You are on page 1of 68

Lecture Notes

Geodesy and Geodynamics


Nico Sneeuw
Geodatisches Institut
Universitat Stuttgart
February 13, 2006

c Nico Sneeuw, 2005, 2006



These are lecture notes in progress. Please contact me (n.sneeuw@gis.uni-stuttgart.de)
for remarks, errors, suggestions, etc.

Contents
1 Introduction
1.1 Physical Geodesy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Links to Earth sciences . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3 Applications in engineering . . . . . . . . . . . . . . . . . . . . . . . . . .

5
5
5
7

2 Spherical and ellipsoidal geometry


2.1 Approximation 1: the sphere . . . . . . . . . .
2.1.1 From planar to spherical trigonometry .
2.1.2 The direct problem . . . . . . . . . . . .
2.1.3 The inverse problem . . . . . . . . . . .
2.2 Approximation 2: the ellipsoid . . . . . . . . .
2.2.1 Curvature . . . . . . . . . . . . . . . . .
2.2.2 The direct and inverse geodetic problem

. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
on the

. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
ellipsoid

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

9
9
10
11
12
14
17
20

3 Gravitation
3.1 Newtonian gravitation . . . . . . . .
3.1.1 Vectorial attraction of a point
3.1.2 Gravitational potential . . . .
3.1.3 Superpositiondiscrete . . .
3.1.4 Superpositioncontinuous .
3.2 Ideal solids . . . . . . . . . . . . . .
3.2.1 Solid homogeneous sphere . .
3.2.2 Spherical shell . . . . . . . .
3.2.3 Solid homogeneous cylinder .
3.3 Tides . . . . . . . . . . . . . . . . . .
3.4 Summary . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

21
21
22
23
24
24
25
26
30
32
36
37

4 Rotation
4.1 Kinematics: acceleration in a rotating frame . . .
4.2 Dynamics: precession, nutation, polar motion . .
4.3 Geometry: defining the inertial reference system
4.3.1 Inertial space . . . . . . . . . . . . . . . .
4.3.2 Transformations . . . . . . . . . . . . . .
4.3.3 Conventional inertial reference system . .
4.3.4 Overview . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

38
38
42
45
45
45
47
49

. . . .
mass
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

Contents
5 Gravity
51
5.1 Gravity attraction and potential . . . . . . . . . . . . . . . . . . . . . . . 51
6 The
6.1
6.2
6.3

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

56
57
59
61
61
62

7 Height Systems
7.1 Height Systems . . . . . . . . . . . . . . . . . . . .
7.1.1 Raw levelled heights . . . . . . . . . . . . .
7.1.2 Geopotential numbers . . . . . . . . . . . .
7.1.3 Dynamic height . . . . . . . . . . . . . . . .
7.1.4 Orthometric height . . . . . . . . . . . . . .
7.1.5 Normal heights . . . . . . . . . . . . . . . .
7.2 Height computations and corrections . . . . . . . .
7.2.1 Dynamic heights . . . . . . . . . . . . . . .
7.2.2 Orthometric heights . . . . . . . . . . . . .
7.2.3 Normal heights . . . . . . . . . . . . . . . .
7.3 Normal vs. true heights . . . . . . . . . . . . . . .
7.4 Vertical reference systems in use in North America

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

64
64
64
64
64
65
65
66
66
66
67
67
68

normal field
Normal potential . . . .
Normal gravity . . . . .
Adopted normal gravity
6.3.1 Formulae . . . .
6.3.2 GRS80 constants

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

1 Introduction
1.1 Physical Geodesy
Geodesy aims at the determination of the geometrical and physical shape of the Earth
and its orientation in space. The branch of geodesy that is concerned with determining
the physical shape of the Earth is called physical geodesy. It does interact strongly with
the other branches, though, as will be seen later.
Physical geodesy is different from other geomatics disciplines in that it is concerned with
field quantities: the scalar potential field or the vectorial gravity and gravitational fields.
These are continuous quantities, as opposed to point fields, networks, pixels, etc., which
are discrete by nature.
Gravity field theory uses a number of tools from mathematics and physics:
Newtonian gravitation theory (relativity is not required for now)
Potential theory
Vector calculus
Special functions (Legendre)
Partial differential equations
Boundary value problems
Signal processing
Gravity field theory is interacting with many other disciplines. A few examples may
clarify the importance of physical geodesy to those disciplines. The Earth sciences
disciplines are rather operating on a global scale, whereas the engineering applications
are more local. This distinction is not fundamental, though.

1.2 Links to Earth sciences


Oceanography. The Earths gravity field determines the geoid, which is the equipotential surface at mean sea level. If the oceans would be at restno waves, no currents,
no tidesthe ocean surface would coincide with the geoid. In reality it deviates by
up to 1 m. The difference is called sea surface topography. It reflects the dynamical
equilibrium in the oceans. Only large scale currents can sustain these deviations.
The sea surface itself can be accurately measured by radar altimeter satellites. If the

1 Introduction
geoid would be known up to the same accuracy, the sea surface topography and consequently the global ocean circulation could be determined. The problem is the insufficient
knowledge of the marine geoid.
Geophysics. The Earths gravity field reflects the internal mass distribution, the determination of which is one of the tasks of geophysics. By itself gravity field knowledge
is insufficient to recover this distribution. A given gravity field can be produced by an
infinity of mass distributions. Nevertheless, gravity is is an important constraint, which
is used together with seismic and other data.
As an example, consider the gravity field over a volcanic island like Hawaii. A volcano by
itself represents a geophysical anomaly already, which will have a gravitational signature.
Over geologic time scales, a huge volcanic mass is piled up on the ocean sphere. This
will cause a bending of the ocean floor. Geometrically speaking one would have a cone
in a bowl. This bowl is likely to be filled with sediment. Moreover the mass load will
be supported by buoyant forces within the mantle. This process is called isostasy. The
gravity signal of this whole mass configuration carries clues to the density structure
below the surface.
Geology. Different geological formations have different density structures and hence
different gravity signals. One interesting example of this is the Chicxulub crater, partially
on the Yucatan peninsula (Mexico) and partially in the Gulf of Mexico. This crater
with a diameter of 180 km was caused by a meteorite impact, which occurred at the K-T
boundary (cretaceous-tertiary) some 66 million years ago. This impact is thought to
have caused the extinction of dinosaurs. The Chicxulub crater was discovered by careful
analysis of gravity data.
Hydrology. Minute changes in the gravity field over timeafter correcting for other
time-variable effects like tides or atmospheric loadingcan be attributed to changes in
hydrological parameters: soil moisture, water table, snow load. For static gravimetry
these are usually nuisance effects. Nowadays, with precise satellite techniques, hydrology
is one of the main aims of spaceborne gravimetry. Despite a low spatial resolution, the
results of satellite gravity missions may be used to constrain basin-scale hydrological
parameters.
Glaciology and sea level. The behaviour of the Earths ice masses is a critical indicator
of global climate change and global sea level behaviour. Thus, monitoring of the melting
of the Greenland and Antarctica ice caps is an important issue. The ice caps are huge
mass loads, sitting on the Earths crust, which will necessarily be depressed. Melting
causes a rebound of the crust. This process is still going on since the last Ice Age, but
there is also an instant effect from melting taking place right now. The change in surface

1.3 Applications in engineering


ice contains a direct gravitational component and an effect, due to the uplift. Therefore,
precise gravity measurements carry information on ice melting and consequently on sea
level rise.

1.3 Applications in engineering


Geophysical prospecting. Since gravity contains information on the subsurface density
structure, gravimetry is a standard tool in the oil and gas industry (and other mineral
resources for that matter). It will always be used together with seismic profiling, test
drilling and magnetometry. The advantages of gravimetry over these other techniques
are:
relatively inexpensive,
non destructive (one can easily measure inside buildings),
compact equipment, e.g. for borehole measurements
Gravimetry is used to localize salt domes or fractures in layers, to estimate depth, and
in general to get a first idea of the subsurface structure.
Geotechnical Engineering. In order to gain knowledge about the subsurface structure,
gravimetry is a valuable tool for certain geotechnical (civil) engineering projects. One
can think of determining the depth-to-bedrock for the layout of a tunnel. Or making
sure no subsurface voids exist below the planned building site of a nuclear power plant.
For examples, see the (micro-)gravity case histories and applications on:
http://www.geop.ubc.ca/ubcgif/casehist/index.html, or
http://www.esci.keele.ac.uk/geophysics/Research/Gravity/.
Geomatics Engineering. Most surveying observables are related to the gravity field.
i) After leveling a theodolite or a total station, its vertical axis is automatically aligned with the local gravity vector. Thus all measurements with these
instruments are referenced to the gravity fieldthey are in a local astronomic
frame. To convert them to a geodetic frame the deflection of the vertical (,)
and the perturbation in azimuth (A) must be known.
ii) The line of sight of a level is tangent to the local equipotential surface. So levelled height differences are really physical height differences. The basic quantity
of physical heights are the potentials or the potential differences. To obtain precise height differences one should also use a gravimeter:
W =

g dx =

gdh

gi hi .

1 Introduction
The hi are the levelled height increments. Using gravity measurements gi
along the way gives a geopotential difference, which can be transformed into a
physical height difference, for instance an orthometric height difference.
iii) GPS positioning is a geometric techniques. The geometric gps heights are
related to physically meaningful heights through the geoid or the quasi-geoid:
h = H + N = orthometric height + geoid height,
h = H n + = normal height + quasi-geoid height.
In geomatics engineering, gps measurements are usually made over a certain
baseline and processed in differential mode. In that case, the above two formulas
become h = N + H, etc. The geoid difference between the baselines
endpoints must therefore be known.
= a, which is integrated twice to
iv) The basic equation of inertial surveying is x
provide the trajectory x(t). The equation says that the kinematic acceleration
equals the specific force vector a: the sum of all forces (per unit mass) acting
on a proof mass). An inertial measurement unit, though, measures the sum of
kinematic acceleration and gravitation. Thus the gravitational field must be
corrected for, before performing the integration.

2 Spherical and ellipsoidal geometry


The Earths surface is a complicated manifold. For many purposes in surveying, navigation and several geosciences, a spherical description is more than sufficient. With a
flattening in the order of 103 a spherical approximation implies errors less than 1 %.
For geodetic applications in which this error level is unacceptable, an ellipsoid of revolution is used as a higher quality approximation. This chapter provides tools to perform
calculations on these surfaces.

2.1 Approximation 1: the sphere


Remark 2.1 In this section, the symbol will be used for the geocentric latitude.
The sphere can be described in a number of ways, see fig. 2.1:
geometrically as the set of points with constant distance (or radius) to a focal point at
the centre, leading to the following algebraic formulation.
algebraically (implicit) x2 + y 2 + z 2 = R2 =

x2
R2

y2
R2

z2
R2

=1

parametrically (explicit)

r cos cos
x
r = x2 + y 2 + z 2


y
y = r cos sin = arctan x

z
r sin
z
= arcsin r

Figure 2.1: Spherical geometry.


read for )

(please

2 Spherical and ellipsoidal geometry

2.1.1 From planar to spherical trigonometry

C
C

a
B

Figure 2.2: Planar and spherical triangle.

Grokreis

When going from the plane to the sphere many trigonometric relationships between
angles and sides are similar. One must be careful, though. In plane trigonometry,
triangle sides are line segments, measured in linear units. On the sphere, however, sides
are great circle segments, or rather angles, expressed in angular units. They may be
converted to linear units, e.g., by sa = aR, with R the spherical radius. The following
relationships existmostly in parallelbetween planar and spherical trigonometry:

angles:
area:

planar

spherical

+ + = 180

+ + = 180 +

2s = a + b + c

2s = a + b + c

A=

s(s a)(s b)(s c)

(Herons formula)

sine:
cosine:

sin
sin
sin
=
=
a
b
c
2
2
2
a = b + c 2bc cos

tan 41 =

sb
sc
tan 2s tan sa
2 tan 2 tan 2

(lHuiliers formula)
A = R2
sin
sin
sin
=
=
sin a
sin b
sin c
cos a = cos b cos c + sin b sin c cos

Further cosine formulas and sine-cosine formulas are obtained by cyclic permutation
a b c a ... and ...
The quantity is called the spherical excess. According to the above formula, the sum
of angles in a spherical triangle is more than 180 . How much more, depends on the area
of the triangle. The formula A = R2 actually tells us that is the solid (geo-)centric
angle, subtended by the spherical triangle. The unit of a solid angle is steradian.

