3.2 Geometric Content of the Implicit Function Theorem

\(\newcommand{\R}{\mathbb R }\)

3.2 Geometric Content of the Implicit Function Theorem

  1. Curves in \(\R^2\)
  2. Surfaces in \(\R^3\)
  3. Curves in \(\R^3\)
  4. The General Case
  5. Problems

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

Curves in \(\R^2\)

There are 3 natural ways to represent a curve \(S\subseteq \R^2\):

  1. As a graph: \[\begin{equation} \left. \begin{array}{c} S = \{ (x,y) \in \R^2: \ y = f(x) \text{ for }x\in I\}\\ \text{OR} \\ S = \{ (x,y) \in \R^2: \ x = f(y) \text{ for }y\in I\} \end{array}\right\} \label{graph} \end{equation}\] for some \(f:I\to \R\), where \(I\subseteq \R\) is an interval.

  2. As a level set, that is, a set of the form \[\begin{equation} S = \{ (x,y)\in U : F(x,y) = c \} \label{locus}\end{equation}\] for some open \(U\subseteq \R^2\), some \(F:U\to \R\), and some \(c\in \R\).
    This is also called the “zero locus” of \(F\) when \(c=0\).

  3. Parametrically, that is, in the form \[\begin{equation} S = \{ \mathbf f(t) : \ t\in I \} \label{parametrically}\end{equation}\] for some interval \(I\subseteq \R\) and some \(\mathbf f:I\to \R^2\).


If \(S\) is a curve in \(\R^2\) that can be represented as the graph of a \(C^1\) function \(f:I\to \R\) for some open interval \(I\subseteq \R\), then

These are true because, if \(f:I\to \R\) is a \(C^1\) function, then for example \[\begin{align} \{ (x,y) \in \R^2: \ x\in I, y = f(x) \} &= \{(x,y)\in \R^2 : F(x,y)= 0\} \nonumber \\ &=\{ \mathbf g(t): t\in I \} \nonumber \end{align}\] for \(F(x,y) = y-f(x)\) and \(\mathbf g(t) = (t, f(t))\).
Similar considerations apply if \(S= \{ (x,y) \in \R^2: \ y\in I, x = f(y)\}\).

On the other hand, if a curve is represented as a zero locus, or parametrically, then it cannot always be represented as a graph.

Example 1.

For example, the Figure Eight curve below shows the set \[ S = \{(x,y)\in \R^2 : y^2 = x^2(4-x^2) \}. \] The curve can also be represented parametrically as \[ S = \{ \mathbf f(t) : t\in \R\} ,\qquad\text{ for }\mathbf f(t) = \binom{2\cos t}{2\sin 2t}. \] But cannot be represented as the graph of a \(C^1\) function. Even worse, for every \(r>0\), the set \(S\cap B(\mathbf 0; r)\), that is, the portion of the curve that is within distance \(r\) from the origin, cannot be represented as a \(C^1\) graph.

drawing
Figure Eight

Example 2.

Similarly, the curve below is (a part of) \[\begin{align} S &= \{(x,y) : x^5-y^2 = 0\} \nonumber \\ &= \{ \mathbf f(t) : t\in \R\} \quad \text{ for }\ \ \mathbf f(t) = \binom{t^2}{t^5}. \nonumber \end{align}\]

drawing

This can be written as \(\{ (x,y) : x = |y|^{2/5}\}\), but the function \(f(y) = |y|^{2/5}\) is not differentiable at \(y=0\). So here too, for every \(r>0\), the set \(S\cap B(\mathbf 0; r)\) cannot be represented as a \(C^1\) graph.

\(C^1\) graphs, and singularity vs regularity

Below we will say that a curve is a \(C^1\) graph as shorthand for can be represented as the graph of a \(C^1\) function.

