next up previous
Next: 4 Manifolds all of Up: Axiomatic and Coordinate Geometry Previous: 2 The axiomatic approach


3 Riemannian Geometry

Let us now consider a small region M of space on which it is possible to put coordinates, say (x1, x2, x3). Moreover, any other choice of coordinates, say (y1, y2, y3) is differentiable (to any order) with respect to the given choice, that is to say the yi are differentiable functions of the xi, and vice versa. Then we have the non-singularity of the Jacobian matrix

J(y, x) = $\displaystyle \begin{pmatrix}
\frac{\partial y_1}{\partial x_1} & \frac{\parti...
...{\partial y_3}{\partial x_2} & \frac{\partial y_3}{\partial x_3}
\end{pmatrix}$.

The basic geometric notion other than that of points of the space is the notion of tangent vectors at points. At any point these represent the physical notion of instantaneous velocities and form a linear space of dimension 3 called the tangent space at the given point. In some sense this space is the Euclidean approximation of our space at the point.

Figure 8: The tangent space to a space
\begin{figure}\epsfbox{tangent.eps}\end{figure}

With respect to the given choice of coordinates (x1, x2, x3) a vector % latex2html id marker 2314
$ \;\stackrel{\rightarrow}{v}\;$ can be represented by a column vector a = (a1, a2, a3)t. A basis for the tangent vector space at all points is given by the standard basis {e1, e2, e3} of $ \bf R^{3}_{}$. We are also interested in tensor fields, which are sums of terms of the form fi1, i2,..., irei1 $ \otimes$ ei2 $ \otimes$ ... $ \otimes$ eir where fi1, i2,..., ir is a differentiable function. The number r is called the order of the corresponding tensor. A tensor of order 0 is thus just a differentiable function and one of order 1 is a (tangent) vector field.

In another system of coordinates (y1, y2, y3) the vector % latex2html id marker 2331
$ \;\stackrel{\rightarrow}{v}\;$ is given by the column vector J(x, y) . a. This can be extended to tensors in an obvious way. With this understanding we now fix a choice of coordinates and a corresponding representation of tangent vectors as column vectors.

In order to measure the magnitude and angle of velocities we are given a symmetric positive definite bilinear pairing < , > on the tangent space at each point. With respect to a given representation of tangent vectors as column vectors, this is given by a positive definite symmetric matrix (gij)i, j = 1i, j = 3 of functions on the space. Moreover, we further assume that the matrix entries gij are differentiable (to any order) functions of the coordinates. Such a pairing is called a Riemannian metric. Having prescribed this we have a (local) Riemannian manifold.

The pairing < , > also extends to pairing on the various types of tensors by induction on the order. In fact we can pair a tensor of order r with a tensor of order s to obtain a tensor of order | r - s|.

For any vector field V we have learned in vector calculus to compute the gradient $ \nabla_{\vec{v}}^{}$(V) with respect to a vector % latex2html id marker 2346
$ \;\stackrel{\rightarrow}{v}\;$ at a point p. We have also learned about the gradient of a function f. These give an example of a derivation; in other words we have the Liebnitz rule

$\displaystyle \nabla_{\vec{v}}^{}$(f . V) = $\displaystyle \nabla_{\vec{v}}^{}$(f ) . A(p) + f (p) . $\displaystyle \nabla_{\vec{v}}^{}$(A)

Now we can clearly extend this to all tensors by applying the Liebnitz rule again as follows,

$\displaystyle \nabla_{\vec{v}}^{}$(A $\displaystyle \otimes$ B) = $\displaystyle \nabla_{\vec{v}}^{}$(A) $\displaystyle \otimes$ B(p) + A(p) $\displaystyle \otimes$ $\displaystyle \nabla_{\vec{v}}^{}$(B)

Finally, we also have the linearity relation

$\displaystyle \nabla_{a\vec{v}+b\vec{w}}^{}$ = a$\displaystyle \nabla_{\vec{v}}^{}$ + b$\displaystyle \nabla_{\vec{w}}^{}$

Now we can also consider other derivations D on tensor fields which satisfy the above rules. It follows easily from the first rule that any two derivations differ by a linear operator. In other words, any derivation D differs from $ \nabla$ by a (linear) map which associates to any tangent vector % latex2html id marker 2371
$ \;\stackrel{\rightarrow}{v}\;$ a linear endomorphism $ \omega$(% latex2html id marker 2376
$ \;\stackrel{\rightarrow}{v}\;$) of the space of tangent vectors; we write this as