10

2.1 Approximation 1: the sphere

NP
12
90-2
90-1
A*21
A12
P2

P2
A21

12

tor

a
equ

12

Figure 2.3:
triangle.

The

polar

spherical

Remark 2.2 Consider the extreme spherical triangle of the following 3 points: Northpole, intersection of Greenwich meridian and equator, and the point on the equator at
90 longitude. All of the angles in this triangle are right angles. Thus + + = 270 ,
i.e. = 90 .
Exercise 2.1 Determine the sides of the triangle in remark 2.2 and check the validity of
all above spherical trigonometric formulas.
The discussion of the direct and inverse problems in the following sections is based on
the so-called polar spherical triangle, see fig. 2.3.

2.1.2 The direct problem


The direct problem is defined as the following initial value problem:

Anfangswertproblem

Given: 1 and 1 of the first point


12 and A12 between the first and second point
Find: 2 and 2 of the second point
and the inverse azimuth A21
Determination of 2
From the spherical cosine formula:
cos (90 2 ) = cos(90 1 ) cos 12 + sin(90 1 ) sin 12 cos A12

11

2 Spherical and ellipsoidal geometry

sin 2 = sin 1 cos 12 + cos 1 sin 12 cos A12

Determination of 2
From the spherical cosine formula:
cos 12 = cos(90 1 ) cos(90 2 ) + sin(90 1 ) sin(90 2 ) cos 12
cos 12 =

cos 12 sin 1 sin 2


cos 1 cos 2

2 = 1 + 12

Determination of reverse azimuth A21


From the spherical cosine formula:
cos(90 1 ) = cos (90 2 ) cos 12 + sin (90 2 ) sin 12 cos A21
cos A21 =

sin 1 sin 2 cos 12


cos 2 sin 12

From the spherical sine formula:


sin A21
sin 12
=

sin (90 1 )
sin 12

sin A21 =

A21 = 360 arctan

sin 12 cos 1
sin 12

sin A21
cos A21

2.1.3 The inverse problem


Randwertproblem

The inverse problem is defined as the following boundary value problem:


Given: 1 and 1 of the first point
2 and 2 of the second point
Find: 12 between the first and second point
The azimuths A12 and A21 in both end points
Determination of spherical distance 12
From the spherical cosine formula:
cos 12 = cos(90 1 ) cos(90 2 ) + sin(90 1 ) sin(90 2 ) cos 12

12

2.1 Approximation 1: the sphere

cos 12 = sin 1 sin 2 + cos 1 cos 2 cos 12

Determination of azimuth A12


From the spherical cosine formula:
cos(90 2 ) = cos(90 1 ) cos 12 + sin(90 1 ) sin 12 cos A12
cos A12 =

sin 2 sin 1 cos 12


cos 1 sin 12

From the spherical sine formula:


sin A12
sin 12
=
sin (90 2 )
sin 12

sin A12 =

sin 12 cos 2
sin 12

sin A12
A12 = arctan cos
A

12

Determination of azimuth A21


From the spherical cosine formula:
cos(90 1 ) = cos (90 2 ) cos 12 + sin (90 2 ) sin 12 cos A21
cos A21 =

sin 1 sin 2 cos 12


cos 2 sin 12

From the spherical sine formula:


sin A21
sin 12
=

sin (90 1 )
sin 12

sin A21 =

A12 = 360 arctan

sin 12 cos 1
sin 12

sin A21
cos A21

Remark 2.3 In the above derivations extra effort has been put into defining the angles
in the right quadrant by determining an angle both with a sine-rule and a cosine-rule.
In many cases, in which the quadrant is clear, simpler formulas like the sine formulas
would be sufficient.

13

2 Spherical and ellipsoidal geometry

2.2 Approximation 2: the ellipsoid


Remark 2.4 In this section, the symbol will be used for the geodetic or ellipsoidal
latitude.
The ellipsoid is described in several ways:

Brennpunkte
zweiachsig

geometrically The ellipse is defined as the set of points whose sum of distances to two
foci is constant. This definition provides a curve in two-dimensional space. The
bi-axial ellipsoid in 3d space is the result of rotating the ellipse around one of its
axes.
Inspection of fig. 2.4, in which we choose a point on the major axis (left panel),
tells us that this sum must be (a + x) + (a x) = 2a, the length of the major axis.
The quantity a is called the semi-major axis.

lange Halbachse

b
a-x

b
ae

a
a+x
Figure 2.4: Planar geometry of the ellipse.

But then, for a point on the minor axis, see right panel, we have a symmetrical
configuration. The distance from this point to each of the foci is a. The length
b is called the semi-minor axis. Knowing
both axes, we can express the distance
to focus and centre of the ellipse. It is a2 b2 . Usually it is expressed as a
proportion e of the semi-major axis a:

kurze Halbachse

(ae)2 + b2 = a2 = e2 =

p
a2 b2
1 e2 a .
,
or
b
=
a2

The proportionality factor e is called the eccentricity; the out-of-centre distance


ae is known as the linear eccentricity.

Exzentrizit
at

algebraically (implicit) In case the axis of symmetry is the z-axis:


x2 y 2 z 2
+ 2 + 2 = 1.
a2
a
b

14

2.2 Approximation 2: the ellipsoid


One can obtain a 2d again by the following substitution:
(

in which p =
plane.

p2 z 2
x = p cos
= 2 + 2 = 1 ,
y = p sin
a
b

x2 + y 2 can be considered the horizontal coordinate in the meridian

parametrically (explicit) For points on the ellipsoid the transformation from ellipsoidal
to Cartesian coordinates reads:

x
N () cos cos

y = N () cos sin
z
N ()(1 e2 ) sin

, with:

N () = q

a
1 e2 sin2

(2.1a)

For points above the ellipsoidal surface, we have to add the ellipsoidal height h in
normal direction as follows:

x
(N + h) cos cos

y = (N + h) cos sin
z
(N (1 e2 ) + h) sin

(2.1b)

Figure 2.5: Ellipsoidal geometry.

A closed analytical solution for the reverse transformation from Cartesian to geodetic
coordinates does exist. Here, however, we will simply apply an iteration. First off,
longitude can be determined by: tan = xy . But geodetic latitude and height must be
solved iteratively together. To that end we introduce the coordinate p again (distance
to z-axis):
p=
iteration equation 1:
z = (N (1 e2 ) + h) sin
iteration equation 2:

x2 + y 2 = (N + h) cos
p
h=
N ()
cos

N (1 e2 ) + h
z
=
tan
p
N +h


z
N +h
= arctan
p N (1 e2 ) + h
15

2 Spherical and ellipsoidal geometry

Figure 2.6: Normal vector f vs.


radial direction r and link between
geodetic latitude and geocentric
latitude z

With the two above equations, the iteration runs as follows:


Starting value i = 0: h0 = 0 (just assume point on surface, if no better information
available).
1
Starting latitude: 0 = arctan( zp (1e
2 ) ) from iteration equation 2.
N (0 ) = . . .
hi+1 = cospi N (i ) from iteration equation 1.
i+1 = arctan

N (i )+hi
z
p N (i )(1e2 )+hi

from iteration equation 2 again.

N (i+1 ) = and so on.


Iteration until convergence is achieved
|hi+1 hi | < h
|i+1 i | <
Geodetic and geocentric latitudes From the implicit formulation of the ellipsoid, we
can derive the surface normal vector simply by taking the gradient:

3D

2D

x2 y 2 z 2
+ 2 + 2 = 1 = f (x, y, z)
a2
a b

p2
z2
+
= 1 = f (p, z)
a2
b2

f = 2

x
a2
y
a2
z

x N cos cos

y = N cos sin

16

N (1 e2 ) sin

f = 2

p

p
a2
z
b2

N cos
N (1

e2 ) sin

2.2 Approximation 2: the ellipsoid

Figure 2.7: Finite and infinitesimal arc


length on the sphere.

From fig. 2.6 the link between geocentric and geodetic latitude becomes clear:
tan z = zp (see figure)
2
tan = bz2 : ap2 = ab2 pz (from f )

= tan z =

z
b2
tan = (1 e2 ) tan .
2
a
p

2.2.1 Curvature
Sphere An infinitesimal arc length ds on the sphere is related to its infinitesimal central
angle simply by multiplying by the spheres radius R, see fig. 2.7:
ds = Rd .
This is more or less the translation of d in angular measure into linear measure. However, it leads to a more fundamental concept, as the quantity
=

1
d
=
R
ds

is called the curvature. The radius R is known as the radius of curvature. In general,
the curvature of a surface is a local quantity, that is, it depends on position. On the
sphere, though, curvature is constant. Thus, in 3.2.1 we might have added surface of
constant curvature as a definition of the sphere.

Kr
ummung
Kr
ummungsradius

Ellipsoid On the ellipsoid, on the other hand, the curvature is a local measure. To be
more precise:
= (, ) ,
that is, the curvature is latitude and direction dependent. It is a function of latitude
and on azimuth . At every point on the ellipsoid there will be a direction in which the
curvature is maximal and a direction in which it is minimal. Each direction spans up
a surface through the local normal vector. Such surfaces are called normal sections, see
fig. 2.9.
As might be expected, the two extremes in curvature take place

17

Azimut
Normalschnitte

2 Spherical and ellipsoidal geometry


i) in the meridian section, and
ii) in the prime vertical normal section, which is perpendicular to the meridian
section and tangent to the local latitude circle.
Note that the plane through a latitude circle by itself is not a normal section.
Let us consider the curvature and its variations in the meridian and in the equator. The
latitude dependence is obvious from fig. 2.8 (left panel). At the equator, the smaller
circle fits the ellipse in an optimal way. Its radius is the radius of curvature. It is clear
that this radius of curvature is smaller than the semi-major axis a. At the pole, though,
the best fitting circle has the largest possible radius, larger than a. Thus the curvature
at the pole, ( = 90 ), is minimum.

Figure 2.8: Latitude dependence of curvature in the meridian plane (left) and azimuth dependence at the equator (right).

At the pole, no directional dependence can exist, as all meridian planes are normal
sections. At the equator, though, there will be a clear difference in curvature between
meridian plane (as discussed above) and in the equator plane. The equatorial normal
section of the ellipsoid is a circle, see fig. 2.8. The radius of curvature at the equator in
East-West direction is therefore a and the curvature ( = 0 , = 90 ) = 1/a. In the
previous paragraph, we already concluded that the radius of curvature at the equator in
North-South direction was smaller than a.

Meridiankr
ummungsradius
Normalkr
ummungsradius

Main radii of curvature This behaviour is not only valid at the equator. At every
latitude we will see the minimum radius of curvature (and hence the maximum curvature)
in the meridian plane and the maximum radius of curvature in the prime vertical normal
section. They are known, respectively, as the meridian radius of curvature M () and
normal radius of curvature N (). The latter radius is exactly the quantity that we know
already from (2.1). The corresponding equations and some examples are given in the
following table.

18

Meridianschnit

2.2 Approximation 2: the ellipsoid

Figure 2.9

in meridian
general
at equator
at pole

M () = a

in prime vertical

1 e2
(1 e2 sin2 )3/2

N () = a

M (0 ) = a(1 e2 )
M (90 ) =

1
(1

e2 sin2 )1/2

N (0 ) = a

a
1 e2

N (90 ) =

a
1 e2

The table indeed confirms that the smallest radius of curvature is in North-South direction: M (0 ) < N (0 ). Moreover, at the poles there is no azimuth dependence:
M (90 ) = N (90 ).
Gauss curvature The radius of a best fitting sphere at a certain latitude is the Gauss
radius of curvature:

a 1 e2
RG = M N =
.
1 e2 sin2
Mean curvature

The mean curvature is defined by:


M =

1
1
=
RM
2

1
1
+
M
N

Curvature in arbitrary direction The mathematician Euler developed a formula that


relates the curvatures in North-South direction ( = 0 ) and in East-West direction

19

2 Spherical and ellipsoidal geometry


( = 90 ) to the curvature in arbitrary direction:
() =

sin2 cos2
1
=
+
.
R
N
M

(2.2)

2.2.2 The direct and inverse geodetic problem on the ellipsoid


geod
atische Linie

The shortest path between two points on a curved surface is called a geodesic. Solving the
direct and inverse geodetic problem on the ellipsoid would require finding and describing
geodesics on the ellipsoid. This is a mathematically demanding topic, particularly if
analytical solutions are attempted. To exemplify the level of complexity on the ellipsoid,
it is remarked that a geodesic is in general not a closed curve, like the great circle on
the sphere. It suffices to say that the geodesic is described by a set of three coupled
ordinary differential equations, that may be solved numerically.

