5.3 Surface Integrals

5.3 Surface Integrals

  1. Surface Area and Surface Integrals
  2. An Invariance Property
  3. Volume and Area
  4. Problems

\(\Leftarrow\)  \(\Uparrow\)  \(\Rightarrow\)

Surface Area and Surface Integrals

Suppose that \(S\) is a surface in \(\R^3\) parametrized by a function \(\mathbf G:T\to S\), where \(T\) is a measurable subset of \(\R^2\), \(\mathbf G\) is \(C^1\) and injective, and that \(\left\{\partial_1 \mathbf G, \partial_2 \mathbf G\right\}\) are linearly independent at every point in \(T\), except possibly on a set of zero content.

If \(|\partial_1 \mathbf G \times \partial_2 \mathbf G|\) is integrable on \(T\), then we define the surface area of \(S\) as \[ \iint_T |\partial_1 \mathbf G \times \partial_2 \mathbf G|\, dA \]
With the same set up, suppose further that \(f:\R^3\to \R\) is continuous function, and that \(\mathbf F\) is a continuous vector field on \(\R^3\). We define the surface integral of \(f\) over \(S\) as \[ \iint_S f\, dA = \iint_T f(\mathbf G(\mathbf u)) |\partial_1 \mathbf G \times \partial_2 \mathbf G|\, du \, dv, \] and the surface integral of \(\mathbf F\) over \(S\) as \[ \iint_S \mathbf F\cdot \mathbf n dA = \iint_T \mathbf F(\mathbf G(\mathbf u)) \cdot \left(\partial_1 \mathbf G \times \partial_2 \mathbf G\right) \, du \, dv . \]

We also sometimes say or write “integral” instead of “surface integral” when referring to the above integrals, if it is clear that we are integrating over a surface.

Note that the area of \(S\) equals \(\iint_S 1\, dA\).

Although we have used a parametrization in the above definitions, we will see below that, exactly as with line integrals,

Example 1

Let \(S\) be the set \(\left\{ (x,y,z)\in \R^3 : x^2+y^2+z^2 = 1, \ z> 0\right\}\). Compute the area of \(S\), as well as the integrals \[ \iint_S z\, dA, \quad\text{ and }\quad \iint_S (0,0,z)\cdot \mathbf n \, dA, \] with \(\mathbf n\) oriented upwards (i.e. \(\mathbf n \cdot [0,0,1]>0\)).

Solution. First we parametrize \(S\). We can do this by spherical coordinates with \(r=1\), \[ \mathbf x = \mathbf G(\theta,\phi)= (\cos\theta \sin\phi , \sin\theta\sin\phi ,\cos \phi), \qquad 0\le\theta\le 2\pi, \ \ 0\le \phi\le \pi/2 . \] Now we compute the integrals. \[\begin{align*} \partial_\theta \mathbf G &= ( -\sin\theta \sin\phi , \cos\theta\sin\phi ,0) \\ \partial_\phi\mathbf G &=( \cos\theta \cos\phi , \sin\theta\cos\phi ,-\sin \phi),\quad\text{ and thus } \\ \partial_\theta \mathbf G\times \partial_\phi \mathbf G &=(-\cos\theta\sin^2\phi, -\sin\theta\sin^2\phi, -\cos\phi\sin \phi) \end{align*}\] Then a short computation shows that \[ |\partial_\theta \mathbf G\times \partial_\phi \mathbf G| = |\sin \phi| = \sin \phi \] since \(\sin \phi\ge 0\) for \(\phi\in [0,\pi/2]\). It follows that \[\begin{align*} &\text{area}(S) = \int_0^{2\pi}\int_0^{\pi/2} \sin \phi \, d\phi\, d\theta = \cdots = 2\pi \\ &\iint_S z \, dA =\int_0^{2\pi}\int_0^{\pi/2} \cos\phi \sin\phi\, d\phi \, d\theta = \pi. \end{align*}\] The substitution \(u = \sin \phi\) is helpful in evaluating the second integral.

