You are on page 1of 29

Differential geometry of surfaces

In mathematics, the differential geometry of surfaces deals with the


differential geometry of smooth surfaces with various additional
structures, most often, a Riemannian metric. Surfaces have been
extensively studied from various perspectives: extrinsically, relating to
their embedding in Euclidean space and intrinsically, reflecting their
properties determined solely by the distance within the surface as
measured along curves on the surface. One of the fundamental concepts
investigated is the Gaussian curvature, first studied in depth by Carl
Friedrich Gauss (articles of 1825 and 1827), who showed that curvature
was an intrinsic property of a surface, independent of its isometric
embedding in Euclidean space.

Surfaces naturally arise as graphs of functions of a pair of variables, and


sometimes appear in parametric form or as loci associated to space curves. Carl Friedrich Gauss in 1828
An important role in their study has been played by Lie groups (in the
spirit of the Erlangen program), namely the symmetry groups of the
Euclidean plane, the sphere and the hyperbolic plane. These Lie groups can be used to describe surfaces of constant
Gaussian curvature; they also provide an essential ingredient in the modern approach to intrinsic differential
geometry through connections. On the other hand, extrinsic properties relying on an embedding of a surface in
Euclidean space have also been extensively studied. This is well illustrated by the non-linear EulerLagrange
equations in the calculus of variations: although Euler developed the one variable equations to understand geodesics,
defined independently of an embedding, one of Lagrange's main applications of the two variable equations was to
minimal surfaces, a concept that can only be defined in terms of an embedding.

Contents
1 Overview
1.1 History of surfaces
2 Curvature of surfaces in Euclidean space
3 Examples
3.1 Surfaces of revolution
3.2 Quadric surfaces
3.3 Ruled surfaces
3.4 Minimal surfaces
3.5 Surfaces of constant Gaussian curvature
4 Local metric structure
4.1 Line and area elements
4.2 Second fundamental form
4.3 Shape operator
5 Geodesic curves on a surface
5.1 Geodesics
5.2 Geodesic curvature
5.3 Isometric embedding problem
5.4 Orthogonal coordinates
6 Geodesic polar coordinates
6.1 Exponential map
6.2 Computation of normal coordinates
6.3 Gauss's lemma
6.4 Theorema Egregium
6.5 GaussJacobi equation
6.6 LaplaceBeltrami operator
7 GaussBonnet theorem
7.1 Geodesic triangles
7.2 GaussBonnet theorem
7.3 Curvature and embeddings
8 Surfaces of constant curvature
8.1 Euclidean geometry
8.2 Spherical geometry
8.3 Hyperbolic geometry
8.4 Uniformization
9 Surfaces of non-positive curvature
9.1 Alexandrov's comparison inequality
9.2 Existence of geodesics
9.3 Von Mangoldt-Hadamard theorem
10 Riemannian connection and parallel transport
10.1 Covariant derivative
10.2 Parallel transport
10.3 Connection 1-form
11 Global differential geometry of surfaces
12 Reading guide
13 See also
14 Notes
15 References

Overview
Polyhedra in the Euclidean space, such as the boundary of a cube, are among the first surfaces encountered in
geometry. It is also possible to define smooth surfaces, in which each point has a neighborhood diffeomorphic to some
open set in E2, the Euclidean plane. This elaboration allows calculus to be applied to surfaces to prove many results.

Two smooth surfaces are diffeomorphic if and only if they are homeomorphic. (The analogous result does not hold for
manifolds of dimension greater than three.) It follows that closed surfaces are classified up to diffeomorphism by their
Euler characteristic and orientability.

Smooth surfaces equipped with Riemannian metrics are of foundational importance in differential geometry. A
Riemannian metric endows a surface with notions of geodesic, distance, angle, and area. An important class of such
surfaces are the developable surfaces: surfaces that can be flattened to a plane without stretching; examples include
the cylinder and the cone.

In addition, there are properties of surfaces which depend on an embedding of the surface into Euclidean space. These
surfaces are the subject of extrinsic geometry. They include

Minimal surfaces are surfaces that minimize the surface area for given boundary conditions; examples include
soap films stretched across a wire frame, catenoids and helicoids.
Ruled surfaces are surfaces that have at least one straight line running through every point; examples include the
cylinder and the hyperboloid of one sheet.
Any n-dimensional complex manifold is, at the same time, a 2n-dimensional real manifold. Thus any complex one-
manifold (also called a Riemann surface) is a smooth oriented surface with an associated complex structure. Every
closed surface admits complex structures. Any complex algebraic curve or real algebraic surface is also a smooth
surface, possibly with singularities.

Complex structures on a closed oriented surface correspond to conformal equivalence classes of Riemannian metrics
on the surface. One version of the uniformization theorem (due to Poincar) states that any Riemannian metric on an
oriented, closed surface is conformally equivalent to an essentially unique metric of constant curvature. This provides
a starting point for one of the approaches to Teichmller theory, which provides a finer classification of Riemann
surfaces than the topological one by Euler characteristic alone.

The uniformization theorem states that every smooth Riemannian surface S is conformally equivalent to a surface
having constant curvature, and the constant may be taken to be 1, 0, or 1. A surface of constant curvature 1 is locally
isometric to the sphere, which means that every point on the surface has an open neighborhood that is isometric to an
open set on the unit sphere in E3 with its intrinsic Riemannian metric. Likewise, a surface of constant curvature 0 is
locally isometric to the Euclidean plane, and a surface of constant curvature 1 is locally isometric to the hyperbolic
plane.

Constant curvature surfaces are the two-dimensional realization of what are known as space forms. These are often
studied from the point of view of Felix Klein's Erlangen programme, by means of smooth transformation groups. Any
connected surface with a three-dimensional group of isometries is a surface of constant curvature.

A complex surface is a complex two-manifold and thus a real four-manifold; it is not a surface in the sense of this
article. Neither are algebraic curves or surfaces defined over fields other than the complex numbers.

History of surfaces
Isolated properties of surfaces of revolution were known already to Archimedes. The development of calculus in the
seventeenth century provided a more systematic way of proving them. Curvature of general surfaces was first studied
by Euler. In 1760[1] he proved a formula for the curvature of a plane section of a surface and in 1771[2] he considered
surfaces represented in a parametric form. Monge laid down the foundations of their theory in his classical memoir
L'application de l'analyse la gometrie which appeared in 1795. The defining contribution to the theory of surfaces
was made by Gauss in two remarkable papers written in 1825 and 1827.[3] This marked a new departure from tradition
because for the first time Gauss considered the intrinsic geometry of a surface, the properties which are determined
only by the geodesic distances between points on the surface independently of the particular way in which the surface
is located in the ambient Euclidean space. The crowning result, the Theorema Egregium of Gauss, established that the
Gaussian curvature is an intrinsic invariant, i.e. invariant under local isometries. This point of view was extended to
higher-dimensional spaces by Riemann and led to what is known today as Riemannian geometry. The nineteenth
century was the golden age for the theory of surfaces, from both the topological and the differential-geometric point of
view, with most leading geometers devoting themselves to their study. Darboux collected many results in his four-
volume treatise Thorie des surfaces (18871896).

The presentation below largely follows Gauss, but with important later contributions from other geometers. For a time
Gauss was Cartographer to George III of Great Britain and Hanover; this royal patronage could explain why these
papers contain practical calculations of the curvature of the earth based purely on measurements on the surface of the
planet.

Curvature of surfaces in Euclidean space


Informally Gauss defined the curvature of a surface in terms of the
curvatures of certain plane curves connected with the surface. He later
found a series of equivalent definitions. One of the first was in terms of the
area-expanding properties of the Gauss map, a map from the surface to a
2-dimensional sphere. However, before obtaining a more intrinsic
definition in terms of the area and angles of small triangles, Gauss needed
to make an in-depth investigation of the properties of geodesics on the
surface, i.e. paths of shortest length between two fixed points on the
surface[4] (see below). The principal curvatures at a point
on a surface
The Gaussian curvature at a point on an embedded smooth surface
given locally by the equation

z = F(x,y)
in Euclidean space (E3), is defined to be the product of the principal
curvatures at the point;[5] the mean curvature is defined to be their
average. The principal curvatures are the maximum and minimum
The Gauss map sends a point on
curvatures of the plane curves obtained by intersecting the surface with
the surface to the outward pointing
planes normal to the tangent plane at the point. If the point is (0, 0, 0)
unit normal vector, a point on S2
with tangent plane z = 0, then, after a rotation about the z-axis setting the
coefficient on xy to zero, F will have the Taylor series expansion

The principal curvatures are k1 and k2. In this case, the Gaussian curvature is given by

and the mean curvature by

Since K and Km are invariant under isometries of E3, in general

and

where the derivatives at the point are given by P = Fx, Q = Fy, R = Fxx, S = Fxy, and T = Fyy.[6]

For every oriented embedded surface the Gauss map is the map into the unit sphere sending each point to the
(outward pointing) unit normal vector to the oriented tangent plane at the point. In coordinates the map sends (x,y,z)
to
Direct computation shows that: the Gaussian curvature is the Jacobian of the Gauss map.[7]

Examples

Surfaces of revolution
A surface of revolution can be obtained by rotating a curve in the xz-plane
about the z-axis, assuming the curve does not intersect the z-axis. Suppose
that the curve is given by

with t lies in (a, b), and is parametrized by arclength, so that

Then the surface of revolution is the point set

The surface of revolution obtained


by rotating the curve x = 2 + cos z
about the z-axis.