Meridianbogen

Meridian arc A meridian arc s is a special geodesic. It is described by a single differential equation:
ds
= M () ,
d
which is of course the reverse of the definition of a differential arc length (compare the
spherical case):
ds = M ()d .
Therefore, the meridian arc length between two points at different latitudes is
s1,2 =

Z2
1

ds =

Z2

M ()d ,

which can be evaluated by numerical quadrature.

20

3 Gravitation
3.1 Newtonian gravitation
In 1687 Newton1 published his Philosophiae naturalis principia mathematica, or Principia in short. The Latin title can be translated as Mathematical principles of natural
philosophy, in which natural philosophy can be read as physics. Although Newton was
definitely not the only physicist working on gravitation in that era, his name is nevertheless remembered and attached to gravity because of the Principia. The greatness
of this work lies in the fact that Newton was able to bring empirical observations on
a mathematical footing and to explain in a unifying manner many natural phenomena:
planetary motion (in particular elliptical motion, as discovered by Kepler2 ),
free fall, e.g. the famous apple from the tree,
tides,
equilibrium shape of the Earth.
Newton made fundamental observations on gravitation:
The force between two attracting bodies is proportional to the individual masses.
The force is inversely proportional to the square of the distance.
The force is directed along the line connecting the two bodies.
Mathematically, the first two are translated into:
F12 = G

m1 m2
,
2
r12

(3.1)

in which G is a proportionality factor. It is called the gravitational constant or Newton


constant. It has a value of G = 6.672 1011 m3 s2 kg1 (or N m2 kg2 ).
1
2

Sir Isaac Newton (16421727).


Johannes Kepler (15711630), German astronomer and mathematician; formulated the famous laws
of planetary motion: i) orbits are ellipses with Sun in one of the foci, ii) the areas swept out by the
line between Sun and planet are equal over equal time intervals (area law), and iii) the ratio of the
cube of the semi-major axis and the square of the orbital period is constant (or n2 a3 = GM ).

21

3 Gravitation
Remark 3.1 (mathematical model of gravitation) Soon after the publication of the Principia Newton was strongly criticized for his law of gravitation, e.g. by his contemporary
Huygens. Equation (3.1) implies that gravitation acts at a distance, and that it acts
instantaneously. Such action is unphysical in a modern sense. For instance, in Einsteins
relativity theory no interaction can be faster than the speed of light. However, Newton
did not consider his formula (3.1) as some fundamental law. Instead, he saw it as a convenient mathematical description. As such, Newtons law of gravitation is still a viable
model for gravitation in physical geodesy.
Equation (3.1) is symmetric: the mass m1 exerts a force on m2 and m2 exerts a force
of the same magnitude but in opposite direction on m1 . From now on we will be
interested in the gravitational field generated by a single test mass. For that purpose
we set m1 := m and we drop the indices. The mass m2 can be an arbitrary mass at an
arbitrary location. Thus we eliminate m2 by a = F/m2 . The gravitational attraction a
of m becomes:
m
a=G 2,
(3.2)
r
in which r is the distance between mass point and evaluation point. The gravitational
attraction has units m/s2 . In geodesy one often uses the unit Gal, named after Galileo3 :
1 Gal = 102 m/s2 = 1 cm/s2
1 mGal = 105 m/s2
1 Gal = 108 m/s2 .
Remark 3.2 (kinematics vs. dynamics) The gravitational attraction is not an acceleration. It is a dynamical quantity: force per unit mass or specific force. Accelerations on
the other hand are kinematic quantities.

3.1.1 Vectorial attraction of a point mass


The gravitational attraction works along the line connecting the point masses. In this
symmetrical situation the attraction at point 1 is equal in size, but opposite in direction, to the attraction at point 2: a12 = a21 . This corresponds to Newtons law:
action = reaction.
In case we have only one point mass m, located in r 1 , whose attraction is evaluated in
point r 2 , this symmetry is broken. The vector a is considered to be the corresponding
attraction.

Galileo Galilei (15641642).

22

x2 x1

r = r 2 r 1 = y2 y1 , and r = |r|
z2 z1

3.1 Newtonian gravitation

z
P1

r = r2 - r1

P2

r1

- r1

r2
r

Figure 3.1: Attraction of a point mass m,


located in point P1 , on P2 .

a = G
= G

mr
m
m
e12 = G 2
= G 3 r
2
r
r r
r
m
[(x2 x1

)2

+ (y2 y1 )2 + (z2 z1 )2 ]3/2

x2 x1

y2 y1 .
z2 z1

3.1.2 Gravitational potential


The gravitational attraction field a is a conservative field. This means that the same
amount of work has to be done to go from point A to point B, no matter which path
you take. Mathematically, this is expressed by the fact that the field a is curl-free:
rot a = a = 0 .

(3.3)

Now from vector analysis it is known that the curl of any gradient field is always equal
to zero: rot grad F = F = 0. Therefore, a can be written as a gradient of some
scalar field. This scalar field is called gravitational potential V . The amount of energy it
takes (or can be gained) to go from A to B is simply VB VA . Instead of having to deal
with a vector field (3 numbers at each point) the gravitational field is fully described by
a scalar field (1 number).
The gravitational potential that would generate a can be derived by evaluating the
amount of work (per unit mass) required to get to location r. We assume that the mass
point is at the origin of our coordinate system. Since the integration in a conservative
field is path independent we can choose our path in a convenient way. We will start at
infinity, where the attraction is zero and go straight along the radial direction to our
point r.
V =

ZB

a dx

V =

Zr

m r
m
m
dr
=
G
=G .

2
r
r
r

(3.4)

23

3 Gravitation
The attraction is generated from the potential by the gradient operator:

a = grad V = V =

V
x
V
y
V
z

That this indeed leads to the same vector field is demonstrated by performing the partial
differentiations, e.g. for the x-coordinate:
1
1 r
V
= Gm r = Gm r
x
x
r x

 
2x
1
m
= Gm 2
= G 3 x ,
r
2r
r
and similarly for y and z.

3.1.3 Superpositiondiscrete
Gravitational formulae were derived for single point masses so far. One important property of gravitation is the so-called superposition principle. It says that the gravitational
potential of a system of masses can be achieved simply by adding the potentials of single
masses. In general we have:
V =

N
X
i=1

Vi = G

N
X
m1
m2
mN
mi
+G
+ ... + G
=G
.
r1
r2
rN
r
i=1 i

(3.5)

The mi are the single masses and the ri are the distances between masspoints and the
evaluation point. The total gravitational attraction is simply obtained by a = V again:
a = V =

X
i

Vi = G

X mi
i

ri3

ri .

(3.6)

3.1.4 Superpositioncontinuous
Real world mass configurations can be thought of as systems of infinitely many and
infinitely close point masses. The discrete formulation will become a continuous one.
N

X
i

ZZ Z

mi dm

24

3.2 Ideal solids

r1
1

r4

dz

dy

dx
y

Figure 3.2: Superposition for discrete (left) and continuous (right) mass distributions.

The body consists of mass elements dm, that are the infinitesimal masses of infinitesimal cubes dxdydz with local density (x, y, z):
dm(x, y, z) = (x, y, z) dxdydz .

(3.7)

Integrating over all mass elements in the continuous equivalent of superposition


gives the potential generated by :
VP = G

ZZ Z

dm
=G
r

ZZ Z

(x, y, z)
dxdydz ,
r

(3.8)

with r the distance between computation point P and mass element dm. Again, the
gravitational attraction of is obtained by applying the gradient operator:
a = V = G

ZZ Z

(x, y, z)
r dxdydz .
r3

(3.9)

The potential (3.8) and the attraction (3.9) can in principle be determined using volume
integrals if the density distribution within the body is known. However, we can
obviously not apply these integrals to the real Earth. The Earths internal density
distribution is insufficiently known. For that reason we will make use of potential theory
to turn the volume integrals into surface integrals in a later chapter.

3.2 Ideal solids


Using the general formulae for potential and attraction, we will investigate the gravitational effect of some ideal solid bodies now.

25

3 Gravitation

r sind
rd
dr

Figure 3.3: One octant of a solid sphere.


The volume element has sides dr in radial
direction, rd in co-latitude direction and
r sin d in longitude direction.

3.2.1 Solid homogeneous sphere


Consider a sphere of radius R with homogeneous density (x, y, z) = . In order to
evaluate the integrals (3.8) we assume a coordinate system with its origin at the centre
of the sphere. Since a sphere is rotationally symmetric we can evaluate the gravitational
potential at an arbitrary point. Our choice is a general point P on the positive z-axis.
Thus we have for evaluation point P and mass point Q the following vectors:

0

r P = 0 , rP = z ,
z

r sin cos

r Q = r sin sin , rQ = r < R ,


r cos

rP Q

r sin cos
p

= r Q r P = r sin sin , rP Q = z 2 + r2 2rz cos .


r cos z

(3.10)

It is easier to integrate in spherical coordinates than in Cartesian4 ones. Thus we use


the radius r, co-latitude and longitude . The integration bounds become

RR R

2
R

r=0 =0 =0

and the volume element dxdydz is replaced by r2 sin dddr. Applying this change of
4

Rene Descartes or Cartesius (15961650), French mathematician, scientist and philosopher whose
work La geometrie (1637), includes his application of algebra to geometry from which we now have
Cartesian geometry.

26

3.2 Ideal solids


coordinates to (3.8) and putting the constant density outside of the integral, yields
the following integration:
VP = G

= G

ZZZ

1
rP Q

dxdydz

ZR Z Z2

r=0 =0 =0

= G

ZR Z Z2

r=0 =0 =0

= 2G

ZR Z

r=0 =0

1
rP Q

r2 sin dddr

r2 sin
dddr
z 2 + r2 2rz cos

r2 sin
ddr .
z 2 + r2 2rz cos

The integration over was trivial, since doesnt appear in the integrand. The integration over is not straightforward, though. A good trick is to change variables. Call
rP Q (3.10) l now. Then

d z 2 + r2 2rz cos
zr sin
l
dl
=
=

dl = sin d
d
d
l
zr
rl
r2 sin d = dl .
z
Thus the integral becomes:
VP = 2G

ZR Zl+

r=0 l=l

The integration bounds of


cases.

r
dldr .
z

(3.11)

have to be determined first. We have to distinguish two

Point P outside the sphere (z > R): From fig. 3.4 (left) the integration bounds for
l become immediately clear:
= 0 l = z r
= l+ = z + r
VP = 2G

Z
ZR z+r

r=0 zr

= 2G

ZR

r=0

r
dldr = 2G
z

ZR 

r=0

r
l
z

z+r

dr

zr

4
R3
2r2
dr = G
.
z
3
z

27

3 Gravitation

P
Q (=0)

z
r

Q (=0)
r P
0

Q (=)

Q (=)

Figure 3.4: Determining the integration limits for variable l when the evaluation point P is
outside (left) or inside (right) the solid sphere.