Suppose that \(S\) is a curve in \(\R^2\) and \(\mathbf a\in S\). If, for every \(r>0\), \(S \cap B(\mathbf a;r)\) is not a \(C^1\) graph (of a function \((a,b)\to \R\)), then \(S\) is singular at \(\mathbf a\). If there exists some \(r>0\) such that \(S \cap B(\mathbf a;r)\) is a \(C^1\) graph (of a function \((a,b)\to \R\)), then \(S\) is regular at \(\mathbf a\).

When is a curve regular?

We first consider a curve defined as a zero locus of a function \(F\).

Suppose that \(F:\R^2\to \R\) is \(C^1\), and let \[ S = \{ \mathbf x\in \R^2 : F(\mathbf x) = 0\} \] If \(\mathbf a\in S\) and \(\nabla F(\mathbf a)\ne \bf0\), then there exists some \(r>0\) such that \(B(\mathbf a;r)\cap S\) is a \(C^1\) graph.
If \(\nabla F(\mathbf a)\ne 0\), then at least one partial derivative must be nonzero. If \(\partial_y F(\mathbf a) \ne 0\), then the Implicit Function Theorem implies that there exists \(r>0\) such that \(B(\mathbf a;r)\cap S\) can be written in the form \(\{ (x,y): y=f(x)\}\) for some \(C^1\) function \(f\) whose domain is a subset of \(\R\). If \(\partial_x F(\mathbf a)\ne 0\), then the Implicit Function Theorem implies that there exists \(r>0\) such that \(B(\mathbf a;r)\cap S\) can be written in the form \(\{ (x,y): x=f(y)\}\) for some \(C^1\) function \(f\) whose domain is a subset of \(\R\). In either case, \(B(\mathbf a;r)\cap S\) is a \(C^1\) graph.

As a result of Theorem 1, if \(\nabla F\) does not vanish at any point of \(S\), then \(S\) is not singular anywhere. By contrast, if \(\mathbf a\) is a point where \(S\) is singular, then \(\nabla F(\mathbf a) = 0\), that is, \(\mathbf a\) is a critical point.

For example, you can easily check that for both Examples 1 and 2 above, the origin is a critical point of the function that appears in the definition of the curve \(S\).

Example 3. (Question 3 on MAT237 Test 3, February 2018)

The picture below shows the set \[ S = \{(x,y)\in \R^2 : y^2 - e^{2x} + 4e^x = \alpha \}, \qquad\text{ for a particular }\alpha\in \R. \]

Determine the value of \(\alpha\).

drawing
Solution. Let \(\mathbf a\) denote the point where the two branches of \(S\) cross. According to Theorem 1, \(\mathbf a\) must be a critical point of \(F(x,y) = y^2 - e^{2x} + 4e^x\). If you check, you will find that \(F\) has exactly one critical point. So this point must be \(\mathbf a\). You can then determine \(\alpha\) by evaluating \(F\) at \(\mathbf a\).

We next consider a curve defined parametrically

Suppose that \(\mathbf f:(a,b)\to \R^2\) is \(C^1\), and let \[ S = \{ \mathbf f(t): t\in (a,b)\}. \] If \(\mathbf f'(c) \ne\bf 0\) for some \(c\in (a,b)\) then there exists some \(r>0\) such that \(\{ \mathbf f(t) : |t-c|<r\}\) is a \(C^1\) graph.

The assumption that \(\mathbf f'(c)\ne \bf0\) implies that at least one component \(f_j'(c)\ne 0\), for \(j=1,2\). Let us suppose for concreteness that \(f_1'(c)\ne 0\); the other case is essentially identical.

Since \(f_1'(c)\ne 0\) and \(f_1'\) is continuous, there is some \(\varepsilon>0\) such that \(f_1'(t)\ne 0\) for \(t\in (c-\varepsilon, c+\varepsilon)\). Thus \(f_1\) is either strictly increasing or strictly decreasing in this interval, hence invertible.

