This chapter is a continuation of the idea of parallel transport in Section 3.4 . I debated whether to move that section to this chapter, but ultimately decided linking covariant derivatives and parallel transport was necessary as motivation. But perhaps it would be worthwhile to read that section again now.
In this chapter we develop the theory of geodesics, which generalise the notion of a ‘straight line’. To put the question provocatively: what does it mean to have a straight line in a curved space? Consider this question for if somebody asked you to fly an aeroplane in a straight line between two cities. A reasonable definition would be a flight path that did not require steering the plane’s control stick. This is the same as the idea of ‘walking without turning’ that we used previously in our thought experiment. Such paths are called geodesics. On the other hand, this is Riemannian geometry. In euclidean space, the shortest path between two points is a straight line. Perhaps this should be taken as the defining feature of straight lines. It turns out that this length-minimising property is also true of geodesics, so that the two possible definitions coincide.
First we will formalise the definition of geodesic. Although we mostly restrict ourselves to consideration of the Levi-Civita connection, we do examine in the the mutual dependence of geodesics and connections. Using the length of a curve we define a distance function on a Riemannian manifold as the infimum of the length of all smooth curves. We will show that geodesics are critical points of the length functional, using a proof that is reminiscent of the proof that minimal surfaces are critical for area from Section 1.6. Showing that they are locally length minimising, surprisingly, is significantly more difficult. This leads us to construct special coordinates, so-called normal coordinates, in which the geodesic structure of the Riemannian manifold is a little clearer.
Adding to the discussion above, we understand that ‘walking without turning’ along a curve means that the tangent of the curve is parallel. The technicality is that is not a vector field on the manifold, but this is not a problem because it is a vector field along the curve and we know that that is sufficient to define its covariant derivative.
In a chart we may write and use the Christoffel coefficients to describe the covariant derivative:
This is generally known as the geodesic equation. It is of course the parallel transport equation (3.37) with . But since the vector field in this case is linked the the curve, we now have a second-order nonlinear system of ODEs. However we can use a trick to write this as a system of first order ODEs:
From this we conclude the local existence and uniqueness of geodesics in a neighbourhood of every point and in every direction. Specifically, given a point and a direction existence implies that there is a curve with and . We will drop the when it is unambiguous from context. We assume that is maximal in the sense that any other solution is a restriction of to some smaller domain . This assumption may always be achieved, because uniqueness makes it possible to glue together any two solutions into a ‘longer’ one. Related to this is the observation that if and , then
Example 4.2 (Helicoid). Consider the helicoid with the inherited metric from and the tangent connection. We consider again in this example two (sets of) curves: radial lines and helices. First there are radial curves for some . The tangent connection is the projection of the derivational derivative in to the tangent plane of the helicoid. The directional derivative in the direction is the derivative with respect to . Hence , because is linear in . Hence radial lines are geodesics of the helicoid.
On the other hand we have the helices for some . As above but this time it is not zero. Now we reuse our previous calculations. As remarked upon in Example 1.23, already lies in the tangent plane. Hence
This shows that the helix is not a geodesic for helicoid, unless (in which case it is the central axis of the helicoid).
How do the geodesics of a manifold depend on the choice of covariant derivative? Following Lemma 3.23 we consider the connections and . In a chart we may write and since is -bilinear these functions completely determine . Moreover . Let be a geodesics of . Then the geodesic equation for reads
Thus is also a geodesic of if and only if this quantity is zero. What is obscured by index notation is the symmetry here. Using a relabelling of summation indices
This can of course be zero for some curve . If is antisymmetric, , then this quantity is zero for all curves, and in particular the two connections have the same geodesics. Conversely, we know that every vector at every point is the tangent vector to some geodesic of . If the two connections have the same geodesics, then this forces to be antisymmetric. Therefore we have proved
Theorem 4.3 (Geodesic Agreement). Two covariant derivatives have exactly the same geodesics if and only if they differ by an antisymmetric .
In particular, if we absorb the torsion of a connection , then this does not change the geodesics. This justifies our comment at the end of Chapter 3 that torsion is often unimportant.
The geodesic equation says that the curve is parallel transporting its own tangent vector. If the connection is metric-compatible then this means that the length of the tangent vector is not changing. In Riemannian geometry, we call the speed of the curve .
Theorem 4.5 (Constant Speed). Let be a Riemannian manifold and a metric-compatible connection. Then the geodesics of have constant speed.
Proof. To remove the square root from the norm, observe that a continuous function is constant if and only if its square is constant. The result now follows from the definition of metric-compatibility:
For any point there is a special geodesic called the constant geodesic. It is the geodesic associated to the zero vector. One can check easily that for all solves the geodesic equation with the initial condition and . The interpretation is that walking in a straight line with initial speed zero means standing still. As a function they are constant, hence the name. This degenerate example is none-the-less important to include.