The Gaussian curvature and mean curvature are given by[8]

Geodesics on a surface of revolution are governed by Clairaut's relation.

Quadric surfaces
Consider the quadric surface defined by[9]

This surface admits a parametrization

A quadric ellipsoid
The Gaussian curvature and mean curvature are given by

Ruled surfaces
A ruled surface is one which can be generated by the motion of a straight
line in E3.[10] Choosing a directrix on the surface, i.e. a smooth unit speed
curve c(t) orthogonal to the straight lines, and then choosing u(t) to be unit
vectors along the curve in the direction of the lines, the velocity vector
v = ct and u satisfy

The surface consists of points

as s and t vary.

Then, if

A single-sheeted quadric
hyperboloid which is a ruled surface
the Gaussian and mean curvature are given by in two different ways.

The Gaussian curvature of the ruled surface vanishes if and only if ut and v are proportional,[11] This condition is
equivalent to the surface being the envelope of the planes along the curve containing the tangent vector v and the
orthogonal vector u, i.e. to the surface being developable along the curve.[12] More generally a surface in E3 has
vanishing Gaussian curvature near a point if and only if it is developable near that point.[13] (An equivalent condition
is given below in terms of the metric.)

Minimal surfaces
In 1760 Lagrange extended Euler's results on the calculus of variations involving integrals in one variable to two
variables.[14] He had in mind the following problem:

Given a closed curve in E3, find a surface having the curve as boundary with minimal area.

Such a surface is called a minimal surface.


In 1776 Jean Baptiste Meusnier showed that the differential equation derived by Lagrange was equivalent to the
vanishing of the mean curvature of the surface:

A surface is minimal if and only if its mean curvature vanishes.

Minimal surfaces have a simple interpretation in real life: they are the shape a soap film will assume if a wire frame
shaped like the curve is dipped into a soap solution and then carefully lifted out. The question as to whether a minimal
surface with given boundary exists is called Plateau's problem after the Belgian physicist Joseph Plateau who carried
out experiments on soap films in the mid-nineteenth century. In 1930 Jesse Douglas and Tibor Rad gave an
affirmative answer to Plateau's problem (Douglas was awarded one of the first Fields medals for this work in 1936).[15]

Many explicit examples of minimal surface are known explicitly, such as the catenoid, the helicoid, the Scherk surface
and the Enneper surface. There has been extensive research in this area, summarised in Osserman (2002). In
particular a result of Osserman shows that if a minimal surface is non-planar, then its image under the Gauss map is
dense in S2.

Surfaces of constant Gaussian curvature


If a surface has constant Gaussian curvature, it is called a surface of
constant curvature.[16]

The unit sphere in E3 has constant Gaussian curvature +1.


The Euclidean plane and the cylinder both have constant Gaussian
curvature 0.
The surfaces of revolution with tt = have constant Gaussian
curvature 1. Particular cases are obtained by taking (t) =C cosh t,
C sinh t and C et.[17] The latter case is the classical pseudosphere
generated by rotating a tractrix around a central axis. In 1868 Beltrami
showed that the geometry of the pseudosphere was directly related to Surfaces with (from l. to r.) constant
that of the hyperbolic plane, discovered independently by negative, zero and positive
Lobachevsky (1830) and Bolyai (1832). Already in 1840, F. Minding, a Gaussian curvature
student of Gauss, had obtained trigonometric formulas for the
pseudosphere identical to those for the hyperbolic plane.[18] This
surface of constant curvature is now better understood in terms of the
Poincar metric on the upper half plane or the unit disc, and has been described by other models such as the
Klein model or the hyperboloid model, obtained by considering the two-sheeted hyperboloid q(x, y, z) = 1 in
three-dimensional Minkowski space, where q(x, y, z) = x2 + y2 z2.[19]
Each of these surfaces of constant curvature has a transitive Lie group of symmetries. This group theoretic fact has far-
reaching consequences, all the more remarkable because of the central role these special surfaces play in the geometry
of surfaces, due to Poincar's uniformization theorem (see below).

Other examples of surfaces with Gaussian curvature 0 include cones, tangent developables, and more generally any
developable surface.

Local metric structure


For any surface embedded in Euclidean space of dimension 3 or higher, it is possible to measure the length of a curve
on the surface, the angle between two curves and the area of a region on the surface. This structure is encoded
infinitesimally in a Riemannian metric on the surface through line elements and area elements. Classically in the
nineteenth and early twentieth centuries only surfaces embedded in R3 were considered and the metric was given as a
22 positive definite matrix varying smoothly from point to point in a local parametrization of the surface. The idea of
local parametrization and change of coordinate was later formalized through the current abstract notion of a manifold,
a topological space where the smooth structure is given by local charts on the manifold, exactly as the planet Earth is
mapped by atlases today. Changes of coordinates between different charts of the same
region are required to be smooth. Just as contour lines on real-life maps encode
changes in elevation, taking into account local distortions of the Earth's surface to
calculate true distances, so the Riemannian metric describes distances and areas "in
the small" in each local chart. In each local chart a Riemannian metric is given by
smoothly assigning a 22 positive definite matrix to each point; when a different chart
is taken, the matrix is transformed according to the Jacobian matrix of the coordinate
change. The manifold then has the structure of a 2-dimensional Riemannian manifold.

Line and area elements


Taking a local chart, for example by projecting onto the xy-plane (z = 0), the line
element ds and the area element dA can be written in terms of local coordinates as A chart for the upper
hemisphere of the 2-
ds2 = E dx2 + 2F dx dy + G dy2 sphere obtained by
projecting onto the xy-
and plane

1
dA = (EG F2) 2 dx dy.
The expression E dx2 + 2F dx dy + G dy2 is called the first fundamental form.[20]

The matrix

Coordinate changes
between different local
charts must be smooth
is required to be positive-definite and to depend smoothly on x and y.

In a similar way line and area elements can be associated to any abstract Riemannian 2-manifold in a local chart.

Second fundamental form


The extrinsic geometry of surfaces studies the properties of
surfaces embedded into a Euclidean space, typically E3. In
intrinsic geometry, two surfaces are "the same" if it is
possible to unfold one surface onto the other without
stretching it, i.e. a map of one surface onto the other
preserving distance. Thus a cylinder is locally "the same" as
the plane. In extrinsic geometry, two surfaces are "the same"
if they are congruent in the ambient Euclidean space, i.e.
there is an isometry of E3 carrying one surface onto the
other. With this more rigid definition of similitude, the
cylinder and the plane are obviously no longer the same. Definition of second fundamental form

Although the primary invariant in the study of the intrinsic


geometry of surfaces is the metric (the first fundamental form) and the Gaussian curvature, certain properties of
surfaces also depend on an embedding into E3 (or a higher dimensional Euclidean space). The most important
example is the second fundamental form, defined classically as follows.[21]
Take a point (x, y) on the surface in a local chart. The Euclidean distance from a nearby point (x + dx, y + dy) to the
tangent plane at (x, y), i.e. the length of the perpendicular dropped from the nearby point to the tangent plane, has the
form

e dx2 + 2f dx dy + g dy2
plus third and higher order corrections. The above expression, a symmetric bilinear form at each point, is the second
fundamental form. It is described by a 2 2 symmetric matrix

which depends smoothly on x and y. The Gaussian curvature can be calculated as the ratio of the determinants of the
second and first fundamental forms:

Remarkably Gauss proved that it is an intrinsic invariant (see his Theorema Egregium below).

One of the other extrinsic numerical invariants of a surface is the mean curvature Km defined as the sum of the
principal curvatures. It is given by the formula[20]

The coefficients of the first and second fundamental forms satisfy certain compatibility conditions known as the
k
Gauss-Codazzi equations; they involve the Christoffel symbols ij associated with the first fundamental form:[22]

These equations can also be succinctly expressed and derived in the language of connection forms due to lie
Cartan.[23] Pierre Bonnet proved that two quadratic forms satisfying the Gauss-Codazzi equations always uniquely
determine an embedded surface locally.[24] For this reason the Gauss-Codazzi equations are often called the
fundamental equations for embedded surfaces, precisely identifying where the intrinsic and extrinsic curvatures come
from. They admit generalizations to surfaces embedded in more general Riemannian manifolds.

