28.12.2013 Views

Appendix B Vector Calculus in Three Dimensions - School of ...

Appendix B Vector Calculus in Three Dimensions - School of ...

Appendix B Vector Calculus in Three Dimensions - School of ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Appendix</strong> B<br />

<strong>Vector</strong><strong>Calculus</strong><strong>in</strong><strong>Three</strong><strong>Dimensions</strong><br />

In this appendix, we review the fundamentals <strong>of</strong> three-dimensional vector calculus.<br />

The student is expected to have already encountered most <strong>of</strong> these topics <strong>in</strong> an <strong>in</strong>troductory<br />

multi-variable calculus course. We shall be deal<strong>in</strong>g with calculus on curves, surfaces<br />

and solid bodies <strong>in</strong> three-dimensional space. The three methods <strong>of</strong> <strong>in</strong>tegration — l<strong>in</strong>e,<br />

surface and volume (triple) <strong>in</strong>tegrals — and the fundamental vector differential operators<br />

— gradient, curl and divergence — are <strong>in</strong>timately related. The differential operators and<br />

<strong>in</strong>tegrals underlie the multivariate versions <strong>of</strong> the fundamental theorem <strong>of</strong> calculus, known<br />

as Stokes’ Theorem and the Divergence Theorem.<br />

All <strong>of</strong> these topics will be reviewed <strong>in</strong> rapid succession, with most details be<strong>in</strong>g relegated<br />

to the exercises. A more detailed development can be found <strong>in</strong> any reasonable<br />

multi-variable calculus text, <strong>in</strong>clud<strong>in</strong>g [9,46,70].<br />

B.1. Dot and Cross Product.<br />

We beg<strong>in</strong> by review<strong>in</strong>g the basic algebraic operations between vectors <strong>in</strong> three-dimensional<br />

space R 3 . We shall cont<strong>in</strong>ue to use column vector notation<br />

v =<br />

⎛<br />

⎝ v ⎞<br />

1<br />

v 2<br />

⎠ = (v 1 ,v 2 ,v 3 ) T ∈ R 3 .<br />

v 3<br />

The standard basis vectors <strong>of</strong> R 3 are<br />

⎛ ⎞<br />

⎛<br />

⎞<br />

⎛<br />

⎞<br />

e 1 = i =<br />

⎝ 1 0<br />

0<br />

⎠, e 2 = j =<br />

⎝ 0 1⎠, e 3 = k =<br />

0<br />

⎝ 0 0<br />

1<br />

⎠. (B.1)<br />

We prefer the former notation, as it easily generalizes to n-dimensional space. Any vector<br />

v =<br />

⎛<br />

⎝ v ⎞<br />

1<br />

v 2<br />

⎠ = v 1 e 1 +v 2 e 2 +v 3 e 3<br />

v 3<br />

isal<strong>in</strong>earcomb<strong>in</strong>ation<strong>of</strong>the basisvectors. The coefficients hev 1 ,v 2 ,v 3 arethecoord<strong>in</strong>ates<br />

<strong>of</strong> the vector with respect to the standard basis.<br />

12/11/12 1230 c○ 2012 Peter J. Olver


Space comes equipped with an orientation — either right- or left-handed. One cannot<br />

alter † the orientation by physical motion, although look<strong>in</strong>g <strong>in</strong> a mirror — or, mathematically,<br />

perform<strong>in</strong>g a reflection — reverses the orientation. The standard basis vectors are<br />

graphed with a right-hand orientation, as <strong>in</strong> Figure rhr . When you po<strong>in</strong>t with your right<br />

hand, e 1 lies <strong>in</strong> the direction <strong>of</strong> your <strong>in</strong>dex f<strong>in</strong>ger, e 2 lies <strong>in</strong> the direction <strong>of</strong> your middle<br />

f<strong>in</strong>ger, and e 3 is <strong>in</strong> the direction <strong>of</strong> your thumb. In general, a set <strong>of</strong> three l<strong>in</strong>early <strong>in</strong>dependent<br />

vectors v 1 ,v 2 ,v 3 is said to have a right-handed orientation if they have the same<br />

orientation as the standard basis. It is not difficult to prove that this is the case if and<br />

only if the determ<strong>in</strong>ant <strong>of</strong> the 3×3 matrix whose columns are the given vectors is positive:<br />

det(v 1 ,v 2 ,v 3 ) > 0. Interchang<strong>in</strong>g the order <strong>of</strong> the vectors may switch their orientation;<br />

for example if v 1 ,v 2 ,v 3 are right-handed, then v 2 ,v 1 ,v 3 is left-handed.<br />

We have already made extensive use <strong>of</strong> the Euclidean dot product<br />

⎛ ⎛<br />

v·w = v 1 w 1 +v 2 w 2 +v 3 w 3 , where v =<br />

along with the Euclidean norm<br />

⎝ v ⎞<br />

1<br />

v 2<br />

v 3<br />

⎠, w =<br />

⎝ w ⎞<br />

1<br />

w 2<br />

w 3<br />

⎠, (B.2)<br />

‖v‖ = √ v·v = √ v 2 1 +v2 2 +v2 3 .<br />

(B.3)<br />

As <strong>in</strong> Def<strong>in</strong>ition 3.1, the dot product is bil<strong>in</strong>ear, symmetric: v ·w = w ·v, and positive.<br />

The Cauchy–Schwarz <strong>in</strong>equality<br />

|v·w| ≤ ‖v‖‖w‖.<br />

(B.4)<br />

implies that the dot product can be used to measure the angle θ between the two vectors<br />

v and w:<br />

v·w = ‖v‖‖w‖ cosθ.<br />

(B.5)<br />

(See also (3.17).)<br />

Remark: In this chapter, we will only use the Euclidean dot product and its associated<br />

norm. Adapt<strong>in</strong>g the constructions to more general norms and <strong>in</strong>ner products is an<br />

<strong>in</strong>terest<strong>in</strong>g exercise, but will not concern us here.<br />

Also <strong>of</strong> great importance — but particular to three-dimensional space — is the cross<br />

product between vectors. While the dot product produces a scalar, the three-dimensional<br />

cross product produces a vector, def<strong>in</strong>ed by the formula<br />

v×w =<br />

⎛<br />

⎝ v 2 w 3 −v 3 w ⎞<br />

2<br />

v 3 w 1 −v 1 w 3<br />

⎠ where v =<br />

v 1 w 2 −v 2 w 1<br />

⎛<br />

⎝ v ⎞<br />

1<br />

v 2<br />

⎠, w =<br />

v 3<br />

⎛<br />

⎝ w ⎞<br />

1<br />

w 2<br />

⎠, (B.6)<br />

w 3<br />

† This assumes that space is identified with the three-dimensional Euclidean space R 3 , or,<br />

more generally, an oriented three-dimensional manifold, [21].<br />

12/11/12 1231 c○ 2012 Peter J. Olver


We have chosen to employ the more modern wedge notation rather the more traditional<br />

cross symbol, v×w, for this quantity. The cross product formula is most easily memorized<br />

as a formal 3×3 determ<strong>in</strong>ant<br />

⎛<br />

v×w = det⎝ v ⎞<br />

1 w 1 e 1<br />

v 2 w 2 e 2<br />

⎠ = (v 2 w 3 −v 3 w 2 )e 1 +(v 3 w 1 −v 1 w 3 )e 2 +(v 1 w 2 −v 2 w 1 )e 3 ,<br />

v 3 w 3 e 3<br />

(B.7)<br />

<strong>in</strong>volv<strong>in</strong>g the standard basis vectors (B.1). We note that, like the dot product, the cross<br />

product is a bil<strong>in</strong>ear function, mean<strong>in</strong>g that<br />

(cu+dv)×w = c(u×w)+d(v×w), u×(cv+dw) = c(u×v)+d(u×w),<br />

(B.8)<br />

for any vectors u,v,w ∈ R 3 and any scalars c,d ∈ R. On the other hand, unlike the dot<br />

product, the cross product is an anti-symmetric quantity<br />

v×w = −w×v,<br />

(B.9)<br />

which changes its sign when the two vectors are <strong>in</strong>terchanged. In particular, the cross<br />

product <strong>of</strong> a vector with itself is automatically zero:<br />

v×v = 0.<br />

Geometrically, the cross product vector u = v×w is orthogonal to the two vectors v<br />

and w:<br />

v·(v×w) = 0 = w·(v×w).<br />

Thus, when v and w are l<strong>in</strong>early <strong>in</strong>dependent, their cross product u = v×w ≠ 0 def<strong>in</strong>es<br />

a normal direction to the plane spanned by v and w. The direction <strong>of</strong> the cross product<br />

is fixed by the requirement that v,w,u = v ×w form a right-handed triple. The length<br />

<strong>of</strong> the cross product vector is equal to the area <strong>of</strong> the parallelogram def<strong>in</strong>ed by the two<br />

vectors, which is<br />

‖v×w‖ = ‖v‖‖w‖|s<strong>in</strong>θ|<br />

(B.10)<br />

where θ is than angle between the two vectors, as <strong>in</strong> Figure para . Consequently, the cross<br />

product vector is zero, v × w = 0, if and only if the two vectors are coll<strong>in</strong>ear (l<strong>in</strong>early<br />

dependent) and hence only span a l<strong>in</strong>e.<br />

The scalar triple product u·(v×w) between three vectors u,v,w is def<strong>in</strong>ed as the<br />

dot product between the first vector with the cross product <strong>of</strong> the second and third vectors.<br />

The parenthesis is <strong>of</strong>ten omitted because there is only one way to make sense <strong>of</strong><br />

u·v × w. Comb<strong>in</strong><strong>in</strong>g (B.2), (B.7), shows that one can compute the triple product by the<br />

determ<strong>in</strong>antal formula ⎛<br />

u·v×w = det⎝ u ⎞<br />

1 v 1 w 1<br />

u 2 v 2 w 2<br />

⎠. (B.11)<br />

u 3 v 3 w 3<br />

By the properties <strong>of</strong> the determ<strong>in</strong>ant, permut<strong>in</strong>g the order <strong>of</strong> the vectors merely changes<br />

the sign <strong>of</strong> the triple product:<br />

u·v×w = −v·u×w = +v·w×u = ··· .<br />

12/11/12 1232 c○ 2012 Peter J. Olver


The triple product vanishes, u · v × w = 0, if and only if the three vectors are l<strong>in</strong>early<br />

dependent, i.e., coplanar or coll<strong>in</strong>ear. The triple product is positive, u·v×w > 0 if and<br />

only if the three vectors form a right-handed basis. Its magnitude |u·v×w| measures<br />

the volume <strong>of</strong> the parallelepiped spanned by the three vectors u,v,w, as <strong>in</strong> Figure ppp .<br />

B.2. Curves.<br />

A space curve C ⊂ R 3 is parametrized by a vector-valued function<br />

⎛<br />

x(t) = ⎝ x(t)<br />

⎞<br />

y(t) ⎠ ∈ R 3 , a ≤ t ≤ b, (B.12)<br />

z(t)<br />

that depends upon a s<strong>in</strong>gle parameter t that varies over some <strong>in</strong>terval. We shall always<br />

assume that x(t) is cont<strong>in</strong>uously differentiable. The curve is smooth provided its tangent<br />

vector is cont<strong>in</strong>uous and everywhere nonzero:<br />

⎛ ⎞<br />

x<br />

dx<br />

dt = x = ⎝ yz ⎠ ≠ 0.<br />

(B.13)<br />

As <strong>in</strong> the planar situation, the smoothness condition (B.13) precludes the formulation <strong>of</strong><br />

corners, cusps or other s<strong>in</strong>gularities <strong>in</strong> the curve.<br />

Physically, we can th<strong>in</strong>k <strong>of</strong> a curve as the trajectory described by a particle mov<strong>in</strong>g <strong>in</strong><br />

space. At each time t, the tangent vector x(t) represents the <strong>in</strong>stantaneous velocity <strong>of</strong> the<br />

particle. Thus, as long as the particle moves with nonzero speed, ‖ x‖ = √ <br />

x 2 + y 2 + z 2 ><br />

0, its trajectory is necessarily a smooth curve.<br />

Example B.1. Acharged particle<strong>in</strong>aconstant magneticfield movesalongthecurve<br />

⎛<br />

x(t) = ⎝ ρ cost<br />

⎞<br />

ρ s<strong>in</strong>t⎠, (B.14)<br />

ct<br />

where c > 0 and ρ > 0 are positive constants. The curve describes a circular helix <strong>of</strong> radius<br />

ρ spiral<strong>in</strong>g up the z axis. The parameter c determ<strong>in</strong>es the pitch <strong>of</strong> the helix, <strong>in</strong>dicat<strong>in</strong>g<br />

how tightly its coils are wound; the smaller c is, the closer the w<strong>in</strong>d<strong>in</strong>g. See Figure B.1 for<br />

an illustration. DNA is, remarkably, formed <strong>in</strong> the shape <strong>of</strong> a (bent and twisted) double<br />

helix. The tangent to the helix at a po<strong>in</strong>t x(t) is the vector<br />

<br />

x(t) =<br />

⎛<br />

⎝<br />

−ρ s<strong>in</strong>t<br />

ρ cost<br />

c<br />

Note that the speed <strong>of</strong> the particle,<br />

√<br />

‖ x‖ = ρ 2 s<strong>in</strong> 2 t+ρ 2 cos 2 t+c 2 = √ ρ 2 +c 2 ,<br />

rema<strong>in</strong>s constant, although the velocity vector x twists around.<br />

⎞<br />

⎠.<br />

(B.15)<br />

12/11/12 1233 c○ 2012 Peter J. Olver


-0.5<br />

-1<br />

-1 -0.5 0.5 1 0 0.5 0<br />

1<br />

10<br />

0<br />

-10<br />

Figure B.1.<br />

A Helix.<br />

Most <strong>of</strong> the term<strong>in</strong>ology <strong>in</strong>troduced <strong>in</strong> Chapter A for planar curves carries over to<br />

space curves without significant alteration. In particular, a curve is simple if it never<br />

crosses itself, and closed if its ends meet, x(a) = x(b). In the plane, simple closed curves<br />

are all topologically equivalent, mean<strong>in</strong>g one can be cont<strong>in</strong>uously deformed to the other.<br />

In space, this is no longer true. Closed curves can be knotted, and thus have nontrivial<br />

topology.<br />

Example B.2. The curve<br />

⎛<br />

x(t) = ⎝ (2+cos3t)cos2t<br />

⎞<br />

(2+cos3t)s<strong>in</strong>2t⎠ for 0 ≤ t ≤ 2π, (B.16)<br />

s<strong>in</strong>3t<br />

describes a closedcurve that is<strong>in</strong>theshape <strong>of</strong> a trefoilknot, asdepicted <strong>in</strong>FigureB.2. The<br />

trefoil is a genu<strong>in</strong>e knot, mean<strong>in</strong>g it cannot be deformed <strong>in</strong>to an unknotted circle without<br />

cutt<strong>in</strong>g and rety<strong>in</strong>g. (However, a rigorous pro<strong>of</strong> <strong>of</strong> this fact is not easy.) The trefoil is the<br />

simplest <strong>of</strong> the “toroidal knots”, <strong>in</strong>vestigated <strong>in</strong> more detail <strong>in</strong> Exercise B.2.6.<br />

The study and classification <strong>of</strong> knots is a subject <strong>of</strong> great historical importance. Indeed,<br />

they were first considered from a mathematical viewpo<strong>in</strong>t <strong>in</strong> the n<strong>in</strong>eteenth century,<br />

when the English applied mathematician William Thompson (later Lord Kelv<strong>in</strong>) proposed<br />

a theory <strong>of</strong> atoms based on knots! In recent years, knot theory has witnessed a tremendous<br />

revival, ow<strong>in</strong>g to its great relevance to modern day mathematics and physics. We refer the<br />

12/11/12 1234 c○ 2012 Peter J. Olver


Figure B.2.<br />

Two Views <strong>of</strong> a Trefoil Knot.<br />

<strong>in</strong>terested reader to the advanced text [115] for details.<br />

B.3. L<strong>in</strong>e Integrals.<br />

In Section A.5, we encountered three different types <strong>of</strong> l<strong>in</strong>e <strong>in</strong>tegrals along plane<br />

curves. Two <strong>of</strong> these — <strong>in</strong>tegrals with respect to arc length, (A.35), and circulation<br />