Notice that a geodesic is a parameterised curve. To give an example, both and are geodesics in . Their images in but they are not equal as parameterised curves because they have different speeds. But our intuition tells us that a straight line is about direction, not speed.
The following lemma tells us how geodesics in the same direction with different initial speeds are related. It says that they have the same image and differ only by a constant rescaling factor. For this reason, geodesics that only differ by speed are often conflated with one another.
Lemma 4.6 (Rescaling Lemma). Let and . Whenever one expression is defined, so is the other and they are equal:
Proof. We deal with the case separately.1 On the left hand side we have for all . On the right we have the constant geodesic , which we know is equal to for all time. Hence the two sides are defined for all and equal.
Now assume . We use the first-order version of the geodesic equation, because it makes the idea a little clearer. As per definition, is the geodesic with an initial direction of . Let . It obeys
Now consider the reparameterised curve . The starting point has not changed: . But the velocity has changed:
In particular . Finally, we see for the second ODE
Thus we see that is a solution of the same ODE and IVP that solves. By uniqueness and maximality of , we conclude that is the restriction of . But we can also run this argument in the reverse direction and conclude that is a restriction of . The conclusion must be that one exists if and only if the other does, and they are equal. □
A particular case is for . The lemma says : the geodesic in the reverse direction is the same as walking backwards. If we choose a time and set and , then we have
This says that if you walk along a geodesic for a certain amount of time , turn around and walk back for the same amount of time, then you end back where you started. These properties might seem obvious, but it is important to question whether our intuition carry over to the general setting.
Exercise 4.7. Show that a reparameterisation of a non-constant geodesic is again a geodesic if and only if is a linear function.
Exercise 4.8. Show that any curve with the property that for some function can be reparameterised to a geodesic.
Finally, what can be said about geodesics in an immersed manifold? For this question to be sensible, we should use the tangent connection on from Definition 3.59. To recap, we have a Riemannian manifold with the Levi-Civita connection and is Riemannian immersed in . The tangent connection on is the projection of to the tangent space of . A curve on is a geodesic iff
Because is a second-order derivative of we interpret it as a type of acceleration. The above equation says that a curve is a geodesic of when its acceleration in is always perpendicular to .
We can relate this back to normal curvature in Section 1.5. In that situation we have and the connection is just ordinary directional derivatives . Therefore . The acceleration of a curve in has components in the and directions, Equation (1.28). If we project this onto the tangent plane then we get
Hence we see that is a geodesic of the surface if and only if is parameterised with constant speed and , the normal of the curve, is perpendicular to . Since any regular curve can be reparameterised by arc-length, the first condition is only a technical point. If is perpendicular to this means that the angle between and the surface normal is zero. In other words, the normal curvature is equal to the curvature of . Traditionally, one defines the geodesic curvature of as
This is a measure of how far an arc-length parameterised curve is from being a geodesic. Alternatively, the observation (since the normal and geodesic curvatures are projections of the curvature vector) leads us to say a curve is a geodesic if its curvature is entirely normal.
Example 4.9 (Tangent Connection). Consider . Consider a geodesic and suppose without loss of generality that is parameterised by arc-length. From the above discussion, a geodesics must have in the normal direction to the sphere, . For the sphere . Therefore . On the other hand , using the Frenet equations, and we can conclude that . This proves that the geodesics of the sphere lie in the plane containing the normal and their tangent vector. These are exactly the great circles.
So far we have seen the example of and . We have the sense that these spaces are special. There is a another important two-dimensional Riemannian manifold: the hyperbolic plane . Unlike the sphere, a theorem of Hilbert proves that the hyperbolic plane cannot be isometrically immersed into .2 Therefore we really need to use the tools of manifolds and charts to understand this space, there can be no resorting to geometric tricks in euclidean space.
This space was first discovered in connection to the ‘parallel postulate’ of Euclid. Euclid gave five postulates3, which we would call axioms,
Clearly one of these is not like the others. The fifth postulate is called the parallel postulate, because it is equivalent to Playfair’s axiom:
There is at exactly one line that can be drawn parallel to another given one through an external point.
There are several interesting geometries that come from replacing this axiom. If there are no parallel lines then we get projective geometry (one intersection) or spherical geometry (two intersections). If there are more than one parallel line then we get hyperbolic geometry.
For our purposes we will introduce the hyperbolic plane by fiat. Like the euclidean plane it can be covered by a single coordinate chart. There are several ‘models’ of the hyperbolic plane, but we will use the ‘half-plane’ model where the geodesics have the easiest formulas.
Definition 4.10. The hyperbolic plane is the following manifold: Let ,
and , so that is its atlas. Additionally, it is a Riemannian manifold with the metric
given in the chart.
