PHYSICAL
GEODESY
Aalto University publication series
Physical geodesy
Martin Vermeer
Aalto University
School of Engineering
Department of Built Environment
Aalto University publication series
SCIENCE + TECHNOLOGY 2/2020
Helsinki 2020
Finland
Abstract
Aalto University, P.O. Box 11000, FI-00076 Aalto www.aalto.fi
Author
Martin Vermeer
Name of the publication
Physical geodesy
Publisher School of Engineering
Unit Department of Built Environment
Series Aalto University publication series SCIENCE + TECHNOLOGY 2/2020
Field of research Geodesy
Language English
Abstract
Physical geodesy studies the large-scale figure and gravity field of the Earth, which are
closely related. Our understanding of the gravity field is based on Newton’s theory of
gravitation. We present field theory, with partial differential equations describing the
behaviour of the field throughout space. Techniques for solving these equations using
boundary conditions on the Earth’s surface are explained. A central concept is the
geopotential.
The figure of the Earth is approximated by an ellipsoid of revolution, after which the
precise figure is described by small deviations from this ellipsoid. Vertical reference
systems are discussed in this context. Extending the approach to the Earth’s gravity
field yields small difference quantities, such as the disturbing potential and gravity
anomalies.
Approaches to modelling the gravity field explained are spectral development of the
field using spherical harmonics, the Stokes equation, numerical techniques based on
the Fast Fourier Transform, the remove-restore technique, and least-squares
collocation. Gravity measurement techniques are discussed, as are the multiple links
with geophysics, such as terrain effects, isostasy, mean sea level and the sea level
equation, and the tides.
Keywords figure of the Earth, gravity field, geopotential, reference ellipsoid, normal field,
disturbing potential, gravity anomaly, geoid, height system, spherical harmonics,
Stokes equation, remove-restore, least-squares collocation, gravimetry, isostasy,
mean sea level, tides
ISBN (pdf) 978-952-60-8940-9
ISSN (PDF) 1799-490X
Location of publisher Helsinki Location of printing Helsinki Year 2020
Pages 524 urn http://urn.fi/URN:ISBN:978-952-60-8940-9
Preface
This book aims to present an overview of the current state of the study
of the Earth’s gravity field and those parts of geophysics closely related
to it, especially geodynamics, the study of the changing Earth. It has
grown out of over two decades of teaching at Helsinki’s two universities:
the Helsinki University of Technology — today absorbed into Aalto
University — and the University of Helsinki. As such, it presents
a somewhat Fennoscandian perspective on a very global subject. In
addition, the author’s own research on gravimetric geoid determination
helped shape the presentation. While there exist excellent textbooks on
all the different parts of what is presented here, I may still hope that
this text will find a niche to fill.
Martin Vermeer
–i–
ii Preface
Acknowledgements
Thanks are due to the many students and colleagues, both in academia
and at the Finnish Geodetic Institute, who have given useful comments
and corrections over the course of many years of lecturing both at the
University of Helsinki and at the Helsinki University of Technology,
today Aalto University.
Special thanks are due to the foreign students at Aalto University,
who forced me during recent years to provide an English version of this
text. The translation work also prompted a basic revision of the Finnish
text, which was long overdue as parts were written in the 1990s before
the author had had the benefit of pedagogical training. Thanks are thus
also due to Aalto University’s pedagogical training programme.
Olivier Francis is gratefully acknowledged for contributing figure
11.8.
Several map images were drawn using the Generic Mapping Tools
(Wessel et al., 2013).
The English language was competently checked by the Finnish Trans-
lation Agency Aakkosto Oy. Tarja Paalanen designed the cover. Laura
Mure and Matti Yrjölä helped with the practicalities of publishing.
This content is licenced under the Creative Commons Attribution 4.0
International (CC BY 4.0) licence, except as noted in the text or otherwise
apparent.
