask, A proof of Schwartz’s kernel theorem Math. Scand. 8 (1960), 327–332.
11. G. Gilboa and S. Osher, Nonlocal operators with applications to image processing,
Multi. Model. Simul. 7 (2008), 1005–1028.
12. M. Gunzburger and R. Lehoucq, A nonlocal vector calculus with application to
nonlocal boundary value problems, Multi. Model. Simul. 8 (2010), 1581–1620,
doi:10.1137/090766607.
13. L. Ignat and J. Rossi, Decay estimates for nonlocal problems via energy methods, J.
Math. Pures Appl. 92 (2009), 163–187.
14. X. Jiang, L.-H. Lim, Y. Yao, and Y. Ye, Statistical ranking and combinatorial Hodge
theory, Math. Prog. 127 (2011), 203–244, doi:10.1007/s10107-010-0419-x.
15. R. Lehoucq and O. von Lilienfeld, Translation of Walter Noll’s “Derivation of the
Fundamental Equations of Continuum Thermodynamics from Statistical Mechanics”,
J. Elast. 100 (2010), 5–24.
16. O. Lézoray, V.-T. Ta, and A. Elmoataz, Partial differences as tools for filtering data
on graphs, Pattern Recog. Lett. 31 (2010), 2201–2213.
17. Walter Noll, Die Herleitung der Grundgleichungen der Thermomechanik der Kon-
tinua aus der statistischen Mechanik, Indiana U. Math. J. 4 (1955), 627–646; original
publishing journal was the J. Rat. Mech. Anal. See the English translation Ref. 15.
18. W. Noll, Thoughts on the concept of stress, J. Elast. 100 (2010), 25–32,
doi:10.1007/s10659-010-9247-8.
19. F. Schuricht, A new mathematical foundation for contact interactions in continuum
physics, Arch Rat. Mech. Anal. 184 (2007), 495–551, doi:10.1007/s00205-006-0032-6.
20. L. Schwartz, Espaces de fonctions differentiables à valerus vectorielles, J. Analyse
Math. 4 (1954-1955), 88–148.
21. S. Silling, Reformulation of elasticity theory for discontinuities and long-range forces,
J. Mech. Phys. Solids 48 (2000), 175–209.
March 15, 2012 14:17 WSPC/INSTRUCTION FILE nlvc-revision

A nonlocal vector calculus, nonlocal volume-constrained problems, and nonlocal balance laws 47

22. , Linearized theory of peridynamic states, J. Elast. 99 (2010), 85–111.


23. S. Silling, M. Epton, O. Weckner, J. Xu, and E. Askari, Peridynamic states and
constitutive modeling, J. Elast. 88 (2007), 151–184.
Sandia National Labs SAND 2010-8353J

24. S. Silling and R. Lehoucq, Peridynamic theory of solid mechanics, Adv. Appl. Mech.
44 (2010), 73–69, doi:10.1016/S0065-2156(10)44002-8.
25. V. Tarasov, Fractional vector calculus and fractional Maxwell’s equations, Annals
Phys. 323 (2008), 2756–2778, doi:10.1016/j.aop.2008.04.005.
26. D. Zhou and B. Schölkopf, Regularization on discrete spaces, Pattern Recognition, Pro-
ceedings of the 27th DAGM Symposium (W.G Kropatsch, R. Sablatnig, and A. Han-
bury, eds.), LNCS, 3663, Springer (2005), pp. 361–368.
27. K. Zhou and Q. Du, Mathematical and numerical analysis of linear peridynamic mod-
els with nonlocal boundary conditions, SIAM J. Numer. Anal. 48 (2010), 1759–1780,
doi:10.1137/090781267.

You might also like