Shape operator
The differential df of the Gauss map f can be used to define a type of extrinsic curvature, known as the shape
operator[25] or Weingarten map. This operator first appeared implicitly in the work of Wilhelm Blaschke and later
explicitly in a treatise by Burali-Forti and Burgati.[26] Since at each point x of the surface, the tangent space is an inner
product space, the shape operator Sx can be defined as a linear operator on this space by the formula

for tangent vectors v, w (the inner product makes sense because df(v) and w both lie in E3).[27] The right hand side is
symmetric in v and w, so the shape operator is self-adjoint on the tangent space. The eigenvalues of Sx are just the
principal curvatures k1 and k2 at x. In particular the determinant of the shape operator at a point is the Gaussian
curvature, but it also contains other information, since the mean curvature is half the trace
of the shape operator. The mean curvature is an extrinsic invariant. In intrinsic geometry,
a cylinder is developable, meaning that every piece of it is intrinsically indistinguishable
from a piece of a plane since its Gauss curvature vanishes identically. Its mean curvature is
not zero, though; hence extrinsically it is different from a plane.

In general, the eigenvectors and eigenvalues of the shape operator at each point determine
the directions in which the surface bends at each point. The eigenvalues correspond to the
principal curvatures of the surface and the eigenvectors are the corresponding principal
directions. The principal directions specify the directions that a curve embedded in the
surface must travel to have maximum and minimum curvature, these being given by the Wilhelm Blaschke
principal curvatures. (1885-1962)

The shape operator is given in terms of the components of the first and second
fundamental forms by the Weingarten equations:[28]

Geodesic curves on a surface


Curves on a surface which minimize length between the endpoints are called geodesics; they are the shape that an
elastic band stretched between the two points would take. Mathematically they are described using partial differential
equations from the calculus of variations. The differential geometry of surfaces revolves around the study of geodesics.
It is still an open question whether every Riemannian metric on a 2-dimensional local chart arises from an embedding
in 3-dimensional Euclidean space: the theory of geodesics has been used to show this is true in the important case
when the components of the metric are analytic.

Geodesics
Given a piecewise smooth path c(t) = (x(t), y(t)) in the chart for t in [a, b], its length is
defined by

and energy by

A geodesic triangle on
the sphere. The
geodesics are great
circle arcs.
The length is independent of the parametrization of a path. By the EulerLagrange
equations, if c(t) is a path minimising length, parametrized by arclength, it must
satisfy the Euler equations

k
where the Christoffel symbols ij are given by
where g11 = E, g12 = F, g22 = G and gij is the inverse matrix to gij. A path satisfying the Euler equations is called a
geodesic. By the CauchySchwarz inequality a path minimising energy is just a geodesic parametrised by arc length;
and, for any geodesic, the parameter t is proportional to arclength.[29]

Geodesic curvature
The geodesic curvature kg at a point of a curve c(t), parametrised by arc length, on an oriented surface is defined to
be[30]

where n(t) is the "principal" unit normal to the curve in the surface, constructed by rotating the unit tangent vector
(t) through an angle of +90.

The geodesic curvature at a point is an intrinsic invariant depending only on the metric near the point.
A unit speed curve on a surface is a geodesic if and only if its geodesic curvature vanishes at all points on the
curve.
A unit speed curve c(t) in an embedded surface is a geodesic if and only if its acceleration vector c(t) is normal to
the surface.
The geodesic curvature measures in a precise way how far a curve on the surface is from being a geodesic.

Isometric embedding problem


A result of Jacobowitz (1972) and Poznjak (1973) shows that every metric structure on a surface arises from a local
embedding in E4. Apart from some special cases, whether this is possible in E3 remains an open question, the so-
called "Weyl problem".[31] In 1926 Maurice Janet proved that it is always possible locally if E, F and G are analytic;
1
soon afterwards lie Cartan generalised this to local embeddings of Riemannian n-manifolds in Em where m = (n2 +
2
n). To prove Janet's theorem near (0,0), the CauchyKowalevski theorem is used twice to produce analytic geodesics
orthogonal to the y-axis and then the x-axis to make an analytic change of coordinate so that E = 1 and F = 0. An
isometric embedding u must satisfy

ux ux = 1, ux uy = 0, uy uy = G.

Differentiating gives the three additional equations

uxx uy = 0, uxx ux = 0, uxx uyy = uxy uxy 12 Gxx.

with u(0,y) and ux(0,y) prescribed. These equations can be solved near (0,0) using the CauchyKowalevski theorem
and yield a solution of the original embedding equations.

Orthogonal coordinates
When F = 0 in the metric, lines parallel to the x- and y-axes are orthogonal and provide orthogonal coordinates. If
1
H = (EG) 2, then the Gaussian curvature is given by[32]
1
If in addition E = 1, so that H = G 2, then the angle at the intersection
between geodesic (x(t),y(t)) and the line y = constant is given by the
equation

The derivative of is given by a classical derivative formula of Gauss:[33]

In orthogonal coordinates is the


angle the tangent L to the geodesic
C makes with the x-axis
Geodesic polar coordinates
Once a metric is given on a surface and a base point is fixed, there is a
unique geodesic connecting the base point to each sufficiently nearby
point. The direction of the geodesic at the base point and the distance
uniquely determine the other endpoint. These two bits of data, a direction
and a magnitude, thus determine a tangent vector at the base point. The
map from tangent vectors to endpoints smoothly sweeps out a
neighbourhood of the base point and defines what is called the
"exponential map", defining a local coordinate chart at that base point. The
neighbourhood swept out has similar properties to balls in Euclidean
space, namely any two points in it are joined by a unique geodesic. This
property is called "geodesic convexity" and the coordinates are called
"normal coordinates". The explicit calculation of normal coordinates can
be accomplished by considering the differential equation satisfied by
Carl Jacobi (18041851)
geodesics. The convexity properties are consequences of Gauss's lemma
and its generalisations. Roughly speaking this lemma states that geodesics
starting at the base point must cut the spheres of fixed radius centred on
the base point at right angles. Geodesic polar coordinates are obtained by
combining the exponential map with polar coordinates on tangent vectors
at the base point. The Gaussian curvature of the surface is then given by
the second order deviation of the metric at the point from the Euclidean
metric. In particular the Gaussian curvature is an invariant of the metric,
Gauss's celebrated Theorema Egregium. A convenient way to understand
the curvature comes from an ordinary differential equation, first
considered by Gauss and later generalized by Jacobi, arising from the
change of normal coordinates about two different points. The Gauss Contour lines tracking the motion of
points on a fixed curve moving
Jacobi equation provides another way of computing the Gaussian
along geodesics towards a
curvature. Geometrically it explains what happens to geodesics from a
basepoint
fixed base point as the endpoint varies along a small curve segment
through data recorded in the Jacobi field, a vector field along the
geodesic.[34] One and a quarter centuries after Gauss and Jacobi, Marston Morse gave a more conceptual
interpretation of the Jacobi field in terms of second derivatives of the energy function on the infinite-dimensional
Hilbert manifold of paths.[35]

Exponential map
dv
The theory of ordinary differential equations shows that if f(t, v) is smooth then the differential equation = f(t,v)
dt
with initial condition v(0) = v0 has a unique solution for |t| sufficiently small and the solution depends smoothly on t
and v0. This implies that for sufficiently small tangent vectors v at a given point p = (x0,y0), there is a geodesic cv(t)
defined on (2,2) with cv(0) = (x0,y0) and v(0) = v. Moreover, if |s| 1, then csv = cv(st). The exponential map is
defined by

expp(v) = cv (1)

and gives a diffeomorphism between a disc v < and a neighbourhood of p; more generally the map sending (p,v) to
expp(v) gives a local diffeomorphism onto a neighbourhood of (p,p). The exponential map gives geodesic normal
coordinates near p.[36]

Computation of normal coordinates


There is a standard technique (see for example Berger (2004)) for computing the change of variables to normal
coordinates u, v at a point as a formal Taylor series expansion. If the coordinates x, y at (0,0) are locally orthogonal,
write

x(u,v) = u + L(u,v) + (u,v) +


y(u,v) = v + M(u,v) + (u,v) +
where L, M are quadratic and , cubic homogeneous polynomials in u and v. If u and v are fixed, x(t) = x(tu,tv) and
y(t) = y(tu, tv) can be considered as formal power series solutions of the Euler equations: this uniquely determines ,
, L, M, and .

Gauss's lemma
In these coordinates the matrix g(x) satisfies g(0) = I and the lines t tv are
geodesics through 0. Euler's equations imply the matrix equation

g(v)v = v,
a key result, usually called the Gauss lemma. Geometrically it states that

the geodesics through 0 cut the circles centred at 0


orthogonally.