í ¤. û
Contents
Chapters
Õ ! 1. Fundamentals of the theory of gravitation . . . . . . . . . . 1
Õ ! 2. The Laplace equation and its solutions . . . . . . . . . . . . 41
Õ ! 3. Legendre functions and spherical harmonics . . . . . . . . . 57
Õ ! 4. The normal gravity field . . . . . . . . . . . . . . . . . . . . . 87
Õ ! 5. Anomalous quantities of the gravity field . . . . . . . . . . . 111
Õ ! 6. Geophysical reductions . . . . . . . . . . . . . . . . . . . . . 129
Õ ! 7. Vertical reference systems . . . . . . . . . . . . . . . . . . . . 163
Õ ! 8. The Stokes equation and other integral equations . . . . . . 191
Õ ! 9. Spectral techniques, FFT . . . . . . . . . . . . . . . . . . . . . 231
Õ ! 10. Statistical methods . . . . . . . . . . . . . . . . . . . . . . . . 255
Õ ! 11. Gravimetric measurement devices . . . . . . . . . . . . . . . 299
Õ ! 12. The geoid, mean sea level, and sea-surface topography . . . 329
Õ ! 13. Satellite altimetry and satellite gravity missions . . . . . . . 351
Õ ! 14. Tides, the atmosphere, and Earth crustal movements . . . . 385
Õ ! 15. Earth gravity field research . . . . . . . . . . . . . . . . . . . 401
– iii –
iv Contents
Preface i
List of Tables xi
Acronyms xvii
í ¤. û
Contents
v
2.2 The Laplace equation in rectangular co-ordinates . . . 43
2.3 The Laplace equation in polar co-ordinates . . . . . . . 48
2.4 Spherical, geodetic, ellipsoidal co-ordinates . . . . . . . 50
2.5 The Laplace equation in spherical co-ordinates . . . . . 53
2.6 Dependence on height . . . . . . . . . . . . . . . . . . . 54
Self-test questions . . . . . . . . . . . . . . . . . . . . . . . . . . 55
í ¤. û
vi Contents
í ¤. û
Contents
vii
7. Vertical reference systems 163
7.1 Levelling, orthometric heights and the geoid . . . . . . 163
7.2 Orthometric heights . . . . . . . . . . . . . . . . . . . . 165
7.3 Normal heights . . . . . . . . . . . . . . . . . . . . . . . 168
7.4 Difference between geoid height and height anomaly . 175
7.5 Difference between orthometric and normal heights . . 178
7.6 Calculating orthometric heights precisely . . . . . . . . 178
7.7 Calculating normal heights precisely . . . . . . . . . . . 181
7.8 Calculation example for heights . . . . . . . . . . . . . 181
7.9 Orthometric and normal corrections . . . . . . . . . . . 183
7.10 A vision for the future: relativistic levelling . . . . . . . 186
Self-test questions . . . . . . . . . . . . . . . . . . . . . . . . . . 188
Exercise 7–1: Calculating orthometric heights . . . . . . . . . 188
Exercise 7–2: Calculating normal heights . . . . . . . . . . . . 188
Exercise 7–3: Difference between orthometric and normal
height . . . . . . . . . . . . . . . . . . . . . . . . 189
í ¤. û
viii Contents
í ¤. û
Contents
ix
11.2 The relative or spring gravimeter . . . . . . . . . . . . . 301
11.3 The absolute or ballistic gravimeter . . . . . . . . . . . 308
11.4 Network hierarchy in gravimetry . . . . . . . . . . . . . 313
11.5 The superconducting gravimeter . . . . . . . . . . . . . 314
11.6 Gravity measurement and the atmosphere . . . . . . . 317
11.7 Airborne gravimetry and GNSS . . . . . . . . . . . . . 319
11.8 Measuring the gravity gradient . . . . . . . . . . . . . . 322
Self-test questions . . . . . . . . . . . . . . . . . . . . . . . . . . 324
Exercise 11–1: Absolute gravimeter . . . . . . . . . . . . . . . 325
Exercise 11–2: Spring gravimeter . . . . . . . . . . . . . . . . . 326
Exercise 11–3: Air pressure and gravity . . . . . . . . . . . . . 326
12. The geoid, mean sea level, and sea-surface topography 329
12.1 Basic concepts . . . . . . . . . . . . . . . . . . . . . . . . 329
12.2 Geoid models and national height datums . . . . . . . 331
12.3 The geoid and post-glacial land uplift . . . . . . . . . . 333
12.4 Determining the sea-surface topography . . . . . . . . 336
12.5 Global sea-surface topography and heat transport . . . 337
12.6 The global behaviour of the sea level . . . . . . . . . . . 340
12.7 The sea-level equation . . . . . . . . . . . . . . . . . . . 342
Self-test questions . . . . . . . . . . . . . . . . . . . . . . . . . . 347
Exercise 12–1: Coriolis force, ocean current . . . . . . . . . . . 348
Exercise 12–2: Land subsidence and the mechanism of land
uplift . . . . . . . . . . . . . . . . . . . . . . . . 349
í ¤. û