Before evaluating the integral of the vector field, we have to pay attention to orientation. The problem says that \(\mathbf n\) should be oriented upwards. What this means is that \(\frac{\partial \mathbf G}{\partial u}\times \frac {\partial \mathbf G}{\partial v}\) should point upwards. “Point upwards” means that the \(z\)-component of the vector should be positive. We can see that for the parametrization we have chosen, the \(z\)-component of \(\frac{\partial \mathbf G}{\partial u}\times \frac {\partial \mathbf G}{\partial v}\)is \(-\cos\phi \sin\phi\), which is negative for \(0\le \phi < \pi/2\), hence points downwards. So to correct for the wrong orientation, we multiply by \(-1\). Thus \[ \iint_S (0,0,z)\cdot \mathbf n \, dA = -\int_0^{2\pi}\int_0^{\pi/2} -\cos^2\phi\sin \phi \, d\phi\,d\theta = +\frac 2 3 \pi, \] using the substitution \(u = \cos\phi\).

The meaning of \(\partial_1 \mathbf G \times \partial_2 \mathbf G\).

When \(\left\{\partial_1 \mathbf G, \partial_2 \mathbf G\right\}\) are linearly independent (as we typically require, except possibly for a set of zero content), then

As a result of these considerations, \[ \frac{ \partial_1 \mathbf G \times \partial_2 \mathbf G } {|\partial_1 \mathbf G \times \partial_2 \mathbf G|} \qquad\text{ and }\qquad \frac{ -\partial_1 \mathbf G \times \partial_2 \mathbf G } {|\partial_1 \mathbf G \times \partial_2 \mathbf G|} = \frac{ \partial_2 \mathbf G \times \partial_1 \mathbf G } {|\partial_2 \mathbf G \times \partial_1 \mathbf G|} \] are unit normal vectors to \(S\) — unit vectors that are normal to \(S\) at the point \(\mathbf G(u,v)\).

We therefore sometimes write \(\dfrac{\partial_1 \mathbf G \times \partial_2 \mathbf G} {|\partial_1 \mathbf G \times \partial_2 \mathbf G|}\) as \(\mathbf n(\mathbf G(\mathbf u))\), where “\(\mathbf n\)” stands for “normal”.

With this notation, for any vector field \(\mathbf F\), the surface integral of the real-valued function \(\mathbf F\cdot \mathbf n\) over \(S\) is \[\begin{align} \iint_S \mathbf F\cdot \mathbf n \,dA &= \iint_T \mathbf F(\mathbf G(\mathbf u))\cdot \left(\partial_1 \mathbf G(\mathbf u) \times \partial_2 \mathbf G(\mathbf u)\right)\, du\, dv\nonumber \\ &= \iint_T \mathbf F(\mathbf G(\mathbf u))\cdot \mathbf n(\mathbf G(\mathbf u)) \left| \partial_1 \mathbf G \times \partial_2 \mathbf G \right| \, du\, dv\nonumber \\ \end{align}\] For this reason, the notation \(\iint_S \mathbf F\cdot \mathbf n\, dA\) for a surface integral of a vector field is entirely consistent with the notation for the surface integral of a real-valued function.

So far we have discussed the direction of \(\partial_1 \mathbf G \times \partial_2 \mathbf G\). What about its magnitude? Properties of the cross product imply that its magnitude is exactly the area of the parallelogram in \(\R^3\) with sides \(\partial_1 \mathbf G\) and \(\partial_2 \mathbf G\). This leads to a non-rigorous but suggestive interpretation, which is that \[\begin{align} | \partial_1 \mathbf G \times \partial_2 \mathbf G|\,du\,dv & = | (\partial_1 \mathbf G du) \times (\partial_2 \mathbf G dv)|\nonumber \\ \end{align}\] regarding this as the area of the infinitesimal parallelogram with sides \(\partial_1 \mathbf G du\) and \(\partial_2 \mathbf G dv\).

Thus, it is traditional to imagine that when we integrate \(\iint_R | \partial_1 \mathbf G \times \partial_2 \mathbf G|\,du\,dv\), we are “adding up the areas of the infinitely many infinitesimal parallelograms that make up the surface \(S\)”. This can be viewed as a sort of justification for our defintion of area. One can also make similar arguments with rectangles of very small but but finite size. Other justifications will be presented below.