In geodesic polar coordinates


Taking polar coordinates (r,), it follows that the metric has the form
the geodesics radiating from the
origin cut the circles of constant
ds2 = dr2 + G(r,) d2. radius orthogonally. The
distances along radii are true
In geodesic coordinates, it is easy to check that the geodesics through zero distances but on the concentric
minimize length. The topology on the Riemannian manifold is then given by a circles small arcs have length
distance function d(p,q), namely the infimum of the lengths of piecewise H(r,) = G(r,)
1
times the
2
smooth paths between p and q. This distance is realised locally by geodesics, angle they subtend.
so that in normal coordinates d(0,v) = v. If the radius is taken small
enough, a slight sharpening of the Gauss lemma shows that the image U of the
disc v < under the exponential map is geodesically convex, i.e. any two points in U are joined by a unique geodesic
lying entirely inside U.[5][37]

Theorema Egregium
Taking x and y coordinates of a surface in E3 corresponding to F(x,y) = k1x2 + k2y2 + , the power series expansion
of the metric is given in normal coordinates (u, v) as

ds2 = du2 + dv2 + K(u dv v du)2 +


This extraordinary result Gauss's Theorema Egregium shows that the Gaussian curvature of a surface can be
computed solely in terms of the metric and is thus an intrinsic invariant of the surface, independent of any embedding
in E3 and unchanged under coordinate transformations. In particular isometries of surfaces preserve Gaussian
curvature.[5]

GaussJacobi equation
Taking a coordinate change from normal coordinates at p to normal coordinates at a nearby point q, yields the Sturm
1
Liouville equation satisfied by H(r,) = G(r,) 2, discovered by Gauss and later generalised by Jacobi,

Hrr = KH

The Jacobian of this coordinate change at q is equal to Hr. This gives another way of establishing the intrinsic nature
of Gaussian curvature. Because H(r,) can be interpreted as the length of the line element in the direction, the
GaussJacobi equation shows that the Gaussian curvature measures the spreading of geodesics on a geometric surface
as they move away from a point.[38]

LaplaceBeltrami operator
On a surface with local metric

and LaplaceBeltrami operator

where H2 = EG F2, the Gaussian curvature at a point is given by the formula[39]

where r is the denotes the geodesic distance from the point. Since is manifestly an intrinsic invariant, this gives yet
another proof that the Gaussian curvature is an intrinsic invariant.

In isothermal coordinates, first considered by Gauss, the metric is required to be of the special form

In this case the LaplaceBeltrami operator is given by

and satisfies Liouville's equation[40]


Isothermal coordinates are known to exist in a neighbourhood of any point on the surface, although all proofs to date
rely on non-trivial results on partial differential equations.[41] There is an elementary proof for minimal surfaces.[42]

GaussBonnet theorem
On a sphere or a hyperboloid, the area of a geodesic triangle, i.e. a triangle all the
sides of which are geodesics, is proportional to the difference of the sum of the
interior angles and . The constant of proportionality is just the Gaussian curvature,
a constant for these surfaces. For the torus, the difference is zero, reflecting the fact
that its Gaussian curvature is zero. These are standard results in spherical, hyperbolic
and high school trigonometry (see below). Gauss generalised these results to an
arbitrary surface by showing that the integral of the Gaussian curvature over the
interior of a geodesic triangle is also equal to this angle difference or excess. His
formula showed that the Gaussian curvature could be calculated near a point as the A triangulation of the torus
limit of area over angle excess for geodesic triangles shrinking to the point. Since any
closed surface can be decomposed up into geodesic triangles, the formula
could also be used to compute the integral of the curvature over the whole
surface. As a special case of what is now called the GaussBonnet theorem,
Gauss proved that this integral was remarkably always 2 times an integer,
a topological invariant of the surface called the Euler characteristic. This
invariant is easy to compute combinatorially in terms of the number of
vertices, edges, and faces of the triangles in the decomposition, also called
a triangulation. This interaction between analysis and topology was the
forerunner of many later results in geometry, culminating in the Atiyah-
Singer index theorem. In particular properties of the curvature impose
restrictions on the topology of the surface.

The area of a spherical triangle on


Geodesic triangles the unit sphere is + + .
Gauss proved that, if is a geodesic triangle on a surface with angles ,
and at vertices A, B and C, then

In fact taking geodesic polar coordinates with origin A and AB, AC the radii at polar angles 0 and :

where the second equality follows from the GaussJacobi equation and the fourth from Gauss' derivative formula in
the orthogonal coordinates (r,).
Gauss' formula shows that the curvature at a point can be calculated as the limit of angle excess + + over
area for successively smaller geodesic triangles near the point. Qualitatively a surface is positively or negatively curved
according to the sign of the angle excess for arbitrarily small geodesic triangles.[20]

GaussBonnet theorem
Since every compact oriented 2-manifold M can be triangulated by small
geodesic triangles, it follows that

where (M) denotes the Euler characteristic of the surface.

In fact if there are F faces, E edges and V vertices, then 3F = 2E and the
left hand side equals 2V F = 2(V E + F) = 2(M).

This is the celebrated GaussBonnet theorem: it shows that the


The Euler characteristic of a sphere,
integral of the Gaussian curvature is a topological invariant of the triangulated like an icosahedron, is
manifold, namely the Euler characteristic. This theorem can be interpreted V - E + F = 12 30 + 20 = 2.
in many ways; perhaps one of the most far-reaching has been as the index
theorem for an elliptic differential operator on M, one of the simplest cases
of the Atiyah-Singer index theorem. Another related result, which can be proved using the GaussBonnet theorem, is
the Poincar-Hopf index theorem for vector fields on M which vanish at only a finite number of points: the sum of the
indices at these points equals the Euler characteristic, where the index of a point is defined as follows: on a small circle
round each isolated zero, the vector field defines a map into the unit circle; the index is just the winding number of this
map.)[20]

Curvature and embeddings


If the Gaussian curvature of a surface M is everywhere positive, then the Euler characteristic is positive so M is
homeomorphic (and therefore diffeomorphic) to S2. If in addition the surface is isometrically embedded in E3, the
Gauss map provides an explicit diffeomorphism. As Hadamard observed, in this case the surface is convex; this
criterion for convexity can be viewed as a 2-dimensional generalisation of the well-known second derivative criterion
for convexity of plane curves. Hilbert proved that every isometrically embedded closed surface must have a point of
positive curvature. Thus a closed Riemannian 2-manifold of non-positive curvature can never be embedded
isometrically in E3; however, as Adriano Garsia showed using the Beltrami equation for quasiconformal mappings,
this is always possible for some conformally equivalent metric.[43]

Surfaces of constant curvature


The simply connected surfaces of constant curvature 0, +1 and 1 are the Euclidean plane, the unit sphere in E3, and
the hyperbolic plane. Each of these has a transitive three-dimensional Lie group of orientation preserving isometries
G, which can be used to study their geometry. Each of the two non-compact surfaces can be identified with the
quotient G / K where K is a maximal compact subgroup of G. Here K is isomorphic to SO(2). Any other closed
Riemannian 2-manifold M of constant Gaussian curvature, after scaling the metric by a constant factor if necessary,
will have one of these three surfaces as its universal covering space. In the orientable case, the fundamental group of
M can be identified with a torsion-free uniform subgroup of G and M can then be identified with the double coset
space \ G / K. In the case of the sphere and the Euclidean plane, the only possible examples are the sphere itself and
tori obtained as quotients of R2 by discrete rank 2 subgroups. For closed surfaces of genus g 2, the moduli space of
Riemann surfaces obtained as varies over all such subgroups, has real dimension 6g 6.[44] By Poincar's
uniformization theorem, any orientable closed 2-manifold is conformally equivalent to a surface of constant curvature
0, +1 or 1. In other words, by multiplying the metric by a positive scaling factor, the Gaussian curvature can be made
to take exactly one of these values (the sign of the Euler characteristic of M).[45]

Euclidean geometry
In the case of the Euclidean plane, the symmetry group is the Euclidean
motion group, the semidirect product of the two dimensional group of
translations by the group of rotations.[46] Geodesics are straight lines and
the geometry is encoded in the elementary formulas of trigonometry, such
as the cosine rule for a triangle with sides a, b, c and angles , , :

A triangle in the plane


Flat tori can be obtained by taking the quotient of R2 by a lattice, i.e. a free
Abelian subgroup of rank 2. These closed surfaces have no isometric
embeddings in E3. They do nevertheless admit isometric embeddings in E4; in the easiest case this follows from the
fact that the torus is a product of two circles and each circle can be isometrically embedded in E2.[47]

Spherical geometry
The isometry group of the unit sphere S2 in E3 is
the orthogonal group O(3), with the rotation group
SO(3) as the subgroup of isometries preserving
orientation. It is the direct product of SO(3) with
the antipodal map, sending x to x.[48] The group
SO(3) acts transitively on S2. The stabilizer
subgroup of the unit vector (0,0,1) can be identified
with SO(2), so that S2 = SO(3)/SO(2).