D% latex2html id marker 2380
$\scriptstyle \;\stackrel{\rightarrow}{v}\;$ = $\displaystyle \nabla_{\vec{v}}^{}$ + $\displaystyle \omega$(% latex2html id marker 2385
$\displaystyle \;\stackrel{\rightarrow}{v}\;$)

where $ \omega$(% latex2html id marker 2390
$ \;\stackrel{\rightarrow}{v}\;$)(f . V) = f (p)$ \omega$(% latex2html id marker 2394
$ \;\stackrel{\rightarrow}{v}\;$)(V) is the connection form of Cartan in this special case. By the second rule such a derivation can be extended to tensor fields. Finally, the last rule says that $ \omega$(% latex2html id marker 2399
$ \;\stackrel{\rightarrow}{v}\;$) is linear in % latex2html id marker 2403
$ \;\stackrel{\rightarrow}{v}\;$.

Given a vector field V we can apply the existence theorem for ordinary differential equations to construct a 1-parameter flow $ \phi_{t}^{}$ on our space M. This satisfies $ \phi_{t+s}^{}$ = $ \phi_{t}^{}$o$ \phi_{s}^{}$ and $ \phi_{0}^{}$ fixes all points. Finally, we have d$ \phi_{t}^{}$(p)/dt = V($ \phi_{t}^{}$(p)); which is the ordinary differential equation we have solved. For any tensor field A, the following limit exists by differentiability of the quantities involved

LV(A)(p) = $\displaystyle \lim_{t\to 0}^{}$$\displaystyle {\frac{A(p) - \phi_{-t} A(\phi_t(p))}{t}}$

and it is called the Lie derivative of A with respect to V. For a pair of vector fields V and W it is not difficult to show that LV(f )(p) = $ \nabla_{V(p)}^{}$(f ) for any differentiable function f and

LV(W)(p) = $\displaystyle \nabla_{V(p)}^{}$(W) - $\displaystyle \nabla_{W(p)}^{}$(V)

for any vector field W (and hence LV(W) = - LW(V)). Note that LV(W) involves the derivatives of V even though the derivative is being taken (in some sense) with respect to V.

Figure 9: A tangent vector field and its flow
\begin{figure}\epsfbox{vecfield.eps}\end{figure}

While the gradient of a function is independent of a choice of coordinates because of the interpretation as Lie derivative, the gradient of a tensor of order at least 1 is clearly dependent upon the choice of coordinates, thus it is natural to ask for a derivation D that is in some sense ``canonical''. This is provided by the condition that the inner product form < , > has no derivative,

D% latex2html id marker 2444
$\scriptstyle \;\stackrel{\rightarrow}{u}\;$( < V, W > ) = < D% latex2html id marker 2447
$\scriptstyle \;\stackrel{\rightarrow}{u}\;$(V), W > + < V, D% latex2html id marker 2450
$\scriptstyle \;\stackrel{\rightarrow}{u}\;$(W) >

where D% latex2html id marker 2454
$\scriptstyle \;\stackrel{\rightarrow}{u}\;$(f )= $ \nabla_{\vec{u}}^{}$(f ) as before for any function f. In addition we demand the condition that the derivation is torsion-free,

LV(W)(p) = DV(p)(W) - DW(p)(V)

This can be translated into the condition on the connection form $ \omega$

$\displaystyle \omega$(% latex2html id marker 2464
$\displaystyle \;\stackrel{\rightarrow}{v}\;$)(% latex2html id marker 2467
$\displaystyle \;\stackrel{\rightarrow}{w}\;$) = $\displaystyle \omega$(% latex2html id marker 2471
$\displaystyle \;\stackrel{\rightarrow}{w}\;$)(% latex2html id marker 2474
$\displaystyle \;\stackrel{\rightarrow}{v}\;$)

There is then a unique such derivation and thus it is independent of any choice of coordinates.

In geometric terms, the usual gradient $ \nabla_{\vec{v}}^{}$(A) measures the deviation of A(p + t% latex2html id marker 2480
$ \;\stackrel{\rightarrow}{v}\;$) from a copy of A(p) moved to the point p + t% latex2html id marker 2485
$ \;\stackrel{\rightarrow}{v}\;$. However, this motion requires a notion of parallel transport or rigid motion--which may be different for our geometry from the Euclidean one provided by the coordinates; the Riemannian metric provides the correct notion of angles and distances and hence of rigid motion. Thus D provides the ``corrected'' gradient.