Piecewise Smooth Surfaces

Example 2

Let \(S\) be the boundary of the set \(\left\{ (x,y,z)\in \R^3 : x^2+y^2\le 1, |z|\le 1\right\}\). Compute \(\iint_S \mathbf F\cdot \mathbf n\, dA\) for \(\mathbf F = (x^2z,y,0)\), and with \(\mathbf n\) oriented outwards.

Solution. A feature of this problem is that \(\partial S\) is not a smooth surface everywhere. However, it is a piecewise smooth surface, that is, a union of smooth surfaces: \[ \begin{aligned} S_{top} &= \left\{(x,y,z)\in \R^3 : \ x^2+y^2\le 1, \ z = 1\right\} \\ S_{side} &= \left\{(x,y,z)\in \R^3 : \ x^2+y^2 = 1, \ |z|\le 1\right\} \\ S_{bottom} &= \left\{(x,y,z)\in \R^3 : \ x^2+y^2\le 1, \ z = - 1\right\} \end{aligned} \] that intersect at sets of lower dimension. For example, \(S_{top}\cap S_{side} = \left\{(x,y,z) : x^2+y^2 = 1, z = 1\right\}\) which is a smooth curve in \(\R^3\), hence has zero area. In cases like this, we simply integrate along each of the surfaces and add up the results.

  1. Integral over \(S_{top}\).. With some practice you can see right away that this integral must equal \(0\), because we can see geometrically that \(\mathbf n = (0,0,1)\) on \(S_{top}\), and hence that \(\mathbf F\cdot \mathbf n = 0\) there, for our specific \(\mathbf F\). But for practice, let’s do it:

    • parametrize by \(\mathbf x = (u,v,1) = \mathbf G(u,v)\) for \((u,v)\in R = \left\{(u,v)\in \R^2 : u^2+v^2\le 1\right\}\).
    • then \(\partial_u \mathbf G = (1,0,0)\) and \(\partial_v \mathbf G = (0,1,0)\), so \(\partial_u \mathbf G \times \partial_v \mathbf G = (0,0,1)\).
    • Thus, after consulting the definition of \(\mathbf F\), we see that \(\mathbf F(\mathbf G(u))\cdot \partial_u \mathbf G \times \partial_v \mathbf G = 0\) for all \((u,v)\in R\). So it is clear that when we integrate, we will get \(0\). (Thus we don’t need to worry about orientation, since if the orientation is wrong, it would just change our answer to \(-0\).)
    • We conclude that \[ \iint_{S_{top}} \mathbf F\cdot \mathbf n \, dA = 0. \]
  2. Integral over \(S_{bottom}\). For exactly the same reason, \[ \iint_{S_{bottom}} \mathbf F\cdot \mathbf n \, dA = 0. \] For practice, let’s not do it. It’s just as important to practice avoiding computations as doing computations.

  3. Integral over \(S_{side}\)..

    • parametrize by \(\mathbf x = (\cos u, \sin u, v) = \mathbf G(u,v)\) for \((u,v)\in R = \left\{(u,v)\in \R^2 : 0\le u\le 2\pi, |v|\le 1\right\}\).
    • then \(\partial_u \mathbf G = (-\sin u,\cos u,0)\) and \(\partial_v \mathbf G = (0,0,1)\), so \(\partial_u \mathbf G \times \partial_v \mathbf G = (\cos u,\sin u,0)\).
    • Check the orientation. We require \(\mathbf n\) to point outward, and we see that \((\cos u,\sin u, 0)\) does this. For example, if we start at a point \(\mathbf G(u,v) = (\cos u,\sin u, v)\) and we move in the \((\cos u, \sin u, 0)\) direction, we move out of the set enclosed by \(S\). So we can use this parametrization as is, and we do not need to multiply by \(-1\).
    • Recalling that \(\mathbf F = (x^2z,y,0)\), we conclude that \[ \iint_{S_{side}} \mathbf F\cdot \mathbf n \, dA = \int_0^{2\pi} \int_{-1}^1 (v \cos^2 u , \sin u, 0)\cdot (\cos u, \sin u, 0) dv\, du = \cdots = 2\pi. \]