The first common misconception to address is the ‘boundary’ at . As a Riemannian manifold, there is no boundary. If you lived in the hyperbolic plane, as you tried to approach at constant speed you would find that the coordinates were changing at an every decreasing rate. If we explain this using vectors in coordinates, for small, a unit-length vector has very small coefficients, and therefore moving at unit-speed makes only a small change in coordinates. Conversely, for large , a unit-length vector has large coefficients. This is analogous to looking at the world in the Mercator map, where a plane on the equator changes its longitude far less than a plane near the poles even at the same speed.
Let us compute the Christoffel coefficients for the Levi-Civita connection. Only two derivatives of the metric are non-zero
and the inverse matrix is
Using Equation (3.55)
and the other four are zero.
This means that we can try to solve the geodesic equation. It is certainly possible to solve it directly, but we will solve it in a special case and then use isometries to obtain the full solution, as this is geometrically more interesting. The special case is the following: let the the starting point of the geodesic and be its initial direction. If the geodesic is then the geodesic formula is
Notice that if is a solution to these ODEs, so too is . Moreover, and , so they have the same initial condition. Therefore they must be equal, in other words , which implies . The second ODE now simplifies to
The initial conditions force . We continue
again using the initial conditions. Thus the unique geodesic through with initial vector is the vertical line .
Already in this argument we already saw a glimpse to the method we will use to find all other geodesics. We used the reflection . This preserved the geodesic equations because this is in fact an isometry of . After we strip away the terminology of Definitions 3.4 and 3.9, what we see is that an isometry is a diffeomorphism, in particular a bijective map, such that the pushforward of vector doesn’t change the inner product. The pushforward of by reflection is
for and , and
because the components of the metric are independent of .
Another obvious isometry is horizontal translation . From this we can conclude that the geodesic through with initial direction is . These are simply all the vertical lines.
The translation is not an isometry. For one thing, it is not even a well-defined map on , which only has the points of the upper half-plane. But even between points where it is well defined, we see that the metric is changed because .
It turns out that simple scaling of the points is an isometry. This is called dilation. We compute
Unfortunately, this does tell us any new geodesics, because it takes vertical lines to vertical lines.
We need a different idea. Notice that because the matrix of the metric is a scalar of the euclidean metric (conformal), hyperbolic angles and euclidean angles are equal:
Therefore we should look for transformations of the upper half-plane that preserve euclidean angles (conformal). Perhaps you have encountered circle inversion before. It restricts to give a bijective map on the upper half-plane , because . We see that, for , ,
The calculation for is exactly the same. That leaves
Therefore is an isometry, not just conformal.
Remarkably circle inversion and the horizontal translations are all the isometries we need to get all the geodesics. By applying to the geodesic we get the geodesic
Looking on a graphing tool we see that they are semicircles centered on the -axis. This is easy to verify algebraically
In particular we see that we have semicircles of every radius. By horizontal translation then, the set of geodesics includes every vertical line as well as every semicircle centered on the -axis (from now on we will just say semicircle, leaving the centering implicit).
But in fact this accounts for every unit-speed geodesic. Choose any point and any direction with . If is , then the desired geodesic is the vertical line. For any other , by straightedge and compass (for coolness factor) one can construct a semicircle through that point with the direction as tangent. This must be the unique geodesic. For non-unit-speed geodesics, one can simply rescale , using Lemma 4.6.
These isometries are not only transitive on the set of geodesics, but also on the set of points. This is easily achieved by dilating to and then a horizontal translation to . Every point can therefore be mapped to any other by first bringing it to and then sending it on its way. Spaces with a transitive set of isometries are called homogeneous. But the hyperbolic plane is also isotropic, it looks the same in every direction. This means that there is a full set of isometric rotations. Begin with . Dilate and translate it to any point on a different vertical line as above. We know that this vertical line is transformed by circle inversion to a semicircle. A semicircle has every possible tangent direction except . You can now translate and dilate it back to . This give every direction except . But this can be achieved through two rotations.
Exercise 4.12. Argue that there is exactly one isometry that fixes and rotates its tangent space by a given angle.
Exercise 4.13. Prove that a homogeneous space is isotropic at one point if and only if it is isotropic at every point.
Now that we have determined the geodesics, we can comment on the similarities and differences to other geometries. The first observation is that between any two points there is a unique geodesic. Secondly any two geodesics can intersect at most once. Both of these properties are similar to the euclidean plane but are in contrast to spherical geometry, where antipodal points have infinitely many geodesics between them and every pair of geodesics intersects twice.
We can also consider the parallel postulate. Choose a geodesic and a point external to it. We know that there is an isometry that will transform the geodesic into the -axis, so we assume this without loss of generality. What are the geodesics through the point that do not cross the -axis? Not only is there the vertical line through , there are infinitely many semicircles. In particular, there is one semicircle through which is tangent to the -axis. This is called a limiting parallel. The vertical line through is also a limiting parallel, based on our earlier explanation that as points increase in value for constant value, the distance between them shrinks.