Now it is conceivable that there is a choice of coordinates in which the $ \omega$(% latex2html id marker 2491
$ \;\stackrel{\rightarrow}{v}\;$)'s vanish or simplify. Thus we need to find some way of checking this possibility. Riemann introduced the Riemannian curvature as a measure of this. This was modified by Christoffel who defined the curvature operator as follows. First of all for a vector field V and a tensor field A let us define DV(A)(p) = DV(p)(A). We then define the operator R by

R(V, W)(X) = DV(DW(X)) - DW(DV(X)) - DLV(W)(X)

One checks that for any functions f, g and h we have

R(fV, gW)(hX) = fgh . R(V, W)(X)

From this it follows that for each point p, R(V, W) defines an endomorphism of the tangent space which depends only on the values V(p) and W(p). We can thus denote this by R(% latex2html id marker 2508
$ \;\stackrel{\rightarrow}{v}\;$,% latex2html id marker 2511
$ \;\stackrel{\rightarrow}{w}\;$). The Riemannian curvature is then defined by

K(% latex2html id marker 2515
$\displaystyle \;\stackrel{\rightarrow}{v}\;$,% latex2html id marker 2518
$\displaystyle \;\stackrel{\rightarrow}{w}\;$,% latex2html id marker 2521
$\displaystyle \;\stackrel{\rightarrow}{x}\;$,% latex2html id marker 2524
$\displaystyle \;\stackrel{\rightarrow}{y}\;$) = < R(% latex2html id marker 2527
$\displaystyle \;\stackrel{\rightarrow}{v}\;$,% latex2html id marker 2530
$\displaystyle \;\stackrel{\rightarrow}{w}\;$)(% latex2html id marker 2533
$\displaystyle \;\stackrel{\rightarrow}{x}\;$),% latex2html id marker 2536
$\displaystyle \;\stackrel{\rightarrow}{y}\;$ >

Bianchi proved that K(% latex2html id marker 2540
$ \;\stackrel{\rightarrow}{v}\;$,% latex2html id marker 2543
$ \;\stackrel{\rightarrow}{w}\;$,% latex2html id marker 2546
$ \;\stackrel{\rightarrow}{x}\;$,% latex2html id marker 2549
$ \;\stackrel{\rightarrow}{y}\;$) = K(% latex2html id marker 2552
$ \;\stackrel{\rightarrow}{x}\;$,% latex2html id marker 2555
$ \;\stackrel{\rightarrow}{y}\;$,% latex2html id marker 2558
$ \;\stackrel{\rightarrow}{v}\;$,% latex2html id marker 2561
$ \;\stackrel{\rightarrow}{w}\;$), and for any three vector fields U, V, W

(DUR)(V, W) + (DWR)(U, V) + (DVR)(W, U) = 0

We now consider the curvature as a function of a pair of tensors of order 2

S(% latex2html id marker 2569
$\displaystyle \;\stackrel{\rightarrow}{v}\;$ $\displaystyle \otimes$ % latex2html id marker 2573
$\displaystyle \;\stackrel{\rightarrow}{w}\;$,% latex2html id marker 2576
$\displaystyle \;\stackrel{\rightarrow}{x}\;$ $\displaystyle \otimes$ % latex2html id marker 2580
$\displaystyle \;\stackrel{\rightarrow}{y}\;$) = K(% latex2html id marker 2583
$\displaystyle \;\stackrel{\rightarrow}{v}\;$,% latex2html id marker 2586
$\displaystyle \;\stackrel{\rightarrow}{w}\;$,% latex2html id marker 2589
$\displaystyle \;\stackrel{\rightarrow}{x}\;$,% latex2html id marker 2592
$\displaystyle \;\stackrel{\rightarrow}{y}\;$)