The geodesics between two points on the sphere are


the great circle arcs with these given endpoints. If
the points are not antipodal, there is a unique
shortest geodesic between the points. The geodesics
can also be described group theoretically: each
geodesic through the North pole (0,0,1) is the orbit
of the subgroup of rotations about an axis through
antipodal points on the equator.

A spherical triangle is a geodesic triangle on the A spherical triangle


sphere. It is defined by points A, B, C on the sphere
with sides BC, CA, AB formed from great circle
arcs of length less than . If the lengths of the sides are a, b, c and the angles between the sides , , , then the
spherical cosine law states that

The area of the triangle is given by

Area = + + .
Using stereographic projection from the North pole, the sphere can be identified with the extended complex plane
C {}. The explicit map is given by

Under this correspondence every rotation of S2 corresponds to a Mbius transformation in SU(2), unique up to
sign.[49] With respect to the coordinates (u, v) in the complex plane, the spherical metric becomes[50]

The unit sphere is the unique closed orientable surface with constant curvature +1. The quotient SO(3)/O(2) can be
identified with the real projective plane. It is non-orientable and can be described as the quotient of S2 by the
antipodal map (multiplication by 1). The sphere is simply connected, while the real projective plane has fundamental
group Z2. The finite subgroups of SO(3), corresponding to the finite subgroups of O(2) and the symmetry groups of
the platonic solids, do not act freely on S2, so the corresponding quotients are not 2-manifolds, just orbifolds.

Hyperbolic geometry
Non-Euclidean geometry[51] was first discussed in letters of Gauss, who made extensive
computations at the turn of the nineteenth century which, although privately circulated, he
decided not to put into print. In 1830 Lobachevsky and independently in 1832 Bolyai, the
son of one Gauss' correspondents, published synthetic versions of this new geometry, for
which they were severely criticized. However it was not until 1868 that Beltrami, followed by
Klein in 1871 and Poincar in 1882, gave concrete analytic models for what Klein dubbed
hyperbolic geometry. The four models of 2-dimensional hyperbolic geometry that
emerged were:
Eugenio Beltrami
the Beltrami-Klein model;
(1835-1899)
the Poincar disk;
the Poincar upper half-plane;
the hyperboloid model of Wilhelm Killing in 3-dimensional Minkowski space.
The first model, based on a disk, has the advantage that geodesics are actually line segments
(that is, intersections of Euclidean lines with the open unit disk).The last model has the
advantage that it gives a construction which is completely parallel to that of the unit sphere
in 3-dimensional Euclidean space. Because of their application in complex analysis and
geometry, however, the models of Poincar are the most widely used: they are
interchangeable thanks to the Mbius transformations between the disk and the upper half-
plane. Felix Klein (1849-
1925)
Let

be the Poincar disk in the complex plane with Poincar metric

In polar coordinates (r, ) the metric is given by


The length of a curve :[a,b] D is given by the formula

The group G = SU(1,1) given by Henri Poincar


(1854-1912)

acts transitively by Mbius transformations on D and the stabilizer


subgroup of 0 is the rotation group

The quotient group SU(1,1)/I is the group of orientation-preserving


isometries of D. Any two points z, w in D are joined by a unique geodesic,
given by the portion of the circle or straight line passing through z and w
A hyperbolic triangle in the Poincar
and orthogonal to the boundary circle. The distance between z and w is
disk model
given by

1
In particular d(0,r) = 2 tanh1 r and c(t) = tanh t is the geodesic through 0 along the real axis, parametrized by
2
arclength.

The topology defined by this metric is equivalent to the usual Euclidean topology, although as a metric space (D,d) is
complete.

A hyperbolic triangle is a geodesic triangle for this metric: any three points in D are vertices of a hyperbolic triangle. If
the sides have length a, b, c with corresponding angles , , , then the hyperbolic cosine rule states that

The area of the hyperbolic triangle is given by[52]

Area = .
The unit disk and the upper half-plane

are conformally equivalent by the Mbius transformations


Under this correspondence the action of SL(2,R) by Mbius transformations on H corresponds to that of SU(1,1) on
D. The metric on H becomes

Since lines or circles are preserved under Mbius transformations, geodesics are again described by lines or circles
orthogonal to the real axis.

The unit disk with the Poincar metric is the unique simply connected oriented 2-dimensional Riemannian manifold
with constant curvature 1. Any oriented closed surface M with this property has D as its universal covering space. Its
fundamental group can be identified with a torsion-free concompact subgroup of SU(1,1), in such a way that

In this case is a finitely presented group. The generators and relations are encoded in a geodesically convex
fundamental geodesic polygon in D (or H) corresponding geometrically to closed geodesics on M.

Examples.

the Bolza surface of genus 2;


the Klein quartic of genus 3;
the Macbeath surface of genus 7;
the First Hurwitz triplet of genus 14.

Uniformization
Given an oriented closed surface M with Gaussian curvature K, the metric on M can be changed conformally by
scaling it by a factor e2u. The new Gaussian curvature K is then given by

where is the Laplacian for the original metric. Thus to show that a given surface is conformally equivalent to a
metric with constant curvature K it suffices to solve the following variant of Liouville's equation:

When M has Euler characteristic 0, so is diffeomorphic to a torus, K = 0, so this amounts to solving

By standard elliptic theory, this is possible because the integral of K over M is zero, by the GaussBonnet theorem.[53]

When M has negative Euler characteristic, K = 1, so the equation to be solved is:

Using the continuity of the exponential map on Sobolev space due to Neil Trudinger, this non-linear equation can
always be solved.[54]
Finally in the case of the 2-sphere, K = 1 and the equation becomes:

So far this non-linear equation has not been analysed directly, although classical results such as the Riemann-Roch
theorem imply that it always has a solution.[55] The method of Ricci flow, developed by Richard S. Hamilton, gives
another proof of existence based on non-linear partial differential equations to prove existence.[56] In fact the Ricci
flow on conformal metrics on S2 is defined on functions u(x, t) by

After finite time, Chow showed that K becomes positive; previous results of Hamilton could then be used to show that
K converges to +1.[57] Prior to these results on Ricci flow, Osgood, Phillips & Sarnak (1988) had given an alternative
and technically simpler approach to uniformization based on the flow on Riemannian metrics g defined by log det g.

A simple proof using only elliptic operators discovered in 1988 can be found in Ding (2001). Let G be the Green's
function on S2 satisfying G = 1 + 4P, where P is the point measure at a fixed point P of S2. The equation
v = 2K 2, has a smooth solution v, because the right hand side has integral 0 by the GaussBonnet theorem. Thus
= 2G + v satisfies = 2K away from P. It follows that g1 = eg is a complete metric of constant curvature 0 on
the complement of P, which is therefore isometric to the plane. Composing with stereographic projection, it follows
that there is a smooth function u such that e2ug has Gaussian curvature +1 on the complement of P. The function u
automatically extends to a smooth function on the whole of S2.[58]

Surfaces of non-positive curvature


In a region where the curvature of the surface satisfies K 0, geodesic triangles satisfy the CAT(0) inequalities of
comparison geometry, studied by Cartan, Alexandrov and Toponogov, and considered later from a different point
of view by Bruhat and Tits; thanks to the vision of Gromov, this characterisation of non-positive curvature in terms of
the underlying metric space has had a profound impact on modern geometry and in particular geometric group theory.
Many results known for smooth surfaces and their geodesics, such as Birkhoff's method of constructing geodesics by
his curve-shortening process or van Mangoldt and Hadamard's theorem that a simply connected surface of non-
positive curvature is homeomorphic to the plane, are equally valid in this more general setting.

Alexandrov's comparison inequality


The simplest form of the comparison inequality, first proved for surfaces
by Alexandrov around 1940, states that

The distance between a vertex of a geodesic triangle and the


midpoint of the opposite side is always less than the
corresponding distance in the comparison triangle in the
plane with the same side-lengths.

The inequality follows from the fact that if c(t) describes a geodesic
parametrized by arclength and a is a fixed point, then The median in the comparison
triangle is always longer than the
actual median.
f(t) = d(a,c(t))2 t2
is a convex function, i.e.
Taking geodesic polar coordinates with origin at a so that c(t) = r(t), convexity is equivalent to

Changing to normal coordinates u, v at c(t), this inequality becomes

u2 + H1Hrv2 1,

where (u,v) corresponds to the unit vector (t). This follows from the inequality Hr H, a consequence of the non-
negativity of the derivative of the Wronskian of H and r from SturmLiouville theory.[59]

Existence of geodesics
On a complete curved surface any two points can be joined by a geodesic.
This is a special case of the Hopf-Rinow theorem, which also applies in
higher dimensions. The completeness assumption is automatically fulfilled
for a surface which is embedded as a closed subset of Euclidean space.
However, it is no longer fulfilled if, for example, we remove an isolated
point from a surface. For example, the complement of the origin in the
Euclidean plane is an example of a non-complete surface; in this example
two points which are diametrically opposite across the origin cannot be
joined by a geodesic without leaving the punctured plane).