<strong>in</strong>tegrals, (A.37) — are directly applicable to space curves. On the other hand, for threedimensional<br />

flows, the analog <strong>of</strong> the flux l<strong>in</strong>e <strong>in</strong>tegral (A.42) is a surface <strong>in</strong>tegral, and will<br />

be discussed later <strong>in</strong> the chapter.<br />

Arc Length<br />

The length <strong>of</strong> the space curve x(t) over the parameter range a ≤ t ≤ b is computed<br />

by <strong>in</strong>tegrat<strong>in</strong>g the norm <strong>of</strong> its tangent vector:<br />

∫ b<br />

∫ L(C) =<br />

dx<br />

b<br />

∥ dt ∥ dt = √ x 2 + y 2 + z 2 dt. (B.17)<br />

a<br />

It is not hard to show that the length <strong>of</strong> the curve is <strong>in</strong>dependent <strong>of</strong> the parametrization<br />

— as it should be.<br />

Start<strong>in</strong>g at the endpo<strong>in</strong>t x(a), the arc length parameter s is given by<br />

s =<br />

∫ t<br />

a<br />

∥<br />

dx<br />

dt<br />

a<br />

∥ dt and so ds = ‖ x‖dt = √ <br />

x 2 + y 2 + z 2 dt. (B.18)<br />

The arc length s measures the distance along the curve start<strong>in</strong>g from the <strong>in</strong>itial po<strong>in</strong>t x(a).<br />

Thus, the length <strong>of</strong> the part <strong>of</strong> the curve between s = α and s = β is exactly β − α. It<br />

is <strong>of</strong>ten convenient to reparametrize the curve by its arc length, x(s). This has the same<br />

effect as mov<strong>in</strong>g along the curve at unit speed, s<strong>in</strong>ce, by the cha<strong>in</strong> rule,<br />

dx<br />

ds = dx<br />

dt<br />

∥<br />

dt x ∥∥∥<br />

ds = ‖ x‖ , so that dx<br />

ds<br />

∥ = 1.<br />

12/11/12 1235 c○ 2012 Peter J. Olver


Therefore dx/ds is the unit tangent vector po<strong>in</strong>t<strong>in</strong>g <strong>in</strong> the direction <strong>of</strong> motion along the<br />

curve.<br />

Example B.3. The length <strong>of</strong> one turn <strong>of</strong> a helix (B.14) is, us<strong>in</strong>g (B.15),<br />

∫ 2π<br />

∫ L(C) =<br />

dx<br />

2π<br />

∥ dt ∥ dt = √<br />

ρ2 +c 2 dt = 2π √ ρ 2 +c 2 .<br />

0<br />

0<br />

The arc length parameter, measured from the po<strong>in</strong>t x(0) = (r,0,0) T is merely a rescal<strong>in</strong>g,<br />

s =<br />

∫ t<br />

0<br />

√<br />

ρ2 +c 2 dt = √ ρ 2 +c 2 t,<br />

<strong>of</strong> the orig<strong>in</strong>al parameter t. When the helix is parametrized by arc length,<br />

(<br />

) T<br />

s<br />

x(s) = ρ cos √<br />

ρ2 +c , ρ s<strong>in</strong> s<br />

√ 2 ρ2 +c , cs<br />

√ ,<br />

2 ρ2 +c 2<br />

we move along it with unit speed. It now takes time s = 2π √ ρ 2 +c 2 to complete on turn<br />

<strong>of</strong> the helix.<br />

Example B.4. To compute the length <strong>of</strong> the trefoil knot (B.16), we beg<strong>in</strong> by comput<strong>in</strong>g<br />

the tangent vector<br />

⎛<br />

⎞<br />

−2(2+cos3t)s<strong>in</strong>2t−3 s<strong>in</strong>3t cos2t<br />

dx<br />

dt = ⎝ 2(2+cos3t)cos2t−3 s<strong>in</strong>3t s<strong>in</strong>2t ⎠.<br />

3 cos3t<br />

After some algebra <strong>in</strong>volv<strong>in</strong>g trigonometric identities, we f<strong>in</strong>d<br />

‖ x‖ = √ 27+16 cos3t+2 cos6t ,<br />

which is never 0. Unfortunately, the result<strong>in</strong>g arc length <strong>in</strong>tegral<br />

∫ 2π<br />

0<br />

‖ x‖dt =<br />

∫ 2π<br />

0<br />

√<br />

27+16 cos3t+2 cos6t dt<br />

cannot be completed <strong>in</strong> elementary terms. Numerical <strong>in</strong>tegration can be used to f<strong>in</strong>d the<br />

approximate value 31.8986 for the length <strong>of</strong> the knot.<br />

The arc length <strong>in</strong>tegral <strong>of</strong> a scalar field u(x) = u(x,y,z) along a curve C is<br />

∫ ∫ l ∫ l<br />

uds = u(x(s))ds = u(x(s),y(s),z(s))ds,<br />

C<br />

0<br />

0<br />

(B.19)<br />

where l is the total length <strong>of</strong> the curve. For example, if ρ(x,y,z) represents ∫ the density at<br />

position x = (x,y,z) <strong>of</strong> a wire bent <strong>in</strong> the shape <strong>of</strong> the curve C, then ρds represents<br />

the total mass <strong>of</strong> the wire. In particular, the <strong>in</strong>tegral<br />

∫ ∫ l<br />

ds = ds = l<br />

C<br />

12/11/12 1236 c○ 2012 Peter J. Olver<br />

0<br />

C


ecovers the length <strong>of</strong> the curve.<br />

If it isnot convenient to work directly with the arc length parametrization, we can still<br />

compute the arc length <strong>in</strong>tegral <strong>in</strong> terms <strong>of</strong> the orig<strong>in</strong>al parametrization x(t) for a ≤ t ≤ b.<br />

Us<strong>in</strong>g the change <strong>of</strong> parameter formula (B.18), we f<strong>in</strong>d<br />

∫ ∫ b ∫ b<br />

uds = u(x(t)) ‖ x‖dt = u(x(t),y(t),z(t)) √ <br />

x 2 + y 2 + z 2 dt. (B.20)<br />

C<br />

a<br />

a<br />

Example B.5. The density <strong>of</strong> a wire that is wound <strong>in</strong> the shape <strong>of</strong> a helix is<br />

proportional to its height. Let us compute the mass <strong>of</strong> one full turn <strong>of</strong> the helical wire.<br />

Thus, the density is given by ρ(x,y,z) = az, where a is the constant <strong>of</strong> proportionality,<br />

and we are assum<strong>in</strong>g z ≥ 0. Substitut<strong>in</strong>g <strong>in</strong>to (B.20), the total mass <strong>of</strong> the wire is<br />

∫ ∫ 2π<br />

L(C) = azds = act √ r 2 +c 2 dt = 2π 2 ac √ r 2 +c 2 .<br />

L<strong>in</strong>e Integrals <strong>of</strong> <strong>Vector</strong> Fields<br />

C<br />

0<br />

As <strong>in</strong> the two-dimensional situation (A.37), the l<strong>in</strong>e <strong>in</strong>tegral <strong>of</strong> a vector field v along a<br />

parametrized curve x(t) isobta<strong>in</strong>ed by <strong>in</strong>tegration<strong>of</strong>itstangential component withrespect<br />

to the arc length. The tangential component <strong>of</strong> v is given by<br />

v·t, where t = dx<br />

ds<br />

is the unit tangent vector to the curve. Thus, the l<strong>in</strong>e <strong>in</strong>tegral <strong>of</strong> v is written as<br />

∫ ∫<br />

∫<br />

v·dx = v 1 (x,y,z)dx+v 2 (x,y,z)dy+v 3 (x,y,z)dz = v·t ds.<br />

C<br />

C<br />

C<br />

(B.21)<br />

We can evaluate the l<strong>in</strong>e <strong>in</strong>tegral <strong>in</strong> terms <strong>of</strong> an arbitrary parametrization <strong>of</strong> the curve by<br />

the general formula<br />

∫<br />

C<br />

v·dx =<br />

=<br />

∫ b<br />

a<br />

∫ b<br />

a<br />

v(x(t))· dx<br />

dt dt<br />

[<br />

v 1 (x(t),y(t),z(t)) dx<br />

dt +v dy<br />

2 (x(t),y(t),z(t))<br />

dt +v dz<br />

3 (x(t),y(t),z(t))<br />

dt<br />

(B.22)<br />

]<br />

dt.<br />

C<br />

L<strong>in</strong>e <strong>in</strong>tegrals <strong>in</strong> three dimensions enjoy all <strong>of</strong> the properties <strong>of</strong> their two-dimensional<br />

sibl<strong>in</strong>gs: Revers<strong>in</strong>g the direction <strong>of</strong> parameterization along the curve changes the sign;<br />

also, the <strong>in</strong>tegral can be decomposed <strong>in</strong>to sums over components:<br />

∫ ∫ ∫ ∫ ∫<br />

v·dx = − v·dx, v·dx = v·dx + v·dx, C = C 1 ∪C 2 .<br />

−C C C C 1 C 2<br />

(B.23)<br />

If f(x) ∫ represents a force field, e.g., gravity, electromagnetic force, etc., then its l<strong>in</strong>e<br />

<strong>in</strong>tegral f ·dx represents the work done by mov<strong>in</strong>g along the curve. As <strong>in</strong> two dimensions,<br />

work is <strong>in</strong>dependent <strong>of</strong> the parametrization <strong>of</strong> the curve, i.e., the particle’s speed <strong>of</strong><br />

traversal.<br />

12/11/12 1237 c○ 2012 Peter J. Olver


Example B.6. Our goal is to move a mass through the force field f = (y,−x,1) T<br />

start<strong>in</strong>gfromthe<strong>in</strong>itialpo<strong>in</strong>t(1,0,1) T andmov<strong>in</strong>gverticallytothef<strong>in</strong>alpo<strong>in</strong>t(1,0,2π) T .<br />

Question: does it require more work to move <strong>in</strong> a straight l<strong>in</strong>e x(t) = (1,0,t) T or along<br />

the spiral helix x(t) = (cost,s<strong>in</strong>t,t) T , where, <strong>in</strong> both cases, 0 ≤ t ≤ 2π? The work l<strong>in</strong>e<br />

<strong>in</strong>tegral has the form<br />

∫ ∫ ∫ 2π<br />

[<br />

f ·dx = ydx−xdy +dz = y dx dy<br />

−x<br />

dt dt + dz ]<br />

dt.<br />

dt<br />

C<br />

Along the straight l<strong>in</strong>e, the amount <strong>of</strong> work is<br />

∫<br />

f ·dx =<br />

As for the spiral helix,<br />

∫<br />

C<br />

C<br />

f ·dx =<br />

C<br />

∫ 2π<br />

0<br />

∫ 2π<br />

0<br />

0<br />

dt = 2π.<br />

[<br />

− s<strong>in</strong> 2 t−cos 2 t+1 ] dt = 0.<br />

Thus, although we travel a more roundabout route, it takes no work to move along the<br />

helix!<br />

Thereasonforthesecondresultisthattheforcevectorfieldf iseverywhereorthogonal<br />

to the tangent to the curve: f ·t = 0, and so there is no tangential force exerted upon the<br />

motion. In such cases, the work l<strong>in</strong>e <strong>in</strong>tegral<br />

∫ ∫<br />

f ·dx = f ·t ds = 0<br />

C<br />

automatically vanishes. In other words, it takes no work whatsoever to move <strong>in</strong> any<br />

direction which is orthogonal to the given force vector.<br />

B.4. Surfaces.<br />

Curves are one-dimensional, and so can be traced out by a s<strong>in</strong>gle parameter. Surfaces<br />

are two-dimensional, and hence require two dist<strong>in</strong>ct parameters. Thus, a surface S ⊂ R 3<br />

is parametrized by a vector-valued function<br />

x(p,q) = (x(p,q),y(p,q),z(p,q)) T<br />

C<br />

(B.24)<br />

that depends on two variables. As the parameters (p,q) ∈ Ω range over a prescribed<br />

plane doma<strong>in</strong> Ω ⊂ R 2 , the locus <strong>of</strong> po<strong>in</strong>ts x(p,q) traces out the surface <strong>in</strong> space. See<br />

Figure surf for an illustration. The parameters are <strong>of</strong>ten thought <strong>of</strong> as def<strong>in</strong><strong>in</strong>g a system<br />

<strong>of</strong> local coord<strong>in</strong>ates on the curved surface.<br />

We shall always assume that the surface is simple, mean<strong>in</strong>g that it does not <strong>in</strong>tersect<br />

itself, so x(p,q) = x(˜p,˜q) if and only if p = ˜p and q = ˜q. In practice, this condition can be<br />

quite hard to check! The boundary<br />

∂S = {x(p,q) | (p,q) ∈ ∂Ω}<br />

(B.25)<br />

<strong>of</strong> a simple surface consists <strong>of</strong> one or more simple curves, as <strong>in</strong> Figure surf . If the<br />

underly<strong>in</strong>g parameter doma<strong>in</strong> Ω is bounded and simply connected, then ∂Ω is a simple<br />

closed plane curve, and so ∂S is also a simple closed curve.<br />

12/11/12 1238 c○ 2012 Peter J. Olver


Example B.7. The simplest <strong>in</strong>stance <strong>of</strong> a surface is a graph <strong>of</strong> a function. The<br />

parameters are the x,y coord<strong>in</strong>ates, and the surface co<strong>in</strong>cides with the portion <strong>of</strong> the<br />

graph <strong>of</strong> the function z = u(x,y) that lies over a fixed doma<strong>in</strong> (x,y) ∈ Ω ⊂ R 2 , as<br />

illustrated <strong>in</strong> Figure gsurf . Thus, a graphical surface has the parametric form<br />

x(p,q) = (p,q,u(p,q)) T , (p,q) ∈ Ω.<br />

Thus, theparametrizationidentifies x = pand y = q, whilez = u(p,q) = u(x,y)represents<br />

the height <strong>of</strong> the surface above the po<strong>in</strong>t (x,y) ∈ Ω.<br />

For example, the upper hemisphere S + r <strong>of</strong> radius r centered at the orig<strong>in</strong> can be<br />

parametrized as a graph<br />

z = √ r 2 −x 2 −y 2 , x 2 +y 2 < r 2 , (B.26)<br />

sitt<strong>in</strong>g over the disk D r = {x 2 +y 2 < r 2 } <strong>of</strong> radius r. The boundary <strong>of</strong> the hemisphere<br />

is the image <strong>of</strong> the circle C r = ∂D r = {x 2 +y 2 = r 2 } <strong>of</strong> radius r, and is itself a circle <strong>of</strong><br />

radius r sitt<strong>in</strong>g <strong>in</strong> the x,y plane: ∂S + r = {x2 +y 2 = r,z = 0}.<br />

Remark: One can <strong>in</strong>terpret the Dirichlet problem (15.1), (15.4) for the two-dimensional<br />

Laplace equation as the problem <strong>of</strong> f<strong>in</strong>d<strong>in</strong>g a surface S that is the graph <strong>of</strong> a<br />

harmonic function with a prescribed boundary ∂S = {z = h(x,y) for (x,y) ∈ ∂Ω}.<br />

Example B.8. Asphere S r <strong>of</strong>radiusrcanbeexplicitlyparametrizedby twoangular<br />

variables ϕ,θ <strong>in</strong> the form<br />

x(ϕ,θ) = (rs<strong>in</strong>ϕcosθ,rs<strong>in</strong>ϕs<strong>in</strong>θ,rcosϕ), 0 ≤ θ < 2π, 0 ≤ ϕ ≤ π. (B.27)<br />

The reader can easily check that ‖x‖ 2 = r 2 , as it should be. As illustrated <strong>in</strong> Figure<br />

sangle , θ measures the azimuthal angle or longitude, while ϕ measures the zenith<br />

angle or latitude. Thus, the upper hemisphere S r + is obta<strong>in</strong>ed by restrict<strong>in</strong>g the zenith<br />

parameter to the range 0 ≤ ϕ ≤ 1 π. Each parameter value ϕ,θ corresponds to a unique<br />

2<br />

po<strong>in</strong>t on the sphere, except when ϕ = 0 or π. All po<strong>in</strong>ts (θ,0)are mapped to the north pole<br />