A triangle in the hyperbolic plane is of course a shape with three geodesic sides, none of which are parallel. Just as for euclidean geometry, we can try to classify triangles up to isometry, which is called congruence in elementary geometry. By isometry, we may bring any edge of the triangle to the -axis, so we assume that one vertex is at and the other . Let be the angle at and the angle at . We see that this information already uniquely determines a triangle, thus we have the ‘angle-side-angle’ rule of triangle congruence.
But more is true. There are restrictions on the angles for two other sides to intersect be able to intersect, namely , just as for euclidean space. But unlike the classical situation, even if this is no guarantee that the two sides meet. For any , there is some such that for the sides are limiting parallels. As ranges from down to , the third angle ranges from (limiting parallel) to (all three corners very close to , so we can approximate the circles by their tangents). Therefore the sum of angles in a hyperbolic triangle is always less that . Also note that the angle increases monotonically. This means there is a 1-to-1 correspondence between and the third angle. In other words, two hyperbolic triangle are congruent if and only if they have corresponding angles: the ‘angle-angle-angle’ rule!
We close this section with a useful calculation method. It is often profitable to use complex numbers to describe the points of . The chief advantage is the ease of writing isometries. Clearly horizontal translation and dilations are and for respectively. But inversion in the unit circle is and reflection in the -axis is . As you can see by the presence of complex conjugation, these later two isometries are orientation reversing; they are both reflections after all. Therefore it is common to combine them to an isometry . This along with translation and dilation generate all orientation preserving isometries. A general orientation preserving isometry is therefore
for with . Transformations of this form are called Möbius transformations. The set of isometries is three-dimensional so it was necessary to normalise the four constants, and using the ‘determinant’ makes the formula for the inverse transform simple.
In the above discussion of hyperbolic triangle, we were careful not to speak of the distance between the vertices of triangle because we simply have not yet defined distance on a Riemannian manifold. Let us remedy that situation. In euclidean space, we had a distance function already and in Theorem 1.6 were able to show that the length of a path as in Definition 1.4 was the integral of its speed. But here we have no prior distance function, but the Riemannian metric does give allows us to determine the speed. Therefore we ‘reverse’ these theorems:
Definition 4.15. In a Riemannian manifold with metric , the (Riemannian) length of a smooth path is defined to be
The distance between two points is defined to be the infimum of the lengths of all smooth paths connecting them.
Example 4.16 (Hyperbolic Plane). With this definition of length we can calculate the lengths of geodesics in the hyperbolic plane. Because we obtained the geodesics by isometries, which preserve the metric and therefore the length, it is sufficient to calculate the length of the vertical geodesic. It has the constant speed parameterisation
Therefore
To put this more geometrically, the length of the geodesic connecting and is .
There are (at least) two equivalent definitions of the distance function on a Riemannian manifold in the literature, with the difference being over which set of paths the infimum is taken. We have chosen ‘smooth curves’. Perhaps more common is to choose ‘piecewise smooth curves’. A curve is piecewise smooth if it is continuous and the set of non-smooth points is finite. These non-smooth points are called corners. Smooth functions are clearly nicer to work with than piecewise smooth function, if you work with the latter you are forever splitting the curve into its smooth intervals and you must provide correction terms for the corners. The advantage of the piecewise approach is clear if you try to prove the triangle inequality for the distance function. In the piecewise approach it is immediate because the concatenation of a piecewise smooth curve from to and from to is again piecewise smooth. However the concatenation of smooth paths will in general not be a smooth path, and some smoothing will be required.
Remark 4.17. We see that the definition of the distance function implicitly assumes that there is a smooth path between every point. For manifolds, connected, path-connected, and smooth-path-connected are all equivalent. When we use the distance function, we assume that the manifold is connected.
We truly need to take the infimum, even for subsets of euclidean space. Consider and two points . The distance between them is . The unique path in the full plane that achieves this is the straight line, but this is not a path in the manifold, because it would pass through the removed origin. However, by taking paths that pass arbitrarily close to the origin, we see that the distance between these points is still
This leads to an important point: even for manifolds that are Riemannian immersed in euclidean space the distance function of Definition 4.15 is not the restriction of the euclidean distance function. Consider . Now the (limiting) shortest path between and skims the boundary of the unit disk An easy estimate shows that any such path is longer than , which is strictly greater than . Therefore the distance function coming from the Riemannian metric and the restriction of the euclidean distance function to are different.
These considerations also apply to a general metric space. Given a distance function, we can define the length of a continuous curve as in Definition 1.4. Then we can define a second distance function, called the intrinsic distance function, as in Definition 4.15, as the infimum of the length of paths. The intrinsic distance between points is easily proved to be greater than or equal to their distance. Metric spaces for which the distance function is equal to the intrinsic distance are called length spaces. For euclidean space, one can calculate that the length of a line segment is equal to the distance between the end points, which proves that euclidean space is a length space.