Let \[ \mathbf F(t,x,y) = \left(\begin{array}{c} f_1(t) - x \\ f_2(t) - y \end{array}\right). \] Since \(f_1'(c)\ne 0\), the matrix \[ \left(\begin{array}{cc} \partial_t F_1& \partial_y F_1 \\ \partial_t F_2& \partial_y F_2 \end{array}\right) \ = \ \left(\begin{array}{cc} f_1'(t) &0 \\ f_2'(t) &-1 \end{array}\right) \] is invertible at \((t,x,y) = (c, f_1(c), f_2(c))\), so the Implicit Function Theorem implies that there exists some \(r_0, r_1>0\) and a function \(\mathbf g: (f_1(c)-r_0,f_1(c)+r_0)\to \R^2\) such that \[\begin{multline}\label{iiift} \text{ For } (t,x,y)\text{ such that }|x - f_1(c)|<r_0\text{ and } |(t,y) - (c, f_2(c))| < r_1 , \\ \mathbf F(t,x,y) = {\bf 0} \text{ if and only if }(t,y) = \mathbf g(x) \end{multline}\]

Since \(\mathbf f\) is continuous, there exists some \(r>0\) such that if \(|t-c|<r\), then \((t,x,y) = (t, f_1(t), f_2(t))\) satisfies \[ |x - f_1(c)|<r_0\text{ and } |(t,y) - (c, f_2(c))| < r_1 . \] Since \(\mathbf F(t, f_1(t), f_2(t)) = \bf0\), it follows from \(\eqref{iiift}\) that the set \[ \{ (f_1(t), f_2(t)) : |t-c|<r \} = \{ (x, g_2(x)) : x\in I\} \] where \(I\) is some subset of \((f_1(c)- r_0, f_1(c)+r_0)\).

What this says is less intuitive than Theorem 1. If \(\mathbf f'(c)\ne \bf 0\), it does not say that \(S\) is regular near \(\mathbf f(c)\). It says instead only that the parametrization is regular near \(t=c\).

For example, consider again Example 1 above, where the curve can be written as \(\{\mathbf f(t) : t\in \R\}\) for \(\mathbf f(t) = \binom{2\cos t}{2\sin 2t}\). It is easy to see that

This is consistent with the theorem. To see this, consider the picture below, which shows the sets \[ \{ \mathbf f(t) : |t-c|<\pi/4\} \quad\text{ for }\begin{cases} c = \pi/2 &\text{ in blue}\\ c = 3\pi/2 &\text{ in purple}. \end{cases} \]

drawing

Both the blue and the purple curves are \(C^1\) graphs. This illustrates how the parametrization can be regular even if the curve is not.

By contrast, in Example 2 above, \(\mathbf f'(t)= 0\) only for \(t=0\), and \(\mathbf f(0)\) is exactly the singular point of the curve, at the origin.

Example 4.

The picture below shows part of the set \[ S = \{ \mathbf f(t) : t \in \R \}, \quad\text { for } \mathbf f(t) = \binom{t^3-4t^2+5t}{t^5+6t^4+4t^3-16t^2-9t}. \] If \(\mathbf a\) denotes the point where the curve forms a cusp, then for which value of \(t\) is it true that \(\mathbf a = \mathbf f(t)\)?

drawing
Solution. By the contrapositive of Theorem 2, the cusp is a point \(\mathbf f(c)\) where \(\mathbf f'(c)= \bf 0\). The first component of \(\mathbf f'(t)\) factors as \((t-1)(3t-5)\), hence we only need to check the second component at \(t=1\) and \(t=\frac 53\). It is \(0\) at \(t=1\), and non-zero at \(t=\frac 53\), since it is strictly increasing on \((1, \infty)\). So the cusp is at \(f(1)=\binom{2}{-14}\).