(0,0,r), while all po<strong>in</strong>ts (θ,π) are mapped to the south pole (0,0,−r). Away from the<br />

poles, the spherical angles provide bona fide coord<strong>in</strong>ates on the sphere. Fortunately, the<br />

polar s<strong>in</strong>gularities do not <strong>in</strong>terfere with the overall smoothness <strong>of</strong> the sphere. Nevertheless,<br />

one must always be careful at or near these two dist<strong>in</strong>guished po<strong>in</strong>ts.<br />

Remark: In terrestrial cartography and navigation, the latitude is measured from the<br />

equator, and equals 1 π−ϕ, with postive values referr<strong>in</strong>g to the northern hemishpere and<br />

2<br />

negative the southern henisphere (or, vice versa, if you are an anitpodean). The longitude<br />

istaken withrespect to the prime meridian, throughGreenwich, England, and equalsπ−θ,<br />

with positive vlaues referr<strong>in</strong>g to the western hemisphere. Of course, <strong>in</strong> practice, both are<br />

measured <strong>in</strong> degrees rather than radians. The curves {ϕ = c} where the zenith angle takes<br />

a prescribed constant value are the circular parallels <strong>of</strong> constant latitude — except for the<br />

north andsouth poles which aremerely po<strong>in</strong>ts. The equatoris atϕ = 1 π, whilethe tropics<br />

2<br />

<strong>of</strong> Cancer and Capricorn are 232◦ 1 ≈ 0.41 radians above and below the equator. The curves<br />

{θ = c} where the meridial angle is constant are the semi-circular meridians <strong>of</strong> constant<br />

longitude stretch<strong>in</strong>g from north to south pole. Note that θ = 0 and θ = 2π describe the<br />

same meridian. In terrestrial navigation, latitude is the angle, <strong>in</strong> degrees, measured from<br />

the equator, while longitude is the angle measured from the Greenwich meridian.<br />

12/11/12 1239 c○ 2012 Peter J. Olver


Example B.9. A torus is a surface <strong>of</strong> the form <strong>of</strong> an <strong>in</strong>ner tube. One convenient<br />

parametrization <strong>of</strong> a particular toroidal surface is<br />

x(ψ,θ) = ((2+cosψ)cosθ,(2+cosψ)s<strong>in</strong>θ,s<strong>in</strong>ψ) T for 0 ≤ ψ,θ ≤ 2π. (B.28)<br />

Note that the parametrization is 2π periodic <strong>in</strong> both ψ and θ. If we <strong>in</strong>troduce cyl<strong>in</strong>drical<br />

coord<strong>in</strong>ates<br />

x = r cosθ, y = r s<strong>in</strong>θ, z,<br />

then the torus is parametrized by<br />

r = 2+cosψ,<br />

z = s<strong>in</strong>ψ.<br />

Therefore, the relevant values <strong>of</strong> (r,z) all lie on the circle<br />

(r −2) 2 +z 2 = 1 (B.29)<br />

<strong>of</strong> radius 1 centered at (2,0). As the polar angle θ <strong>in</strong>creases from 0 to 2π, the circle rotates<br />

around the z axis, and thereby sweeps out the torus.<br />

Remark: The sphere and the torus are examples <strong>of</strong> closed surfaces. The requirements<br />

for a surface to be closed are that it be simple and bounded, and, moreover, have no<br />

boundary. In general, a subset S ⊂ R 3 is bounded provided it does not stretch <strong>of</strong>f <strong>in</strong>f<strong>in</strong>itely<br />

far away. More precisely, boundedness is equivalent to the existence <strong>of</strong> a fixed number<br />

R > 0 which bounds the norm ‖x‖ < R <strong>of</strong> all po<strong>in</strong>ts x ∈ S.<br />

Tangents to Surfaces<br />

Consider a surface S parameterized by x(p,q) where (p,q) ∈ Ω. Each parametrized<br />

curve(p(t),q(t))<strong>in</strong>theparameter doma<strong>in</strong>ΩwillbemappedtoaparametrizedcurveC ⊂ S<br />

conta<strong>in</strong>ed <strong>in</strong> the surface. The curve C is parametrized by the composite map<br />

The tangent vector<br />

x(t) = x(p(t),q(t)) = (x(p(t),q(t)),y(p(t),q(t)),z(p(t),q(t))) T .<br />

dx<br />

dt = ∂x<br />

∂p<br />

dp<br />

dt + ∂x<br />

∂q<br />

dq<br />

dt<br />

(B.30)<br />

to such a curve will be tangent to the surface. The set <strong>of</strong> all possible tangent vectors to<br />

curves pass<strong>in</strong>g through a given po<strong>in</strong>t <strong>in</strong> the surface traces out the tangent plane to the<br />

surface at that po<strong>in</strong>t, as <strong>in</strong> Figure tp . Note that the tangent vector (B.30) is a l<strong>in</strong>ear<br />

comb<strong>in</strong>ation <strong>of</strong> the two basis tangent vectors<br />

x p = ∂x ( ∂x<br />

∂p = ∂p , ∂y<br />

∂p , ∂z ) T<br />

, x<br />

∂p q = ∂x ( ∂x<br />

∂q = ∂q , ∂y<br />

∂q , ∂z ) T<br />

, (B.31)<br />

∂q<br />

which therefore span the tangent plane to the surface at the po<strong>in</strong>t x(p,q) ∈ S. The first<br />

basis vector is tangent to the curves where q = constant, while the second is tangent to<br />

the curves where p = constant.<br />

12/11/12 1240 c○ 2012 Peter J. Olver


Example B.10. Consider the torus T parametrized as <strong>in</strong> (B.28). The basis tangent<br />

vectors are<br />

⎛<br />

∂x<br />

∂ψ = ⎝ −(2+cosθ)s<strong>in</strong>ψ<br />

⎞ ⎛<br />

∂x<br />

⎠,<br />

∂θ = ⎝<br />

(2+cosθ)cosψ<br />

0<br />

⎞<br />

−s<strong>in</strong>θ cosψ<br />

−s<strong>in</strong>θ s<strong>in</strong>ψ ⎠. (B.32)<br />

cosθ<br />

They serve to span the tangent plane to the torus at the po<strong>in</strong>t x(θ,ψ). For example, at<br />

the po<strong>in</strong>t x(0,0) = (3,0,0) T correspond<strong>in</strong>g to the particular parameter values θ = ψ = 0,<br />

the basis tangent vectors are<br />

x ψ (0,0) = (0,3,0) T = 3e 2 , x θ (0,0) = (0,0,1) T = e 3 ,<br />

andsothetangentplaneatthisparticularpo<strong>in</strong>tisthe(y,z)–planespanned bythestandard<br />

basis vectors e 2 ,e 3 .<br />

The tangent to any curve conta<strong>in</strong>ed with<strong>in</strong> the torus at the given po<strong>in</strong>t will be a l<strong>in</strong>ear<br />

comb<strong>in</strong>ation <strong>of</strong> these two vectors. For <strong>in</strong>stance, the toroidal knot (B.16) corresponds to<br />

the straight l<strong>in</strong>e<br />

ψ(t) = 2t, 0 ≤ t ≤ 2π, θ(t) = 3t,<br />

<strong>in</strong> the parameter space. Its tangent vector<br />

⎛<br />

−(4+2cos3t)s<strong>in</strong>2t−3 s<strong>in</strong>3t cos2t<br />

dx<br />

dt = ⎝ (4+2cos3t)cos2t−3 s<strong>in</strong>3t s<strong>in</strong>2t<br />

3 cos3t<br />

lies <strong>in</strong> the tangent plane to the torus at each po<strong>in</strong>t. In particular, at t = 0, the knot passes<br />

through the po<strong>in</strong>t x(0,0) = (3,0,0) T , and has tangent vector<br />

⎛<br />

dx<br />

dt = ⎝ 0 ⎞<br />

6⎠ dψ<br />

= 2x ψ (0,0)+3x θ (0,0) s<strong>in</strong>ce<br />

dt = 2, dθ<br />

dt = 3.<br />

3<br />

A po<strong>in</strong>t x(p,q) ∈ S on the surface is said to be nons<strong>in</strong>gular provided the basis tangent<br />

vectors x p (p,q),x q (p,q) arel<strong>in</strong>early <strong>in</strong>dependent. Thus the po<strong>in</strong>t is nons<strong>in</strong>gular ifand only<br />

if the tangent vectors span a full two-dimensional subspace <strong>of</strong> R 3 — the tangent plane to<br />

the surface at the po<strong>in</strong>t. Nons<strong>in</strong>gularity ensures the smoothness <strong>of</strong> the surface at each<br />

po<strong>in</strong>t, which is a consequence <strong>of</strong> the general Implicit Function Theorem, [159]. S<strong>in</strong>gular<br />

po<strong>in</strong>ts, where the tangent vectors are l<strong>in</strong>early dependent, can take the form <strong>of</strong> corners,<br />

cusps and folds <strong>in</strong> the surface. From now on, we shall always assume that our surface is<br />

nons<strong>in</strong>gular mean<strong>in</strong>g every po<strong>in</strong>t is a nons<strong>in</strong>gular po<strong>in</strong>t.<br />

L<strong>in</strong>ear <strong>in</strong>dependence <strong>of</strong> the tangent vectors is equivalent to the requirement that their<br />

cross product is a nonzero vector:<br />

N = ∂x<br />

∂p × ∂x ( ∂(y,z)<br />

∂q = ∂(p,q) , ∂(z,x)<br />

∂(p,q) , ∂(x,y) ) T<br />

≠ 0. (B.33)<br />

∂(p,q)<br />

12/11/12 1241 c○ 2012 Peter J. Olver<br />

⎞<br />


In this formula, we have adopted the standard notation<br />

( )<br />

∂(x,y)<br />

∂(p,q) = det xp x q<br />

= ∂x<br />

y p y q ∂p<br />

∂y<br />

∂q − ∂x<br />

∂q<br />

∂y<br />

∂p<br />

(B.34)<br />

for the Jacobian determ<strong>in</strong>ant <strong>of</strong> the functions x,y with respect to the variables p,q, which<br />

we already encountered <strong>in</strong> the change <strong>of</strong> variables formula (A.52) for double <strong>in</strong>tegrals.<br />

The cross-product vector N <strong>in</strong> (B.33) is orthogonal to both tangent vectors, and hence<br />

orthogonal to the entire tangent plane. Therefore, N def<strong>in</strong>es a normal vector to the surface<br />

at the given (nons<strong>in</strong>gular) po<strong>in</strong>t.<br />

Example B.11. Consider a surface S parametrized as the graph <strong>of</strong> a function z =<br />

u(x,y), and so, as <strong>in</strong> Example B.7<br />

x(x,y) = (x,y,u(x,y)) T , (x,y) ∈ Ω.<br />

The tangent vectors<br />

(<br />

∂x<br />

∂x = 1,0, ∂u ) T<br />

,<br />

∂x<br />

(<br />

∂x<br />

∂y = 0,1, ∂u ) T<br />

,<br />

∂y<br />

span the tangent plane sitt<strong>in</strong>g at the po<strong>in</strong>t (x,y,u(x,y) on S. The normal vector is<br />

N = ∂x<br />

∂x × ∂x (<br />

∂y = − ∂u ) T<br />

∂u<br />

,−<br />

∂x ∂y ,1 ,<br />

and po<strong>in</strong>ts upwards, as <strong>in</strong> Figure graphN . Note that every po<strong>in</strong>t on the graph is nons<strong>in</strong>gular.<br />

The unit normal to the surface at the po<strong>in</strong>t is a unit vector orthogonal to the tangent<br />

plane, and hence given by<br />

n = N<br />

‖N‖ = x p ×x q<br />

‖x p ×x q ‖ .<br />

(B.35)<br />

In general, the direction <strong>of</strong> the normal vector N depends upon the order <strong>of</strong> the two parameters<br />

p,q. Comput<strong>in</strong>g the cross product <strong>in</strong> the reverse order, x q ×x p = −N, reverses the<br />

sign <strong>of</strong> the normal vector, and hence switches its direction. Thus, there are two possible<br />

unit normals to the surface at each po<strong>in</strong>t, namely n and −n. For a closed surface, one<br />

normal po<strong>in</strong>ts outwards and one po<strong>in</strong>ts <strong>in</strong>wards.<br />

When possible, a consistent (mean<strong>in</strong>g cont<strong>in</strong>uously vary<strong>in</strong>g) choice <strong>of</strong> a unit normal<br />

serves to def<strong>in</strong>e an orientation <strong>of</strong> the surface. All closed surfaces, and most other surfaces<br />

can be oriented. The usual convention for closed surfaces is to choose the orientation<br />

def<strong>in</strong>ed by the outward normal. The simplest example <strong>of</strong> a non-orientable surface is the<br />

Möbius strip obta<strong>in</strong>ed by glu<strong>in</strong>g together the ends <strong>of</strong> a twisted strip <strong>of</strong> paper; see Exercise<br />

.<br />

12/11/12 1242 c○ 2012 Peter J. Olver


Example B.12. For the sphere <strong>of</strong> radius r parametrized by the spherical angles as<br />

<strong>in</strong> (B.27), the tangent vectors are<br />

⎛<br />

∂x<br />

∂ϕ = ⎝ rcosϕcosθ<br />

⎞ ⎛<br />

∂x<br />

⎠,<br />

∂θ = ⎝ −rs<strong>in</strong>ϕs<strong>in</strong>θ<br />

⎞<br />

⎠.<br />

rs<strong>in</strong>ϕcosθ<br />

−r s<strong>in</strong>ϕ<br />

rs<strong>in</strong>ϕcosθ<br />

0<br />

These vectors are tangent to, respectively, the meridians <strong>of</strong> constant longitude, and the<br />

parallels <strong>of</strong> constant latitude. The normal vector is<br />

⎛<br />

N = ∂x<br />

∂ϕ × ∂x<br />

∂θ = ⎝ r2 s<strong>in</strong> 2 ⎞<br />

ϕcosθ<br />

r 2 s<strong>in</strong> 2 ϕs<strong>in</strong>θ ⎠ = rs<strong>in</strong>ϕ x. (B.36)<br />

r 2 cosϕs<strong>in</strong>ϕ<br />

Thus N is a non-zero multiple <strong>of</strong> the radial vector x, except at the north or south poles<br />

when ϕ = 0 or π. This reconfirms our earlier observation that the poles are problematic<br />

po<strong>in</strong>ts for the spherical angle parametrization. The unit normal<br />

n = N<br />

‖N‖ = x r<br />

determ<strong>in</strong>ed by the spherical coord<strong>in</strong>ates ϕ,θ is the outward po<strong>in</strong>t<strong>in</strong>g normal. Revers<strong>in</strong>g<br />

the order <strong>of</strong> the angles, θ,ϕ, would lead to the outwards normal −n = −x/r.<br />

Remark: As we already saw <strong>in</strong> the example <strong>of</strong> the hemisphere, a given surface can be<br />

parametrized <strong>in</strong> many different ways. In general, to change parameters<br />

p = g(˜p,˜q),<br />

q = h(˜p,˜q),<br />

requires a smooth, <strong>in</strong>vertible map between the two parameter doma<strong>in</strong>s ˜Ω → Ω. Many<br />

<strong>in</strong>terest<strong>in</strong>g surfaces, particularly closed surfaces, cannot be parametrized <strong>in</strong> a s<strong>in</strong>gle consistent<br />

manner that satisfies the smoothness constra<strong>in</strong>t (B.33) on the entire surface. In<br />

such cases, one must assemble the surface out <strong>of</strong> pieces, each parametrized <strong>in</strong> the proper<br />

manner. The key problem <strong>in</strong> cartography is to f<strong>in</strong>d convenient parametrizations <strong>of</strong> the<br />

globe that do not significantly distort the geographical features <strong>of</strong> the planet.<br />

A surface is piecewise smooth if it can be constructed by glu<strong>in</strong>g together a f<strong>in</strong>ite<br />

number <strong>of</strong> smooth parts, jo<strong>in</strong>ed along piecewise smooth curves. For example, a cube is<br />

a piecewise smooth surface, consist<strong>in</strong>g <strong>of</strong> squares jo<strong>in</strong>ed along straight l<strong>in</strong>e segments. We<br />

shall rely on the reader’s <strong>in</strong>tuition to formalize these ideas, leav<strong>in</strong>g a rigorous development<br />

to a more comprehensive treatment <strong>of</strong> surface geometry, e.g., [60].<br />