Special Cases

The above formulas simplify if \(S\) is the graph of some function \(\phi\), i.e. \(S = \left\{ (x,y, \phi(x,y) ): (x,y)\in T\right\}\). Then we can use the parametrization \[ \mathbf x = (u , v , \phi(u,v)) = \mathbf G(u,v), \qquad (u,v)\in T. \] With this choice, \[ \partial_u\mathbf G = (1, 0, \partial_u \phi), \qquad \partial_v\mathbf G = ( 0,1, \partial_v \phi), \qquad \partial_u\mathbf G\times \partial_v\mathbf G=(-\partial_u \phi, -\partial_v\phi, 1). \] and thus \(|\partial_u\mathbf G\times \partial_v\mathbf G| = \sqrt {1+|\nabla\phi|^2}\). So we get the simpler formulas \[ \text{area}(S) = \iint_T \sqrt {1+|\nabla\phi|^2}\, dA \] \[ \iint_S f\, dA = \iint_T f(u,v,\phi(u,v)) \sqrt {1+|\nabla\phi|^2} \, dA \] and \[ \iint_S \mathbf F\cdot \mathbf n \, dA = \iint_T \mathbf F(u,v,\phi(u,v)) \cdot (-\partial_u\phi, -\partial_v\phi, 1) \, dA \] if \(S\) is oriented with the unit normal pointing upwards; for the downward unit normal, one would have to multiply by \(-1\).

Since \(u=x\) and \(v=y\), these formulas are often written in terms of variables \((x,y)\) rather than \((u,v)\).

An Invariance Property

Similarly to the line integral, we will show that the integral of a real-valued function \(f\) over a surface \(S\) is independent of the parametrization.

Suppose that \(S\) is a surface parametrized by \[ (x,y,z) = \mathbf G(u,v) \ \ \ \text{ for }(u,v) \in T, \] and that \(\varphi:W\to T\) is a function that is one-to-one, onto, of class \(C^1\), and with \(C^1\) inverse. Here both \(W\) and \(T\) are supposed to be measurable subsets of \(\R^2\). Let us say that the variables in \(W\) are called \((s,t)\), so that the \((s,t)\) and \((u,v)\) variables are related by \((u,v) = \varphi(s,t)\).

Now define \(\mathbf H(s,t) = \mathbf G\circ\varphi(s,t)\). Then \(S\) is also parametrized by \(\mathbf x = \mathbf H(s,t), (s,t)\in W\).

Our assumptions about \(\varphi\) imply that \(\det D\varphi(s,t)\ne 0\) for all \((s,t)\in W\). Also, by a short computation that uses the chain rule, one can check that

\[\begin{equation}\label{cvsurf} \frac{\partial \mathbf H}{\partial s}\times \frac{\partial \mathbf H}{\partial t} = \left(\frac{\partial \mathbf G}{\partial u}\times \frac{\partial \mathbf G}{\partial v}\right) \cdot \det D\varphi \end{equation}\]

where derivatives of \(\mathbf H, \varphi\) are evalated at \(s,t\) and derivatives of \(\mathbf G\) are evaluated at \((u,v) = \varphi(s,t)\). It follows that

\[ \begin{aligned} \det D\varphi >0 & \text{ everywhere in }W \\ &\ \iff \ \frac{\partial \mathbf H}{\partial s}\times \frac{\partial \mathbf H}{\partial t} \text{ and } \frac{\partial \mathbf G}{\partial u}\times \frac{\partial \mathbf G}{\partial v}\text{ point in the same direction} \\ &\ \iff \ \text{ parametrizations }\mathbf G\text{ and }\mathbf H\text{ induce the same orientation}, \end{aligned} \] where the orientation induced by a parametrization \(\mathbf G\) is the orientation of \(\mathbf n = \partial_u\mathbf G\times \partial_v \mathbf G/|\partial_u\mathbf G\times \partial_v \mathbf G|\).