We chose the evaluation point P arbitrary on the z-axis. In general, we can replace z
by r now because of radial symmetry. Thus we obtain:
1
4
(3.12)
V (r) = GR3 .
3
r
Recognizing that the mass M of a sphere equals 34 R3 , we simply obtain V = GM
r . So
the potential of a solid sphere equals that of a point mass, at least outside the sphere.
Point P within sphere (z < R): For this situation we must distinguish between mass
points below the evaluation point (r < z) and mass point outside (z < r < R). The
former configuration would be a sphere of radius z. Its potential in point P (= [0, 0, z])
is
4
z3
4
VP = G = Gz 2 .
(3.13)
3
z
3
For the masses outside P we have the following integration bounds for l:
= 0 l = r z ,
= l+ = r + z .
The integration over r runs from z to R. With the same change of variables we obtain
VP = 2G

Z
ZR r+z

r=z rz

= 2G

ZR

r=z

28

r
dldr = 2G
z

ZR 

r=z

h iR

2r dr = 2G r2

r
l
z

r+z

dr

rz

= 2G(R2 z 2 ) .

3.2 Ideal solids


The combined effect of the smaller sphere (r < z) and spherical shell (z < r < R) is:
4
1
VP = Gz 2 + 2G(R2 z 2 ) = 2G(R2 z 2 ) .
3
3

(3.14)

Again we can replace z now by r. In summary, the gravitational potential of a sphere


of radius R reads
outside:
inside:

V (r > R) =

4
1
GR3 ,
3
r

1
V (r < R) = 2G(R2 r2 ) .
3

(3.15a)
(3.15b)

Naturally, at the boundary the potential will be continuous. This is verified by putting
r = R in both equations, yielding:
4
V (R) = GR2 .
3

(3.16)

This result is visualized in fig. 3.5. Not only is the potential continuous across the surface
of the sphere, it is also smooth.
Attraction. It is very easy now to find the attraction of a solid sphere. It simply is the
radial derivative. Since the direction is radially towards the spheres center (the origin)
we only need to deal with the radial component:
outside:
inside:

4
1
a(r > R) = GR3 2 ,
3
r
4
a(r < R) = Gr .
3

(3.17a)
(3.17b)

Continuity at the boundary is verified by


4
a(R) = GR .
3

(3.18)

Again, the result is visualized in fig. 3.5. Although the attraction is continuous across
the boundary, it is not differentiable anymore.
Exercise 3.1 Given the gravitation a = 981 Gal on the surface of a sphere of radius
R = 6378 km, calculate the mass of the Earth ME and its mean density E .
Exercise 3.2 Try to find more general gravitational formulae for V and a for the case
that the density is not constant but depends on the radial distance: = (r). First, set
up the integrals and then try to solve them.

29

3 Gravitation

V(r)

a(r)
-R

Figure 3.5: Potential V and attraction


a as a function of r, due to a solid homogeneous sphere of radius R.

3.2.2 Spherical shell


A spherical shell is a hollow sphere with inner radius R1 and outer radius R2 . The
gravitational potential of it may be found analogous to the derivations in 3.2.1. Of
course the proper integration bounds should be used. However, due to the superposition
principle, we can simply consider a spherical shell to be the difference between two solid
spheres. Symbolically, we could write:
spherical shell(R1 , R2 ) = sphere(R2 ) sphere(R1 ) .

(3.19)

By substracting equations (3.15) and (3.17) with the proper radii, one arrives at the
potential and attraction of the spherical shell. One has to be careful in the area R1 <
r < R2 , though. We should pick the outside formula for sphere (R1 ) and the inside
formula for sphere (R2 ).
4
1
G(R23 R13 ) ,
3
r
1
1
4
in shell: V (R1 < r < R2 ) = 2G(R22 r2 ) GR13 ,
3
3
r
2
2
inner part
V (r < R1 ) = 2G(R2 R1 ) .
outer part:

V (r > R2 ) =

(3.20a)
(3.20b)
(3.20c)

Note that the potential in the inner part is constant.


Remark 3.3 The potential outside the spherical shell with radi R1 and R2 and density
could also have been generated by a point mass with M = 34 (R23 R13 ). But also by a
solid sphere of radius R2 and density 0 = (R23 R13 )/R23 . If we would not have seismic
data we could never tell if the Earth was hollow or solid.

30

3.2 Ideal solids

V(r)

a(r)

-R2

R1

-R1

R2

Figure 3.6: Potential V and attraction


a as a function of r, due to a homogeneous spherical shell with inner radius
R1 and outer radius R2 .

Remark 3.4 Remark 3.3 can be generalized. If the density structure within a sphere of
radius R is purely radially dependent, the potential outside is of the form GM/r:
= (r) V (r > R) =

GM
.
r

Similarly, for the attraction we obtain:


1
4
a(r > R2 ) = G(R23 R13 ) 2 ,
3
r
4
1
in shell: a(R1 < r < R2 ) = G(r3 R13 ) 2 ,
3
r
inner part
a(r < R1 ) = 0 .
outer part:

(3.21a)
(3.21b)
(3.21c)

Since the potential is constant within the shell, the gravitational attraction vanishes
there. The resulting potential and attraction are visualized in fig. 3.6.
Exercise 3.3 Check the continuity of V and a at the boundaries r = R1 and r = R2 .
Exercise 3.4 The basic structure of the Earth is radial: inner core, outer core, mantle,
crust. Assume the following simplified structure:
core: Rc = 3500 km , c = 10 500 kg m3
mantle: Rm = 6400 km , m = 4500 kg m3 .

31

3 Gravitation

z
P

h/2

dr
dz

rPQ

ZP

rd

r
z

R
h/2

Figure 3.7: Cylinder (left) of radius R and height h. The origin of the coordinate system is
located in the center of the cylinder. The evaluation point P is located on the z-axis (symmetry
axis). The volume element of the cylinder (right) has sides dr in radial direction, dz in vertical
direction and rd in longitude direction.

Write down the formulae to evaluate potential and attraction. Calculate these along a
radial profile and plot them.

3.2.3 Solid homogeneous cylinder


The gravitational attraction of a cylinder is useful for gravity reductions (Bouguer corrections), isostasy modelling and terrain modelling. Assume a configuration with the
origin in the center of the cylinder and the z-axis coinciding with the symmetry axis.
The cylinder has radius R and height h. Again, assume the evaluation point P on the
positive z-axis. As before in 3.2.1 we switch from Cartesian to suitable coordinates. In
this case that would be cylinder coordinates (r, , z):

x
r cos

y = r sin .
z
z

32

(3.22)

3.2 Ideal solids


For the vector from evaluation point P to mass point Q we can write down:
rP Q

r cos
r cos
0


= r Q r P = r sin 0 = r sin ,
z zP
z
zP

rP Q =

r2 + (zP z)2 .

(3.23)
The volume element dxdydz becomes rddrdz and the the integration bounds are
h/2

RR 2
R

. The integration process for the potential of the cylinder turns out to

z=h/2 r=0 =0

be somewhat cumbersome. Therefore we integrate the attraction (3.9) directly:

Zh/2

aP = G

ZR Z2

z=h/2 r=0 =0

Zh/2

= 2G

z=h/2

ZR

r cos
1

r sin rddrdz
rP3 Q
z zP

(z zP )

r=0

0
1
0 rdrdz .
rP3 Q
1

On the symmetry axis the attraction will have a vertical component only. So we can
continue with a scalar aP now. Again a change of variables brings us further. Calling
rP Q (3.23) l again gives:
p

dl
d r2 + (zP z)2
r
r
=
=p 2
=
dr
dr
l
r + (zP z)2

r
dr = dl .
l

(3.24)

 l+

(3.25)

Thus the integral becomes:


aP = 2G

Zh/2

z=h/2

(z zP )

Zl+

l=l

1
dldz = 2G
l2

Zh/2

z=h/2

(z zP )

1
l

dz .

Indeed, the integrand is much easier now at the cost of more difficult integration bounds
of l, which must be determined now. Analogous to 3.2.1 we could distinguish between P
outside (above) and P inside the cylinder. It will be shown later, though, that the latter
case can be derived from the former. So with zP > h/2 we get the following bounds:
r = 0 l = zP z
r = R l+ =
With these bounds we arrive at:
aP = 2G

Zh/2

z=h/2

R2 + (zP z)2 .
!

z zP
p
+ 1 dz
2
R + (zP z)2
33

3 Gravitation

= 2G

q

R2 + (zP z)2 + z

= 2G h +

h/2

h/2

R2 + (zP h/2)2

R2 + (zP + h/2)2

Now that the integration over z has been performed we can use the variable z again to
replace zP and get:
a(z >

h/2)

= 2G h +

R2

+ (z

h/2)2

R2

+ (z +

h/2)2

(3.26)

Recall that this formula holds outside the cylinder along the positive z-axis (symmetry
axis).
Negative z-axis The corresponding attraction along the negative z-axis (z < h/2)
can be found by adjusting the integration bounds of l. Alternatively, we can replace z
by z and change the overall sign.


a(z < h/2) = +2G h +




R2 + (z h/2)2

= 2G h +

R2

+ (z

h/2)2

R2 + (z + h/2)2

R2

+ (z +

h/2)2

(3.27)

P within cylinder First, we need to know the attraction at the top and at the base of
the cylinder. Inserting z = h/2 in (3.26) and z = h/2 in (3.27) we obtain


a(h/2) = 2G h + R


a(h/2) = 2G h +

R2 + h2 ,

(3.28a)

(3.28b)

R2 + h2 R .

Notice that a(h/2) = a(h/2) indeed.


In order to calculate the attraction inside the cylinder, we separate the cylinder into two
cylinders exactly at the evaluation point. So the evaluation point is at the base of a
cylinder of height (h/2 z) and at the top of cylinder of height (h/2 + z). Replacing the
heights h in (3.28) by these new heights gives:


base of upper cylinder : 2G (h/2 z) +




R2 + (h/2 z)2 R

top of lower cylinder : 2G (h/2 + z) + R


=

34

a(h/2

<z<

h/2)

= 2G 2z +

R2

+ (z

R2 + (h/2 + z)2

h/2)2

R2

+ (z +

h/2)2

3.2 Ideal solids


Summary The attraction of a cylinder of height h and radius R along its symmetry
axis reads:

q
q
h

a(z) = 2G 2z + R2 + (z h/2)2 R2 + (z + h/2)2

z>h/2
h/2<z<h/2

(3.29)

z<h/2

This result is visualized in fig. 3.8.

a(z)
-h/2

h/2

Figure 3.8: Attraction a as a function of z, due to a solid homogeneous cylinder of radius R and
height h. Note that the horizontal axis is the z-axis in this visualization.

Exercise 3.5 Find out the formulae for a cylindrical shell, i.e. a hollow cylinder with
inner radius R1 and outer radius R2 .
Infinite plate of thickness h. Using the above formulae one can easily derive the
attraction of an infinite plate. If we let the radius R go to infinity, we will get:

lim a(z) = 2G

h , above
2z , within .

h , below

(3.30)

This formula is remarkable. First, by taking the limits, the square root terms have
vanished. Second, above and below the plate the attraction does not depend on z
anymore. It is constant there. The above infinite plate formula is often used in gravity
reductions, as will be seen in 5.
Exercise 3.6 Calgary lies approximately 1000 m above sealevel. Calculate the attraction
of the layer between the surface and sealevel. Think of the layer as a homogeneous plate
of infinite radius with density = 2670 kg/m3 .

35

3 Gravitation
Exercise 3.7 Simulate a volcano by a cone with top angle 90 , i.e. its height equals the
radius at the base. Derive the corresponding formulae for the attraction by choosing
the proper coordinate system (hint: z = 0 at base) and integration bounds. Do this in
particular for Mount Fuji (H = 3776 m) with = 3300 kg m3 .