Von Mangoldt-Hadamard theorem


George Birkhoff (1884-1944)
For closed surfaces of non-positive curvature, von Mangoldt (1881) and
Hadamard (1898) proved that the exponential map at a point is a covering
map, so that the universal covering space of the manifold is E2. This result was generalised to higher dimensions by
Cartan and is usually referred to in this form as the CartanHadamard theorem. For surfaces, this result follows
from three important facts:[60]

The exponential map has non-zero Jacobian everywhere for non-positively curved surfaces, a consequence of
the non-vanishing of Hr.
Every geodesic is infinitely extendible, a result known as the Hopf-Rinow theorem for n-dimensional manifolds. In
two dimensions, if a geodesic tended at infinity towards a point x, a closed disc D centred on a nearby point y
with x removed would be contractible to y along geodesics, a topological impossibility.
Every two points in a homotopy class are connected by a unique geodesic (see above).

Riemannian connection and parallel transport


The classical approach of Gauss to the differential geometry of surfaces was the standard elementary approach[61]
which predated the emergence of the concepts of Riemannian manifold initiated by Bernhard Riemann in the mid-
nineteenth century and of connection developed by Tullio Levi-Civita, lie Cartan and Hermann Weyl in the early
twentieth century. The notion of connection, covariant derivative and parallel transport gave a more conceptual and
uniform way of understanding curvature, which not only allowed generalisations to higher dimensional manifolds but
also provided an important tool for defining new geometric invariants, called characteristic classes.[62] The approach
using covariant derivatives and connections is nowadays the one adopted in more advanced textbooks.[63]

Covariant derivative
Connections on a surface can be defined from various equivalent but equally important
points of view. The Riemannian connection or Levi-Civita connection[20] is perhaps
most easily understood in terms of lifting vector fields, considered as first order
differential operators acting on functions on the manifold, to differential operators on the
tangent bundle or frame bundle. In the case of an embedded surface, the lift to an operator
on vector fields, called the covariant derivative, is very simply described in terms of
orthogonal projection. Indeed, a vector field on a surface embedded in R3 can be regarded
as a function from the surface into R3. Another vector field acts as a differential operator
component-wise. The resulting vector field will not be tangent to the surface, but this can
be corrected taking its orthogonal projection onto the tangent space at each point of the Tullio Levi-Civita
surface. As Ricci and Levi-Civita realised at the turn of the twentieth century, this process (1873-1941)
depends only on the metric and can be locally expressed in terms of the Christoffel
symbols.

Parallel transport
Parallel transport of tangent vectors along a curve in the surface was the
next major advance in the subject, due to Levi-Civita.[20] It is related to the
earlier notion of covariant derivative, because it is the monodromy of the
ordinary differential equation on the curve defined by the covariant derivative
with respect to the velocity vector of the curve. Parallel transport along
geodesics, the "straight lines" of the surface, can also easily be described
directly. A vector in the tangent plane is transported along a geodesic as the
unique vector field with constant length and making a constant angle with the
velocity vector of the geodesic. For a general curve, this process has to be
modified using the geodesic curvature, which measures how far the curve Parallel transport of a vector
departs from being a geodesic.[5] around a geodesic triangle on
the sphere. The length of the
A vector field v(t) along a unit speed curve c(t), with geodesic curvature kg(t), transported vector and the angle
is said to be parallel along the curve if it makes with each side remain
constant.
it has constant length
the angle (t) that it makes with the velocity vector (t) satisfies

This recaptures the rule for parallel transport along a geodesic or piecewise geodesic curve, because in that case
kg = 0, so that the angle (t) should remain constant on any geodesic segment. The existence of parallel transport
follows because (t) can be computed as the integral of the geodesic curvature. Since it therefore depends continuously
on the L2 norm of kg, it follows that parallel transport for an arbitrary curve can be obtained as the limit of the parallel
transport on approximating piecewise geodesic curves.[64]

The connection can thus be described in terms of lifting paths in the manifold to paths in the tangent or orthonormal
frame bundle, thus formalising the classical theory of the "moving frame", favoured by French authors.[65] Lifts of
loops about a point give rise to the holonomy group at that point. The Gaussian curvature at a point can be recovered
from parallel transport around increasingly small loops at the point. Equivalently curvature can be calculated directly
at an infinitesimal level in terms of Lie brackets of lifted vector fields.

Connection 1-form
The approach of Cartan and Weyl, using connection 1-forms on the frame
bundle of M, gives a third way to understand the Riemannian connection.
They noticed that parallel transport dictates that a path in the surface be
lifted to a path in the frame bundle so that its tangent vectors lie in a
special subspace of codimension one in the three-dimensional tangent
space of the frame bundle. The projection onto this subspace is defined by
a differential 1-form on the orthonormal frame bundle, the connection
form. This enabled the curvature properties of the surface to be encoded
in differential forms on the frame bundle and formulas involving their
exterior derivatives.

This approach is particularly simple for an embedded surface. Thanks to a


result of Kobayashi (1956), the connection 1-form on a surface embedded
in Euclidean space E3 is just the pullback under the Gauss map of the
connection 1-form on S2.[66] Using the identification of S2 with the
homogeneous space SO(3)/SO(2), the connection 1-form is just a
component of the MaurerCartan 1-form on SO(3).[67] lie Cartan in 1904

Global differential geometry of


surfaces
Although the characterisation of curvature involves only the local geometry of a surface, there are important global
aspects such as the GaussBonnet theorem, the uniformization theorem, the von Mangoldt-Hadamard theorem, and
the embeddability theorem. There are other important aspects of the global geometry of surfaces.[68] These include:

Injectivity radius, defined as the largest r such that two points at a distance less than r are joined by a unique
geodesic. Wilhelm Klingenberg proved in 1959 that the injectivity radius of a closed surface is bounded below by

the minimum of = and the length of its smallest closed geodesic. This improved a theorem of Bonnet
sup K
who showed in 1855 that the diameter of a closed surface of positive Gaussian curvature is always bounded
above by ; in other words a geodesic realising the metric distance between two points cannot have length
greater than .
Rigidity. In 1927 Cohn-Vossen proved that two ovaloids closed surfaces with positive Gaussian curvature
that are isometric are necessarily congruent by an isometry of E3. Moreover, a closed embedded surface with
positive Gaussian curvature and constant mean curvature is necessarily a sphere; likewise a closed embedded
surface of constant Gaussian curvature must be a sphere (Liebmann 1899). Heinz Hopf showed in 1950 that a
closed embedded surface with constant mean curvature and genus 0, i.e. homeomorphic to a sphere, is
necessarily a sphere; five years later Alexandrov removed the topological assumption. In the 1980s, Wente
constructed immersed tori of constant mean curvature in Euclidean 3-space.
Carathodory conjecture: This conjecture states that a closed convex three times differentiable surface admits
at least two umbilic points. The first work on this conjecture was in 1924 by Hans Hamburger, who noted that it
follows from the following stronger claim: the half-integer valued index of the principal curvature foliation of an
isolated umbilic is at most one.
Zero Gaussian curvature: a complete surface in E3 with zero Gaussian curvature must be a cylinder or a plane.
Hilbert's theorem (1901): no complete surface with constant negative curvature can be immersed isometrically in
E3.
The Willmore conjecture. This conjecture states that the integral of the square of the mean curvature of a torus
immersed in E3 should be bounded below by 22. It is known that the integral is Moebius invariant. It was solved
in 2012 by Fernando Cod Marques and Andr Neves.[69]
Isoperimetric inequalities. In 1939 Schmidt proved that the classical isoperimetric inequality for curves in the
Euclidean plane is also valid on the sphere or in the hyperbolic plane: namely he showed that among all closed
curves bounding a domain of fixed area, the perimeter is minimized by when the curve is a circle for the metric. In
one dimension higher, it is known that among all closed surfaces in E3 arising as the boundary of a bounded
domain of unit volume, the surface area is minimized for a Euclidean ball.
Systolic inequalities for curves on surfaces. Given a closed surface,
its systole is defined to be the smallest length of any non-contractible
closed curve on the surface. In 1949 Loewner proved a torus inequality
for metrics on the torus, namely that the area of the torus over the square
3
of its systole is bounded below by , with equality in the flat (constant
2
curvature) case. A similar result is given by Pu's inequality for the real
2
projective plane from 1952, with a lower bound of also attained in the

constant curvature case. For the Klein bottle, Blatter and Bavard later
8
obtained a lower bound of . For a closed surface of genus g, Hebda

1
and Burago showed that the ratio is bounded below by . Three years Shortest loop on a torus
2
1
later Mikhail Gromov found a lower bound given by a constant times g 2,
although this is not optimal. Asymptotically sharp upper and lower bounds
g
given by constant times are due to Gromov and Buser-Sarnak, and can be found in Katz (2007). There is
(log g)2
also a version for metrics on the sphere, taking for the systole the length of the smallest closed geodesic. Gromov
1 1
conjectured a lower bound of in 1980: the best result so far is the lower bound of obtained by Regina
23 8
Rotman in 2006.[70]

Reading guide
One of the most comprehensive introductory surveys of the subject, charting the historical development from before
Gauss to modern times, is by Berger (2004). Accounts of the classical theory are given in Eisenhart (2004), Kreyszig
(1991) and Struik (1988); the more modern copiously illustrated undergraduate textbooks by Gray, Abbena & Salamon
(2006), Pressley (2001) and Wilson (2008) might be found more accessible. An accessible account of the classical
theory can be found in Hilbert & Cohn-Vossen (1952). More sophisticated graduate-level treatments using the
Riemannian connection on a surface can be found in Singer & Thorpe (1967), do Carmo (1976) and O'Neill (1997).