Surfaces in \(\R^3\)

The discussion here closely follows our discussion of curves in \(\R^2\), starting with the 3 common ways to represent a surface \(S\subseteq \R^3\):
  1. As a graph: \[ \left. \begin{array}{c} S = \{ (x,y,z) \in \R^3: \ z = f(x,y) \text{ for }(x,y)\in T\}\\ \text{OR} \\ S = \{ (x,y,z) \in \R^3: \ y = f(x,z) \text{ for }(x,z)\in T\}\\ \text{OR} \\ S = \{ (x,y,z) \in \R^3: \ x = f(y,z) \text{ for }(y,z)\in T\}\\ \end{array}\right\} \] for some \(f:T\to \R\), where \(T\) is an open subset of \(\R^2\).

  2. As a level set, that is, a set of the form \[ S \ = \ \{ (x,y,z)\in U : F(x,y,z) = c \} \] for some open \(U\subseteq \R^3\), some \(F:U\to \R\), and some \(c\in \R\).
    This is also called the “zero locus” of \(F\) when \(c=0\).

  3. Parametrically, that is, in the form \[ S \ = \ \{ \mathbf f(\mathbf t) : \ \mathbf t\in T \} \] for some open \(T\subseteq \R^2\) and some \(\mathbf f:T\to \R^3\).

Similar to curves, if \(S\) is a surface in \(\R^3\) that can be represented as the graph of a \(C^1\) function \(f:T\to \R\) for some open interval \(T\subseteq \R^2\), then

On the other hand, if a curve is represented as a zero locus, or parametrically, then it cannot always be represented as a graph.

Example 5.

Below is a picture of the surface \[ S = \{ (x,y,z)\in \R^3 : y^2 + z^2 = x^2(4-x^2)\} \] It can be represented parametrically as \[ \left\{ \left( \begin{array}{c}2\cos t \\ 2\cos s \, \sin 2t \\ 2\sin s\, \sin 2t \end{array} \right) : (s,t)\in \R^2 \right\} \]

drawing

We can show that \(S\cap B(\mathbf 0;r)\) is not a \(C^1\) graph for any \(r>0\).

The same is true for the surface pictured below, which is \[ S = \{ (x,y,z) : x^5 - y^4-z^2=0 \} \] It can also be written parametrically as (for example) \[ \left\{ \left( f_1(s,t) , s^5 , t^5 \right) : (s,t)\in \R^2 \right\} \quad \text{ for }f_1(s,t) = \begin{cases} (s^{20}+t^{10})^{1/5}&\text{ if }(s,t)\ne (0,0)\\ 0&\text{ if }(s,t) = (0,0). \end{cases} \] It is not obvious that \(f_1\) is \(C^1\) - check that it is!

drawing

\(C^1\) graphs, and singularity vs regularity

Suppose that \(S\) is a surface in \(\R^3\) and \(\mathbf a\in S\). If, for every \(r>0\), \(S \cap B(\mathbf a;r)\) is not a \(C^1\) graph (of a function \(U\to \R\) for some open set \(U\subseteq \R^2\)), then \(S\) is singular at \(\mathbf a\). If there exists some \(r>0\) such that \(S \cap B(\mathbf a;r)\) is a \(C^1\) graph (of a function \(U\to \R\) for some open set \(U\subseteq \R^2\)), then \(S\) is regular at \(\mathbf a\).

When is a surface regular?

We first consider a surface defined as a zero locus of a function \(F\).

Suppose that \(U\) is an open subset of \(\R^3\) and \(F:U\to \R\) is \(C^1\), and let \[ S = \{ \mathbf x\in U : F(\mathbf x) = 0\} \] If \(\mathbf a\in S\) and \(\nabla F(\mathbf a)\ne \bf0\), then there exists some \(r>0\) such that \(B(\mathbf a;r)\cap S\) is a \(C^1\) graph.
As with Theorem 1, this is essentially a special case of the Implicit Function Theorem.