3.3 Tides
Sorry, its ebb.

36

3.4 Summary

3.4 Summary
point mass in origin
V (r) = Gm

1
r

a(r) = Gm

1
r2

solid sphere of radius R and constant density , centered in origin

V (r) =

4
31

GR

2G(R2 1 r 2 )

a(r) =

4
3 1

GR 2 , outside

4 Gr

, inside

spherical shell with inner and outer radii R1 and R2 , resp., and constant density

V (r) =

4
3
3 1

G(R

R
)

2
1

2G(R22 r2 ) GR13

3
3
r

a(r) =

2G(R22 R12 )

4
3 1
3

G(R

R
) 2 , outside

2
1

G(r3 R13 ) 2 , in shell

3
r

, inside

cylinder with height h and radius R, centered at origin, constant density

, above
q
h
q

a(z) = 2G 2z + R2 + (z h/2)2 R2 + (z + h/2)2 , within

, below

infinite plate of thickness h and constant density


a(z) = 2G

h , above
2z , within .

h , below

37

4 Rotation
kinematics Gravity related measurements take generally place on non-static platforms:
sea-gravimetry, airborne gravimetry, satellite gravity gradiometry, inertial navigation.
Even measurements on a fixed point on Earth belong to this category because of the
Earths rotation. Accelerated motion of the reference frame induces inertial accelerations, which must be taken into account in physical geodesy. The rotation of the Earth
causes a centrifugal acceleration which is combined with the gravitational attraction
into a new quantity: gravity. Other inertial accelerations are usually accounted for by
correcting the gravity related measurements, e.g. the E
otv
os correction. For these and
other purposes we will start this chapter by investigating velocity and acceleration in a
rotating frame.

dynamics One of geodesys core areas is determining the orientation of Earth in space.
This goes to the heart of the transformation between inertial and Earth-fixed reference
systems. The solar and lunar gravitational fields exert a torque on the flattened Earth,
resulting in changes of the polar axis. We need to elaborate on the dynamics of solid
body rotation to understand how the polar axis behaves in inertial and in Earth-fixed
space.

geometry Newtons laws of motion are valid in inertial space. If we have to deal with
satellite techniques, for instance, the satellites ephemeris is most probably given in
inertial coordinates. Star coordinates are by default given in inertial coordinates: right
ascension and declination . Moreover, the law of gravitation is defined in inertial
space. Therefore, after understanding the kinematics and dynamics of rotation, we will
discuss the definition of inertial reference systems and their realizations. An overview
will be presented relating the conventional inertial reference system to the conventional
terrestrial one.

4.1 Kinematics: acceleration in a rotating frame


Let us consider the situation of motion in a rotating reference frame and let us associate
this rotating frame with the Earth-fixed frame. The following discussion on velocities
and accelerations would be valid for any rotating frame, though.

38

4.1 Kinematics: acceleration in a rotating frame


Inertial coordinates, velocities and accelerations will be denoted with the index i. Earthfixed quantities get the index e. Now suppose that a time-dependent rotation matrix
R = R((t)), applied to the inertial vector r i , results in the Earth-fixed vector r e .
We would be interested in velocities and accelerations in the rotating frame. The time
derivations must be performed in the inertial frame, though.
From Rr i = r e we get:
r i = RT r e

(4.1a)

time derivative

r i = RT r e + R T r e

(4.1b)

multiply by R

Rr i = r e + RR T r e
= r e + r e

(4.1c)

The matrix = RR T is called Cartan1 matrix. It describes the rotation rate, as can be
seen from the following simple 2D example with (t) = t:
R=

cos t sin t
sin t cos t

cos t sin t
sin t cos t

!
!

sin t cos t
cos t sin t

0
0

It is useful to introduce . In the next time differentiation step we can now distinguish
between time dependent rotation matrices and time variable rotation rate. Lets pick
up the previous derivation again:
multiply by RT

r i = RT r e + RT r e

(4.1d)

time derivative

e + RT r e
ri = RT r e + R T r e + R T r e + RT r
e
= RT r e + 2R T r e + R T r e + RT r

(4.1e)

multiply by R

e
R
r i = r e + 2r e + r e + r
or the other way around

e
r e = R
r i 2r e r e r

(4.1f)

This equation tells us that acceleration in the rotating e-frame equals acceleration in the
inertial i-framein the proper orientation, thoughwhen 3 more terms are added. The
1

Elie Joseph Cartan (18691951), French mathematician.

39

4 Rotation
additional terms are called inertial accelerations Analyzing (4.1f) we can distinguish the
four terms at the right hand side:
R
r i is the inertial acceleration vector, expressed in the orientation of the rotating
frame.
2r e is the so-called Coriolis 2 acceleration, which is due to motion in the rotating
frame.
r e is the centrifugal acceleration, determined by the position in the rotating
frame.
e is sometimes referred to as Euler 3 acceleration or inertial acceleration of rotar
tion. It is due to a non-constant rotation rate.
Remark 4.1 Equation (4.1f) can be generalized to moving frames with time-variable
origin. If the linear acceleration of the e-frames origin is expressed in the i-frame with
i , the only change to be made to (4.1f) is R
i ).
b
r i R(
ri b
Properties of the Cartan matrix . Cartan matrices are skew-symmetric, i.e. T =
. This can be seen in the simple 2D example above already. But it also follows from
the orthogonality of rotation matrices:
RRT = I =

d
T
(RRT ) = RR
R T} = 0 = T = .
| {z } + R
| {z
dt

(4.2)

A second interesting property is the fact that multiplication of a vector with the Cartan
matrix equals the cross product of the vector with a corresponding rotation vector:
r = r

(4.3)

This property becomes clear from writing out the 3 Cartan matrices, corresponding to
the three independent rotation matrices:

0 0 0

R1 (1 t) 1 = 0 0 1

0 1 0

0 0 2

R2 (2 t) 2 = 0 0 0

2 0 0

2
3

0 3 0

R3 (3 t) 3 = 3 0 0

0 0 0

Gaspard Gustave de Coriolis (17921843).


Leonhard Euler (17071783).

40

general

0 3 2

= 3 0 1 .
2 1 0

(4.4)

4.1 Kinematics: acceleration in a rotating frame


Indeed, when a general rotation vector = (1 , 2 , 3 )T is defined, we see that:

x
x
1
0 3 2

0 1 y = 2 y .
3
z
3
2 1 0
z

The skew-symmetry (4.2) of is related to the fact r = r .


Exercise 4.1 Convince yourself that the above Cartan matrices i are correct, by doing
the derivation yourself. Also verify (4.3) by writing out lhs and rhs.
Using property (4.3), the velocity (4.1c) and acceleration (4.1f) may be recast into the
perhaps more familiar form:
r e = Rr i r e

(4.5a)

r e = R
r i 2 r e ( r e ) r e

(4.5b)

Inertial acceleration due to Earth rotation


Neglecting precession, nutation and polar motion, the transformation from inertial to
Earth-fixed frame is given by:
or

r e = R3 (gast)r i r e = R3 (t)r i .

(4.6)

The latter is allowed here, since we are only interested in the acceleration effects, due
to the rotation. We are not interested in the rotation of position vectors. With great
precision, one can say that the Earths rotation rate is constant: = 0 The corresponding
Cartan matrix and its time derivative read:

0 0

= 0 0
0 0 0

and = 0 .

The three inertial accelerations, due to the rotation of the Earth, become:
Coriolis:

centrifugal:
Euler:

y e

2r e = 2 x e
0

xe

2
r e = ye
0
e =0
r

(4.7a)

(4.7b)
(4.7c)

The Coriolis acceleration is perpendicular to both the velocity vector and the Earths
rotation axis. It will be discussed further in 5.1. The centrifugal acceleration is perpendicular to the rotation axis and is parallel to the equator plane, cf. fig. 5.2.

41

4 Rotation
Exercise 4.2 Determine the direction and the magnitude of the Coriolis acceleration if
you are driving from Calgary to Banff with 100 km/h.
Exercise 4.3 How large is the centrifugal acceleration in Calgary? On the equator? At
the North Pole? And in which direction?

4.2 Dynamics: precession, nutation, polar motion


Instead of linear velocity (or momentum) and forces we will have to deal with angular
momentum and torques. Starting with the basic definition of angular momentum of a
point mass, we will step by step arrive at the angular momentum of solid bodies and
their tensor of inertia. In the following all vectors are assumed to be given in an inertial
frame, unless otherwise indicated.
Angular momentum of a point mass The basic definition of angular momentum of a
point mass is the cross product of position and velocity: L = mr v. It is a vector
quantity. Due to the definition the direction of the angular momentum is perpendicular
to both r and v.
In our case, the only motion v that exists is due to the rotation of the point mass. By
substituting v = r we get:
L = mr ( r)

(4.8a)

x
x
1


= m y 2 y
z
3
z

1 y 2 2 xy 3 xz + 1 z 2

= m 2 z 2 3 yz 1 yx + 2 x2
3 x2 1 zx 2 zy + 3 y 2

y 2 + z 2 xy
xz
1

= m xy x2 + z 2 yz 2
xz
yz x2 + y 2
3
= M .

(4.8b)

The matrix M is called the tensor of inertia. It has units of [kg m2 ]. Since M is not
an ordinary matrix, but a tensor, which has certain transformation properties, we will
indicate it by boldface math type, just like vectors.
Compare now the angular momentum equation L = M with the linear momentum
equation p = mv, see also tbl. 4.1. It may be useful to think of m as a mass scalar and
of M as a mass matrix. Since the mass m is simply a scalar, the linear momentum p

42

4.2 Dynamics: precession, nutation, polar motion


will always be in the same direction as the velocity vector v. The angular momentum L,
though, will generally be in a different direction than , depending on the matrix M .
Exercise 4.4 Consider yourself a point mass and compute your angular momentum, due
to the Earths rotation, in two ways:
straightforward by (4.8a), and
by calculating your tensor of inertia first and then applying (4.8b).
Is L parallel to in this case?

Angular momentum of systems of point masses The concept of tensor of inertia is


easily generalized to systems of point masses. The total tensor of inertia is just the
superposition of the individual tensors. The total angular momentum reads:
L=

N
X

n=1

mn r n v n =

N
X

M n .

(4.9)

n=1

Angular momentum of a solid body We will now make the transition from a discrete
to a continuous mass distribution, similar to the gravitational superposition case in 3.1.4.
Symbolically:
lim

N
X

mn . . . =

n=1

ZZ Z

. . . dm .

Again, the angular momentum reads L = M . For a solid body, the tensor of inertia
M is defined as:

M =

ZZ Z

RRR

y 2 + z 2 xy
xz

2
2
yz dm
xy x + z
xz
yz x2 + y 2

(y 2 + z 2 )dm

symmetric

RRR

RRR

xydm

(x2 + z 2 )dm

RRR

RRR

RRR

xzdm
yzdm

(x2 + y 2 )dm

The diagonal elements of this matrix are called moments of inertia. The off-diagonal
terms are known as products of inertia.
Exercise 4.5
Show that in vector-matrix notation the tensor of inertia M can be written
RRR T
as: M = (r rI rr T )dm .

43

4 Rotation
Torque If no external torques are applied to the rotating body, angular momentum is
conserved. A change in angular momentum can only be effected by applying a torque
T :
dL
=T =rF .
(4.10)
dt
Equation (4.10) is the rotational equivalent of p = F , see tbl. 4.1. Because of the crossproduct, the change in the angular momentum vector is always perpendicular to both r
and F . Try to intuitively change the axis orientation of a spinning wheel by applying a
force to the axis and the axis will probably go a different way. If no torques are applied
(T = 0) the angular momentum will be constant, indeed.
Table 4.1: Comparison between linear and rotational dynamics

linear
linear momentum
force

p = mv
dp
=F
dt

rotational
point mass
solid body
L = mr v
dL
=rF
dt

L = M
dL
=T
dt

angular momentum
torque

Three cases will be distinguished in the following:


T is constant precession,
which is a secular motion of the angular momentum vector in inertial space,
T is periodic nutation (or forced nutation),
which is a periodic motion of L in inertial space,
T is zero
free nutation, polar motion,
which is a motion of the rotation axis in Earth-fixed space.