See also
Zoll surface

Notes
1. Euler 1760
2. Euler 1771
3. Gauss 1825 and 1827
4. This is the final position into which a rubber band stretched between two fixed points on the surface would fall.
5. Berger 2004
6. Eisenhart 2004, p. 123
7. Singer & Thorpe 1967, p. 223
8. do Carmo 1976, pp. 161162
9. Eisenhart 2004, pp. 228229
10. Eisenhart 2004, pp. 241250; do Carmo 1976, pp. 188197.
11. do Carmo 1976, p. 194.
12. Eisenhart 2004, pp. 6165.
13. Eisenhart 2004
14. Eisenhart 2004, pp. 250269; do Carmo 1976, pp. 197213.
15. Douglas' solution is described in Courant (1950).
16. Eisenhart 2004, pp. 270291; O'Neill, pp. 249251; Hilbert & Cohn-Vossen 1952.
17. O'Neill, pp. 249251; do Carmo, pp. 168170; Gray, Abbena & Salamon 2006.
18. Stillwell 1996, pp. 15.
19. Wilson 2008.
20. Levi-Civita 1917.
21. Eisenhart 2004, pp. 114115; Pressley 2001, pp. 123124; Wilson 2008, pp. 123124.
22. Eisenhart 2004, p. 156
23. O'Neill 1997, p. 257
24. do Carmo 1976, pp. 309314
25. O'Neill 1997, pp. 195216; do Carmo 1976, pp. 134153; Singer & Thorpe 1967, pp. 216224.
26. Gray, Abbena & Salamon 2006, p. 386.
27. Note that in some more recent texts the symmetric bilinear form on the right hand side is referred to as the
second fundamental form; however, it does not in general correspond to the classically defined second
fundamental form.
28. Gray, Abbena & Salamon 2006, p. 394.
29. Berger 2004; Wilson 2008; Milnor 1963.
30. Eisenhart 2002, p. 131; Berger 2004, p. 39; do Carmo 1976, p. 248; O'Neill 1997, p. 237
31. Han & Hong 2006
32. Eisenhart 2004; Taylor 1996a, Appendix C.
33. Eisenhart 2004; Berger 2004.
34. doCarmo 1976, p. 357
35. Milnor 1963
36. Wilson 2008
37. do Carmo 1976, pp. 303305
38. O'Neill 1997, p. 395
39. Helgason 1978, p. 92
40. O'Niell 1997, p. 286
41. do Carmo 1976, p. 227
42. Osserman 2002, pp. 3132
43. Singer & Thorpe 1967; Garsia, Adriano M. (1961), "An imbedding of closed Riemann surfaces in Euclidean
space", Comment. Math. Helv., 35: 93110, doi:10.1007/BF02567009 (https://doi.org/10.1007%2FBF02567009)
44. Imayoshi & Taniguchi 1992, pp. 4749
45. Berger 1977; Taylor 1996.
46. Wilson 2008, pp. 123, Chapter I, Euclidean geometry.
47. do Carmo 1976.
48. Wilson 2008, pp. 2549, Chapter II, Spherical geometry.
49. Wilson 2008, Chapter 2.
50. Eisenhart 2004, p. 110.
51. Stillwell 1990; Bonola, Carslaw & Enriques 1955.
52. Wilson 2008, Chapter 5.
53. Taylor 1996b, p. 107; Berger 1977, pp. 341343.
54. Berger 1977, pp. 222225; Taylor 1996b, pp. 101108.
55. Taylor 1996b
56. Chow 1991
57. Chen, Lu & Tian (2006) pointed out and corrected a missing step in the approach of Hamilton and Chow; see also
Andrews & Bryan (2009).
58. This follows by an argument involving a theorem of Sacks & Uhlenbeck (1981) on removable singularities of
harmonic maps of finite energy.
59. Berger 2004; Jost, Jrgen (1997), Nonpositive curvature: geometric and analytic aspects, Lectures in
Mathematics, ETH Zurich, Birkhuser, ISBN 0-8176-5736-3
60. do Carmo 1976; Berger 2004.
61. Eisenhart 2004; Kreyszig 1991; Berger 2004; Wilson 2008.
62. Kobayashi & Nomizu 1969, Chapter XII.
63. do Carmo 1976; O'Neill 1997; Singer & Thorpe 1967.
64. Arnold 1989, pp. 301306, Appendix I.; Berger 2004, pp. 263264.
65. Darboux & 1887,1889,1896
66. Kobayashi & Nomizu 1969
67. Ivey & Landsberg 2003.
68. Berger 2004, pp. 145161; do Carmo 1976; Chern 1967; Hopf 1989.
69. Cod Marques, Fernando; Neves, Andr, Min-Max theory and the Willmore conjecture, Annals of Mathematics
179 (2014), 683-782 https://dx.doi.org/10.4007/annals.2014.179.2.6
70. Rotman, R. (2006) "The length of a shortest closed geodesic and the area of a 2-dimensional sphere", Proc.
Amer. Math. Soc. 134: 3041-3047. Previous lower bounds had been obtained by Croke, Rotman-Nabutovsky and
Sabourau.