B.5. Surface Integrals.<br />

As with spatial l<strong>in</strong>e <strong>in</strong>tegrals, there are two important types <strong>of</strong> surface <strong>in</strong>tegral. The<br />

first is the <strong>in</strong>tegration <strong>of</strong> a scalar field with respect to surface area. A typical application<br />

is to compute the area <strong>of</strong> a curved surface or the mass and center <strong>of</strong> mass <strong>of</strong> a curved shell<br />

<strong>of</strong> possibly variable density. The second type is the surface <strong>in</strong>tegral that computes the flux<br />

12/11/12 1243 c○ 2012 Peter J. Olver


associated with a vector field through an oriented surface. Applications appear <strong>in</strong> fluid<br />

mechanics, electromagnetism, thermodynamics, gravitation, and many other fields.<br />

Surface Area<br />

Accord<strong>in</strong>g to (B.10), the length <strong>of</strong> the cross product <strong>of</strong> two vectors measures the area<br />

<strong>of</strong> the parallelogram they span. This observation underlies the pro<strong>of</strong> that the length <strong>of</strong> the<br />

normal vector to a surface (B.35), namely<br />

‖N‖ = ‖x p ×x q ‖,<br />

is a measure <strong>of</strong> the <strong>in</strong>f<strong>in</strong>itesimal element <strong>of</strong> surface area, denoted<br />

dS = ‖N‖dpdq = ‖x p ×x q ‖dpdq.<br />

(B.37)<br />

The total area <strong>of</strong> the surface is found by summ<strong>in</strong>g up these <strong>in</strong>f<strong>in</strong>itesimal contributions,<br />

and is therefore given by the double <strong>in</strong>tegral<br />

∫∫ ∫∫<br />

area S = dS = ‖x p ×x q ‖dpdq<br />

S Ω<br />

√<br />

∫∫ (∂(y,z) ) 2 ( ) 2 ( ) 2<br />

(B.38)<br />

∂(z,x) ∂(x,y)<br />

= + + dpdq.<br />

∂(p,q) ∂(p,q) ∂(p,q)<br />

Ω<br />

The surface’s area does not depend upon the parametrizationused to compute the <strong>in</strong>tegral.<br />

In particular, if the surface is parametrized by x,y as the graph z = u(x,y) <strong>of</strong> a function<br />

over a doma<strong>in</strong> (x,y) ∈ Ω, then the surface area <strong>in</strong>tegral reduces to the familiar form<br />

∫∫<br />

area S =<br />

S<br />

∫∫<br />

dS =<br />

Ω<br />

√<br />

1+<br />

( ) 2 ∂u<br />

+<br />

∂x<br />

( ) ∂u<br />

∂y<br />

A detailed justification <strong>of</strong> these formulae can be found <strong>in</strong> [9,70].<br />

dxdy.<br />

(B.39)<br />

Example B.13. The well-known formula for the surface area <strong>of</strong> a sphere is a simple<br />

consequence <strong>of</strong> the <strong>in</strong>tegral formula (B.38). Us<strong>in</strong>g the parametrization by spherical angles<br />

(B.27) and the formula (B.36) for the normal, we f<strong>in</strong>d<br />

∫∫ ∫ 2π ∫ π<br />

area S r = dS = r 2 s<strong>in</strong>ϕ dϕdθ = 4πr 2 . (B.40)<br />

S r 0 0<br />

Fortunately, the problematic poles do not cause any difficulty <strong>in</strong> the computation, s<strong>in</strong>ce<br />

they contribute noth<strong>in</strong>g to the surface area <strong>in</strong>tegral.<br />

Alternatively, we can compute the area <strong>of</strong> one hemisphere S r + by realiz<strong>in</strong>g it as a<br />

graph<br />

z = √ r 2 −x 2 −y 2 for x 2 +y 2 ≤ 1,<br />

over the disk <strong>of</strong> radius r, and so, by (B.39),<br />

∫∫<br />

√<br />

area S r + = x<br />

1+<br />

2<br />

Ω r 2 −x 2 −y + y 2<br />

2 r 2 −x 2 −y dxdy 2<br />

∫∫<br />

∫<br />

r<br />

r ∫ 2π<br />

= √<br />

r2 −x 2 −y dxdy = rρ<br />

√ 2 r2 −ρ dθdρ = 2 2πr2 ,<br />

Ω<br />

12/11/12 1244 c○ 2012 Peter J. Olver<br />

0<br />

0


where we used polar coord<strong>in</strong>ates x = ρ cosθ,y = ρ s<strong>in</strong>θ to evaluate the f<strong>in</strong>al <strong>in</strong>tegral. The<br />

area <strong>of</strong> the entire sphere is twice the area <strong>of</strong> the hemisphere.<br />

Example B.14. Similarly, to compute the surface area <strong>of</strong> the torus T parametrized<br />

<strong>in</strong> (B.28), we use the tangent vectors <strong>in</strong> (B.32) to compute the normal to the torus:<br />

⎛<br />

N = x ψ ×x θ = ⎝ (2+cosψ)cosψcosθ<br />

⎞<br />

(2+cosψ)cosψs<strong>in</strong>θ ⎠, with ‖x ψ ×x θ ‖ = 2+cosψ.<br />

(2+cosψ)s<strong>in</strong>ψ<br />

Therefore,<br />

area T =<br />

∫ 2π ∫ 2π<br />

0<br />

0<br />

(2+cosψ)dψdθ = 8π 2 .<br />

If S ⊂ R 3 is a surface with f<strong>in</strong>ite area, the mean or average <strong>of</strong> a scalar function<br />

f(x,y,z) over S is given by<br />

M S [f ] = 1 ∫∫<br />

f dS.<br />

(B.41)<br />

area S S<br />

For example, the mean <strong>of</strong> a function over a sphere S r = {‖x‖ = r} <strong>of</strong> radius r is explicitly<br />

given by<br />

M Sr<br />

[f ] = 1<br />

4πr 2 ∫∫‖x‖=r<br />

f(x)dS = 1<br />

4π<br />

∫ 2π ∫ π<br />

0<br />

0<br />

F(r,ϕ,θ)s<strong>in</strong>ϕdϕdθ,<br />

(B.42)<br />

where F(r,ϕ,θ) is the spherical coord<strong>in</strong>ate expression for the scalar function f. As usual,<br />

the mean lies between the maximum and m<strong>in</strong>imum values <strong>of</strong> the function on the surface:<br />

m<strong>in</strong><br />

S<br />

f ≤ M S [f ] ≤ max<br />

S f.<br />

In particular, the center <strong>of</strong> mass C <strong>of</strong> a surface (assum<strong>in</strong>g it has constant density) is equal<br />

to the mean <strong>of</strong> the coord<strong>in</strong>ate functions x = (x,y,z) T , so<br />

C = (M S [x],M S [y],M S [z]) T = 1 (∫∫ ∫∫ ∫∫ T<br />

xdS, ydS, zdS)<br />

. (B.43)<br />

area S S S S<br />

More generally, the <strong>in</strong>tegral <strong>of</strong> a scalar field u(x,y,z) over the surface is given by<br />

∫∫ ∫∫<br />

udS = u(x(p,q),y(p,q),z(p,q))‖x p ×x q ‖dpdq. (B.44)<br />

S<br />

Ω<br />

If S represents a th<strong>in</strong> curved shell, and u = ρ(x) the density <strong>of</strong> the material at position<br />

x ∈ S, then the surface <strong>in</strong>tegral (B.44) represents the total mass <strong>of</strong> the shell. For example,<br />

the <strong>in</strong>tegral <strong>of</strong> u(x,y,z) over a hemisphere S r + <strong>of</strong> radius r can be evaluated by either <strong>of</strong><br />

the formulae<br />

∫∫ ∫ 2π ∫ π/2<br />

udS = u(rcosθs<strong>in</strong>ϕ,rs<strong>in</strong>θs<strong>in</strong>ϕ,rcosϕ) r 2 s<strong>in</strong>ϕ dϕdθ<br />

S + r<br />

0<br />

∫∫<br />

=<br />

0<br />

x 2 +y 2 ≤r 2<br />

r<br />

√<br />

r2 −x 2 −y 2 u(x,y,√ r 2 −x 2 −y 2 ) dxdy,<br />

(B.45)<br />

12/11/12 1245 c○ 2012 Peter J. Olver


depend<strong>in</strong>g upon whether one prefers spherical or graphical coord<strong>in</strong>ates.<br />

Flux Integrals<br />

Now assumethat S isanoriented surface withchosen unit normaln. Ifv = (u,v,w) T<br />

is a vector field, then the surface <strong>in</strong>tegral<br />

⎛<br />

∫∫ ∫∫<br />

∫∫<br />

( )<br />

v·n dS = v·xp ×x q dpdq = det⎝ u x ⎞<br />

p x q<br />

v y p y q<br />

⎠dpdq (B.46)<br />

S Ω<br />

Ω<br />

w z p z q<br />

<strong>of</strong> the normal component <strong>of</strong> v over the entire surface measures its flux through the surface.<br />

An alternative common notation for the flux <strong>in</strong>tegral is<br />

∫∫<br />

S<br />

∫∫<br />

v·n dS =<br />

∫∫<br />

=<br />

S<br />

Ω<br />

udydz +vdzdx+wdxdy<br />

(B.47)<br />

(<br />

u(x,y,z) ∂(y,z)<br />

)<br />

∂(z,x) ∂(x,y)<br />

+v(x,y,z) +w(x,y,z) dxdy,<br />

∂(p,q) ∂(p,q) ∂(p,q)<br />

Note how the Jacobian determ<strong>in</strong>ant notation (B.34) seamlessly <strong>in</strong>teracts with the <strong>in</strong>tegration.<br />

In particular, if the surface is the graph <strong>of</strong> a function z = h(x,y), then the surface<br />

<strong>in</strong>tegral reduces to the particularly simple form<br />

∫∫ ∫∫ (<br />

v·n dS = u(x,y,z) ∂z<br />

)<br />

∂z<br />

+v(x,y,z)<br />

∂x ∂y +w(x,y,z) dpdq (B.48)<br />

S<br />

Ω<br />

The flux surface <strong>in</strong>tegral relies upon the consistent choice <strong>of</strong> an orientation or unit<br />

normal on the surface. Thus, flux only makes sense through an oriented surface — it<br />

doesn’t make sense to speak <strong>of</strong> “flux through a Möbius band”. If we switch normals,<br />

us<strong>in</strong>g, say, the <strong>in</strong>ward <strong>in</strong>stead <strong>of</strong> the outward normal, then the surface <strong>in</strong>tegral changes<br />

sign — just like a l<strong>in</strong>e <strong>in</strong>tegral if we reverse the orientation <strong>of</strong> a curve. Similarly, if we<br />

decompose a surface <strong>in</strong>to the union <strong>of</strong> two or more parts, with only their boundaries <strong>in</strong><br />

common, then the surface <strong>in</strong>tegral similarly decomposes <strong>in</strong>to a sum <strong>of</strong> surface <strong>in</strong>tegrals.<br />

Thus,<br />

∫∫ ∫∫<br />

v·n dS = − v·n dS,<br />

−S<br />

S<br />

∫∫ ∫∫ ∫∫<br />

(B.49)<br />

v·n dS = v·n dS + v·n dS, S = S 1 ∪S 2 .<br />

S 1 S 2<br />

S<br />

In the first formula, −S denotes the surface S with the reverse orientation. In the second<br />

formula, S 1 and S 2 are only allowed to <strong>in</strong>tersect along their boundaries; moreover, they<br />

must be oriented <strong>in</strong> the same manner as S, i.e., have the same unit normal direction.<br />

Example B.15. Let S denote the triangular surface given by that portion <strong>of</strong> the<br />

plane x + y + z = 1 that lies <strong>in</strong>side the positive orthant {x ≥ 0, y ≥ 0, z ≥ 0}, as <strong>in</strong><br />

12/11/12 1246 c○ 2012 Peter J. Olver


Figure tri3 . The flux <strong>of</strong> the vector field v = (y,xz,0) T through S equals the surface<br />

<strong>in</strong>tegral<br />

∫∫<br />

ydydz +xzdzdx,<br />

S<br />

where we orient S by choos<strong>in</strong>g the upwards po<strong>in</strong>t<strong>in</strong>g normal. To compute, we note that<br />

S can be identified as the graph <strong>of</strong> the function z = 1 − x − y ly<strong>in</strong>g over the triangle<br />

T = {0 ≤ x ≤ 1,0 ≤ y ≤ 1−x}. Therefore, by (B.47),<br />

∫∫<br />

S<br />

∫∫<br />

ydydz +xzdzdx =<br />

=<br />

∫ 1 ∫ 1−x<br />

0<br />

0<br />

(1−x)(y +x)dydx =<br />

T<br />

[<br />

y ∂(y,1−x−y)<br />

∂(x,y)<br />

∫ 1<br />

0<br />

+x(1−x−y) ∂(1−x−y,x)<br />

∂(x,y)<br />

( 1<br />

2 + 1 2 x− 1 2 x2 + 1 2 x3) dx = 17<br />

24 .<br />

]<br />

dxdy<br />

If v represents the velocity vector field for a steady state fluid flow, then its flux<br />

<strong>in</strong>tegral (B.46) tells us the total volume <strong>of</strong> fluid pass<strong>in</strong>g through S per unit time. Indeed,<br />

at each po<strong>in</strong>t on S, the volume fluid that flows across a small part the surface <strong>in</strong> unit time<br />

will fill a th<strong>in</strong> cyl<strong>in</strong>der whose base is the surface area element dS and whose height v ·n<br />

is the normal component <strong>of</strong> the fluid velocity v, as pictured <strong>in</strong> Figure fluxs . Summ<strong>in</strong>g<br />

(<strong>in</strong>tegrat<strong>in</strong>g)allthesefluxcyl<strong>in</strong>dervolumesoverthesurfaceresults<strong>in</strong>theflux<strong>in</strong>tegral. The<br />

choice <strong>of</strong> orientation or unit normal specifies the convention for measur<strong>in</strong>g the direction <strong>of</strong><br />

positive flux through the surface. If S is a closed surface, and we choose n to be the unit<br />

outward normal, then the flux <strong>in</strong>tegral (B.46) represents the net amount <strong>of</strong> fluid flow<strong>in</strong>g<br />

out <strong>of</strong> the solid region bounded by S per unit time.<br />

Example B.16. The vector field v = (0,0,1) T represents a fluid mov<strong>in</strong>g with constant<br />

velocity <strong>in</strong> the vertical direction. Let us compute the fluid flux through a hemisphere<br />

S r<br />

{z + = = √ ∣ } ∣∣<br />

r 2 −x 2 −y 2 x 2 +y 2 ≤ 1 ,<br />

sitt<strong>in</strong>g over the disk D r <strong>of</strong> radius r <strong>in</strong> the x,y plane. The flux <strong>in</strong>tegral over S r +<br />

us<strong>in</strong>g (B.48), so<br />

∫∫ ∫∫ ∫∫<br />

v·n dS = dx×dy = dxdy = πr 2 .<br />

D r<br />

S + r<br />

S + r<br />

is computed<br />

The result<strong>in</strong>g double <strong>in</strong>tegral is just the area <strong>of</strong> the disk. Indeed, <strong>in</strong> this case, the value <strong>of</strong><br />

the flux <strong>in</strong>tegral is the same for all surfaces z = h(x,y) sitt<strong>in</strong>g over the disk D r .<br />

This example provides a particular case <strong>of</strong> a surface-<strong>in</strong>dependent flux <strong>in</strong>tegral, which<br />

aredef<strong>in</strong>ed <strong>in</strong>analogywiththepath-<strong>in</strong>dependent l<strong>in</strong>e<strong>in</strong>tegralsthat weencountered earlier.<br />

In general, a flux <strong>in</strong>tegral is called surface-<strong>in</strong>dependent if<br />

∫∫<br />

v·n dS =<br />

S 1<br />

∫∫<br />

v·n dS<br />

S 2<br />

(B.50)<br />

whenever the surfaces S 1 and S 2 have a common boundary ∂S 1 = ∂S 2 . In other words,<br />

the value <strong>of</strong> the <strong>in</strong>tegral depends only upon the boundary <strong>of</strong> the surface. The veracity <strong>of</strong><br />