Now that we have enough examples to make us cautious, let us begin to prove that the Riemannian distance function is reasonably behaved. First we show that locally it is equivalent to the euclidean metric of a chart. Then we show that has the three properties of a distance function. The proof that is symmetric is trivial and the proof that it has positivity will be addressed in Corollary 4.20. The triangle inequality Corollary 4.27 requires Corollary 4.26, the ‘corner rounding’ lemma.
Proof. Choose a point and a chart that contains it. Within this chart, there is a convex compact neighbourhood of (for example, a closed ball). In this chart we have the metric as a matrix of functions, but we can also consider the euclidean metric on this chart . Consider all the vectors on that are unit-length with respect to . This is a compact set . Because is positive definite and continuous, it obtains a positive maximum and minimum on this set of vectors. By decreasing or increasing we may assume that . This is not necessary, but makes the results pleasingly symmetric. For any vector , write with , then
If we apply this to paths in , we obtain
It remains to lift this inequality of lengths to an inequality of distance. As mentioned above, the euclidean distance and its intrinsic distance are equal. Hence for any smooth path from to
This proves that is a lower bound for . Because is convex, any two points can be joined with a straight-line , which achieves the euclidean distance between those points. It follows
For small balls measured with , there is a euclidean ball within it and containing it. Since balls generate the topology, this shows the two topologies are equal. □
Proof. One direction is easy: if take the constant path . This has length zero.
We prove the other direction in the contrapositive. Assume that . Choose a chart containing and a compact neighbourhood as in the previous proof. In fact, choose to be a small closed ball (with respect to the euclidean distance) of such that it does not contain . Any path between and must leave this ball at some time. Let be the first time for which . Because the integral defining length is positive, any restriction of a curve decreases its length. Hence
Since this inequality holds for every path, it holds for the distance . □
Hopefully this is illustrative of the general strategy in this topic: We try to reduce the situation to a single chart, and then in the chart we can compare our metric with the easily understood euclidean metric on the chart.
We now prepare to prove the triangle inequality for . First we give an example of blending two smooth curves.
This formula for also works for smooth curves and in euclidean space of any dimension, so long as is defined for and is defined for . In fact, it is not even necessary for the two curves to meet at .
Let us calculate how much the blending increases the length. For convenience, assume that the curves are parameterised by arc-length. The speed can be bounded from above:
which gives
Considering that are unit length parameterised, so the concatenated curve has length , the length of the blended curve is only slightly larger than the distance between the curves at . We can give a more concrete bound in terms of and that distance:
Together then
It is important to note again that the formula for blending requires that is defined for and is defined for . So then how can we know that two smooth curves meeting at a point can both be extended past the corner? Well one answer is that Definition 1.2 defines a smooth path as the restriction of a smooth curve to a closed interval. However we give a stronger result that applies to piecewise smooth curves. This demonstrates that our definition is equivalent to the other common one in the literature. The idea is to not try to extend the curves at the corner, we allow ourselves to ‘graft on’ at a nearby point.
Now we can use three blends to remove a corner from a piecewise smooth curve. In fact, performing this operation on smaller pieces of the curve ever closer to the corner results in curves that have almost the same length.
Lemma 4.24 (Corner Rounding, Euclidean). Let be a piecewise smooth curve with a single corner at . Then for every there exists a smooth curve that agrees with for . Moreover, .
Proof. The previous two examples have shown us how to construct . First, on both sides of the corner, use Example 4.23 to extend and past . Then use Example 4.21 to blend the two together at . In more detail, for in we blended with its tangent at . Similarly for in we blended with its tangent at . And for we blended the two tangents together.
It only remains to show that this process hasn’t made the approximation too long. The only part of the curve that has changed is for
Over each integral we blend two curves together. From the above length estimate (4.22), we have that this is bounded from above by
The certainly tends to zero. But what about the distance between the two tangent lines at . We can handle this with some triangle inequalities:
and since is continuous, this goes to zero also. □
Exercise 4.25. Above we used two tangent lines, but we only used the only property that they intersect at respectively. This explains why the the bound is so large. Improve this bound if you can. Alternatively modify the proof to use the line segment that connects and .
Another approach I considered for the corner rounding process was to use the curve shortening flow. Geometric flows are an important area of research. Most famously, there is the Ricci flow, which was used to solve the Poincare conjecture. In the curve shortening flow, one considers a family of arc-length parameterised curves such that
for the curvature and normal . If we think of as a time parameter, it says that the curve is moving fastest where it is most curved, towards the center of the osculating circle. From our knowledge of curves and curvature, we know that this is equal to
We see that this is a heat equation. We know that the heat equation has very good regularity properties. Given a continuous initial condition, which in this case would be a curve , the solution is smooth in and analytic in . Thus if we begin with a continuous curve, then we obtain a smoothing through this flow. I recommend Andrews et al to students who are intrigued by this.
Our ‘elementary’ corner rounding construction, though it seems as if it is particular to , can also be performed in a Riemannian manifold.