x Contents
í ¤. û
List of Tables
xi
Self-test questions . . . . . . . . . . . . . . . . . . . . . . . . . . 437
Exercise B–1: Orthonormality of the Fourier basis functions . 437
Bibliography 459
Index 479
List of Tables
3.1 Legendre polynomials . . . . . . . . . . . . . . . . . . . . 58
3.2 Associated Legendre functions . . . . . . . . . . . . . . . 60
3.3 Semi-wavelengths for different degrees and orders . . . . 64
3.4 Plotting a surface spherical-harmonic map . . . . . . . . . 66
3.5 EGM96 coefficients and mean errors . . . . . . . . . . . . 78
3.6 Legendre functions of the second kind . . . . . . . . . . . 79
í ¤. û
xii List of Figures
List of Figures
1.1 Gravitation is universal . . . . . . . . . . . . . . . . . . . . 2
1.2 A thin spherical shell consists of rings . . . . . . . . . . . 7
1.3 Dependence of potential and attraction on distance . . . 9
1.4 A double mass-density layer . . . . . . . . . . . . . . . . . 20
1.5 A graphical explanation of the Gauss divergence theorem 23
1.6 A little rectangular box . . . . . . . . . . . . . . . . . . . . 25
1.7 Eight-unit cube . . . . . . . . . . . . . . . . . . . . . . . . 27
1.8 Green’s third theorem for an exterior point . . . . . . . . 30
1.9 Green’s third theorem for an interior point . . . . . . . . . 31
1.10 Green’s third theorem for the space external to a body . . 32
1.11 Iron ore body . . . . . . . . . . . . . . . . . . . . . . . . . 38
í ¤. û
List of Figures
xiii
4.1 The normal gravity field of the Earth . . . . . . . . . . . . 88
4.2 Gravitation and centrifugal force . . . . . . . . . . . . . . 89
4.3 The curvature of level surfaces . . . . . . . . . . . . . . . 93
4.4 The curvature of the plumb line . . . . . . . . . . . . . . . 94
4.5 Natural co-ordinates . . . . . . . . . . . . . . . . . . . . . 96
4.6 Meridian ellipse and latitude types . . . . . . . . . . . . . 100
4.7 The normal field’s potential over the equator . . . . . . . 103
í ¤. û
xiv List of Figures
í ¤. û
List of Figures
xv
11.1 Jean Richer’s report . . . . . . . . . . . . . . . . . . . . . . 300
11.2 Autograv CG-5 spring gravimeter . . . . . . . . . . . . . . 301
11.3 Operating principle of a spring gravimeter . . . . . . . . 304
11.4 The idea of astatisation . . . . . . . . . . . . . . . . . . . . 307
11.5 Operating principle of a ballistic absolute gravimeter . . 309
11.6 Absolute gravimeter of type FG5 . . . . . . . . . . . . . . 310
11.7 The idea of operation of an atomic gravimeter . . . . . . . 314
11.8 International intercomparison of absolute gravimeters . . 315
11.9 Operating principle of a superconducting gravimeter . . 316
í ¤. û
xvi List of Figures
í ¤. û
Acronyms
ABCDEFGHIJKLMNOPRSTUW
A
ACS Advanced Camera for Surveys, instrument on the Hubble Space Telescope
2 ^
AGU American Geophysical Union 402 ^
B
BGI Bureau Gravimétrique International, International Gravity Bureau 126, 136,
154, 401, 402 ^
BVP boundary-value problem 34 ^
C
CHAMP Challenging Minisatellite Payload for Geophysical Research and Ap-
plications 77, 249, 377 ^
D
DMA Defense Mapping Agency (USA) 76 ^
DORIS Doppler Orbitography and Radiopositioning Integrated by Satellite, a
French satellite positioning system 352, 353 ^
DTM digital terrain model 135 ^
E
EGM2008 Earth Gravity Model 2008 63, 77, 120, 126, 136 ^
EGM96 Earth Gravity Model 1996 76–78 ^
– xvii –
xviii ABCDEFGHIJKLMNOPRSTUW Acronyms
F
FAS member of the French Academy of Sciences (Académie des sciences) 433,
434 ^
FFT Fast Fourier Transform 15, 153, 198, 214, 219, 234, 236–239, 241–247, 249,
250, 252, 253, 288, 291, 332, 439, 440 ^
FGI Finnish Geospatial Research Institute, formerly Finnish Geodetic Institute
404 ^
FIN2000 geoid model (Finland) 247, 248, 332 ^
FIN2005N00 geoid model (Finland) 247, 403 ^
FRAS Fellow of the Royal Astronomical Society 300 ^
FRS Fellow of the Royal Society (of London) 4, 17, 226, 299, 300, 390, 391, 433,
434 ^
FRSE Fellow of the Royal Society of Edinburgh 17, 390, 419, 434 ^
G
GDR geophysical data record 356, 373 ^
GFZ Geoforschungszentrum (Potsdam, Germany), German Research Centre for
Geosciences 377 ^
GIA glacial isostatic adjustment 333, 346 ^
GNSS Global Navigation Satellite Systems, which comprise, besides American
GPS, the satellite positioning systems of other countries, such as the