12/11/12 1247 c○ 2012 Peter J. Olver


(B.50) requires that the surfaces be oriented <strong>in</strong> the “same manner”. For <strong>in</strong>stance, if they<br />

do not cross, then the comb<strong>in</strong>ed surface S = S 1 ∪S 2 is closed, and one uses the outward<br />

po<strong>in</strong>t<strong>in</strong>g normal on one surface and the <strong>in</strong>ward po<strong>in</strong>t<strong>in</strong>g normal on the other. In more<br />

complex situations, one checks that the two surfaces <strong>in</strong>duce the same orientation on their<br />

commonboundary. (Wedefer adiscussion <strong>of</strong>theboundaryorientationuntil later.) F<strong>in</strong>ally,<br />

apply<strong>in</strong>g (B.49) to the closed surface S = S 1 ∪S 2 and us<strong>in</strong>g the prescribed orientations,<br />

we deduce an alternative characterization <strong>of</strong> surface-<strong>in</strong>dependent vector fields.<br />

Proposition B.17. A vector field leads to a surface-<strong>in</strong>dependent flux <strong>in</strong>tegral if and<br />

only if<br />

∫∫<br />

v·n dS = 0<br />

(B.51)<br />

for every closed surface S conta<strong>in</strong>ed <strong>in</strong> the doma<strong>in</strong> <strong>of</strong> def<strong>in</strong>ition <strong>of</strong> v.<br />

S<br />

A fluid is <strong>in</strong>compressible when its volume is unaltered by the flow. Therefore, <strong>in</strong> the<br />

absence <strong>of</strong> sources or s<strong>in</strong>ks, there cannot be any net <strong>in</strong>flow or outflow across a simple closed<br />

surfacebound<strong>in</strong>garegionoccupiedbythefluid. ∫∫<br />

Thus, theflux<strong>in</strong>tegraloveraclosedsurface<br />

must vanish: v·n dS = 0. Proposition B.17 implies that the fluid velocity vector<br />

S<br />

field def<strong>in</strong>es a surface-<strong>in</strong>dependent flux <strong>in</strong>tegral. Thus, the flux <strong>of</strong> an <strong>in</strong>compressible fluid<br />

flow through any surface depends only on the (oriented) boundary curve <strong>of</strong> the surface!<br />

B.6. Volume Integrals.<br />

Volume or triple <strong>in</strong>tegrals take place over doma<strong>in</strong>s Ω ⊂ R 3 represent<strong>in</strong>g solid threedimensional<br />

bodies. A simple example <strong>of</strong> such a doma<strong>in</strong> is a ball<br />

B r (a) = {x | ‖x−a‖ < r }<br />

(B.52)<br />

<strong>of</strong> radius r > 0 centered at a po<strong>in</strong>t a ∈ R 3 . Other examples <strong>of</strong> doma<strong>in</strong>s <strong>in</strong>clude solid cubes,<br />

solid cyl<strong>in</strong>ders, solid tetrahedra, solid tori (doughnuts and bagels), solid cones, etc.<br />

In general, a subset Ω ⊂ R 3 is open if, for every po<strong>in</strong>t x ∈ Ω, a small open ball<br />

B ε (a) ⊂ Ω centered at a <strong>of</strong> radius ε = ε(a) > 0, which may depend upon a, is also<br />

conta<strong>in</strong>ed <strong>in</strong> Ω. In particular, the ball (B.52) is open. The boundary ∂Ω <strong>of</strong> an open subset<br />

Ω consists <strong>of</strong> all limit po<strong>in</strong>ts which are not <strong>in</strong> the subset. Thus, the boundary <strong>of</strong> the open<br />

ball B r (a) is the sphere S r (a) = {‖x−a‖ = r} <strong>of</strong> radius r centered at the po<strong>in</strong>t a. An<br />

open subset is called a doma<strong>in</strong> if its boundary ∂Ω consists <strong>of</strong> one or more simple, piecewise<br />

smooth surfaces. We are allow<strong>in</strong>g corners and edges <strong>in</strong> the bound<strong>in</strong>g surfaces, so that an<br />

open cube will be a perfectly valid doma<strong>in</strong>.<br />

A subset Ω ⊂ R 3 is bounded provided it fits <strong>in</strong>side a sphere <strong>of</strong> some (possibly large)<br />

radius. For example, the solid ball B r = {‖x‖ < R} is bounded, while its exterior E r =<br />

{‖x‖ > R} is an unbounded doma<strong>in</strong>. The sphere S R = {‖x‖ = R} is the common<br />

boundary <strong>of</strong> the two doma<strong>in</strong>s: S R = ∂B r = ∂E R . Indeed, any simple closed surface<br />

separates R 3 <strong>in</strong>to two doma<strong>in</strong>s that have a common boundary — its <strong>in</strong>terior, which is<br />

bounded, and its unbounded exterior.<br />

12/11/12 1248 c○ 2012 Peter J. Olver


The boundary <strong>of</strong> a bounded doma<strong>in</strong> consists <strong>of</strong> one or more closed surfaces. For<br />

<strong>in</strong>stance, the solid annular doma<strong>in</strong><br />

A r,R = { 0 < r < ‖x‖ < R }<br />

(B.53)<br />

consist<strong>in</strong>g <strong>of</strong> all po<strong>in</strong>ts ly<strong>in</strong>g between two concentric spheres <strong>of</strong> respective radii r and R<br />

has boundary given by the two spheres: ∂A r,R = S r ∪ S R . On the other hand, sett<strong>in</strong>g<br />

r = 0 <strong>in</strong> (B.53) leads to a punctured ball <strong>of</strong> radius R whose center po<strong>in</strong>t has been removed.<br />

A punctured ball is not a doma<strong>in</strong>, s<strong>in</strong>ce the center po<strong>in</strong>t is part <strong>of</strong> the boundary, but is<br />

not a bona fide surface.<br />

If the doma<strong>in</strong> Ω ⊂ R 3 represents a solid body, and the scalar field ρ(x,y,z) represents<br />

its density at a po<strong>in</strong>t (x,y,z) ∈ Ω, then the triple <strong>in</strong>tegral<br />

∫∫∫<br />

Ω<br />

ρ(x,y,z)dxdydz<br />

(B.54)<br />

equals the total mass <strong>of</strong> the body. In particular, the volume <strong>of</strong> Ω is equal to<br />

∫∫∫<br />

volΩ = dxdydz.<br />

Ω<br />

(B.55)<br />

Triple <strong>in</strong>tegrals can be directly evaluated when the doma<strong>in</strong> has the particular form<br />

Ω = { ξ(x,y) < z < η(x,y), ϕ(x) < y < ϕ(x), a < x < b }<br />

(B.56)<br />

where the z coord<strong>in</strong>ate lies between two graphical surfaces sitt<strong>in</strong>g over a common doma<strong>in</strong><br />

<strong>in</strong> the (x,y)–plane that is itself <strong>of</strong> the form <strong>of</strong> (A.47) used to evaluate double <strong>in</strong>tegrals;<br />

see Figure triple . In such cases we can evaluate the triple <strong>in</strong>tegral by iterated <strong>in</strong>tegration<br />

first with respect to z, then with respect to y and, f<strong>in</strong>ally, with respect to x:<br />

∫∫∫<br />

Ω<br />

u(x,y,z)dxdydz =<br />

∫ b<br />

a<br />

( ∫ (<br />

ψ(x) ∫ ) )<br />

η(x,y)<br />

u(x,y,z)dz dy dx.<br />

ϕ(x)<br />

ξ(x,y)<br />

A similar result holds for other order<strong>in</strong>gs <strong>of</strong> the coord<strong>in</strong>ates.<br />

(B.57)<br />

Fub<strong>in</strong>i’s Theorem, [159,158], assures us that the result <strong>of</strong> iterated <strong>in</strong>tegration does<br />

not depend upon the order <strong>in</strong> which the variables are <strong>in</strong>tegrated. Of course, the doma<strong>in</strong><br />

must be <strong>of</strong> the requisite type <strong>in</strong> order to write the volume <strong>in</strong>tegral as repeated s<strong>in</strong>gle<br />

<strong>in</strong>tegrals. More general triple <strong>in</strong>tegrals can be evaluated by chopp<strong>in</strong>g the doma<strong>in</strong> up <strong>in</strong>to<br />

disjo<strong>in</strong>t pieces that have the proper form.<br />

Example B.18. The volume <strong>of</strong> a solid ball B R <strong>of</strong> radius R can be computed as<br />

follows. We express the doma<strong>in</strong> <strong>of</strong> <strong>in</strong>tegration x 2 +y 2 +z 2 < R 2 <strong>in</strong> the form<br />

−R < x < R, − √ R 2 −x 2 < y < √ R 2 −x 2 , − √ R 2 −x 2 −y 2 < z < √ R 2 −x 2 −y 2 .<br />

12/11/12 1249 c○ 2012 Peter J. Olver


Therefore, <strong>in</strong> accordance with (B.57),<br />

∫∫∫ ∫ (<br />

R ∫<br />

√ (<br />

R 2 −x ∫<br />

√<br />

2 R 2 −x 2 −y 2<br />

dxdydz =<br />

B R −R − √ √ R 2 −x 2 − R 2 −x 2 −y 2<br />

∫ (<br />

R ∫<br />

√<br />

R 2 −x 2<br />

=<br />

−R − √ 2 √ )<br />

R 2 −x 2 −y 2 dy dx<br />

R 2 −x 2<br />

∫ R<br />

(<br />

= y √ R 2 −x 2 −y 2 +(R 2 −x 2 )s<strong>in</strong> −1<br />

=<br />

−R<br />

∫ R<br />

−R<br />

) ∣<br />

π(R 2 −x 2 )dx = π<br />

(R 2 x− x3 ∣∣∣<br />

R<br />

= 4 3<br />

x=−R<br />

3 πR3 ,<br />

recover<strong>in</strong>g the standard formula, as it should.<br />

Change <strong>of</strong> Variables<br />

dz<br />

)<br />

dy<br />

)<br />

dx<br />

) ∣√ y ∣∣∣ R 2 −x 2<br />

√ dx<br />

R2 −x 2 y=− √ R 2 −x 2<br />

If<br />

Sometimes, an <strong>in</strong>spired change <strong>of</strong> variables can be used to simplify a volume <strong>in</strong>tegral.<br />

x = f(p,q,r), y = g(p,q,r), z = h(p,q,r), (B.58)<br />

is an <strong>in</strong>vertible change <strong>of</strong> variables — mean<strong>in</strong>g that each po<strong>in</strong>t (x,y,z) corresponds to a<br />

unique po<strong>in</strong>t (p,q,r) — then<br />

∫∫∫ ∫∫∫<br />

u(x,y,z)dxdydz = U(p,q,r)<br />

∂(x,y,z)<br />

∣ ∂(p,q,r) ∣ dpdqdr. (B.59)<br />

Here<br />

Ω<br />

D<br />

U(p,q,r) = u(x(p,q,r),y(p,q,r),z(p,q,r))<br />

is the expression for the <strong>in</strong>tegrand <strong>in</strong> the new coord<strong>in</strong>ates, while D is the doma<strong>in</strong> consist<strong>in</strong>g<br />

<strong>of</strong> all po<strong>in</strong>ts (p,q,r) that map to po<strong>in</strong>ts (x,y,z) ∈ Ω <strong>in</strong> the orig<strong>in</strong>al doma<strong>in</strong>. Invertibility<br />

requires that each po<strong>in</strong>t <strong>in</strong> D corresponds to a unique po<strong>in</strong>t <strong>in</strong> Ω. The change <strong>in</strong> volume<br />

is governed by the absolute value <strong>of</strong> the three-dimensional Jacobian determ<strong>in</strong>ant<br />

⎛<br />

∂(x,y,z)<br />

∂(p,q,r) = det ⎝ x ⎞<br />

p x q x r<br />

y p y q y r<br />

⎠ = x p ·x q ×x r (B.60)<br />

z p z q z r<br />

for the change <strong>of</strong> variables. The identification <strong>of</strong> the vector triple product (B.60) with<br />

an (<strong>in</strong>f<strong>in</strong>itesimal) volume element lies beh<strong>in</strong>d the justification <strong>of</strong> the change <strong>of</strong> variables<br />

formula; see [9,70] for a detailed pro<strong>of</strong>.<br />

By far, the two most important cases are cyl<strong>in</strong>drical and spherical coord<strong>in</strong>ates. Cyl<strong>in</strong>drical<br />

coord<strong>in</strong>ates correspond to replac<strong>in</strong>g the x and y coord<strong>in</strong>ates by their polar counterparts,<br />

whilereta<strong>in</strong><strong>in</strong>gthe verticalz coord<strong>in</strong>ateunchanged. Thus, thechange <strong>of</strong> coord<strong>in</strong>ates<br />

has the form<br />

x = r cosθ, y = r s<strong>in</strong>θ, z = z. (B.61)<br />

12/11/12 1250 c○ 2012 Peter J. Olver


The Jacobian determ<strong>in</strong>ant for cyl<strong>in</strong>drical coord<strong>in</strong>ates is<br />

⎛<br />

∂(x,y,z)<br />

∂(r,θ,z) = det ⎝ x ⎞ ⎛ ⎞<br />

r x θ x z cosθ −r s<strong>in</strong>θ 0<br />

y r y θ y z<br />

⎠ = det⎝s<strong>in</strong>θ r cosθ 0⎠ = r.<br />

z r z θ z z 0 0 1<br />

(B.62)<br />

Therefore, the general change <strong>of</strong> variables formula (B.59) tells us the formula for a triple<br />

<strong>in</strong>tegral <strong>in</strong> cyl<strong>in</strong>drical coord<strong>in</strong>ates:<br />

∫∫∫ ∫∫∫<br />

f(x,y,z)dxdydz = f(r cosθ,r s<strong>in</strong>θ,z)rdrdθdz. (B.63)<br />

Example B.19. For example, consider an ice cream cone<br />

C h = { x 2 +y 2 < z 2 , 0 < z < h } = { r < z, 0 < z < h }<br />

<strong>of</strong> height h plotted <strong>in</strong> Figure cone . To compute its volume, we express the doma<strong>in</strong> <strong>in</strong><br />

terms <strong>of</strong> the cyl<strong>in</strong>drical coord<strong>in</strong>ates, lead<strong>in</strong>g to<br />

∫∫∫ ∫ h ∫ 2π ∫ z ∫ h<br />

dxdydz = rdrdθdz = πz 2 dz = 1 3 πh3 .<br />

C h 0 0 0<br />

Spherical coord<strong>in</strong>ates are denoted by r,ϕ,θ, where<br />

0<br />

x = r s<strong>in</strong>ϕ cosθ, y = r s<strong>in</strong>ϕ s<strong>in</strong>θ, z = r cosϕ. (B.64)<br />

Here r = ‖x‖ = √ x 2 +y 2 +z 2 represents the radius, 0 ≤ ϕ ≤ π is the zenith angle or<br />

latitude, while 0 ≤ θ < 2π is the azimuthal angle or longitude. The reader may recall that<br />

we already encountered these coord<strong>in</strong>ates <strong>in</strong> our parametrization (B.27) <strong>of</strong> the sphere. It is<br />

important to dist<strong>in</strong>guish between the spherical r,θ and the cyl<strong>in</strong>drical r,θ — even though<br />

the same symbols are used, they represent different quantities.<br />

Warn<strong>in</strong>g: In many books, particularly those <strong>in</strong> physics, the roles <strong>of</strong> θ and ϕ are<br />

reversed, lead<strong>in</strong>g to much confusion when perus<strong>in</strong>g the literature. We prefer the mathematical<br />

convention as the azimuthal angle θ agrees with its cyl<strong>in</strong>drical counterpart. You<br />

need to be very careful to determ<strong>in</strong>e which convention is be<strong>in</strong>g used when consult<strong>in</strong>g any<br />

reference!<br />

A short computation proves that the spherical coord<strong>in</strong>ate Jacobian determ<strong>in</strong>ant is<br />

⎛<br />

∂(x,y,z)<br />

∂(r,ϕ,θ) = det ⎝ x ⎞<br />

r x ϕ x θ<br />

y r y ϕ y θ<br />

⎠<br />

z r z ϕ z θ<br />

⎛<br />

⎞ (B.65)<br />

s<strong>in</strong>ϕ cosθ r cosϕ cosθ −r s<strong>in</strong>ϕ s<strong>in</strong>θ<br />