Corollary 4.26 (Corner Rounding, Riemannian). Let be a piecewise smooth curve in a Riemannian manifold with a single corner at . Then for every there exists a smooth curve that agrees with for . Moreover, .
Proof. The statement is local, in that we only need to change in a neighbourhood of . Choose a chart containing . For sufficiently small, for all . Then by applying Lemma 4.24 we construct a smooth curve ; outside of the curve is unchanged. We know that the length of the modification converges to zero in the euclidean metric of the chart, but not in the Riemannian metric . By taking even smaller if necessary, we can ensure that the modified part lies in a compact neighbourhood , so that the estimate (4.19) applies. Then
Corollary 4.27 (Triangle Inequality). On a Riemannian manifold, the distance function obeys the triangle inequality .
Proof. As we already indicated, the proof comes down the class of curves used to define the distance function. We defined it with smooth curves. Consider any sequences of smooth paths from to and from to . Assume that as their lengths approach the respective distances between the points.
Now concatenate the paths to get a piecewise smooth path with possibly a corner at . Apply the corner rounding procedure of Corollary 4.26 to obtain smooth paths from to : let be the -blending of and . Because the distance is infimum of the lengths of paths, it is lower than any limits of this set. Therefore
We close this section with a classic result: that geodesics are critical points of the length functional. The proof of this statement is reminiscent of the variational approach in Section 1.6 to show that minimal surfaces have vanishing mean curvature. In that situation, we considered graphs, so we could model a variation of the surface by adding another function. In the present case, we don’t want to make a similar simplification and instead we will work directly with smooth families of paths.
Theorem 4.28 (First Variation of Length). Suppose that is a smooth family of paths. Let and so that all paths have the same endpoints. Thus this family of paths represents a variation of the path . Without loss of generality, assume that is parameterised by arc-length. Then
Therefore is a geodesic if and only if it is a critical point of .
Proof. We compute
When we set into the above, because we assumed that was arc-length parameterised. Next we can use Lemma 3.49 to turn the covariant derivative in the into on in the direction. Additionally,
Putting these together gives
The first term vanishes because all the terms have the same endpoint, hence . □
That geodesics are critical points of length, not minima is significant.
Example 4.29. Consider the case of and take points that are not antipodes of one another. There is a unique great circle through these two points, and these points break it into two (arc-length parameterised) geodesic paths, one long and one short. The shorter geodesic is the minimum (we will prove this in the next section) but the longer geodesic is a saddle point in the space of smooth paths between these points. We can see that it is a saddle point by imagining variations. If most of the path is fixed, and we just add a variation in one small area then the geodesic will have the lowest length of this family of paths. On the other hand, consider all the planes that contain . You can do this by rotating a plane on the line through . If you take the family of paths that are the intersection of these planes and the sphere, then the long geodesic is the longest and the short geodesic is the shortest path between in this family.
Although we have given a reasonable definition of the distance on a Riemannian manifold, it is often very difficult in practice to understand this function. We have not given an example of the distance between two points yet because the prospect of searching for the infimum of length over every possible path is daunting. Even in a well-understood space such as the euclidean plane, where would you even begin? While at first glance it seems as if Theorem 4.28 simplifies the search to geodesics, this is only a complete answer if you already know that there exists a length-minimising path between the two points. Again, the example of a punctured plane shows that a length-minimising path may not exist. The example of a punctured sphere and two points either side of the puncture, such that the shorter geodesic is blocked, shows that even though the two points are connected by a geodesic (the longer geodesic), its length is not the distance.
However, what we will show in this chapter is every point has a special ball around it: every point in that ball has a unique geodesic to the center and the length of this geodesic is the distance to the center. The key gadget for this construction is the exponential map, which can be summarised as ‘follow the geodesic out from the center’. The name is due to a relationship with the exponential map in Lie group theory and actual exponentials will not be relevant to us here.
Definition 4.30. For any let the exponential map be the function defined by
where is the unique maximal geodesic with and . Because the solutions of ODEs depend smoothly on their initial conditions, is a smooth function of .
Recall, due to the Rescaling Lemma 4.6 geodesics that only differ by speed have the same image. In the exponential map we have removed this duplication by only considering . This is the only advantage of defining the exponential map over using geodesics directly. In fact the two are equivalent, since from the Rescaling Lemma we have that
The exponential map is a partially defined function; its value only exists if the corresponding geodesic exists at time . Geodesics are defined by an ODE with short-time existence, but we know nothing about their long-time existence. It is natural therefore to ask about . Because of the constant curve we know that . Likewise, short-time existence tells us that contains a neighbourhood of . The final thing we can say is that is star-shaped around : For clarity, write in the above equation: . If exists then exists for because is maximal. This shows that if then for .