Russian GLONASS and the European Galileo 116, 130, 187, 246, 247,
320, 321, 332, 333, 337, 352, 356, 372, 376, 380, 393, 394, 398, 403 ^
GOCE Geopotential and Steady-state Ocean Circulation Explorer 77, 249, 288,
322, 339, 341, 376, 379–382 ^
GPS Global Positioning System 70, 101, 173, 319, 325, 352, 377, 378 ^
GRACE Gravity Recovery And Climate Experiment 77, 249, 377–380 ^
GRAVSOFT Geopotential determination software, mainly developed in Den-
mark 247 ^
GRS80 Geodetic Reference System 1980 6, 101, 104, 109, 112, 119, 247, 318, 332,
333 ^
í ¤. û
Acronyms ABCDEFGHIJKLMNOPRSTUW
xix
GWR20 superconducting gravimeter built by GWR Instruments 315 ^
H
HUT Helsinki University of Technology, today part of Aalto University 404 ^
I
IAG International Association of Geodesy 247, 315, 401, 402 ^
IB inverted barometer 329 ^
ICET International Center for Earth Tides 402 ^
ICGEM International Center for Global Earth Models 402 ^
IDEMS International Digital Elevation Model Service 402 ^
IGeC International Geoid Commission (obsolete) 402 ^
IGeS International Geoid Service (obsolete) 402 ^
IGETS International Geodynamics and Earth Tide Service 315 ^
IGFS International Gravity Field Service 401, 402 ^
IMGC-02 transportable absolute rise-and-fall gravimeter, built by the Istituto
di Metrologia «G. Colonnetti», formerly the Istituto Nazionale di Ricerca
Metrologica in Torino, Italy 312 ^
IOC Initial Operating / Operational Capability 325 ^
ISG International Service for the Geoid 401, 402 ^
IUGG International Union of Geodesy and Geophysics 401 ^
J
J2 The dynamic flattening (“gravitational flattening”) of the Earth 77, 101, 104,
367, 368, 382 ^
Jason American-French-European radar altimetry satellite series, successors
of TOPEX/Poseidon 352–354, 369, 382 ^
JHU Johns Hopkins University 2 ^
K
KKJ National Grid Co-ordinate System (Finland) 124 ^
L
LAGEOS LAser GEOdynamic Satellite 366 ^
Lego™ “Leg Godt”, Engl. “Play Well”, Danish toy brand 26 ^
LLR Lunar laser ranging 308 ^
í ¤. û
xx ABCDEFGHIJKLMNOPRSTUW Acronyms
M
Mf Moon, fortnightly tide 390 ^
N
N2000 height system (Finland) 165, 168, 247, 331 ^
N60 height system (Finland) 164, 165, 247, 331, 332 ^
NAP Normaal Amsterdams Peil, height system (Western Europe) 165 ^
NASA National Aeronautics and Space Administration (USA) 2 ^
NAVD88 North American Vertical Datum 1988 165 ^
NC normal correction 185 ^
NGA National Geospatial-Intelligence Agency (USA, formerly NIMA) 76,
401 ^
NIMA National Imagery and Mapping Agency (USA, formerly DMA) 76 ^
NKG Nordiska Kommissionen för Geodesi, Nordic Geodetic Commission 403 ^
NKG2004 geoid model (Nordic area) 403 ^
NKG2015 geoid model (Nordic area) 403 ^
NN height system (Finland) 165 ^
NOAA National Oceanic and Atmospheric Administration (USA) 341 ^
O
OCorthometric correction 183, 184 ^
OSU Ohio State University, Columbus, Ohio, USA 76 ^
P
ppb parts per billion 120 ^
ppm parts per million 120 ^
PRARE Precise Range And Range-Rate Equipment, not operational 352 ^
PRS President of the Royal Society 3, 122, 146, 419 ^
R
RTM residual terrain modelling 215, 217, 218, 229 ^
S
SI Système international d’unités, International System of Units 11, 37, 101, 120,
126 ^
í ¤. û
Acronyms ABCDEFGHIJKLMNOPRSTUW
xxi
SK-42 Reference system of the Soviet Union, also known as the Krasovsky 1940
reference ellipsoid 102 ^
SLR satellite laser ranging 352, 366 ^
SRAL Synthetic Aperture Radar Altimeter 353 ^
Ssa Sun, semi-annual tide 390 ^
SST satellite-to-satellite tracking 379 ^
STScI Space Telescope Science Institute 2 ^
SWH significant wave height 355, 356 ^
T
TCterrain correction 134, 135, 137–140, 250–252 ^
TOPEX/Poseidon Topography Experiment / Poseidon, American-French radar
altimetry satellite xix, 331, 352–354, 369, 382 ^
U
UCO University of California Observatories 2 ^
W
WGS84 World Geodetic System 1984 70, 101 ^
x
x
í ¤. û
^ Fundamentals of the theory of
gravitation
^ 1.1 General
1
In this chapter we present the foundations of Newton’s theory of
gravitation. Intuitively, the theory of gravitation is easiest to understand
as “action at a distance,” Latin actio ad distans, where the force between
two masses is proportional to the masses themselves, and inversely
proportional to the square of the distance between them. This is the
form of Newton’s general law of gravitation familiar to all.