= det⎝s<strong>in</strong>ϕ s<strong>in</strong>θ r cosϕ s<strong>in</strong>θ r s<strong>in</strong>ϕ cosθ ⎠ = r 2 s<strong>in</strong>ϕ.<br />

cosϕ −r s<strong>in</strong>ϕ 0<br />

Therefore, a triple <strong>in</strong>tegral is evaluated <strong>in</strong> spherical coord<strong>in</strong>ates accord<strong>in</strong>g to the formula<br />

∫∫∫ ∫∫∫<br />

f(x,y,z)dxdydz = F(r,ϕ,θ)r 2 s<strong>in</strong>ϕ drdϕdθ, (B.66)<br />

12/11/12 1251 c○ 2012 Peter J. Olver


where we rewrite the <strong>in</strong>tegrand<br />

F(r,ϕ,θ) = f(r s<strong>in</strong>ϕ cosθ,r s<strong>in</strong>ϕ s<strong>in</strong>θ,r cosϕ)<br />

as a function <strong>of</strong> the spherical coord<strong>in</strong>ates.<br />

(B.67)<br />

Example B.20. The <strong>in</strong>tegration required <strong>in</strong> Example B.18 to compute the volume<br />

<strong>of</strong> a ball B R <strong>of</strong> radius R can be considerably simplified by switch<strong>in</strong>g over to spherical<br />

coord<strong>in</strong>ates. The ball is given by B R = { 0 ≤ r < R,0 ≤ ϕ ≤ π,0 ≤ θ < 2π } . Thus, us<strong>in</strong>g<br />

(B.66), we compute<br />

∫∫∫ ∫ R ∫ π ∫ 2π ∫ R<br />

dxdydz = r 2 s<strong>in</strong>ϕ dθdϕdr = 4πr 2 dr = 4 3 πR3 . (B.68)<br />

B R<br />

0<br />

0<br />

0<br />

The reader may note that the next-to-last <strong>in</strong>tegrand represents the surface area <strong>of</strong> the<br />

sphere <strong>of</strong> radius R. Thus, we are, <strong>in</strong> effect, comput<strong>in</strong>g the volume by summ<strong>in</strong>g up (i.e.,<br />

<strong>in</strong>tegrat<strong>in</strong>g) the surface areas <strong>of</strong> concentric th<strong>in</strong> spherical shells.<br />

Remark: Sometimes, we will be sloppy and use the same letter for a function <strong>in</strong> an<br />

alternative coord<strong>in</strong>ate system. Thus, we may use f(r,ϕ,θ) to represent the spherical<br />

coord<strong>in</strong>ate form (B.67) <strong>of</strong> a function f(x,y,z). Technically, this is not correct! However,<br />

the clarity and <strong>in</strong>tuitionsometimes outweighs the pedantic use <strong>of</strong> a new letter each time we<br />

change coord<strong>in</strong>ates. Moreover, <strong>in</strong> geometry and modern physical theories, [21], the symbol<br />

“f” represents an <strong>in</strong>tr<strong>in</strong>sic scalar field, and f(x,y,z) and f(r,ϕ,θ) merely its <strong>in</strong>carnations<br />

<strong>in</strong> two different coord<strong>in</strong>ate charts on R 3 . Hopefully, this will be clear from the context.<br />

B.7. Gradient, Divergence, and Curl.<br />

There are three important vector differential operators that play a ubiquitous role <strong>in</strong><br />

three-dimensional vector calculus, known as the gradient, divergence and curl. We have<br />

already encountered their two-dimensional counterparts <strong>in</strong> Chapter A.<br />

The Gradient<br />

We beg<strong>in</strong> with the three-dimensional version <strong>of</strong> the gradient operator<br />

⎛<br />

∇u = ⎝ u ⎞<br />

x<br />

u y<br />

⎠. (B.69)<br />

u z<br />

The gradient def<strong>in</strong>es a l<strong>in</strong>ear operator that maps a scalar function u(x,y,z) to the vector<br />

fieldwhosecomponentsareitspartialderivativeswithrespect totheCartesiancoord<strong>in</strong>ates.<br />

If x(t) = (x(t),y(t),z(t)) T is any parametrized curve, then the rate <strong>of</strong> change <strong>in</strong> the<br />

function u as we move along the curve is given by the <strong>in</strong>ner product<br />

d ∂u<br />

u(x(t),y(t),z(t))=<br />

dt ∂x<br />

dx<br />

dt + ∂u<br />

∂y<br />

0<br />

dy<br />

dt + ∂u<br />

∂z<br />

dz<br />

dt<br />

= ∇u· x (B.70)<br />

between the gradient and the tangent vector to the curve. Therefore, as we reasoned<br />

earlier <strong>in</strong> the planar case, the gradient ∇u po<strong>in</strong>ts <strong>in</strong> the direction <strong>of</strong> steepest <strong>in</strong>crease <strong>in</strong><br />

12/11/12 1252 c○ 2012 Peter J. Olver


the function u, while its negative −∇u po<strong>in</strong>ts <strong>in</strong> the direction <strong>of</strong> steepest decrease. For<br />

example, if u(x,y,z) represents the temperature at a po<strong>in</strong>t (x,y,z) <strong>in</strong> space, then ∇u<br />

po<strong>in</strong>ts <strong>in</strong> the direction <strong>in</strong> which temperature is gett<strong>in</strong>g the hottest, while −∇u po<strong>in</strong>ts <strong>in</strong><br />

the directionit gets the coldest. Therefore, ifone wants to cool down asrapidly as possible,<br />

one should move <strong>in</strong> the direction <strong>of</strong> −∇u at each <strong>in</strong>stant, which is the direction <strong>of</strong> the<br />

flow <strong>of</strong> heat energy. Thus, the path x(t) to be followed for the fastest cool down will be a<br />

solution to the gradient flow equations<br />

<br />

x = −∇u,<br />

(B.71)<br />

or, explicitly,<br />

dx<br />

dt = − ∂u<br />

∂x (x,y,z),<br />

dy<br />

dt = − ∂u<br />

∂y (x,y,z),<br />

dz<br />

dt = − ∂u<br />

∂z (x,y,z).<br />

A solution x(t) to such a system <strong>of</strong> ord<strong>in</strong>ary differential equations will experience cont<strong>in</strong>uously<br />

decreas<strong>in</strong>g temperature. In Chapter 19, we will learn how to use such gradient flows<br />

to locate and numerically approximate the m<strong>in</strong>ima <strong>of</strong> functions.<br />

The set <strong>of</strong> all po<strong>in</strong>ts where a scalar field u(x,y,z) has a given value,<br />

u(x,y,z) = c<br />

(B.72)<br />

for some fixed constant c, is known as a level set <strong>of</strong> u. If u measures temperature, then<br />

its level sets are the isothermal surfaces <strong>of</strong> equal temperature. If u is sufficiently smooth,<br />

most <strong>of</strong> its level sets are smooth surfaces. In fact, if ∇u ≠ 0 at a po<strong>in</strong>t, then one can prove<br />

that all nearby level sets are smooth surfaces near the po<strong>in</strong>t <strong>in</strong> question. This important<br />

fact is a consequence <strong>of</strong> the general Implicit Function Theorem, [159]. Thus, if ∇u ≠ 0 at<br />

all po<strong>in</strong>ts on a level set, then the level set is a smooth surface, and, if bounded, a simple<br />

closed surface. (On the other hand, f<strong>in</strong>d<strong>in</strong>g an explicit parametrization <strong>of</strong> a level set may<br />

be quite difficult!)<br />

Theorem B.21. If nonzero, the gradient vector ∇u ≠ 0 def<strong>in</strong>es the normal direction<br />

to the level set {u = c} at each po<strong>in</strong>t.<br />

Pro<strong>of</strong>: Indeed, suppose x(t) is any curve conta<strong>in</strong>ed <strong>in</strong> the level set, so that<br />

u(x(t),y(t),z(t)) = c for all t.<br />

S<strong>in</strong>ce c is constant, the derivative with respect to t is zero, and hence, by (B.70),<br />

d<br />

dt u(x(t),y(t),z(t))= ∇u· x = 0,<br />

which implies that the gradient vector ∇u is orthogonal to the tangent vector x to the<br />

curve. S<strong>in</strong>ce this holds for all such curves conta<strong>in</strong>ed with<strong>in</strong> the level set, the gradient must<br />

be orthogonal to the entire tangent plane at the po<strong>in</strong>t, and hence, if nonzero, def<strong>in</strong>es a<br />

normal direction to the level surface.<br />

Q.E.D.<br />

12/11/12 1253 c○ 2012 Peter J. Olver


Physically,Theorem B.21tellsusthatthedirection<strong>of</strong>steepest <strong>in</strong>crease<strong>in</strong>temperature<br />

is perpendicular to the isothermal surfaces at each po<strong>in</strong>t. Consequently, the solutions to<br />

the gradient flow equations (B.71) form an orthogonal system <strong>of</strong> curves to the level set<br />

surfaces <strong>of</strong> u, and one should follow these curves to m<strong>in</strong>imize the temperature as rapidly<br />

as possible. Similarly, <strong>in</strong> a steady state fluid flow, the fluid potential is represented by a<br />

scalar field ϕ(x,y,z). Its gradient v = ∇ϕ determ<strong>in</strong>es the fluid velocity at each po<strong>in</strong>t. The<br />

streaml<strong>in</strong>es followed by the fluid particles are the solutions to the gradient flow equations<br />

<br />

x = v = ∇ϕ, while the level sets <strong>of</strong> ϕ are the equipotential surfaces. Thus, fluid particles<br />

flow <strong>in</strong> a direction orthogonal to the equipotential surfaces.<br />

Example B.22. The level sets <strong>of</strong> the radial function u = x 2 + y 2 + z 2 are the<br />

concentric spheres centered at the orig<strong>in</strong>. Its gradient ∇u = (2x,2y,2z) T = 2x po<strong>in</strong>ts <strong>in</strong><br />

the radial direction, orthogonal to each spherical level set. Note that ∇u = 0 only at the<br />

orig<strong>in</strong>, which is a level set, but not a smooth surface.<br />

The radial vector also specifies the direction <strong>of</strong> fastest <strong>in</strong>crease (decrease) <strong>in</strong> the function<br />

u. Indeed, the solution to the associated gradient flow system (B.71), namely<br />

<br />

x = −2x is x(t) = x 0 e −2t ,<br />

where x 0 = x(0) is the <strong>in</strong>itial position. Therefore, to decrease the function u as rapidly as<br />

possible, one should follow a radial ray <strong>in</strong>to the orig<strong>in</strong>.<br />

Example B.23. An implicit equation for the torus (B.28) is obta<strong>in</strong>ed by replac<strong>in</strong>g<br />

r = √ x 2 +y 2 <strong>in</strong>(B.29). In thismanner, we areledto consider the levelsets <strong>of</strong> thefunction<br />

u(x,y,z) = x 2 +y 2 +z 2 −4 √ x 2 +y 2 = c,<br />

(B.73)<br />

with the particular value c = −3 correspond<strong>in</strong>g to (B.28). The gradient <strong>of</strong> the function is<br />

∇u(x,y,z) =<br />

(<br />

2x−<br />

4x 4y<br />

√ 2y −<br />

x2 +y2, √<br />

x2 +y 2, 2z ) T<br />

,<br />

(B.74)<br />

which is well-def<strong>in</strong>e except on the z axis, where x = y = 0. Note that ∇F ≠ 0 unless z = 0<br />

and x 2 +y 2 = 4. Therefore, the level sets <strong>of</strong> u are smooth, toroidal surfaces except for z<br />

axis and the circle <strong>of</strong> radius 2 <strong>in</strong> the (x,y) plane.<br />

Divergence and Curl<br />

The second important vector differential operator is the divergence,<br />

divv = ∇·v = ∂v 1<br />

∂x + ∂v 2<br />

∂y + ∂v 3<br />

∂z .<br />

(B.75)<br />

The divergence maps a vector field v = (v 1 ,v 2 ,v 3 ) T to a scalar field f = ∇ · v. For<br />

example, the radial vector field v = (x,y,z) T has constant divergence ∇·v = 3.<br />

In fluid mechanics, the divergence measures the local, <strong>in</strong>stantaneous change <strong>in</strong> the<br />

volume <strong>of</strong> a fluid packet as it moves. Thus, a steady state fluid flow is <strong>in</strong>compressible, with<br />

12/11/12 1254 c○ 2012 Peter J. Olver


unchang<strong>in</strong>g volume, if and only if its velocity vector field is divergence-free: ∇·v ≡ 0. The<br />

connection between <strong>in</strong>compressibility and the earlier zero-flux condition will be addressed<br />

<strong>in</strong> the Divergence Theorem B.35 below.<br />

As <strong>in</strong> the two-dimensional situation, the composition <strong>of</strong> divergence and gradient produces<br />

the Laplacian operator:<br />

∇·∇u = ∆u = u xx +u yy +u zz .<br />

(B.76)<br />

Indeed, as we shall see, except for the miss<strong>in</strong>g m<strong>in</strong>us sign and the all-important boundary<br />

conditions, this is effectively the same as the self-adjo<strong>in</strong>t form <strong>of</strong> the three-dimensional<br />

Laplacian:<br />

∇ ∗ ◦∇u = −∇·∇u = −∆u.<br />

Thethirdimportantvectordifferentialoperatoristhecurl, which, <strong>in</strong>threedimensions,<br />

maps vector fields to vector fields. It is most easily memorized <strong>in</strong> the form <strong>of</strong> a (formal)<br />

3×3 determ<strong>in</strong>ant<br />

⎛ ⎞<br />

∂v 3<br />

∂y − ∂v 2<br />

∂z<br />

curlv = ∇×v =<br />

∂v 1<br />

⎜ ∂z − ∂v 3<br />

∂x<br />

⎝ ∂v 2<br />

∂x − ∂v 1<br />

∂y<br />

⎛<br />

= det⎝ ∂ ⎞<br />

x v 1 e 1<br />

∂ y v 2 e 2<br />

⎠, (B.77)<br />

⎟<br />

⎠ ∂ z v 3 e 3<br />

<strong>in</strong> analogy with the determ<strong>in</strong>antal form (B.6) <strong>of</strong> the cross product. For <strong>in</strong>stance, the radial<br />

vector field v = (x,y,z) T has zero curl:<br />

⎛<br />

∇×v = det⎝ ∂ ⎞<br />

x x e 1<br />

∂ y y e 2<br />

⎠ = 0.<br />

∂ z z e 3<br />

This is <strong>in</strong>dicative <strong>of</strong> the lack or any rotational effect <strong>of</strong> the <strong>in</strong>duced flow.<br />

If v represents the velocity vector field <strong>of</strong> a steady state fluid flow, its curl ∇ × v<br />

measures the <strong>in</strong>stantaneous rotation <strong>of</strong> the fluid flow at a po<strong>in</strong>t, and is known as the<br />

vorticity <strong>of</strong> the flow. When non-zero, the direction <strong>of</strong> the vorticity vector represents the<br />

axis <strong>of</strong> rotation, while its magnitude ‖∇×v‖ measures the <strong>in</strong>stantaneous angular velocity<br />

<strong>of</strong> the swirl<strong>in</strong>g flow. Physically, if we place a microscopic turb<strong>in</strong>e <strong>in</strong> the fluid so that its<br />

shaft po<strong>in</strong>ts <strong>in</strong> the direction specified by a unit vector n, then its rate <strong>of</strong> sp<strong>in</strong> will be<br />

proportional to component <strong>of</strong> the vorticity vector ∇×v <strong>in</strong> the direction <strong>of</strong> its shaft. This<br />

is equal to the dot product<br />

n·(∇×v) = ‖∇×v‖ cosϕ,<br />

where ϕ is the angle between n and the curl vector. Therefore, the maximal rate <strong>of</strong> sp<strong>in</strong><br />

will occur when ϕ = 0, and so the shaft <strong>of</strong> the turb<strong>in</strong>e l<strong>in</strong>es up with the direction <strong>of</strong> the<br />

vorticity vector ∇ × v. In this orientation, the angular velocity <strong>of</strong> the turb<strong>in</strong>e will be<br />

proportional to its magnitude ‖∇×v‖. On the other hand, if the axis <strong>of</strong> the turb<strong>in</strong>e is<br />

12/11/12 1255 c○ 2012 Peter J. Olver


orthogonal to the direction <strong>of</strong> the vorticity, then it will not rotate. If ∇×v ≡ 0, then there<br />

is no net motion <strong>of</strong> a turb<strong>in</strong>e, not matter which orientation it is placed <strong>in</strong> the fluid flow.<br />