Precession The word precession is related to the verb to precede, indicating a steady,
secular motion. In general, precession is caused by constant external torques. In the
case of the Earth, precession is caused by the constant gravitational torques from Sun
and Moon. The Suns (or Moons) gravitational pull on the nearest side of the Earth is
stronger than the pull on the. At the same time the Earth is flattened. Therefore, if the
Sun or Moon is not in the equatorial plane, a torque will be produced by the difference
in gravitational pull on the equatorial bulges. Note that the Sun is only twice a year in
the equatorial plane, namely during the equinoxes (beginning of Spring and Fall). The
Moon goes twice a month through the equator plane.
Thus, the torque is produced because of three simultaneous facts:
the Earth is not a sphere, but rather an ellipsoid,
the equator plane is tilted with respect to the ecliptic by 23. 5 (the obliquity ) and
also tilted with respect to the lunar orbit,

44

4.3 Geometry: defining the inertial reference system


the Earth is a spinning body.
If any of these conditions were absent, no torque would be generated by solar or lunar
gravitation and precession would not take place.
As a result of the constant (or mean) part of the lunar and solar torques, the angular
momentum vector will describe a conical motion around the northern ecliptical pole
(nep) with a radius of . The northern celestial pole (ncp) slowly moves over an ecliptical
latitude circle. It takes the angular momentum vector 25 765 years to complete one
revolution around the nep. That corresponds to 50.00 3 per year.
Nutation The word nutation is derived from the Latin for to nod. Nutation is a periodic
(nodding) motion of the angular momentum vector in space on top of the secular precession. There are many sources of periodic torques, each with its own frequency:
The orbital plane of the moon rotates once every 18.6 years under the influence of
the Earths flattening. The corresponding change in geometry causes also a change
in the lunar gravitational torque of the same period. This effect is known as Bradley
nutation.
The sun goes through the equatorial plane twice a year, during the equinoxes. At
those time the solar torque is zero. Vice versa, during the two solstices, the torque
is maximum. Thus there will be a semi-annual nutation.
The orbit of the Earth around the Sun is elliptical. The gravitational attraction of
the Sun, and consequently the gravitational torque, will vary with an annual period.
The Moon passes the equator twice per lunar revolution, which happens roughly
twice per month. This gives a nutation with a fortnightly period.

4.3 Geometry: defining the inertial reference system


4.3.1 Inertial space
The word true must be understood in the sense that precession and nutation have not
been modelled away. The word mean refers to the fact that nutation effects have been
taken out. Both systems are still time dependent, since the precession has not been
reduced yet. Thus, they are actually not inertial reference systems.

4.3.2 Transformations
Precession The following transformation describes the transition from the mean inertial reference system at epoch T0 to the mean instantaneous one i0 :
r = P r i0 = R3 (z)R2 ()R3 (0 )r i0 .

(4.11)

45

4 Rotation
Figure 4.1 explains which rotations need to be performed to achieve this transformation.
First, a rotation around the north celestial pole at epoch T0 (ncp0 ) shifts the mean
) over the mean equator at T . This is R ( ). Next, the
equinox at epoch T0 (
0
0
3
0
ncp0 is shifted along the cone towards the mean pole at epoch T (ncpT ). This is a
rotation R2 (), which also brings the mean equator at epoch T0 is brought to the mean
equator at epoch T . Finally, a last rotation around the new pole, R3 (z) brings the
) back to the ecliptic. The required precession angles are
mean equinox at epoch T (
T
00
given with a precision of 1 by:
0 = 2306.00 2181 T + 0.00 301 88 T 2
= 2004.00 3109 T 0.00 426 65 T 2

z = 2306.00 2181 T + 1.00 094 68 T 2

The time T is counted in Julian centuries (of 36 525 days) since J2000.0, i.e. January
1, 2000, 12h ut1. It is calculated from calendar date and universal time (ut1) by first
converting to the so-called Julian day number (jd), which is a continuous count of the
number of days. In the following Y, M, D are the calendar year, month and day
Julian days

jd = 367Y floor(7(Y + floor((M + 9)/12))/4)


+ floor(275M/9) + D + 1721 014 + ut1/24 0.5

time since J2000.0 in days


same in Julian centuries

d = jd 2451 545.0
d
T =
36 525

Exercise 4.6 Verify that the equinox moves approximately 5000 per year indeed by projecting the precession angles 0 , , z onto the ecliptic. Use T = 0.01, i.e. one year.
Nutation The following transformation describes the transition from the mean instantaneous inertial reference system to the true instantaneous one i:
r i = N r = R1 ( )R3 ()R1 ()r .

(4.12)

Again, fig. 4.1 explains the individual rotations. First, the mean equator at epoch T is
. This rotation, R (), brings the mean north pole
rotated into the ecliptic around
1
T
towards the nep. Next, a rotation R3 () lets the mean equinox slide over the ecliptic
towards the true instantaneous epoch. Finally, the rotation R1 ( ) brings us back
to an equatorial system, to the true instantaneous equator, to be precise. The nutation
angles are known as nutation in obliquity () and nutation in (ecliptical) longitude
(). Together with the obliquity itself, they are given with a precision of 100 by:
= 84 381.00 448 46.00 8150 T
=

46

0. 0026 cos(f1 ) + 0. 0002 cos(f2 )

4.3 Geometry: defining the inertial reference system


= 0. 0048 sin(f1 ) 0. 0004 sin(f2 )
with
f1 = 125. 0 0. 052 95 d

f2 = 200. 9 + 1. 971 29 d

The obliquity is given in seconds of arc. Converted into degrees we would have 23. 5
indeed. On top of that it changes by some 4700 per Julian century. The nutation angles
are not exact. The above formulae only contain the two main frequencies, as expressed
by the time-variable angles f1 and f2 . The coefficients to the variable d are frequencies
in units of degree/day:
f1 : frequency = 0.052 95 /day
f2 : frequency = 1.971 29 /day

period = 18.6 years


period = 0.5 years

The angle f1 describes the precession of the orbital plane of the moon, which rotates
once every 18.6 years. The angle f2 describes a half-yearly motion, caused by the fact
that the solar torque is zero in the two equinoxes and maximum during the two solstices.
The former has the strongest effect on nutation, when we look at the amplitudes of the
sines and cosines.
GAST For the transformation from the instantaneous true inertial system i to the instantaneous Earth-fixed sytem e we only need to bring the true equinox to the Greenwich
meridian. The angle between the x-axes of both systems is the Greenwich Actual Siderial
Time (gast). Thus, the following rotation is required for the transformation i e:

r e = R3 (gast)r e .

(4.13)

The angle gast is calculated from the Greenwich Mean Siderial Time (gmst) by applying a correction for the nutation.
gmst = ut1 + (24 110.548 41 + 8640 184.812 866 T + 0.093 104 T 2 6.2 106 T 3 )/3600
+ 24n
gast = gmst + ( cos( + ))/15
Universal time ut1 is in decimal hours and n is an arbitrary integer that makes 0
gmst < 24.

4.3.3 Conventional inertial reference system


Not only is the International Earth Rotation and Reference Systems Service (iers)
responsible for the definition and maintenance of the conventional terrestrial coordinate
system itrs (International Terrestrial Reference System) and its realizations itrf. The
iers also defines the conventional inertial coordinate system, called icrs (International
Celestial Reference System), and maintains the corresponding realizations icrf.

47

4 Rotation
system The icrs constitutes a set of prescriptions, models and conventions to define
at any time a triad of inertial axes.
origin:
barycentre of the solar system (6= Suns centre of mass),
at epoch J2000.0,
orientation:
mean equator and mean equinox
0
time system:
barycentric dynamic time tdb,
time evolution: formulae for P and N .

frame A coordinate system like the icrs is a set of rules. It is not a collection of
points and coordinates yet. It has to materialize first. The International Celestial
Reference Frame (icrf) is realized by the coordinates of over 600 that have been observed
by Very Long Baseline Interferometry (vlbi). The position of the quasars, which are
extragalactic radio sources, is determined by their right ascension and declination .
Classically, star coordinates have been measured in the optical waveband. This has
resulted in a series of fundamental catalogues, e.g. FK5. Due to atmospheric refraction,
these coordinates cannot compete with vlbi-derived coordinates. However, in the early
nineties, the astrometry satellite Hipparcos collected the coordinates of over 100 000
stars with a precision better than 1 milliarcsecond. The Hipparcos catalogue constitutes
the primary realization of an inertial frame at optical wavelengths. It has been aligned
with the icrf.

48

4.3 Geometry: defining the inertial reference system

4.3.4 Overview

global

inertial

0

global
geodetic

gg
datum0

i0
ci

conventional
inertial (T0 )

precession

0 , , z

datum

mean
inertial (T )
ra,mean
nutation

ct

conventional
terrestrial

polar
motion

, ,

instantaneous
inertial (T )
ra,true

r 0 , i

e0

gast

local
geodetic

lg
r 00 , a, f
, , h

local
geodetic
lg

defl. of vertical
geoid

, , H

,
N

g0
la

local
astronomic

xP , yP

e
it

0 , 0 , h0

r 00 , 0i
a, f
global
geodetic

gg

local

instantaneous
terrestrial

, , H

instantaneous
local
astronomic
h

49

4 Rotation

ncp0

nep

ncpT

ecliptic
0

T
+

z
true equator
@ epoch T

mean
equator
@ epoch T0
mean
equator
@ epoch T

Figure 4.1: Motion of the true and mean equinox along the ecliptic under the influence of
precession and nutation. This graph visualizes the rotation matrices P and N of 4.3.2. Note
that the drawing is incorrect or misleading to the extent that i) The precession and nutation
angles are grossly exaggerated compared to the obliquity , and ii) ncp0 and ncpT should be on
an ecliptical latitude circle 90 . That means that they should be on a curve parallel to the
ecliptic, around nep.

50

5 Gravity
5.1 Gravity attraction and potential
Suppose we are doing gravitational measurements at a fixed location on the surface of the
Earth. So r e = 0 and the Coriolis acceleration in (4.7) vanishes. The only remaining term
is the centrifugal acceleration ac , specified in the e-frame by: ac = 2 (xe , ye , 0)T . Since
this acceleration is always present, it is usually added to the gravitational attraction.
The sum is called gravity:
gravity = gravitational attraction + centrifugal acceleration
g = a + ac .
The gravitational attraction field was seen to be curl-free ( a = 0) in chapter 3. If
the curl of the centrifugal acceleration is zero as well, the gravity field would be curl-free,
too.

ze

ac
a
g

xe

Figure 5.1: Gravity is the sum of gravitational attraction


and centrifugal acceleration. Note that ac is hugely exaggerated. The centrifugal acceleration vector is about 3 orders
of magnitude smaller than the gravitational attraction.

Applying the curl operator () to the centrifugal acceleration field ac = 2 (xe , ye , 0)T
obviously yields a zero vector. In other words, the centrifugal acceleration is conservative.
Therefore, a corresponding centrifugal potential must exist. Indeed, it is easy to see that
this must be Vc , defined as follows:

xe
1

Vc = 2 (x2e + ye2 ) = ac = Vc = 2 ye .
2
0

(5.1)

51

5 Gravity
Correspondingly a gravity potential is defined
gravity potential = gravitational potential + centrifugal potential
W = V + Vc .

ze

ze

xt (North)
zt (up)
r sin

ac

r
xe

xe

Figure 5.2: Centrifugal acceleration in Earth-fixed and in topocentric frames.