With the above set up, \[ \iint_T f(\mathbf G(u,v))\left|\frac{\partial \mathbf G}{\partial u}\times \frac{\partial \mathbf G}{\partial v}\right| du\,dv = \iint_W f(\mathbf H(s,t))\left|\frac{\partial \mathbf H}{\partial s}\times \frac{\partial \mathbf H}{\partial t}\right| ds\,dt \] and \[ \iint_T \mathbf F(\mathbf G(u,v))\cdot\left(\frac{\partial \mathbf G}{\partial u}\times \frac{\partial \mathbf G}{\partial v}\right) du\,dv = \pm \iint_W \mathbf F(\mathbf H(s,t))\cdot \left(\frac{\partial \mathbf H}{\partial s}\times \frac{\partial \mathbf H}{\partial t}\right) ds\,dt \] with “\(+\)” if \(\mathbf G,\mathbf H\) induce the same orientation and “\(-\)” if they induce the opposite orientation.
This can be checked using and the change of variables theorem for multiple integrals.

The special case \(f=1\) reduces to the area of a surface \(S\) is independent of the parametrization

This is nice to know, because if our definition of “area” depended on the parametrization, then it would probably not be a good definition.

For integrals of vector fields, as already discussed, we have to pay attention to the orientation, which is the choice of the direction in which the unit normal \(\mathbf n\) points, as we have already seen in examples. That is, the integral of a vector field \(\mathbf F\) over a surface \(S\) depends on the orientation of \(S\) but is otherwise independent of the parametrization. In fact, changing the orientation of a surface (which amounts to multiplying the unit normal \(\mathbf n\) by \(-1\), changes the sign of the surface integral of a vector field.

Because of the above facts, if we want to integrate a real-valued function over a surface, or (in particular) to find the area of a surface, we do not need to know how the surface is oriented. But to compute the surface integral of a vector field, we need to specify the orientation of the surface. So if someone asks you to compute \(\iint_S \mathbf F\cdot \mathbf n\, dA\), it is their job to specify not only the vector field \(\mathbf F\) and the surface \(S\), but also the orientation of the unit normal \(\mathbf n\). One that is done, we can speak unambiguously about \(\iint_S \mathbf F\cdot \mathbf n \, dA\), without having to say precisely which parametrization we have in mind.

Volume and Area

Surface area is a subtler concept than volume. If you are asked to measure the volume of a physical object, it is (in principle) easy to do: dunk the object in a bucket of water, and measure how much water it displaces.

Measuring surface area is much more difficult. How might you do it?

One possibility is to paint the object on one side with a thin coat of paint. Then the area of the surface should be proportional to the volume of paint used, which we can measure. For example, if we are able to apply paint with a uniform thickness of exactly \(h\) cm, (a realistic number might be \(h = 0.1\, cm\)) and if we use \(V\) \(cm^3\) of paint, then the area should be about \[ \frac{ V \, cm^3}{h \, cm} = \frac V h \, cm^2. \]

Example 3.

What is the surface area of \(S\), the unit sphere in \(\R^3\)? That is \[ S =\left\{\mathbf x \in \R^3 : |\mathbf x| = 1\right\}. \]

If we coat \(S\) with a layer of paint of thickness \(h\), the painted sphere occupies the region \[ S_h = \left\{ \mathbf x\in \R^3 : 1 \le |\mathbf x| \le 1+h\right\}. \] The volume is easily computed and equals \[ \text{Vol}(S_h) = \frac 43 \pi [(1+h)^3 - 1^3] = \frac 43\pi(3h +3h^2+h^3). \] So the area of the sphere should be approximately \[ \text{area}(S) \approx \frac {\frac 43\pi(3h^2 +3h+h^3)}h = 4\pi + 4\pi h + \frac 43\pi h^2. \] If \(h\) is small, this is very close to the number you have probably computed in first-year calculus or elsewhere, \(4\pi\).