As a result of Theorem 3, if \(\mathbf a\) is a point where \(S\) is singular, then \(\nabla F(\mathbf a) = 0\), that is, \(\mathbf a\) is a critical point. And if there are no critical points of \(F\) on \(S\), then \(S\) is regular everywhere.

For example, you can easily check that for both Examples 5 above, the origin is a critical point of the function that appears in the definition of the surface \(S\).

Example 6.


The picture below shows the set \[ S = \{(x,y,z)\in \R^3 : x^2-4x+y^4+4y^3+6y^2+4y-z^4 = \alpha \} \] for a particular \(\alpha\in \R\). Determine the value of \(\alpha\).

drawing

The idea of the solution is very similar to that of Example 3.

We next consider a parameterized surface.

Suppose that \(T\) is an open subset of \(\R^2\) and that \(\mathbf f:T\to \R^3\) is \(C^1\), and let \[ S = \{ \mathbf f(\mathbf t): \mathbf t\in T \}. \] If \(D\mathbf f(\mathbf c)\) has rank \(2\) at some \(\mathbf c\in T\) (that is, if \(\{ \partial_1 \mathbf f , \partial_2\mathbf f\}\) are linearly independent) then there exists some \(r>0\) such that \(\{ \mathbf f(\mathbf t) : \mathbf t \in B(\mathbf c;r) \}\) is a \(C^1\) graph.

The proof is similar in spirit to that of Theorem 2. Filling in the details is an exercise (probably a little on the hard side).

Like Theorem 2, the conclusion of the theorem is not that \(S\) is regular near \(\mathbf f(\mathbf c)\), but only that the parametrization is regular near \(\mathbf c\).

Curves in \(\R^3\)

Just like with curves in \(\R^2\), there are 3 common ways to represent a curve \(S\subseteq \R^3\):

  1. As a graph: \[ \left. \begin{array}{c} S = \{ (x,y,z) \in \R^3: \ (x,y) = \mathbf f(z) \text{ for }z\in I\}\\ \text{OR} \\ S = \{ (x,y,z) \in \R^3: \ (x,z) = \mathbf f(y) \text{ for }y\in I\}\\ \text{OR} \\ S = \{ (x,y,z) \in \R^3: \ (y,z) = \mathbf f(x) \text{ for }x\in I\}\\ \end{array}\right\} \] for some \(\mathbf f:I\to \R^2\), where \(I\) is some interval.

  2. As a level set, that is, a set of the form \[ S \ = \ \{ (x,y,z)\in U : \mathbf F(x,y,z) = \mathbf c \} \] for some open \(U\subseteq \R^3\), some \(\mathbf F:U\to \R^2\), and some \(\mathbf c\in \R^2\).
    This is also called the “zero locus” of \(\mathbf F\) when \(\mathbf c=\bf0\).

  3. Parametrically, that is, in the form \[ S \ = \ \{ \mathbf f(t) : t\in(a,b) \} \] for some interval \((a,b)\subseteq \R\) and some \(\mathbf f:(a,b)\to \R^3\).

If \(S\) is a curve in \(\R^3\) that can be represented as the graph of a \(C^1\) function \(\mathbf f:I\to \R^2\) for some open interval \(I\subseteq \R\), then

These assertions can be proved by adapting the arguments from our discussion of curves in \(\R^2\).

On the other hand, if a curve is represented as a zero locus, or parametrically, then it cannot always be represented as a graph.

Example 7.

Below is a picture of the curve \[ S = \{ (x,y,z) : \mathbf F(x,y,z) = {\bf 0}\}\quad\text{ for } \mathbf F(x,y,z) = \binom{x^2-z}{y^3-z} \] It can be represented parametrically as \[ \left\{ \left( \begin{array}{c}t^3 \\t^2 \\ t^6 \end{array} \right) : t\in \R \right\} \]

drawing

It is intuitively clear that \(S\cap B(\mathbf 0;r)\) is not a \(C^1\) graph for any \(r>0\).