Centrifugal acceleration in the local frame. Since geodetic observations are usually
made in a local frame, it makes sense to express the centrifugal acceleration in the
following topocentric frame (t-frame):
x-axis tangent to the local meridian, pointing North,
y-axis tangent to shperical latitude circle, pointing East, and
z-axis complementary in left-handed sense, point up.
Note that this is is a left-handed frame. Since it is defined on a sphere, the t-frame can
be considered as a spherical approximation of the local astronomic g-frame. Vectors in
the Earth-fixed frame are transformed into this frame by the sequence:

cos cos cos sin sin

cos
0 re ,
r t = P1 R2 ()R3 ()r e = sin
sin cos sin sin cos

(5.2)

in which is the longitude and the co-latitude. The mirroring matrix P1 = diag(1, 1, 1)
is required to go from a right-handed into a left-handed frame. Note that we did not include a translation vector to go from geocenter to topocenter. We are only interested in
directions here. Applying the transformation now to the centrifugal acceleration vector

52

5.1 Gravity attraction and potential


in the e-frame yields:

cos
cos sin
sin cos

2
0 .
0
=
r
sin

ac,t = P1 R2 ()R3 ()r 2 sin sin = r 2

sin
sin2
0
(5.3)
The centrifugal acceleration in the local frame shows no East-West component. On the
Northern hemisphere the centrifugal acceleration has a South pointing component. For
gravity purposes, the vertical component r 2 sin2 is the most important. It is always
pointing up (thus reducing the gravitational attraction). It reaches its maximum at the
equator and is zero at the poles.
Remark 5.1 This same result could have been obtained by writing the centrifugal potential in spherical coordinates: Vc = 21 2 r2 sin2 and applying the gradient operator in
spherical coordinates, corresponding to the local North-East-Up frame:


,
,
t Vc =
r r sin r

T 

1 2 2 2
r sin
2

Exercise 5.1 Calculate the centrifugal potential and the zenith angle of the centrifugal
acceleration in Calgary ( = 39 ). What is the centrifugal effect on a gravity measurement?
Exercise 5.2 Space agencies prefer launch sites close to the equator. Calculate the
weight reduction of a 10-ton rocket at the equator relative to a Calgary launch site.
The E
otv
os correction
As soon as gravity measurements are performed on a moving platform the Coriolis acceleration plays a role. In the e-frame it is given by (4.7a) as aCor.,e = 2(y e , x e , 0)T .
Again, in order to investigate its effect on local measurements, it makes sense to transform and evaluate the acceleration in a local frame. Let us first write the velocity in
spherical coordinates:

sin cos

r e = r sin sin
cos

(5.4a)

cos cos
sin sin

r e = r cos sin + r sin cos


sin
0

cos cos vN sin vE

= cos sin vN + cos vE


sin vN

(5.4b)

53

5 Gravity
It is assumed here that there is no radial velocity, i.e. r = 0. The quantities vN and vE
are the velocities in North and East directions, respectively, given by:
vN = r

and vE = r sin .

Now the Coriolis acceleration becomes:


aCor.,e

cos sin vN + cos vE

= 2 cos cos vN + sin vE .


0

Although we use North and East velocities, this acceleration is still in the Earth-fixed
frame. Similar to the previous transformation of the centrifugal acceleration, the Coriolis
acceleration is transformed into the local frame according to (5.2):

aCor.,t = P1 R2 ()R3 ()aCor.,e

cos vE

= 2 cos vN .
sin vE

(5.5)

The most important result of this derivation is, that horizontal motionto be precise the
velocity component in East-West directioncauses a vertical acceleration. This effect
can be interpreted as a secondary centrifugal effect. Moving in East-direction the actual
rotation would be faster than the Earths: 0 = + d with d = vE /(r sin ). This
would give the following modification in the vertical centrifugal acceleration:
a0c = 02 r sin2 = ( + d)2 r sin2
2 r sin2 + 2dr sin2 = ac + 2

vE
r sin2
r sin

= ac + 2vE sin .
The additional term 2vE sin is indeed the vertical component of the Coriolis acceleration. This effect must be accounted for when doing gravity measurements on a moving
platform. The reduction of the vertical Coriolis effect is called E
otv
os1 correction.
As an example suppose a ship sails in East-West direction at a speed of 11 knots (
20 km/h) at co-latitude = 60 . The E
otv
os correction becomes 70 105 m/s2 =
70 mGal, which is significant. A North- or Southbound ship is not affected by this
effect.
Remark 5.2 The horizontal components of the Coriolis acceleration are familiar from
weather patterns and ocean circulation. At the northern hemisphere, a velocity in East
direction produces an acceleration in South direction; a North velocity produces an
acceleration in East direction, and so on.
1

L
or
and (Roland) E
otv
os (18481919), Hungarian experimental physicist, widely known for his gravitational research using a torsion balance.

54

5.1 Gravity attraction and potential

NP

aCor
vE

aCor
- vN

vN

aCor
equator

aCor
- vN

vN
aCor
aCor

vE

SP
longitude

Figure 5.3: Horizontal components of the Coriolis acceleration.

Remark 5.3 The Coriolis acceleration can be interpreted in terms of angular momentum
conservation. Consider a mass of air sitting on the surface of the Earth. Because of Earth
rotation it has a certain angular momentum. When it travels North, the mass would
get closer to the spin axis, which would imply a reduced moment of inertia and hence
reduced angular momentum. Because of the conservation of angular momentum the air
mass needs to accelerate in East direction.
Exercise 5.3 Suppose you are doing airborne gravimetry. You are flying with a constant
400 km/h in Eastward direction. Calculate the horizontal Coriolis acceleration. How large
is the E
otv
os correction? How accurately do you need to determine your velocity to
have the Eotv
os correction precise down to 1 mGal? Do the same calculations for a
Northbound flight-path.

55

6 The normal field


Geodetic observables depend on the geometry (r) and the gravity field (W ) of the Earth.
In general the functional relation will be nonlinear:
f = f (r, W ) .
The standard procedure is to develop the observable into a Taylor1 series and to truncate
after the linear term. As a result we obtain a linear observation equation. In order to
perform this linearization we need a proper approximation of both geometry and gravity
field of the Earth. Approximation by a sphere would be too inaccurate. The equatorial
radius of the Earth is some 21.5 km larger than its polar radius. A rotationally symmetric
ellipsoid is accurate enough, though. The geoid, which represents the physical shape of
the Earth, doesnt deviate more than 100 m from the ellipsoid.
The potential and the gravity field that are consistent with such an ellipsoid are called
normal potential and normal gravity. Thus, the normal field is an ellipsoidal approximation to the real gravity field. For the actual gravity potential we have the following
linearization:
W = U +T

(6.1)

with: W = full gravity potential


U = normal potential (W0 )
T = disturbing potential (W )

Physical geodesy is a global discipline by nature. Therefore one has to make sure that the
same normal field is used by everybody everywhere. This has been strongly advocated
by the International Association of Geodesy (iag) over the past century. As a result a
number of commonly accepted so-called Geodetic Reference Systems (grs) have evolved
over the years: grs30, grs67 and the current grs80, see tbl. 6.1. The parameters of the
latter have been adopted by many global and regional coordinate systems and datums.
For instance the wgs84 useswith some insignificant changesthe grs80 parameters.
1

Brook Taylor (16851731).

56

6.1 Normal potential

Table 6.1: Basic parameters of normal fields

[m]

f 1
GM0

[1014 m3 s2 ]
[105 rad s1 ]

grs30

grs67

grs80

6378 388
297
3.986 329
7.292 1151

6378 160
298.247 167
3.986 030
7.292 115 1467

6378 137
298.257 222
3.986 005
7.292 115

6.1 Normal potential


The geometry of the normal field, i.e. the ellipsoid is determined by two parameters for
size and shape. We will choose the semi-major axis (a) and the flattening (f ). The
description of the physical field, i.e. the normal gravity potential, requires two further
parameters. The strength is given by the geocentric gravitational constant (GM0 ). And
since were dealing with gravity the Earth rotation rate () must be involved, too.
This basic set of 4 parameters defines the normal field fully. See tbl. 6.1 for some
examples. But vice versa, any set of 4 independent parameters will do. For grs80 one
uses the dynamical form factor J2 (to be explained later on) instead of the geometric
form factor f . The set a, J2 , GM0 , are called the defining constants.
The normal potential is defined to have the following properties:
it is rotationally symmetric (zonal),
it has equatorial symmetry,
it is constant on the ellipsoid.
The latter property is the most fundamental. It defines the rotating Earth ellipsoid to
be an equipotential surface or level surface.
This set of properties provides an algorithm to derive the normal potential and gravity
formulae. Starting point is the sh development of a potential like (??), together with
the centrifugal potential from 5.1. Together they give the normal potential U :
U (r, , ) =

l
 l+1 X
R
1
GM0 X
Plm (cos ) (clm cos m + slm sin m) + 2 r2 sin2 .
R l=0 r
2
m=0

Since we only want to represent the normal potential on and outside the ellipsoid, its
mass distribution is irrelevant. For the following development it will be useful to assume
all masses to be contained in a sphere of radius a. With the first property, rotational
symmetry, we get the following simplification:
U (r, ) =

 l+1
1
a
GM0 X
cl,0 Pl (cos ) + 2 r2 sin2 .
a l=0 r
2

(6.2)

57

6 The normal field


The next propertyequatorial symmetryreduces the series to even degrees only. Actually only terms up to degree 8 are required. We thus get:
 l+1
8
a
1
GM0 X
cl,0 Pl (cos ) + 2 r2 sin2 .
U (r, ) =
a l=0,[2] r
2

(6.3)

For a better understandingand easier calculuswe will continue now with only the
degree 0 and 2 terms. The purpose is to derive an approximation, linear in f and c2,0 . We
must keep in mind, though, that the actual development should run to degree 8. With
the Legendre polynomial P2 written out, and the centrifugal potential within brackets,
we get:
GM0
U (r, ) =
a

" 

a
c0,0 +
r

 3

a
r

1 2 r2 a
1
sin2 .
c2,0 (3 cos2 1) +
2
2 GM0

(6.4)

In order to impose the main requirementthat of an equipotential ellipsoidwe must


evaluate (6.4) on the ellipsoid. Thus we need an expression for the radius of the ellipsoid
as a function of . The exact form is:
r() = p

ab
a2 cos2

b2 sin2

a
,
1 + e02 cos2

which is easily verified by inserting x = r sin cos , etc. in the equation of the ellipsoid:
x2 + y 2 z 2
+ 2 = 1.
a2
b
Since our goal is a formulation, linear in f , we expand (a/r) in a binomial series:
q(q 1) 2
x + ...
2!

1
1
1 + x = 1 + x x2 + . . .
2
8
p
1
1 + e02 cos2 = 1 + e02 cos2 . . .

2
a

= 1 + f cos2 + O(f 2 ) .
r
(1 + x)q = 1 + qx +

(6.5a)

The last step was due to the fact that


e02 =

a2 b2
ab a+ba
=
= 2f + O(f 2 ) .
2
b
a
b
b
| {z }
f

Similarly, we can derive:

 2

a
r

58

= 1 + 2f cos2 + O(f 2 ) .

(6.5b)

6.2 Normal gravity


The coefficient c0,0 1. If we would insert (6.5a) into (6.4) we recognize that U depends
on 3 small quantities, all of the same order of magnitude:

m=

f 0.003 ,

(6.6a)

c2,0 0.001 ,

(6.6b)

0.003 .

(6.6c)

2 a3
GM0

The quantity m is the relative strength of the centrifugal acceleration (at the equator)
compared to gravitation. Now we will insert (6.5a) into (6.4) indeed, making use of f ,
c2,0 and m. We will neglect all terms that are quadratic in these quantities, i.e. f 2 , f c2,0 ,
f m, c22,0 and c2,0 m. We then get


GM0
1
1
(1 + f cos2 ) + c2,0 (3 cos2 1) + m sin2 ,
a
2
2




1
1
3
1
GM0
2
1 c2,0 + m + f + c2,0 m cos .
=
a
2
2
2
2

U =

(6.7)

This normal potential still depends on , which contradicts the requirement of a constant
potential. The latitude dependence only disappears if the following condition between
f , c2,0 and m holds:
f + 32 c2,0 = 21 m ,

(6.8)

which means that the three quantities (6.6) cannot be independent. Using (6.8), we can
eliminate one of the three small quantities. The constant normal potential value U0 on
the ellipsoid can be written as:


GM0
1
1
1
GM0
1
GM0
U0 =
1 c2,0 + m =
1+ f + m =
1 + f + c2,0
a
2
2
a
3
3
a

. (6.9)

Outside the ellipsoid one has to make use of (6.4) and apply the condition (6.8) to
eliminate one of the three small quantities, e.g. c2,0 :
"

GM0 a
U (r, ) =
+
a
r

 3 

a
r

1
mf
2



1
cos
3
2

1
+
2

 2

r
a

m sin .

(6.10)

Keep in mind that this is a linear approximation. For precise calculations one should
revert to (6.3).