There exists an alternative but equivalent presentation, field theory,
which portrays gravitation as a phenomenon propagating through
space, a field. The propagation is expressed in field equations. The field
approach is not quite as intuitive, but is a powerful theoretical tool.1 1
–1–
2 1 Fundamentals of the theory of gravitation
í Õ ! ¤. û
Gravitation between two masses 1.2
3
of Gauss and Green, with the aid of which we may infer the whole
potential field in space from field values given only on a certain surface.
Other similar examples are the Chasles theorem and the solution to
Dirichlet’s problem.
In chapters 2 and 3, we apply these fundamentals of potential theory
to derive a spectral representation of the Earth’s gravitational field, a
spherical-harmonic expansion. pallofunktio-
kehitelmä
Here, at the beginning of the text, we derive an extensive set of
mathematical equations, such as well-known integral equations. This is
unfortunately necessary preliminary work. These equations, however,
are not an end in themselves and are not worth committing to memory.
Try rather to understand their logic, and how these various results
have been arrived at historically. Try as well to acquire an intuition, a
fingertip feeling, about the nature of this theory.
m1 m2
F=G . (1.1)
ℓ2
F is the attractional force between bodies 1 and 2, m1 and m2 are the
masses of the bodies, and ℓ is the distance between them. We assume the
masses to be points. The constant G, the universal gravitational constant,
has the value
G = 6.674 · 10−11 m3/kg s2 .
The value of G was determined for the first time by Henry Cavendish3 3
2 SirIsaac Newton PRS (1642–1727) was an English universal genius who derived the
mathematical underpinning of astronomy, and much of geophysics, in his major work,
Philosophiæ Naturalis Principia Mathematica (Mathematical Foundations of Physics).
í Õ ! ¤. û
4 1 Fundamentals of the theory of gravitation
Both the force F and the acceleration a have the same direction as the
line connecting the bodies. For this reason one often writes equation
1.1 as a vector equation, which is more expressive:
r−R
a = −GM , (1.2)
ℓ3
in which the three-dimensional vectors of place of both the attracting
6 and attracted masses are defined as follows in rectangular co-ordinates:6
3 Henry Cavendish FRS (1731–1810) was a British natural scientist from a wealthy, noble
background. He did also pioneering work in chemistry. He was extremely shy, and
the renowned neurologist Oliver Sacks retrodiagnosed him as living with Asperger
syndrome (Sacks, 2001).
4 Atleast in a vacuum. The Apollo astronauts showed impressively how on the Moon
a feather and a hammer fall equally fast! YouTube, Hammer vs. Feather.
5 AlbertEinstein (1879–1955) was a theoretical physicist of Jewish German descent,
who created both the special and general theories of relativity, applying the latter to
cosmology, and did fundamental work in quantum theory, a theory which he however
never fully accepted.
í Õ ! ¤. û
The potential of a point mass 1.3
5
r = xi + yj + zk, R = Xi + Yj + Zk,
{︁ }︁
where the triad of unit vectors i, j, k is an orthonormal basis7 in ortonormaali
Euclidean space and kanta
√︂ 7
í Õ ! ¤. û
6 1 Fundamentals of the theory of gravitation
The constant GM has in the case of the Earth (according to the GRS80,
conventionally) the value
This is an integral over the mass elements dm of the body, where every
such mass element is located at its own place R. The potential V is
evaluated at location r, and the distance ℓ = ∥r − R∥.
We now derive the equation for the potential of a thin spherical shell,
see figure 1.2, where we have placed the centre of the sphere at the
origin O.
Because the circumference of a narrow ring, width b · dθ, is 2πb sin θ,
its surface area is
(2πb sin θ) (b · dθ) .