The next thing we have to understand is tricky because it breaks the usual hierarchy of concepts. We need to think about the tangent map . In particular, for it is a map between and itself. Recall the definition of a manifold in Chapter 2. A chart is an open subset of and the tangent space is (with an equivalence relation between these for different charts). Likewise we can think of as an open subset of euclidean space and as . Given we think of it in as the curve . For any we have
Hence .
Because is an isomorphism, the inverse function theorem says that there is a neighbourhood such that is an diffeomorphism onto its image. We can further restrict so that the image is entirely within the chart 4. Therefore we have a transition function . Because is entirely within , it will obey the cocycle conditions with every other transition function of . Because it is defined on all of adding this chart to the atlas does not create any new points of . In the literature this construction is called normal coordinates at . We will call this chart a normal chart at .
A normal chart has two metrics on it. Because it is a chart of , of course it has the metric . But is also a subset of and is an inner-product space using the inner product . Therefore also has the metric that just uses at every point. Clearly the two metrics are equal at . The coefficients of this second metric are constants, and the second metric can be thought of as a euclidean metric. The natural question is therefore how and compare.
A normal chart is centered on , in the sense that . The most useful property of a normal chart is that rays , which are geodesics of , are geodesics of . This is because is constant, so its geodesics are straight lines in . On the other hand, if we view these rays in the chart , we have , which are by definition geodesics of . For this reason we call these rays radial geodesics, without needing to specify which metric. Similarly we define geodesic balls and geodesic spheres centered at to be the sets
By restricting we may assume that it is a geodesic ball. Of course rays and sphere are orthogonal with respect to but remarkably:
Lemma 4.31 (Gauss’ Lemma). Radial geodesics are orthogonal to geodesic spheres with respect to the metric .
Proof. Choose any smooth function with and consider the family of curves in
This family has the property that every main curve is a radial geodesic and that every transverse curve lies in a geodesic sphere. This means that is a tangent vector to the geodesic sphere . Conversely, given any tangent vector to a geodesic sphere arises in this way.
The lemma comes down to showing that and are orthogonal for all with respect to . Because it is enough to prove that is constant. We compute, for the Levi-Civita connection of ,
The facts we have used here are that a geodesic parallel transports its tangent vector (Definition 4.1), a geodesic is constant speed (Lemma 4.5), and symmetry of covariant derivatives for families of curves (Lemma 3.49). □
Example 4.32 (Hyperbolic Plane). How does this apply to the hyperbolic plane? The hyperbolic plane is covered by a single chart . Consider the set of geodesics through any point . We know that they are defined for all time, positive and negative. This tells us that is defined on the whole tangent space. We also know that for any point there is a geodesic from to . This tells us that is surjective onto . Finally, these geodesics only intersect at . This means that is diffeomorphism from to , and the normal chart at covers all of .
Let us be explicit for the point . The geodesics though this point have the form or
We should try to combine these into a single formula. In the limit we have . Therefore the problem is that is not parameterised correctly in some sense. Let . Then
and
Now we see that we have a nice parameterisation of the geodesics through , with giving the angle of the tangent with . For example gives and gives . More generally we see that
In fact, are polar coordinates for the normal chart at .
This gives us a formula for geodesic spheres. In normal coordinates of course, geodesic spheres around are just . In coordinates, they are
for . These are euclidean circles with centers at and radii .
Example 4.33. We can apply the same analysis to the sphere. Consider the set of geodesics through any point . We know that they are defined for all time, positive and negative. This tells us that is defined on the whole tangent space. We also know that for any point there is a geodesic from to . This tells us that is surjective onto . However, this time all geodesics from intersect at the antipode of . Therefore is not injective. To construct a normal chart we therefore have to restrict to a subset of .
For the south pole, to choose a definite point, we know the geodesics are the lines of longitude. Therefore the normal chart at this point has to have lines of longitude as rays. The normal chart is in fact just a rescaling of (stereographic projection) so that the rays are unit speed geodesics.
Theorem 4.34. For any consider any other point that lies in a geodesic ball of . Let be any smooth path from to , not necessarily lying in the geodesic ball. Then the radial geodesic from to is the unique length-minimiser from to .
Proof. We again use a normal chart at . Any unit speed radial geodesic has the form for some . The point lies on the geodesic sphere . Since is tangent vector of at and geodesics are constant speed
Therefore
Now suppose first that stays entirely within the geodesic ball . This means we can write , where and with . So notice that is tangent to the geodesic sphere and is tangent to a radial geodesic. Then with the help of Gauss’ Lemma 4.31
Therefore
Finally, if does not stay entirely within the geodesic ball , there there must be some first time that it crosses . Then
Thus the radial geodesic is a length-minimiser from to .
For the converse, observe that it is only possible to have equality if lies entirely within the geodesic ball and for all . This implies that and that is radial. □
Remark 4.35. An immediate consequence of this is that geodesic balls and metric balls are the same sets, and the of the geodesic ball really is its radius.