6.2 Normal gravity


Within the linear approximation of the previous section we can define normal gravity
as the negative of the radial derivative of the normal potential. From (6.10) we get the

59

6 The normal field


following normal gravity outside the ellipsoid:
"

GM0 a
3
U
=
+
(r, ) =
2
r
a
r
r

 3 

a
r

1
mf
2



1
cos
3
2

r
2 m sin2 . (6.11)
a

On the surface of the ellipsoid, using the same approximations as in the previous section,
we will have:


5
GM0
2
(6.12)
() = 2 1 + m + (f m) sin .
a
2
We cannot have a constant normal gravity on the surface of the ellipsoid simultaneously
with a constant normal potential. Thus the -dependency remains. If we evaluate (6.12)
on the equator and on the pole, we get the values:
equator:
poles:

GM0
(1 + f 32 m) ,
a2
GM0
b =
(1 + m) .
a2

a =

(6.13a)
(6.13b)

Note that b > a since the pole is closer to the Earths center of mass. Similar to the
geometric flattening f = (a b)/a we now define the gravity flattening:
b a
.
a

f =

(6.14)

Numerically f is approximately 0.005, i.e. the same size as the other three small
quantities. If we now insert the normal gravity on equator and pole (6.13) into this
gravity flattening formula we end up with:
f =

5
2m

f
5
mf ,
3
2
1 + f 2m

leading to the remarkable result:


f + f = 25 m .

(6.15)

This result is known as the theorem of Clairaut2 . It is remarkable, because it relates a


dynamic quantity (f ) to a geometric quantity (f ) and the Earths rotation (through
m) in such a simple way.
Although (6.11) describes normal gravity outside the ellipsoid, it would be more practical
to be able to upward continue the normal gravity value on the ellipsoid, i.e. to have a
formula like (h, ) = ()g(h), in which g(h) is some function of the height of above
the ellipsoid. This is achieved by a Taylor series:

Alexis Claude Clairaut (17131765).

60


1 2
(h) = (h = 0) +
h
+
h2 . . . .

h h=0
2 h2 h=0

6.3 Adopted normal gravity


After some derivation one ends up with the following upward continuation:


3
2
(h, ) = () 1 (1 + f + m 2f cos2 )h + 2 h2 .
a
a

(6.16)

6.3 Adopted normal gravity


Until (6.4) the development of the normal field was strictly valid. Starting with (6.4)
approximations were introduced, such that we ended up with expressions in which all
quadratic terms in f , m and c2,0 were neglected. Even worse: for the upward continuation
of the normal gravity a Taylor expansion was introduced.
Below, the exact analytical and precise numerical formulae are given. Formulae for an
exact upward continuation do exist, but are not treated here.

6.3.1 Formulae
The theory of the equipotential ellipsoid was first given by Pizzetti3 in 1894. It was
further elaborated by Somigliana4 in 1929. The following formula for normal gravity is
generally valid. It is called the Somigliana-Pizzetti normal gravity formula:

with

aa cos2 + bb sin2
1 + k sin2
() = q
= a q
,
a2 cos2 + b2 sin2
1 e2 sin2
e2 =

a2 b2
a2

and k =

(6.17)

bb aa
.
aa

The variable is the geodetic latitude. In case of grs80 the constants a, b, a , b , e2 and
k can be taken from the list in the following section.
For grs80 the following series expansion is used:


() = a 1 + 0.005 279 0414 sin2


+ 0.000 023 2718 sin4
+ 0.000 000 1262 sin6


+ 0.000 000 0007 sin8 .

(6.18a)

It has a relative error of 1010 , corresponding to 104 mGal. For most applications, the
following formula, which has an accuracy of 0.1 mGal, will be sufficient:


() = 9.780 327 1 + 0.005 3024 sin2 0.000 0058 sin2 2


3
4

[m s2 ] .

(6.18b)

Paolo Pizzetti (1860-1918), Italian geodesist.


Carlo Somigliana (1860-1955), Italian mathematical physicist.

61

6 The normal field

GRS80 normal gravity


9.84

9.83

[m/s ]

9.82
9.81
9.8
9.79
9.78

9.77

30

60

90
120
colatitude [deg]

150

180

Figure 6.1: grs80 normal gravity.

Conversion between GRS30, GRS67 and GRS80. For converting gravity anomalies
from the International Gravity Formula (1930) to the Gravity Formula 1980 we can use:


1980 1930 = 16.3 + 13.7 sin2

[mGal] ,

(6.19a)

where the main part comes from a change of the Potsdam reference value by -14 mGal.
For the conversion from the Gravity Formula 1967 to the Gravity Formula 1980, a more
accurate formula, corresponding to the precise expansion given above, is:


1980 1967 = 0.8316 + 0.0782 sin2 0.0007 sin4

[mGal] .

(6.19b)

6.3.2 GRS80 constants


Defining constants (exact)
a = 6378 137 m

semi-major axis

GM0 = 3.986 005 1014 m3 s2 geocentric gravitational constant


J2 = 0.001 082 63

dynamic form factor

= 7.292 115 105 rad s1 angular velocity


The following derived constants are accurate to the number of decimal places given. In
case of doubt or in those cases where a higher accuracy is required, these quantities are
to be computed from the defining constants.

62

6.3 Adopted normal gravity


Derived geometric constants
b = 6356 752.3141 m

semi-minor axis

E = 521 854.0097 m

linear eccentricity

c = 6399 593.6259 m

polar radius of curvature

e2 = 0.006 694 380 022 90 first eccentricity (e)


e02 = 0.006 739 496 775 48 second eccentricity (e0 )
f = 0.003 352 810 681 18 flattening
f 1 = 298.257 222 101

reciprocal flattening

Q = 10 001 965.7293 m

meridian quadrant

R1 = 6371 008.7714 m

mean radius R1 = (2a + b)/3

R2 = 6371 007.1810 m

radius of sphere of same surface

R3 = 6371 000.7900 m

radius of sphere of same volume

Derived physical constants


U0 = 62 636 860.850 m2 s2 normal potential on ellipsoid
J4 = 0.000 002 370 912 22 spherical
J6 =

0.000 000 006 083 47

J8 = 0.000 000 000 014 27


m=

harmonic
coefficients

0.003 449 786 003 08 m = 2 a2 b/GM0

a = 9.780 326 7715 m s2 normal gravity at equator


b = 9.832 186 3685 m s2 normal gravity at poles
m = 9.797 644 656 m s2

average of normal gravity over ellipsoid

45 = 9.806 199 203 m s2

normal gravity = 45

f = 0.005 302 440 112


k = 0.001 931 851 353

f = (b a )/a
k = (bb aa )/aa

63

7 Height Systems
7.1 Height Systems
7.1.1 Raw levelled heights
Levelled heights are non-unique. Depending on the levelling path one ends up with
different heights for the same point. They have no physical relevance.
l0P =

n
X

li

i=1

with: l0P = levelled height difference between geoid and point P


li = levelled height differences (increments)

7.1.2 Geopotential numbers


Geopotential numbers are the basic height information. They are unique, i.e. pathindependent.
CP = W0 WP =
with:

ZP
0

g dx =

ZP

gdH

CP = geopotential number of point P


W0 = potential on geoid
WP = potential in point P
g, g = vector and scalar gravity

7.1.3 Dynamic height


Dynamic heights H dyn are just geopotential numbers, scaled into a metric unit. Differences with respect to raw levelled heights can become large, depending on how well 0
represents gravity in the levelled area.
HPdyn =

64

CP
0

7.1 Height Systems


0 = constant normal gravity value, e.g. (45 )

with:

7.1.4 Orthometric height


The orthometric height HP is the length of the plumbline between surface point P and
the geoid, loosely speaking the height above the geoid.
CP
g
g = mean gravity along plumbline between P and geoid.

HP =
with:

Orthometric heights require knowledge or assumptions of g inside the Earth. Gravity


will behave rather linearly between P and geoid. Thus the average would be close to the
value at half height ( 12 H). Assuming a crustal density of = 2670 kg/m3 and a free-air
gradient of 0.3086 mGal/m, we get the so-called
1
Prey-reduction: g = g( H) = gP + 0.0424H ,
2
in which the coefficient of H, which is (bo + 21 fa), is in mGal/m. The procedure is to
remove a Bouguer plate of thickness 12 H, downward continue over a vertical distance
1
2 H and restore the plate again. The orthometric heights that use this mean gravity are
called Helmert heights:
CP
HP =
gP + 0.0424H
We see that H occurs left and right. But since the Prey gradient is small, the accuracy
of H in the denominator doesnt matter. One could use the raw levelled height andif
necessarydo an iteration.
So three assumptions are used in deriving Helmert orthometric heights:
i) linear behaviour of gravity between surface and geoid,
ii) a constant crustal density of = 2670 kg/m3 ,
iii) a fixed free-air gradient of 0.3086 mGal/m.

7.1.5 Normal heights


Normal heights H n do not have this problem. They are defined without any assumption
as:
CP

= mean normal gravity along normal plumbline between P and quasi-geoid.

HPn =
with:

Normal gravity can be calculated at any point without hypothesis. Again, the normal
gravity behaviour is very linear. The mean gravity value along the normal plumbline can

65

7 Height Systems
therefore be approximated by the normal gravity valued at the point which has height
1 n
2 H above the ellipsoid. Neglecting second order terms in height we get:
1
= (P , HPn ) = (P , 0) 0.1543HPn
2
again with the coefficient in

mGal/m.

Thus we end up with

HPn =

P0

CP
,
0.1543H n

which may be solved by inserting raw levelling heights in the denominator or by iteration.
The normal height HPn is interpreted as the height of the surface point P above the quasigeoid. Alternatively it is interpreted as the height of the telluroid above the ellipsoid.
They are related to Molodenskiis gravity field theory.

7.2 Height computations and corrections


Discretization Levelling and gravimetry produce a discrete set of data. The integral
gdH needs to be discretized, which introduces a small error:

CQP = CP CQ =

ZP

gdH

n
X

gi li

i=1

Instead of using the above equations for height calculations, it would be easier to be
able to use levelled height differences and correct them in some way. For a given height
system, we would have formulas of the type:
HQP = HP HQ = lQP + small correction term.

7.2.1 Dynamic heights

dyn
HQP
= lQP + dcQP

dynamic correction:

dcQP =

X gi 0
i

7.2.2 Orthometric heights

HQP = lQP + ocQP

66

li

7.3 Normal vs. true heights


orthometric correction:

ocQP = dcQP + dcP0 P dcQ0 Q


X gi 0
gQ 0
gP 0
li
HQ +
HP
=

0
0
0
i

7.2.3 Normal heights

n
= lQP + ncQP
HQP

normal correction:

ncQP =

X gi 0

li

Q 0 n P 0 n
HQ +
HP
0
0

In all these equations gi is the actual gravity along the levelling line, 0 is the fixed normal
gravity used in the dynamic heights and gP and P respectively denote the mean gravity
and normal gravity along (normal) plumbline.

7.3 Normal vs. true heights


The above height systems are true heights in the sense that they make use of gravity
information,
together with levelling. The only approximation made is the discretization
R
of gdH. If gravity is not available, the best guess would be to use normal gravity
along the levelling line. This can be done for all height systems. It would produce nor
mal geopotential numbers CP , normal dynamic heights HPdyn and normal orthometric
heights HP , together with their corrections ndc, noc.
As an example the normal orthometric height is defined as
normal geopotential number

CP

ZP
0

normal orthometric height


normal orth. height diff.
normal orth. correction

HP =

(, h)dH

(i , hi )li =

i l i

CP
P

HQP
= lQP + nocQP

nocQP =

X i 0
i

li

Q 0
P 0
HQ +
HP
0
0

After some manipulation the normal orthometric correction between two successive levelling points i and j is simplified to:
normal orth. correction
with

ij ij sin 2ij
nocij 0.0053H
ij = 1 (Hi + Hj )
H
2

from raw heights

67

7 Height Systems
ij = j i

1
ij = (i + j )
2

7.4 Vertical reference systems in use in North America


CVGS28
Canadian Vertical Geodetic System 1928
normal orthometric heights
official Canadian height system
overconstrained datum: 4 tide gauges at the East coast and 2 at the West coast
have been assigned the height 0.
system has become obsolete, a new system is underway
NAVD88
North American Vertical Datum 1988
Helmert orthometric heights
official US height system
IGLD55
International Great Lakes Datum 1955
dynamic heights
height system in use around Great Lakes and St. Lawrence river
main purpose: water management, civil/hydraulic engineering

68

You might also like