Let the thickness of the shell be p (small) and its density ρ. We obtain
for the total mass of the ring
Because every point of the ring is at the same distance ℓ from point P,
we may write for the potential of the ring at point P:
2πGpρb2 sin θ dθ
VP = .
ℓ
í Õ ! ¤. û
Potential of a spherical shell 1.4
7
b dθ
p b
P
r
O
θ
ℓ
b
we obtain, using equation 1.5, for the potential of the whole shell
w
2 sin θ dθ
VP = 2πGρpb √ .
r + b2 − 2rb cos θ
2
ℓ dℓ = rb sin θ dθ,
√
and remembering that ℓ = r2 + b2 − 2rb cos θ we obtain
w ℓ2
dℓ
VP = 2πGρpb2 .
ℓ1 rb
In the case that point P is outside the shell, the integration bounds of ℓ
are ℓ1 = r − b and ℓ2 = r + b, and we obtain for the potential of point P
[︂ ]︂ℓ=r+b
2 ℓ 4πGρpb2
VP = 2πGρpb = r .
rb ℓ=r−b
í Õ ! ¤. û
8 1 Fundamentals of the theory of gravitation
Because the mass of the whole shell is Mb = 4πb2 ρp, it follows that the
potential of the shell is the same as that of an equal-sized mass in its centre
O:
GM
VP = r b ,
where r is the distance of computation point P from the centre of the
sphere O. We see that this is identical to equation 1.4.
In the same way, the attraction, or rather, acceleration, caused by the
8 spherical shell is8
r P − rO r −r
aP = ∇V|P = −4πGρpb2 3
= −GMb P 3 O ,
r r
in which r = ∥rP − rO ∥ . This result is identical to the acceleration
caused by an equal-sized point mass located in point O, equation 1.2.
In the case that point P is inside the shell, ℓ1 = b − r and ℓ2 = b + r
and the above integral changes to the following:
]︂ℓ=b+r
ℓ
[︂
VP = 2πGρpb2 = 4πGρpb.
rb ℓ=b−r
í Õ ! ¤. û
Computing the attraction from the potential 1.5
9
Acceleration
b2
4πGρp 2
r
4πGρpb
b
4πGρpb
r
Potential
0
0 b →r
the Earth and many celestial bodies), then inside the body only those
layers of mass “interior” to — meaning closer to the centre than — the
observation point participate in causing attraction, and this attraction is
the same as it would be if the mass of all these layers was concentrated
in the centre of the body. This case, where the distribution of mass
density inside a body only depends on the distance from its centre, is
called an isotropic density distribution.
∂V ∂V ∂V
a = ∇V = grad V = i+ j+ k. (1.7)
∂x ∂y ∂z
í Õ ! ¤. û
10 1 Fundamentals of the theory of gravitation
∂ ∂ ∂
∇=i +j +k .
∂x ∂y ∂z
{︁ }︁
kanta As before, i, j, k is a basis of mutually orthogonal unit vectors in
Euclidean space parallel to the (x, y, z) axes.
Let us try this differentiation in the case of the potential field of
the point mass M. Substitute the above equations for potential V and
9 distance ℓ, 1.3 and 1.4:9
∂V ∂V ∂ℓ 1 x−X x−X
= = GM · − 2 · = −GM 3 .
∂x ∂ℓ ∂x ℓ ℓ ℓ
Similarly we compute the y and z components:
∂V y−Y ∂V z−Z
= −GM 3 , = −GM 3 .
∂y ℓ ∂z ℓ
These are the components of the gravitational acceleration or attraction
vector when the source of the field is one point mass M. So, in this
concrete case, vector equation 1.7 given above applies:
a = grad V = ∇V.
í Õ ! ¤. û
Potential of a solid body 1.6
11
as it is known to always be. However, the potential energy of body
m inside the field V of mass M is negative! More precisely, the
potential energy of body m is
Epot = −Vm.
One calls the vector of gravitational acceleration more briefly the gravi-
tation or attraction vector.
dm(x, y, z)
ρ(x, y, z) = ,
dV(x, y, z)
in which dm is a mass element and dV is the corresponding element of
spatial volume. The dimension of ρ is density, its unit in the SI system,
kg/m3 .
í Õ ! ¤. û
12 1 Fundamentals of the theory of gravitation
y y
dm ρ
V=G =G dV. (1.8)
body ℓ body ℓ
As we showed already above for mass points, the first derivative with
respect to place or gradient of the potential V of a solid body,
grad V = ∇V = a, (1.9)
is also the acceleration vector caused by the attraction of the body. This
applies generally.
∥r∥ → ∞ =⇒ V(r) → 0,
and thus
∥r∥ → ∞ 1 ℓ → 0.