Example 4.36 (Hyperbolic Plane). From Example 4.32 we know that for any point the normal chart covers the entire of the hyperbolic plane. Therefore the distance between any two points is given by the length of the geodesic path between them. This was calculated for vertical geodesics in Example 4.16. We can develop this into a general distance formula.
Consider two points . We will use isometries to bring them both to the -axis. Firstly we translate so that for . Next, we construct an isometry using the complex number form. Suppose that transforms to and to . Then the inverse transform obeys
From the left equation, we see that and . Continuing with the right equation, we have
Thus the distance between and is given by . In the case that we see that so that the distance formula reduces to
as expected from Exercise 4.16.
One can continue to torture the distance formula until it yields to geometric interpretation:
We have the root of the sum of squares, so this must be some euclidean distance in the hyperbolic plane. Indeed, the left square root is the distance between and . For the other square root, if we take , and the semicircle they lie on, and reflect those in the -axis to make a full circle then the second square root is the distance between and the reflection of .
There is one more property of normal charts that we will need in the next chapter, namely that the metric in a normal chart has a nice form at . For any chart, because is non-degenerate we can always find an orthonormal basis of . We can then make a linear change of coordinates such that this basis is a coordinate vector basis near . Doing this will diagonalise the metric, so . But typically are non-constant, so this property only holds at and . A special feature of normal charts is that the partial derivatives of the metric coefficients are zero, so that this procedure simplifies at up to second order.
Lemma 4.37 (Metric in Normal Chart). In a normal chart at , all first partial derivatives of the metric coefficients vanish at , i.e. .
Proof. In a normal chart at , the point is the origin and we know that the radial lines are geodesics. Let be the geodesic in the direction. If we put this into the geodesic equation,
In particular, if we put into this equation and consider all possible , this equation can only hold if for all indices. From the formula (3.55) for the Christoffel coefficients of the Levi-Civita connection, this in turn is only possible if
Permute the indices
and add the two expressions together to obtain
Example 4.38 (Hyperbolic Plane). Consider . In the usual coordinates, so . But so .
Recall Example 4.32. In that exercise we constructed polar coordinates for a normal chart at . Namely were related to the usual coordinates by
Let be (cartesian) normal coordinates at with . Then we can calculate the metric in these normal coordinates using that change of chart formula from Section 3.1:
What we want to see is that and . The direct approach, while elementary, is ugly. To make our point it is sufficient to give Taylor series for the coefficients of the metric up to first order, which requires in turn the Taylor series of the change of coordinates up to second order. The key function to understand is the denominator
Both and are even analytic functions of . Therefore inserting gives
an analytic function of the normal coordinates. As , we can easily write down the Taylor series for
For itself, clearly and
So the Taylor series of up to second order is
Next so up to second order
Finally we have up to first order
We will finish this chapter by indicating various directions in which one could continue. One can investigate further with Riemannian manifolds as metric spaces. We have encountered several times the example of the punctured plane and how it ‘blocks’ geodesics. The Hopf-Rinow theorem states that a connected Riemannian manifold is complete as a metric space if and only if every geodesic exists for all time. One can also develop the theory of the exponential map further. We have mentioned its connection to the exponential map in Lie group theory, so this could be expounded. We can also ask at every what is the largest geodesic ball on which is injective, called the injectivity radius. Relatedly, we have the cut locus at which asks when geodesics from stop being the length-minimisers. We can also consider families of geodesics, using the Jacobi field, of which the radial geodesics are one example. Another example would be geodesics beginning on some hypersurface. A typical question is to ask at what rate these geodesics are moving apart.
Naturally one can look for special manifolds. We found the geodesics of the hyperbolic plane using isometries. As mentioned, spaces whose isometries are transitive are called homogeneous, and one whose isometries are transitive on the unit sphere of are called isotropic at . Spaces such that for every point there is an isometry that acts as on are called symmetric spaces, and there is a complete classification.
These are all relatively ‘pure’ Riemannian geometry questions, in that they try to understand the intrinsic structure of a Riemannian manifold. But we can also take Riemannian manifolds as the setting to investigate all types of geometric problems. Geometric flows are one example. The final direction we will mention however is harmonic maps. Both minimal surfaces and geodesics are extremal for a functional, surface area and length respectively. If we model both of these problems as embedding stretched rubber objects and seeing the ‘minimal tension’ configuration, then this motivates the definition of a harmonic map. The name is due to the defining equation being a generalisation of the Laplacian to the context of Riemannian manifolds (compare the the Laplacian in Equation (1.41)). My PhD work was on harmonic maps from a torus into .
1In the proof of Lee Lemma 5.8 any is allowed but certain places in the proof use .
2It is unknown whether it can be isometrically immersed into , but it can be into
3Here I am using the translation of Richard Fitzpatrick, based on the authoritative Greek edition of J L Heiberg. A beautiful version designed for teaching can be found here, but it changes the numbering.
4It is not necessary to do this step, but it makes things simpler.