References
Aleksandrov, A.D.; Zalgaller, V.A. (1967), Instrinsic Geometry of Surfaces, Translations of Mathematical
Monographs, 15, American Mathematical Society
Andrews, Ben; Bryan, Paul (2010), "Curvature bounds by isoperimetric comparison for normalized Ricci flow on
the two-sphere", Calc. Var. Partial Differential Equations, 39: 419428
Arnold, V.I. (1989), Mathematical methods of classical mechanics., Graduate Texts in Mathematics, 60 (2nd ed.),
New York: Springer-Verlag, ISBN 0-387-90314-3; translated from the Russian by K. Vogtmann and A. Weinstein.
Berger, Marcel (2004), A Panoramic View of Riemannian Geometry, Springer-Verlag, ISBN 3-540-65317-1
Berger, Melvyn S. (1977), Nonlinearity and Functional Analysis, Academic Press, ISBN 0-12-090350-4
Bonola, Roberto; Carslaw, H. S.; Enriques, F. (1955), Non-Euclidean Geometry: A Critical and Historical Study of
Its Development, Dover, ISBN 0-486-60027-0
Brendle, Simon (2010), Ricci flow and the sphere theorem, Graduate Studies in Mathematics, 111, American
Mathematical Society, ISBN 978-0-8218-4938-5
Cartan, lie (1983), Geometry of Riemannian Spaces (https://books.google.com/?id=-YvvVfQ7xz4C&printsec=fro
ntcover&dq=geometry+of+riemannian+spaces+cartan), Math Sci Press, ISBN 978-0-915692-34-7; translated
from 2nd edition of Leons sur la gomtrie des espaces de Riemann (1951) by James Glazebrook.
Cartan, lie (2001), Riemannian Geometry in an Orthogonal Frame (from lectures delivered by Cartan at the
Sorbonne in 1926-27) (http://ebooks.worldscinet.com/mathematics/9789812799715/preserved-docs/9789812799
715.pdf) (PDF), World Scientific, ISBN 981-02-4746-X; translated from Russian by V. V. Goldberg with a foreword
by S. S. Chern.
Chen, Xiuxiong; Lu, Peng; Tian, Gang (2006), "A note on uniformization of Riemann surfaces by Ricci flow", Proc.
AMS, 134 (11): 33913393, doi:10.1090/S0002-9939-06-08360-2 (https://doi.org/10.1090%2FS0002-9939-06-08
360-2)
Chern, S. S. (1967), Curves and Surfaces in Euclidean Spaces, MAA Studies in Mathematics, Mathematical
Association of America
Chow, B. (1991), "The Ricci flow on a 2-sphere", J. Diff. Geom., 33: 325334
Courant, Richard (1950), Dirichlet's Principle, Conformal Mapping and Minimal Surfaces, John Wiley & Sons,
ISBN 0-486-44552-6
Darboux, Gaston (1887,1889,1896), Leons sur la thorie gnrale des surfaces: Volume I (http://www.hti.umich.
edu/cgi/t/text/text-idx?c=umhistmath;idno=ABV4153.0001.001), Volume II (http://www.hti.umich.edu/cgi/t/text/text-
idx?c=umhistmath;idno=ABV4153.0002.001), Volume III (http://www.hti.umich.edu/cgi/t/text/text-idx?c=umhistmat
h;idno=ABV4153.0003.001), Volume IV (http://www.hti.umich.edu/cgi/t/text/text-idx?c=umhistmath;idno=ABV415
3.0004.001), Gauthier-Villars Check date values in: |date= (help); External link in |title= (help)
Ding, W. (2001), "A proof of the uniformization theorem on S2", J. Partial Differential Equations, 14: 247250
do Carmo, Manfredo P. (1976), Differential Geometry of Curves and Surfaces, Prentice-Hall, ISBN 0-13-212589-7
Eisenhart, Luther P. (2004), A Treatise on the Differential Geometry of Curves and Surfaces, Dover, ISBN 0-486-
43820-1 Full 1909 text (https://archive.org/details/treatonthediffer00eiserich) (now out of copyright)
Eisenhart, Luther P. (1947), An Introduction to Differential Geometry with Use of the Tensor Calculus, Princeton
Mathematical Series, 3, Princeton University Press, ISBN 1-4437-2293-6
Euler, Leonhard (1760), "Recherches sur la courbure des surfaces" (http://math.dartmouth.edu/~euler/pages/E33
3.html), Mmoires de l'acadmie des sciences de Berlin (published 1767), 16: 119143.
34 (https://worldwide.espacenet.com/textdoc?DB=EPODOC&IDX=), "De solidis quorum superficiem in planum
explicare licet", published 1772.
Gauss, Carl Friedrich (1825 and 1827), General Investigations of Curved Surfaces (http://quod.lib.umich.edu/cgi/t/
text/text-idx?c=umhistmath;idno=ABR1255), New York: Raven Press (published 1965), ISBN 0-486-44645-X
Check date values in: |date= (help) translated by A.M.Hiltebeitel and J.C.Morehead; "Disquisitiones generales
circa superficies curvas" (http://www-gdz.sub.uni-goettingen.de/cgi-bin/digbib.cgi?PPN35283028X_0006_2NS),
Commentationes Societatis Regiae Scientiarum Gottingesis Recentiores Vol. VI (1827), pp. 99146.
Gray, Alfred; Abbena, Elsa; Salamon, Simon (2006), Modern Differential Geometry of Curves And Surfaces With
Mathematica, CRC Press, ISBN 1-58488-448-7
Han, Qing; Hong, Jia-Xing (2006), Isometric Embedding of Riemannian Manifolds in Euclidean Spaces, American
Mathematical Society, ISBN 0-8218-4071-1

Helgason, Sigurdur (1978), Differential Geometry Lie Groups, and Symmetric Spaces, Academic Press, New
York, ISBN 0-12-338460-5
Hilbert, David; Cohn-Vossen, Stephan (1952), Geometry and the Imagination (2nd ed.), New York: Chelsea,
ISBN 978-0-8284-1087-8
Hopf, Heinz (1989), Lectures on Differential Geometry in the Large, Lecture Notes in Mathematics, 1000,
Springer-Verlag, ISBN 3-540-51497-X
Imayoshi, Y.; Taniguchi, M. (1992), An Introduction to Techmller spaces, Springer-Verlag, ISBN 0-387-70088-9
Ivey, Thomas A.; Landsberg, J.M. (2003), Cartan for Beginners: Differential Geometry via Moving Frames and
Exterior Systems, Graduate Studies in Mathematics, 61, American Mathematical Society, ISBN 0-8218-3375-8
Jacobowitz, Howard (1972), "Local Isometric Embeddings of Surfaces into Euclidean Four Space" (http://www.iu
mj.indiana.edu/IUMJ/FULLTEXT/1972/21/21019), Indiana Univ. Math. J., 21 (3): 249254,
doi:10.1512/iumj.1971.21.21019 (https://doi.org/10.1512%2Fiumj.1971.21.21019)
Katz, Mikhail G. (2007), Systolic geometry and topology, Mathematical Surveys and Monographs, 137, American
Mathematical Society, ISBN 0-8218-4177-7
Kobayashi, Shoshichi (1956), "Induced connections and imbedded Riemannian space", Nagoya Math. J., 10: 15
25
Kobayashi, Shoshichi (1957), "Theory of connections", Annali di Matematica Pura ed Applicata, Series 4, 43:
119194, doi:10.1007/BF02411907 (https://doi.org/10.1007%2FBF02411907),
Kobayashi, Shoshichi; Nomizu, Katsumi (1963), Foundations of Differential Geometry, Vol. I, Wiley Interscience,
ISBN 0-470-49648-7
Kobayashi, Shoshichi; Nomizu, Katsumi (1969), Foundations of Differential Geometry, Vol. II, Wiley Interscience,
ISBN 0-470-49648-7
Kreyszig, Erwin (1991), Differential Geometry, Dover, ISBN 0-486-66721-9
Khnel, Wolfgang (2006), Differential Geometry: Curves - Surfaces - Manifolds, American Mathematical Society,
ISBN 0-8218-3988-8
Levi-Civita, Tullio (1917), "Nozione di parallelismo in una variet qualunque", Rend. Circ. Mat. Palermo, 42: 173
205, doi:10.1007/BF03014898 (https://doi.org/10.1007%2FBF03014898)
O'Neill, Barrett (1997), Elementary Differential Geometry, Academic Press, ISBN 0-12-526750-9
Osgood, B.; Phillips, R.; Sarnak, P. (1988), "Extremals of determinants of Laplacians", J. Funct. Anal., 80: 148
211
Osserman, Robert (2002), A Survey of Minimal Surfaces, Dover, ISBN 0-486-49514-0
Ian R. Porteous (2001) Geometric Differentiation: for the intelligence of curves and surfaces, Cambridge
University Press ISBN 0-521-00264-8.
Poznjak, E.G. (1973), "Isometric imbedding of two-dimensional Riemannian metrics in Euclidean spaces",
Russian Math. Surveys, 28 (4): 4777, doi:10.1070/RM1973v028n04ABEH001591 (https://doi.org/10.1070%2FR
M1973v028n04ABEH001591)
Pressley, Andrew (2001), Elementary Differential Geometry, Springer Undergraduate Mathematics Series,
Springer-Verlag, ISBN 1-85233-152-6
Sacks, J.; Uhlenbeck, Karen (1981), "The existence of minimal immersions of 2-spheres", Ann. of Math., 112 (1):
124, doi:10.2307/1971131 (https://doi.org/10.2307%2F1971131), JSTOR 1971131 (https://www.jstor.org/stable/1
971131)
Singer, Isadore M.; Thorpe, John A. (1967), Lecture Notes on Elementary Topology and Geometry, Springer-
Verlag, ISBN 0-387-90202-3
Stillwell, John (1996), Sources of Hyperbolic Geometry, American Mathematical Society, ISBN 0-8218-0558-4
Struik, Dirk Jan (1988), Lectures on classical differential geometry: Second Edition, Dover, ISBN 0-486-65609-8
Taylor, Michael E. (1996a), Partial Differential Equations II: Qualitative Studies of Linear Equations, Springer-
Verlag, ISBN 1-4419-7051-7
Taylor, Michael E. (1996b), Partial Differential Equations III: Nonlinear equations, Springer-Verlag, ISBN 1-4419-
7048-7
Toponogov, Victor A. (2005), Differential Geometry of Curves and Surfaces: A Concise Guide, Springer-Verlag,
ISBN 0-8176-4384-2
Valiron, Georges (1986), The Classical Differential Geometry of Curves and Surfaces, Math Sci Press, ISBN 0-
915692-39-2 Full text of book (https://books.google.com/books?id=IQXstKvWsHMC&printsec=frontcover&dq=vali
ron+surfaces&source=gbs_summary_r&cad=0)
Wilson, Pelham (2008), Curved Space: From Classical Geometries to Elementary Differential Geometry,
Cambridge University Press, ISBN 978-0-521-71390-0

Retrieved from "https://en.wikipedia.org/w/index.php?title=Differential_geometry_of_surfaces&oldid=811795364"

This page was last edited on 24 November 2017, at 01:07.

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using
this site, you agree to the Terms of Use and Privacy Policy. Wikipedia is a registered trademark of the Wikimedia
Foundation, Inc., a non-profit organization.

You might also like