/︁
=⇒
10 Unfortunatelyalmost the same symbols V and V are used here for potential and
volume, respectively.
í Õ ! ¤. û
Potential of a solid body 1.6
13
For the acceleration of gravitation, the same applies for all three compo-
nents, and thus also for the length of this vectorial quantity:
∥r∥ → ∞ =⇒ ∥∇V∥ → 0.
and thus
1 1 1 1 1/︁ 1 1 1/︁
⩽ ⩽ =⇒ ⩽ ⩽ .
∥r∥ + ϵ ℓ ∥r∥ − ϵ ∥r∥ 1 + ϵ ∥r∥ ℓ ∥r∥ 1 − ϵ ∥r∥
r→∞ 1 ℓ → 1 r.
/︁ /︁
=⇒
When we substitute this into integral 1.8, it follows that for large
distances r → ∞:
y y
ρ G GM
V=G dV ≈ r ρ dV = r ,
body ℓ body
11 The
only important exception is formed by the forces between a planet and its
moons, due both to the flattening of the planet and tidal effects.
í Õ ! ¤. û
14 1 Fundamentals of the theory of gravitation
in which (X, Y) is the location in the plane of the mass line, (x, y, z) is
the location of the point at which the potential is being evaluated, and
the mass line extends from sea level Z = 0 to height Z = H.
Firstly we write ∆x = X − x, ∆y = Y − y, and ∆z = Z − z, and the
potential becomes
w H−z
1
V(∆x, ∆y, ∆z) = G √︁ d(∆z).
−z ∆x + ∆y2 + ∆z2
2
∂2 V ∂ G ∂ℓ−1 ∂ℓ
(︂ )︂
= = G · · =
∂H2 ∂H ℓ ∂ℓ ∂H
H−z
= G · −ℓ−2 · 21 ℓ−1 · 2 (H − z) = −G .
ℓ3
í Õ ! ¤. û
Example: The potential of a line of mass 1.7
15
The third derivative, obtained in the same way:
3 (H − z)2
(︃ )︃ (︃ )︃
∂3 V ∂ H−z 1
= −G =G − 3 =
∂H3 ∂H ℓ3 ℓ5 ℓ
3 (H − z)2 − ℓ2
=G ,
ℓ5
and so on. The Taylor expansion is
∂3H V |
∂2H V | ⏟ ⏞⏞H=0 ⏟
V|H=0 ∂H⏟⏞⏞⏟
V|H=0
⏟⏞⏞⏟ H=0
⏞ ⏟⏟ ⏞ 1 z 3z2 − ℓ20 3
V = 0 + G H + 21 G 3 H2 + 16 G H + ··· , (1.11)
ℓ0 ℓ0 ℓ50
√︂
in which ℓ0 = (X − x)2 + (Y − y)2 + z2 , so the values of the deriva-
tives used in this expansion are obtained by substituting H = 0.
Question How can we exploit this result for computing the gravita-
tional potential of a complete, realistic terrain or topography?
Answer In this expansion, the coefficients 1 ℓ0 , 12 z ℓ30 , . . . , like ℓ0 ,
/︁ /︁
í Õ ! ¤. û
16 1 Fundamentals of the theory of gravitation
def
div a = ⟨∇ · a⟩ = ⟨∇ · (∇V)⟩ = ⟨∇ · ∇⟩ V =
∂2 ∂2 ∂2
= ∆V = V + V + V, (1.12)
∂x2 ∂y2 ∂z2
in which
def ∂2 ∂2 ∂2
∆ = ⟨∇ · ∇⟩ = + +
∂x2 ∂y2 ∂z2
12 is a well-known symbol called the Laplace operator.12
In equation 1.4 for the potential of a point mass we may show, by
performing all partial derivations 1.12, that
∆V = 0, (1.13)
í Õ ! ¤. û
Gauge invariance 1.9
17
In the case where the mass density does not vanish everywhere, we
have a different equation, with ρ the mass density:
∆V = −4πGρ. (1.14)
are known as the field equations of the gravitational field. They play the
same role as Maxwell’s14 field equations in electromagnetism. Unlike 14
in Maxwell’s equations, however, in the above there is no time co-
ordinate. Because of this, it is not possible to derive an equation
describing the propagation in space of gravitational waves, like the one
for electromagnetic waves in Maxwell’s theory.
We know today that these “Newton field equations” are only approx-
imate, and that Einstein’s general theory of relativity is a more precise
theory. Nevertheless, in physical geodesy Newton’s theory is generally
precise enough and we shall use it exclusively.