As usual we introduce the notation

% latex2html id marker 2596
$\displaystyle \;\stackrel{\rightarrow}{v}\;$ $\displaystyle \wedge$ % latex2html id marker 2600
$\displaystyle \;\stackrel{\rightarrow}{w}\;$ = $\displaystyle {\textstyle\frac{1}{2}}$(% latex2html id marker 2604
$\displaystyle \;\stackrel{\rightarrow}{v}\;$ $\displaystyle \otimes$ % latex2html id marker 2608
$\displaystyle \;\stackrel{\rightarrow}{w}\;$ - % latex2html id marker 2611
$\displaystyle \;\stackrel{\rightarrow}{w}\;$ $\displaystyle \otimes$ % latex2html id marker 2615
$\displaystyle \;\stackrel{\rightarrow}{v}\;$)

for skew-symmetric tensors of order 2. Then we have

S(% latex2html id marker 2619
$\displaystyle \;\stackrel{\rightarrow}{v}\;$ $\displaystyle \otimes$ % latex2html id marker 2623
$\displaystyle \;\stackrel{\rightarrow}{w}\;$,% latex2html id marker 2626
$\displaystyle \;\stackrel{\rightarrow}{x}\;$ $\displaystyle \otimes$ % latex2html id marker 2630
$\displaystyle \;\stackrel{\rightarrow}{y}\;$) = S(% latex2html id marker 2633
$\displaystyle \;\stackrel{\rightarrow}{v}\;$ $\displaystyle \wedge$ % latex2html id marker 2637
$\displaystyle \;\stackrel{\rightarrow}{w}\;$,% latex2html id marker 2640
$\displaystyle \;\stackrel{\rightarrow}{x}\;$ $\displaystyle \wedge$ % latex2html id marker 2644
$\displaystyle \;\stackrel{\rightarrow}{y}\;$)

Moreover S(% latex2html id marker 2648
$ \;\stackrel{\rightarrow}{v}\;$ $ \wedge$ % latex2html id marker 2652
$ \;\stackrel{\rightarrow}{w}\;$,% latex2html id marker 2655
$ \;\stackrel{\rightarrow}{x}\;$ $ \wedge$ % latex2html id marker 2659
$ \;\stackrel{\rightarrow}{y}\;$) is a symmetric bilinear pairing on the space of skew-symmetric tensors of order 2. Now one such is given by the pairing of tensors

< % latex2html id marker 2663
$\displaystyle \;\stackrel{\rightarrow}{v}\;$ $\displaystyle \wedge$ % latex2html id marker 2667
$\displaystyle \;\stackrel{\rightarrow}{w}\;$,% latex2html id marker 2670
$\displaystyle \;\stackrel{\rightarrow}{x}\;$ $\displaystyle \wedge$ % latex2html id marker 2674
$\displaystyle \;\stackrel{\rightarrow}{y}\;$ > = $\displaystyle {\textstyle\frac{1}{2}}$( < % latex2html id marker 2678
$\displaystyle \;\stackrel{\rightarrow}{v}\;$,% latex2html id marker 2681
$\displaystyle \;\stackrel{\rightarrow}{x}\;$ > < % latex2html id marker 2684
$\displaystyle \;\stackrel{\rightarrow}{w}\;$,% latex2html id marker 2687
$\displaystyle \;\stackrel{\rightarrow}{y}\;$ > - < % latex2html id marker 2690
$\displaystyle \;\stackrel{\rightarrow}{v}\;$,% latex2html id marker 2693
$\displaystyle \;\stackrel{\rightarrow}{y}\;$ > < % latex2html id marker 2696
$\displaystyle \;\stackrel{\rightarrow}{w}\;$,% latex2html id marker 2699
$\displaystyle \;\stackrel{\rightarrow}{x}\;$ > )

A comparison of the associated quadratic forms is then given by the ratio

$\displaystyle \kappa$(% latex2html id marker 2704
$\displaystyle \;\stackrel{\rightarrow}{v}\;$,% latex2html id marker 2707
$\displaystyle \;\stackrel{\rightarrow}{w}\;$) = - $\displaystyle {\frac{K(\vec{v},\vec{w},\vec{v},\vec{w})}{%%
<\vec{v},\vec{v}><\vec{w},\vec{w}> -
<\vec{v},\vec{w}>^2}}$

This is what is called the sectional curvature of the Riemannian manifold (the sign is for reasons of convention). This is by far the most important invariant associated with a Riemannian manifold. It is clear from the above discussion that the sectional curvature, the Riemannian curvature and the Riemann-Christoffel operator on a Riemannian manifold determine each other.

An important result of Cartan and Hadamard (see [3]) states that any Riemannian space M as above with the property that its sectional curvature is constant can be identified with an open subset of either the Euclidean space or Hyperbolic (Bolyai-Lobachevsky) space or Projective space in an isometric way.

We will see in the next section how to apply this to the study of axiomatic geometries.


next up previous
Next: 4 Manifolds all of Up: Axiomatic and Coordinate Geometry Previous: 2 The axiomatic approach
Kapil Hari Paranjape 2002-11-21