If we had (somehow) painted \(S\) on the inside, we would have gotten a different approximation, area\((S)\approx 4\pi - 4\pi h + \frac 43 \pi h^2\). But these two different approximations would yield the same number for the area if we take the limit \(h\to 0\).

These considerations suggest a way of computing area of a surface \(S\) that is a mathematical idealization of the procedure of measuring the volume of a thin coat of paint.

  1. Choose an orientation \(\mathbf n\) for the surface. (It will not matter which you choose.)

  2. For \(h>0\), define \[\begin{equation}\label{Sh.def} S_h = \left\{ \mathbf x + s \mathbf n : \mathbf x \in S, 0 \le s \le h \right\}. \end{equation}\] This is the set of all points on the side of the surface in the \(\mathbf n\) direction, that are a distance as most \(h\) from the surface, or if you prefer, the very thin region occupied by a coat of paint of thickness \(h\).

  3. Then we expect that \[ \text{area}(S) = \lim_{h\to 0}\frac{\text{Vol}(S_h)}{h}. \] As the example of the sphere suggests, this procedure should yield a result that does not depend on the orientation of \(\mathbf n\).

In fact this is true:

Let \(S\) be a surface parametrized by \(\mathbf G:R\to \R^3\), where \(R\) is compact and measurable, and \(\mathbf G\) is \(C^2\) on an open set containing \(R\). Also suppose that \(\left\{ \partial_u \mathbf G, \partial_v\mathbf G\right\}\) are linearly independent everywhere in \(W\), and hence induce an orientation \(\mathbf n\) (as discussed above). Then \[ \text{area}(S) = \lim_{h\to 0}\frac{\text{Vol}(S_h)}{h} \] for \(S_h\) as defined in (using the orientation induced by \(\mathbf G\)).

We will never use the theorem; its only role for us is as a motivation/justification for the definition of area. You may find it to be a more convincing justification than the hand-waving arguments about adding up areas on infinitesimal parallelograms alluded to earlier.

The proof is a little too involved for this class, but we can at least describe the outline.

Sketch of the proof.

For \(s>0\) define \[ \widetilde{\mathbf G} (u,v,s) = \mathbf G (u,v) + s \mathbf n (\mathbf G (u,v)), \] where \[ \mathbf n(\mathbf G (u,v)) = \frac{\partial_u\mathbf G \times \partial_v \mathbf G}{\left|\partial_u\mathbf G \times \partial_v \mathbf G \right|}(u,v). \] Also, let \(R_h = \left\{(u,v,s) : (u,v)\in R, 0\le s \le h\right\}\). Then

  1. Check that \(\widetilde{\mathbf G}\) is a transformation of \(R_h\) onto \(S_h\) (that is, one-to-one and onto, with \(C^1\) inverse.)

  2. Then in principle we can compute the volume of \(S_h\) using the Change of Variables Theorem, leading to \[ \text{Vol}(S_h) = \iint_{R_h} |\det \widetilde{\mathbf G} (u,v,s)|\, du\,dv\,ds. \]

  3. Next, one can prove that \[ \lim_{h\to 0}|\det \widetilde{\mathbf G} (u,v,s)| = \left|\partial_u\mathbf G \times \partial_v\mathbf G \right|(u,v). \]

  4. Once this is known, then completing the proof of the theorem is relatively straightforward.

The details we have omitted are not at all easy, but they are in principle within reach for MAT237 students who have gotten this far. That is, you know all the mathematical techniques, definitions and theorems needed for the proofs; it’s just that the arguments are long and complicated, and we have to cover other content, so we choose not to spend the time on this.

Problems

Basic

In all these questions, the most important part is to set up the integrals correctly, since this is the new part. We already know (or don’t know, as the case may be) how to evaluate them.

  1. Compute the area of the following surfaces.

  2. Evaluate the following integrals.

  3. Compute the following integrals.

Advanced

  1. Prove \(\eqref{cvsurf}\).

  2. Use \(\eqref{cvsurf}\) to prove Theorem 1.

\(\Leftarrow\)  \(\Uparrow\)  \(\Rightarrow\)

Creative Commons Licence
This work is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 2.0 Canada License.