Thus, a flow with zero curl is irrotational. The precise connection between this def<strong>in</strong>ition<br />

and the earlier zero circulation condition will be expla<strong>in</strong>ed shortly.<br />

Example B.24. Consider a helical fluid flow with velocity vector<br />

v = (−y,x,1) T .<br />

Integrat<strong>in</strong>g the ord<strong>in</strong>ary differential equations x = v, namely<br />

<br />

x = −y,<br />

<br />

y = x,<br />

<br />

z = 1,<br />

with <strong>in</strong>itial conditions x(0) = x 0 , y(0) = y 0 , z(0) = z 0 gives the flow<br />

x(t) = x 0 cost−y 0 s<strong>in</strong>t, y(t) = x 0 s<strong>in</strong>t+y 0 cost, z(t) = z 0 +t. (B.78)<br />

Therefore, the fluid particles move along helices spiral<strong>in</strong>g up the z axis, as illustrated <strong>in</strong><br />

Figure hel .<br />

The divergence <strong>of</strong> the vector field v is<br />

∇·v = ∂<br />

∂x (−y)+ ∂ ∂y x+ ∂ ∂z 1 = 0,<br />

and hence the flow is <strong>in</strong>compressible. Indeed, any fluid packet will spiral up the z axis<br />

unchanged <strong>in</strong> shape, and so its volume does not change.<br />

The vorticity or curl <strong>of</strong> the velocity is<br />

⎛<br />

∂<br />

⎜ ∂y 1− ∂ ∂z x ⎟ ⎛ ⎞<br />

⎞<br />

∇×v =<br />

∂<br />

⎜ ∂z (−y)− ∂<br />

∂x 1<br />

=<br />

⎝ ∂<br />

∂x x− ∂ ⎟<br />

⎠<br />

∂y (−y)<br />

which po<strong>in</strong>ts along the z-axis. This reflects the fact that the flow is spiral<strong>in</strong>g up the z-axis.<br />

If a turb<strong>in</strong>e is placed <strong>in</strong> the fluid at an angle ϕ with the z-axis, then its rate <strong>of</strong> rotation<br />

will be proportional to 2cosϕ.<br />

Example B.25. Any planar vector field v = (v 1 (x,y),v 2 (x,y)) T can be identified<br />

with a three-dimensional vector field<br />

v = (v 1 (x,y),v 2 (x,y),0) T<br />

that has no vertical component. If v represents a fluid velocity, then the fluid particles<br />

rema<strong>in</strong> on horizontal planes {z = c}, and the <strong>in</strong>dividual planar flows are identical. Its<br />

three-dimensional curl<br />

(<br />

∇×v = 0,0, ∂v 2<br />

∂x − ∂v ) T<br />

1<br />

∂y<br />

12/11/12 1256 c○ 2012 Peter J. Olver<br />

⎝ 0 0<br />

2<br />

⎠,


is a purely vertical vector field, whose third component agrees with the scalar two-dimensional<br />

curl (A.18) <strong>of</strong> v. This provides the direct identification between the two- and threedimensional<br />

versions <strong>of</strong> the curl operation. Indeed, our analysis <strong>of</strong> flows around airfoils<br />

<strong>in</strong> Chapter 16 directly relied upon this identification between two- and three-dimensional<br />

flows.<br />

Interconnections and Connectedness<br />

The three basic vector differential operators — gradient, curl and divergence — are<br />

<strong>in</strong>timately <strong>in</strong>ter-related. The pro<strong>of</strong> <strong>of</strong> the key identities relies on the equality <strong>of</strong> mixed partial<br />

derivatives, which <strong>in</strong> turn requires that the functions <strong>in</strong>volved are sufficiently smooth.<br />

We leave the explicit verification <strong>of</strong> the key result to the reader.<br />

Proposition B.26. If u is a smooth scalar field, then ∇×∇u ≡ 0. If v is a smooth<br />

vector field, then ∇·(∇×v) ≡ 0.<br />

Therefore, thecurl<strong>of</strong>anygradientvectorfieldisautomaticallyzero. Asaconsequence,<br />

allgradientvectorfieldsrepresent irrotationalflows. Also,thedivergence<strong>of</strong>anyvectorfield<br />

thatisacurl isalsoautomaticallyzero. Thus, allcurl vectorfields represent <strong>in</strong>compressible<br />

flows. On the other hand, the divergence <strong>of</strong> a gradient vector field is the Laplacian <strong>of</strong> the<br />

underly<strong>in</strong>g potential, as we previously noted, and hence is zero if and only if the potential<br />

is a harmonic function.<br />

The converse statements are almost true. As <strong>in</strong> the two-dimensional case, the precise<br />

statement <strong>of</strong> this result depends upon the topology <strong>of</strong> the underly<strong>in</strong>g doma<strong>in</strong>. In two<br />

dimensions, we only had to worry about whether or not the doma<strong>in</strong> conta<strong>in</strong>ed any holes,<br />

i.e., whether or not the doma<strong>in</strong> was simply connected. Similar concerns arise <strong>in</strong> three<br />

dimensions. Moreover, there are two possible classes <strong>of</strong> “holes” <strong>in</strong> a solid doma<strong>in</strong> — called<br />

tunnels and voids – and so there are two different types <strong>of</strong> connectivity. For lack <strong>of</strong> a<br />

better term<strong>in</strong>ology, we <strong>in</strong>troduce the follow<strong>in</strong>g def<strong>in</strong>ition.<br />

Def<strong>in</strong>ition B.27. A doma<strong>in</strong> Ω ⊂ R 3 is said to be<br />

(a) 0–connected or pathwise connected if there is a curve C ⊂ Ω connect<strong>in</strong>g any two po<strong>in</strong>ts<br />

x 0 ,x 1 ∈ Ω, so that † ∂C = {x 0 ,x 1 }.<br />

(b) 1–connected if every unknotted simple closed curve C ⊂ Ω is the boundary, C = ∂S<br />

<strong>of</strong> an oriented surface S ⊂ Ω.<br />

(c) 2–connected if every simple closed surface S ⊂ Ω is the boundary, S = ∂D <strong>of</strong> a<br />

subdoma<strong>in</strong> D ⊂ Ω.<br />

Remark: The unknotted condition is to avoid consider<strong>in</strong>g “wild” curves that fail to<br />

bound any oriented surface S ⊂ R 3 whatsoever.<br />

For example, R 3 is both 0, 1 and 2–connected, as are all solid balls, cubes, tetrahedra,<br />

solid cyl<strong>in</strong>ders, and so on. A disjo<strong>in</strong>t union <strong>of</strong> balls is not 0–connected, although it does<br />

rema<strong>in</strong>both1and2–connected. Thedoma<strong>in</strong>Ω = { 0 ≤ r < √ x 2 +y 2 < R } ly<strong>in</strong>gbetween<br />

† We use the notation ∂C to denote the endpo<strong>in</strong>ts <strong>of</strong> a curve C.<br />

12/11/12 1257 c○ 2012 Peter J. Olver


two cyl<strong>in</strong>ders is not 1–connected s<strong>in</strong>ce it has a “one-dimensional” hole drilled through it.<br />

Indeed, if C ⊂ Ω is any closed curve that encircles the <strong>in</strong>ner cyl<strong>in</strong>der, then every bound<strong>in</strong>g<br />

surface S with ∂S = C must pass across the <strong>in</strong>ner cyl<strong>in</strong>der and hence will not lie entirely<br />

with<strong>in</strong>thedoma<strong>in</strong>. Ontheotherhand, thiscyl<strong>in</strong>dricaldoma<strong>in</strong>Ωisboth0and2–connected<br />

— even an annular surface that encircles the <strong>in</strong>ner cyl<strong>in</strong>der will bound a solid annular<br />

doma<strong>in</strong> conta<strong>in</strong>ed <strong>in</strong>side Ω. Similarly, the doma<strong>in</strong> Ω = {0 ≤ r < ‖x‖ < R} between two<br />

concentric spheres is 0 and 1–connected, but not 2–connected ow<strong>in</strong>g to the spherical cavity<br />

<strong>in</strong>side. Any closed curve C ⊂ Ω will bound a surface S ⊂ Ω; for <strong>in</strong>stance, a circle go<strong>in</strong>g<br />

around the equator <strong>of</strong> the <strong>in</strong>ner sphere will still bound a hemispherical surface that does<br />

not pass through the spherical cavity. On the other hand, a sphere that lies between the<br />

<strong>in</strong>ner and outer spheres will not bound a solid doma<strong>in</strong> conta<strong>in</strong>ed with<strong>in</strong> the doma<strong>in</strong>. A<br />

full discussion <strong>of</strong> the topology underly<strong>in</strong>g the various types <strong>of</strong> connectivity, the nature <strong>of</strong><br />

tunnels and voids or cavities, and their connection with the existence <strong>of</strong> scalar and vector<br />

potentials, must be deferred to a more advanced course <strong>in</strong> differential topology, [23,86].<br />

We cannow statethebasic theorem relat<strong>in</strong>gthe connectivity<strong>of</strong> doma<strong>in</strong>sto thekernels<br />

<strong>of</strong> the fundamental vector differential operators; see [23] for details.<br />

Theorem B.28. Let Ω ⊂ R 3 be a doma<strong>in</strong>.<br />

(a) If Ω is 0–connected, then a scalar field u(x,y,z) def<strong>in</strong>ed on all <strong>of</strong> Ω has vanish<strong>in</strong>g<br />

gradient, ∇u ≡ 0, if and only if u(x,y,z) = constant.<br />

(b) If Ω is 1–connected, then a vector field v(x,y,z) def<strong>in</strong>ed on all <strong>of</strong> Ω has vanish<strong>in</strong>g<br />

curl, ∇×v ≡ 0, if and only if there is a scalar field ϕ, known as a scalar potential<br />

for v, such that v = ∇ϕ.<br />

(c) If Ω is 2–connected, then a vector field v(x,y,z) def<strong>in</strong>ed on all <strong>of</strong> Ω has vanish<strong>in</strong>g<br />

divergence, ∇ · v ≡ 0, if and only if there is a vector field w, known as a vector<br />

potential for v, such that v = ∇×w.<br />

If v represents the velocity vector field <strong>of</strong> a steady-state fluid flow, then the curl-free<br />

condition ∇×v ≡ 0 corresponds to an irrotational flow. Thus, on a 2–connected doma<strong>in</strong>,<br />

every irrotational flow field v has a scalar potential ϕ with ∇ϕ = v. The divergence-free<br />

condition ∇·v ≡ 0 corresponds to an <strong>in</strong>compressible flow. If the doma<strong>in</strong> is 1–connected,<br />

every <strong>in</strong>compressible flow field v has a vector potential w that satisfies ∇×w = v. The<br />

vector potential can be viewed as the three-dimensional analog <strong>of</strong> the stream function for<br />

planar flows. If the fluid is both irrotational and <strong>in</strong>compressible, then its scalar potential<br />

satisfies<br />

0 = ∇·v = ∇·∇ϕ = ∆ϕ,<br />

which is Laplace’s equation! Thus, just as <strong>in</strong> the two-dimensional case, the scalar potential<br />

to an irrotational, <strong>in</strong>compressible fluid flow is a harmonic function. This fact is used <strong>in</strong><br />

model<strong>in</strong>g many problems aris<strong>in</strong>g <strong>in</strong> physical fluids, <strong>in</strong>clud<strong>in</strong>g water waves, [126]. Unfortunately,<br />

<strong>in</strong> three dimensions there is no counterpart <strong>of</strong> complex function theory to represent<br />

the solutions <strong>of</strong> the Laplace equation, or to connect the vector and scalar potentials.<br />

Example B.29. The vector field<br />

v = (−y,x,1) T<br />

12/11/12 1258 c○ 2012 Peter J. Olver


that generates the helical flow (B.78) satisfies ∇·v = 0, and so is divergence-free, reconfirm<strong>in</strong>g<br />

our observation that the flow is <strong>in</strong>compressible. S<strong>in</strong>ce v is def<strong>in</strong>ed on all <strong>of</strong> R 3 ,<br />

Theorem B.28 assures us that there is a vector potential w that satisfies ∇×w = v. One<br />

candidate for the vector potential is<br />

w = ( y,0, 1 2 x2 + 1 2 y2) T<br />

.<br />

The helical flow is not irrotational, and so it does not admit a scalar potential.<br />

Remark: The construction <strong>of</strong> a vector potential is not entirely straightforward, but we<br />

will not dwell on this problem. Unlike a scalar potential which, when it exists, is uniquely<br />

def<strong>in</strong>ed up to a constant, there is, <strong>in</strong> fact, quite a bit <strong>of</strong> ambiguity <strong>in</strong> a vector potential.<br />

Add<strong>in</strong>g <strong>in</strong> any gradient,<br />

˜w = w+∇ϕ<br />

will give an equally valid vector potential. Indeed, us<strong>in</strong>g Proposition B.26, we have<br />

Thus, any vector field <strong>of</strong> the form<br />

∇× ˜w = ∇×w+∇×∇ϕ = ∇×w.<br />

w =<br />

(<br />

y + ∂ϕ<br />

∂x , ∂ϕ<br />

∂y , x 2<br />

2 + y2<br />

2 + ∂ϕ ) T<br />

,<br />

∂z<br />

where ϕ(x,y,z) is an arbitrary function, is also a valid vector potential for the helical<br />

vector field v = (−y,x,1) T .<br />

B.8. The Fundamental Integration Theorems.<br />

In three-dimensional vector calculus there are 3 fundamental differential operators<br />

— gradient, curl and divergence. There are also 3 types <strong>of</strong> <strong>in</strong>tegration — l<strong>in</strong>e, surface<br />

and volume <strong>in</strong>tegrals. And, not co<strong>in</strong>cidentally, there are 3 basic theorems that generalize<br />

the Fundamental Theorem <strong>of</strong> <strong>Calculus</strong> to l<strong>in</strong>e, surface and volume <strong>in</strong>tegrals <strong>in</strong> threedimensional<br />

space. In all three results, the <strong>in</strong>tegral <strong>of</strong> some differentiated quantity over a<br />

curve, surface, or doma<strong>in</strong> is related to an <strong>in</strong>tegral <strong>of</strong> the quantity over its boundary. The<br />

first theorem relates the l<strong>in</strong>e <strong>in</strong>tegral <strong>of</strong> a gradient over a curve to the values <strong>of</strong> the function<br />

at the boundary or endpo<strong>in</strong>ts <strong>of</strong> the curve. Stokes’ Theorem relates the surface <strong>in</strong>tegral <strong>of</strong><br />

the curl <strong>of</strong> a vector field to the l<strong>in</strong>e <strong>in</strong>tegral <strong>of</strong> the vector field around the boundary curve<br />

<strong>of</strong> the surface. F<strong>in</strong>ally, the Divergence Theorem, also known as Gauss’ Theorem, relates<br />

the volume <strong>in</strong>tegral <strong>of</strong> the divergence <strong>of</strong> a vector field to the surface <strong>in</strong>tegral <strong>of</strong> that vector<br />

field over the boundary <strong>of</strong> the doma<strong>in</strong>.<br />

The Fundamental Theorem for L<strong>in</strong>e Integrals<br />

We beg<strong>in</strong> with the Fundamental Theorem for l<strong>in</strong>e <strong>in</strong>tegrals. This is identical to the<br />

planar version, as stated earlier <strong>in</strong> Theorems A.20 and A.22. We do not need to reproduce<br />

its pro<strong>of</strong> aga<strong>in</strong> here.<br />

12/11/12 1259 c○ 2012 Peter J. Olver


Theorem B.30. Let C ⊂ R 3 be a curve that starts at the endpo<strong>in</strong>t a and goes to<br />

the endpo<strong>in</strong>t b. Then the l<strong>in</strong>e <strong>in</strong>tegral <strong>of</strong> a gradient <strong>of</strong> a function along C is given by<br />

∫<br />

∇u· dx = u(b)−u(a).<br />

(B.79)<br />

C<br />

S<strong>in</strong>ce its value only depends upon the endpo<strong>in</strong>ts, the l<strong>in</strong>e <strong>in</strong>tegral <strong>of</strong> a gradient is<br />