\(C^1\) graphs, and singularity vs regularity

We use identical definitions as curves in \(\R^2\).

Suppose that \(S\) is a curve in \(\R^3\) and \(\mathbf a\in S\). If, for every \(r>0\), \(S \cap B(\mathbf a;r)\) is not a \(C^1\) graph (of a function \((a,b)\to \R^2\)), then \(S\) is singular at \(\mathbf a\). If there exists some \(r>0\) such that \(S \cap B(\mathbf a;r)\) is a \(C^1\) graph (of a function \((a,b)\to \R^2\)), then \(S\) is regular at \(\mathbf a\).

When is a curve in \(\R^3\) regular?

We first consider a curve defined as a zero locus of a function \(\mathbf F\).

Suppose that \(U\) is an open subset of \(\R^3\) and \(\mathbf F:U\to \R^2\) is \(C^1\) with components denoted \((F_1,F_2)\), and let \[ S = \{ \mathbf x\in U : \mathbf F(\mathbf x) = {\bf 0} \} \] If \(\mathbf a\in S\) and rank \((D \mathbf F(\mathbf a))= 2\) (that is, if \(\{ DF_1(\mathbf a), DF_2(\mathbf a)\}\) are linearly independent) then there exists some \(r>0\) such that \(B(\mathbf a;r)\cap S\) is a \(C^1\) graph.
As with Theorems 1 and 3, this is essentially a special case of the Implicit Function Theorem. Here, the point is that if rank\((D F(\mathbf a))= 2\), then at least one of \[ \left( \begin{array}{cc} \partial_x F_1 & \partial_y F_1 \\ \partial_x F_2 & \partial_y F_2 \end{array}\right), \quad \left( \begin{array}{cc} \partial_x F_1 & \partial_z F_1 \\ \partial_x F_2 & \partial_z F_2 \end{array}\right), \quad \left( \begin{array}{cc} \partial_y F_1 & \partial_z F_1 \\ \partial_y F_2 & \partial_z F_2 \end{array}\right), \] must be invertible, and hence the Implicit Function Theorem implies that at least one pair of variables can be written as a function of the remaining variable near \(\mathbf a\). When written out carefully, this implies the conclusion of the theorem.

As a result of Theorem 5, if \(\mathbf a\) is a point where \(S\) is singular, then \(D \mathbf F(\mathbf a)\) has rank less than \(2\).

Example 8

Show that \(D\mathbf F({\bf 0})\) has rank \(1\) for Example 7.

We finally consider a parametrized curve.

Suppose that \(I\) is an open interval and that \(\mathbf f:I\to \R^3\) is \(C^1\), and let \[ S = \{ \mathbf f(t): t\in I \}. \] If \(\mathbf f'(c)\ne \bf 0\) at some \(c\in I\) then there exists some \(r>0\) such that \(\{ \mathbf f(t) : |t-c|<r \}\) is a \(C^1\) graph.

The proof is similar to those of Theorems 2 and 4. Filling in the details is an exercise (probably a little on the hard side).

Like Theorems 2 and 4, the conclusion of the theorem is not that \(S\) is regular near \(\mathbf f(c)\), but only that the parametrization is regular near \(\mathbf c\).

Note that in Example 7, \(\mathbf f'(0) = \bf 0\), consistent with Theorem 6.

The General Case

All the above examples fall into the category of a \(m\)-dimensional “object” in \(\R^M\), where \(1\le m < M\).

For completely general \(m,M\), we can represent such an object as a level set, parametrically, or (sometimes) as a \(C^1\) graph.

Theorems 1, 3, and 5 are special cases of a general theorem. Similarly, Theorems 2, 4, and 6 are special cases of a slightly different general theorem. If you are interested, it is possible to extrapolate these general theorems from the cases considered above.

Problems

Basic

  1. You may want to use this result in later questions. Let \(F_1\) and \(F_2\) be \(C^1\) functions defined on some open set \(U\subseteq \R^2\). Further define \[ F_3 = F_1 F_2\qquad F_4 = F_1^2 + F_2^2, \] and for \(j=1,2,3,4\), let \(S_j = \{(x,y)\in \R^2 : F_j(x,y)= 0\}\).
  1. For each of the following functions \(F(x,y)\), define \(S = \{ (x,y)\in \R^2 : F(x,y)= 0\}\). Then
  1. Let \(f(x,y) = 2x^4 - 8xy +4y^2 +2.\)

To do this by hand, you can think of the equation \(f(x,y)= c\) as a quadratic equation for \(y\), by writing it as \(4 y^2 - (8x) y + (2x^4+2-c) = 0\). This can be solved for \(y\), by using the quadratic formula. This may be time-consuming to plot by hand for several different choices of \(c\), but if you rely entirely on a computer, you will probably get some details wrong. Perhaps it is best to rely on a combination of computer and analysis.

  1. Let \(F(x,y) = 2(x^2 - 1)(y^2-1)\)
  1. Let \(F(x,y) = \dfrac 14(5x^2 + y^2-4)(x^2+5y^2-4)\). Below is a contour plot showing the level sets \[ \{ (x,y) : F(x,y) = c\} \qquad\text{ for }c = -5, -4, -3, \cdots , 10. \] Unfortunately, the level sets are not labelled.
drawing
  1. Let \(F(x,y) =(2y-x^2)(2x-y^2).\) Below is a contour plot showing the level sets \(\{ (x,y) : F(x,y) = c\} \qquad\text{ for }c = -5, -4.5, -4, \cdots ,9.5 , 10.\) Unfortunately, the level sets are not labelled.
    drawing
  1. Below is a contour plot showing the level sets \[ \{ (x,y) : F(x,y) = c\} \qquad\text{ for }c = -45, -44,\ldots, 45 \] for a \(C^1\) function \(F\). Unfortunately, the level sets are not labelled and the formula for \(F\) has been lost.
drawing

On the plot, approximately determine the location of as many critical points as you can. The plot is not completely accurate, and in some places you might have to guess what it would look like if it were.

  1. Given \(\mathbf f:(a,b) \to \R^2\), we say that \(\mathbf f\) is a regular parametrization of \(S = \{ \mathbf f(t) : a < t < b \}\) if \(\mathbf f'(t) \ne {\bf 0}\) for all \(t\in (a,b)\). For each \(\mathbf f:\R\to \R^2\) below,

  1. A variant of the question from Example 6: The picture below shows the set \[ S = \{(x,y,z)\in \R^3 : x^2-2x - y^2+2yz-z^2-z^3 = \alpha \} \] for a particular \(\alpha\in \R\). Determine the value of \(\alpha\).
drawing
  1. Suppose that \[F(x,y,z) = \left((x^2+y^2) -\alpha\right)^2 - z^2\] for some \(\alpha\in \R\), and let \[ S = \{(x,y,z)\in \R^3 : F(x,y,z) = 0\}. \] Find all points in \(S\) where \(\nabla F = \bf 0\).
drawing
  1. For each of the functions \(F:\R^3\to \R\) below, let \(S = \{(x,y,z)\in \R^3 : F(x,y,z) = \alpha \}\), for some \(\alpha\in \R\). Determine for which values of \(\alpha\) it is true that \(S\) is regular.

Advanced

  1. Let \(F(x,y) = \frac{y^3-x^8y}{x^6+y^2}.\) In the exercises in Section 2.1, you have shown that \(F\) is differentiable everywhere in \(\R^2\), and that \(\nabla F(0,0) = (0,1)\). Thus \(\partial_y F(0,0)\ne 0\).
  1. Write out the proof of Theorem 3, which is a modification of the proof of Theorem 1.

  2. Write out the proof of Theorem 4, which is a modification of the proof of Theorem 2. (It may help to consult the proof of Theorem 5.)

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

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