<strong>in</strong>dependent <strong>of</strong> path. In particular, if C is a closed curve, then a = b, and so the endpo<strong>in</strong>t<br />

contributions cancel out: ∮<br />

∇u· dx = 0.<br />

C<br />

Conversely, if v is any vector field with the property that its <strong>in</strong>tegral around any closed<br />

curve vanishes, ∮<br />

v· dx = 0, (B.80)<br />

C<br />

then v = ∇ϕ admits a potential. Indeed, as long as the doma<strong>in</strong> is 0–connected, one can<br />

construct a potential ϕ(x) by <strong>in</strong>tegrat<strong>in</strong>g over any convenient curve C connect<strong>in</strong>g a fixed<br />

po<strong>in</strong>t a ∈ Ω to the po<strong>in</strong>t x<br />

ϕ(x) =<br />

∫ x<br />

a<br />

v· dx.<br />

The pro<strong>of</strong> that this is a well-def<strong>in</strong>ed potential is similar to the planar version discussed <strong>in</strong><br />

Chapter A.<br />

If v represents the velocity vector field <strong>of</strong> a three-dimensional steady state fluid flow,<br />

then its l<strong>in</strong>e <strong>in</strong>tegral around a closed curve C, namely<br />

∮ ∮<br />

v·dx = v·t ds<br />

C<br />

is the <strong>in</strong>tegral <strong>of</strong> the tangential component <strong>of</strong> the velocity vector field. This represents the<br />

circulation <strong>of</strong> the fluid around the curve C. In particular, if the circulation l<strong>in</strong>e <strong>in</strong>tegral is 0<br />

for every closed curve, then the fluid flow will be irrotational because ∇×v = ∇×∇ϕ ≡ 0.<br />

Stokes’ Theorem<br />

The second <strong>of</strong> the three fundamental <strong>in</strong>tegration theorems is known as Stokes’ Theorem.<br />

This important result relates the circulation l<strong>in</strong>e <strong>in</strong>tegral <strong>of</strong> a vector field around<br />

a closed curve with the <strong>in</strong>tegral <strong>of</strong> its curl over any bound<strong>in</strong>g surface. Stokes’ Theorem<br />

first appeared <strong>in</strong> an 1850 letter from Lord Kelv<strong>in</strong> (William Thompson) written to George<br />

Stokes, who made it <strong>in</strong>to an undergraduate exam question for the Smith Prize at Cambridge<br />

University <strong>in</strong> England.<br />

Theorem B.31. Let S ⊂ R 3 be an oriented, bounded surface whose boundary ∂S<br />

consists <strong>of</strong> one or more piecewise smooth simple closed curves. Let v be a smooth vector<br />

field def<strong>in</strong>ed on S. Then<br />

∮ ∫∫<br />

v·dx = (∇×v)·n dS. (B.81)<br />

∂S<br />

S<br />

C<br />

12/11/12 1260 c○ 2012 Peter J. Olver


To make sense <strong>of</strong> Stokes’ formula (B.81), we need to assign a consistent orientation to<br />

the surface — mean<strong>in</strong>g a choice <strong>of</strong> unit normal n — and to its boundary curve — mean<strong>in</strong>g<br />

a direction to go around it. The proper choice is described by the follow<strong>in</strong>g left hand<br />

rule: If we walk along the boundary ∂S with the normal vector n on S po<strong>in</strong>t<strong>in</strong>g upwards,<br />

then the surface should be on our left hand side; see Figure Stokes . For example, if<br />

S ⊂ {z = 0} is a planar doma<strong>in</strong> and we choose the upwards normal n = (0,0,1) T , then<br />

C should be oriented <strong>in</strong> the usual, counterclockwise direction. Indeed, <strong>in</strong> this case, Stokes’<br />

Theorem B.31 reduces to Green’s Theorem A.26!<br />

Stokes’ formula (B.81) can be rewritten us<strong>in</strong>g the alternative notations (B.21), (B.47),<br />

for surface and l<strong>in</strong>e <strong>in</strong>tegrals <strong>in</strong> the form<br />

∮<br />

udx+vdy +wdz =<br />

∂S<br />

∫∫<br />

S<br />

( ∂w<br />

∂y − ∂v<br />

∂z<br />

)<br />

dydz +<br />

( ∂u<br />

∂z − ∂w ) ( ∂v<br />

dzdx+<br />

∂x ∂x − ∂u )<br />

dxdy.<br />

∂y<br />

(B.82)<br />

Recall that a closed surface is one without boundary: ∂S = ∅. In this case, the left<br />

hand side <strong>of</strong> Stokes’ formula (B.81) is zero, and we f<strong>in</strong>d that <strong>in</strong>tegrals <strong>of</strong> curls vanish on<br />

closed surfaces.<br />

∫∫<br />

Proposition B.32. If the vector field v = ∇×w is a curl, then v·n dS = 0<br />

for every closed surface S.<br />

Thus, every curl vector field def<strong>in</strong>es a surface-<strong>in</strong>dependent <strong>in</strong>tegral.<br />

Example B.33. Let S = {x+y +z = 1, x > 0, y > 0, z > 0} denote the triangular<br />

surface considered <strong>in</strong> Example B.15. Its boundary ∂S = L x ∪ L y ∪ L z is a triangle<br />

composed <strong>of</strong> three l<strong>in</strong>e segments<br />

To compute the l<strong>in</strong>e <strong>in</strong>tegral<br />

∮<br />

L x = {x = 0, y +z = 1, y ≥ 0, z ≥ 0},<br />

L y = {y = 0, x+z = 1, x ≥ 0, z ≥ 0},<br />

L z = {z = 0, x+y = 1, x ≥ 0, y ≥ 0}.<br />

∂S<br />

∮<br />

v·dx =<br />

∂S<br />

y 2 dx+xz 2 dy<br />

<strong>of</strong> the vector field v = ( y 2 ,xz 2 ,0 ) T<br />

, we could proceed directly, but this would require<br />

evaluat<strong>in</strong>gthreeseparate<strong>in</strong>tegralsoverthethreesides<strong>of</strong>thetriangle. Asanalternative, we<br />

can use Stokes formula (B.81), and compute the <strong>in</strong>tegral <strong>of</strong> its curl ∇×v = (2y,2xz,0) T<br />

over the triangle, which is<br />

∮<br />

∂S<br />

∫∫<br />

v·dx =<br />

S<br />

∫∫<br />

(∇×v)·n dS =<br />

S<br />

2ydydz +2xzdzdx = 17<br />

12 ,<br />

where this particular surface <strong>in</strong>tegral was already computed <strong>in</strong> Example B.15.<br />

12/11/12 1261 c○ 2012 Peter J. Olver<br />

S


We remark that Stokes’ Theorem B.31 is consistent with Theorem B.28. Suppose<br />

that v is a curl-free vector field, so ∇×v = 0, which is def<strong>in</strong>ed on a 1–connected doma<strong>in</strong><br />

Ω ⊂ R 3 . S<strong>in</strong>ce every simple (unknotted) closed curve C ⊂ Ω bounds a surface, C = ∂S,<br />

with S ⊂ Ω also conta<strong>in</strong>ed <strong>in</strong>side the doma<strong>in</strong>, then, Stokes’ formula (B.81) implies<br />

∮ ∫∫<br />

v·dx = (∇×v)·n dS = 0.<br />

C<br />

S<br />

S<strong>in</strong>ce this happens for every † C ⊂ Ω, then the path-<strong>in</strong>dependence condition (B.80) is<br />

satisfied, and hence v = ∇ϕ admits a potential.<br />

Example B.34. The Newtonian gravitational force field<br />

v(x) = x<br />

‖x‖ 3 =<br />

(x,y,z) T<br />

(x 2 +y 2 +z 2 ) 3/2<br />

is well def<strong>in</strong>ed on Ω = R 3 \{0}, and is divergence-free: divv ≡ 0. Nevertheless, this vector<br />

field does not admit a vector potential. Indeed, on the sphere S a = {‖x‖ = a} <strong>of</strong> radius<br />

a, the unit normal vector at a po<strong>in</strong>t x ∈ S a is n = x/‖x‖. Therefore,<br />

∫∫ ∫∫<br />

v·n dS =<br />

S a<br />

S a<br />

x<br />

‖x‖ 3 ·<br />

∫∫<br />

x<br />

‖x‖ dS =<br />

S a<br />

1<br />

‖x‖ 2 dS = 1 a 2 ∫∫<br />

S a<br />

dS = 4π,<br />

s<strong>in</strong>ce S a has surface area 4πa 2 . Note that this result is <strong>in</strong>dependent <strong>of</strong> the radius <strong>of</strong> the<br />

sphere. If v = ∇×w, this would contradict Proposition B.32.<br />

The problem is, <strong>of</strong>course, thatthe doma<strong>in</strong>Ωisnot 2–connected, andso Theorem B.28<br />

does not apply. However, it would apply to the vector field v on any 2–connected subdoma<strong>in</strong>,<br />

for example the doma<strong>in</strong> ˜Ω = R 3 \ {x = y = 0,z ≤ 0} obta<strong>in</strong>ed by omitt<strong>in</strong>g the<br />

negative z-axis. Exercise asks you to construct a vector potential <strong>in</strong> this case.<br />

We further note that v is curl free: ∇ × v ≡ 0. S<strong>in</strong>ce the doma<strong>in</strong> <strong>of</strong> def<strong>in</strong>ition Ω<br />

is 1–connected, Theorem B.28 tells us that v admits a scalar potential — the Newtonian<br />

gravitational potential. Indeed, ∇ ( ‖x‖ −1) = v, as the reader can check.<br />

The Divergence Theorem<br />

The last <strong>of</strong> the three fundamental <strong>in</strong>tegral theorems is the Divergence Theorem, also<br />

known as Gauss’ Theorem. This result relates a surface flux <strong>in</strong>tegral over a closed surface<br />

to a volume <strong>in</strong>tegral over the doma<strong>in</strong> it bounds.<br />

Theorem B.35. Let Ω ⊂ R 3 be a bounded doma<strong>in</strong> whose boundary ∂Ω consists<br />

<strong>of</strong> one or more piecewise smooth simple closed surfaces. Let n denote the unit outward<br />

normal to the boundary <strong>of</strong> Ω. Let v be a smooth vector field def<strong>in</strong>ed on Ω and cont<strong>in</strong>uous<br />

up to its boundary. Then<br />

∫∫ ∫∫∫<br />

v·n dS = ∇·v dxdydz. (B.83)<br />

∂Ω<br />

Ω<br />

† It suffices to know this for unknotted curves to conclude it for arbitrary closed curves.<br />

12/11/12 1262 c○ 2012 Peter J. Olver


In terms <strong>of</strong> the alternative notation (B.47) for surface <strong>in</strong>tegrals, the divergence formula<br />

(B.83) can be rewritten <strong>in</strong> the form<br />

∫∫<br />

∫∫∫ ( ∂u<br />

u dydz +vdzdx+wdxdy =<br />

∂x + ∂v<br />

∂y + ∂w )<br />

dxdydz. (B.84)<br />

∂z<br />

S<br />

Example B.36. Let us compute the surface <strong>in</strong>tegral<br />

∫∫<br />

xydzdx+zdxdy<br />

S<br />

<strong>of</strong> the vector field v = (0,xy,z) T over the sphere S = {‖x‖ = 1} <strong>of</strong> radius 1. A direct<br />

evaluation<strong>in</strong>eithergraphicalorsphericalcoord<strong>in</strong>atesisnotsopleasant. Butthedivergence<br />

formula (B.84) immediately gives<br />

∫∫<br />

∫∫∫ ( ∂(xy)<br />

xydzdx+zdxdy = + ∂z )<br />

dxdydz<br />

S<br />

Ω ∂y ∂z<br />

∫∫∫ ∫∫∫ ∫∫∫<br />

= (x+1)dxdydz = xdxdydz + dxdydz = 4 3 π,<br />

Ω<br />

where Ω = {‖x‖ < 1} is the unit ball with boundary ∂Ω = S. The f<strong>in</strong>al two <strong>in</strong>tegrals are,<br />

respectively, the x coord<strong>in</strong>ate <strong>of</strong> the center <strong>of</strong> mass <strong>of</strong> the sphere multiplied by its volume,<br />

which is clearly 0, plus the volume <strong>of</strong> the spherical ball.<br />

Example B.37. Suppose v(t,x) is the velocity vector field <strong>of</strong> a time-dependent<br />

fluid flow. Let ρ(t,x) ∫∫ represent the density <strong>of</strong> the fluid at time t and position x. Then the<br />

surface flux <strong>in</strong>tegral (ρv)·ndS represents the mass flux <strong>of</strong> fluid through the surface<br />

S<br />

S ⊂ R 3 . In particular, if S = ∂Ω represents a closed surface bound<strong>in</strong>g a doma<strong>in</strong> Ω, then,<br />

by the Divergence Theorem B.35,<br />

∫∫ ∫∫∫<br />

(ρv)·ndS = ∇·(ρv)dxdydz<br />

∂Ω<br />

represents the net mass flux out <strong>of</strong> the doma<strong>in</strong> Ω at time t. On the other hand, this must<br />

equal the rate <strong>of</strong> change <strong>of</strong> mass <strong>in</strong> the doma<strong>in</strong>, namely<br />

− ∂ ∫∫∫ ∫∫∫<br />

∂ρ<br />

ρdxdydz = −<br />

∂t<br />

∂t dxdydz,<br />

Ω<br />

the m<strong>in</strong>us sign com<strong>in</strong>g from the fact that we are measur<strong>in</strong>g net mass loss due to outflow.<br />

Equat<strong>in</strong>g these two, we discover that<br />

∫∫∫ ( ) ∂ρ<br />

∂t +∇·(ρv) dxdydz = 0<br />

Ω<br />

for every doma<strong>in</strong> occupied by the fluid. S<strong>in</strong>ce the doma<strong>in</strong>is arbitrary, this can only happen<br />

if the <strong>in</strong>tegrand vanishes, and hence<br />

∂ρ<br />

∂t<br />

Ω<br />

Ω<br />

Ω<br />

+∇·(ρv) = 0.<br />

(B.85)<br />

Ω<br />

Ω<br />

12/11/12 1263 c○ 2012 Peter J. Olver


The latter is the basic cont<strong>in</strong>uity equation <strong>of</strong> fluid mechanics, which takes the form <strong>of</strong> a<br />

conservation law.<br />

For a steady state fluid flow, the left hand side <strong>of</strong> the divergence formula (B.83)<br />

measures the fluid flux through the boundary <strong>of</strong> the doma<strong>in</strong> ∂Ω, while the left hand side<br />

<strong>in</strong>tegrates the divergence over the doma<strong>in</strong> Ω. As a consequence, the divergence must<br />

represent the net local change <strong>in</strong> fluid volume at a po<strong>in</strong>t under the flow. In particular, if<br />

∇v = 0, then there is no net flux, and the fluid flow is <strong>in</strong>compressible.<br />

The Divergence Theorem B.35 is also consistent with Theorem B.28. Let v is a<br />

divergence-free vector field, ∇ · v = 0, def<strong>in</strong>ed on a 2–connected doma<strong>in</strong> Ω ⊂ R 3 . Every<br />

simple closed surface S ⊂ Ω bounds a subdoma<strong>in</strong>, so S = ∂D, with D ⊂ Ω also conta<strong>in</strong>ed<br />

<strong>in</strong>side the doma<strong>in</strong> <strong>of</strong> def<strong>in</strong>ition <strong>of</strong> v. Then, by the divergence formula (B.83),<br />

∫∫ ∫∫∫<br />

v·n dS = ∇·v dxdydz = 0.<br />

S<br />

Therefore, by Theorem B.28, v = ∇×w admits a vector potential.<br />

Remark: The pro<strong>of</strong> <strong>of</strong> all three <strong>of</strong> the fundamental <strong>in</strong>tegral theorems, can, <strong>in</strong> fact, be<br />

reduced to the Fundamental Theorem <strong>of</strong> (one-variable) <strong>Calculus</strong>. They are, <strong>in</strong> fact, all<br />

special cases <strong>of</strong> the general Stokes’ Theorem, which forms the foundation <strong>of</strong> the pr<strong>of</strong>ound<br />

theory <strong>of</strong> <strong>in</strong>tegration on manifolds, [2,23,70]. Stokes’ Theorem has deep and beautiful<br />

connections with topology — and is <strong>of</strong> fundamental importance <strong>in</strong> modern mathematics<br />

and physics. However, the full ramifications lie beyond the scope <strong>of</strong> this <strong>in</strong>troductory text.<br />

Ω<br />

12/11/12 1264 c○ 2012 Peter J. Olver

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!