í Õ ! ¤. û
18 1 Fundamentals of the theory of gravitation
Here again ℓ is the distance between the point at which the potential
is to be calculated and the moving mass element in integration dm —
í Õ ! ¤. û
Single mass-density layer 1.10
19
or surface element dS. The dimension of the mass surface density κ is
kg/m2 , different from the dimension of ordinary or volume mass density,
which is kg/m3 .
This case is theoretically interesting, although physically unrealistic.
The function V is everywhere continuous, also at the surface S, however
its first derivatives with place are already discontinuous. The discon-
tinuity appears in the direction perpendicular to the surface, in the
normal derivative.
Let us look at the simple case where a sphere of radius R has been
coated with a layer of constant surface density κ. By computing the
above integral 1.15 we may prove — in a complicated way, see section
1.4 — that the exterior potential is the same as it would be if all of the
mass of the coating were concentrated in the sphere’s centre. Also in
section 1.4 we proved that the potential interior to the sphere is constant.
Thus, the exterior attraction (r > R) , with r the distance of the point
of computation from the centre of the sphere, is
(︂ )︂2
M κ · 4πR2 R
aext (r) = G 2 = G 2
= 4πGκ r .
r r
The interior attraction (r < R) is
aint (r) = 0.
This means that on the surface of the sphere, ℓ = R, the attraction is
discontinuous:
aext (R) − aint (R) = 4πGκ.
In this symmetric case we see that
∂V
a = ∥a∥ = , (1.16)
∂n
in which the differentiation variable n represents the normal direction,
the direction perpendicular to the surface S. If the surface S is an
equipotential surface of the potential V, equation 1.16 applies generally.
Then, the attraction vector — more precisely, the acceleration vector —
is perpendicular to the surface S, and its magnitude is equal to that of
the normal derivative of the potential.
í Õ ! ¤. û
20 1 Fundamentals of the theory of gravitation
P n
ℓ
κ
−κ
δ
dD
µ= ,
dS
in which dD is a “dipole-layer element”. This layer may be seen as
being made up of two single layers. If we have a positive layer at density
κ and a negative layer at density −κ, and the distance between them is
δ, we get for small values of δ an approximate correspondence:
µ ≈ κδ. (1.17)
í Õ ! ¤. û
Double mass-density layer 1.11
21
The combined potential of the two single mass-density layers computed
as explained in the previous section, equation 1.15, is
x
1 1
(︂ )︂
V=G κ − dS.
surface ℓ1 ℓ2
The following relationship exists between the quantities ℓ1 , ℓ2 and δ
(Taylor expansion of the function 1 ℓ ):
/︁
1 1 ∂ 1
(︂ )︂
= +δ· + ··· ,
ℓ1 ℓ2 ∂n ℓ
∂
in which ∂n is again the derivative of the quantity in the normal direction
of the surface.
Substitution into the equation yields
x x
∂ 1 ∂ 1
(︂ )︂ (︂ )︂
V≈G κδ dS = G µ dS. (1.18)
surface ∂n ℓ surface ∂n ℓ
í Õ ! ¤. û
22 1 Fundamentals of the theory of gravitation
^ 1.12.1 Presentation
15 The Gauss divergence theorem,15 famous from physics, looks in vector
form like this:
y x ⟨︁ ⟩︁
div a dV = a · n dS, (1.19)
V ∂V
15 Johann Carl Friedrich Gauss (1777–1855) was a German mathematician and universal
í Õ ! ¤. û
The Gauss divergence theorem 1.12
23
Field line, field a
Normal n
Body
Sources surface
gravitational field.
The situation is analogous to the flow pattern of a liquid: positive
charges correspond to points where liquid is added to the flow, nega-
tive charges17 correspond to “sinks” through which liquid disappears. 17
The vector a is in this metaphor the velocity of flow; in the absence
of “sources” and “sinks” it satisfies the condition div a = 0, which
expresses the conservation and incompressibility of matter.
On the other hand, the function
⟨︁ ⟩︁ ∂V
a·n =
∂n
is often called the flux, in other words, how much field stuff “flows out” vuo
17 But the ”charges” for gravitation, the masses, are always positive.
í Õ ! ¤. û
24 1 Fundamentals of the theory of gravitation
— just like a liquid flow — from the space inside the surface ∂V through
the surface to the outside.
The Gauss divergence theorem states the two amounts to be equal: it
is in a way a book-keeping statement demanding that everything which is
produced inside a surface — div a — also has to come out through the
⟨︁ ⟩︁
surface — a · n .
In figure 1.5 it is graphically explained that the sum of “sources” over
∑︁
the inner space of the body, (+ + + · · · ), has to be the same as the
∑︁
sum of “flux” (↑ ↑ ↑ · · · ) over the whole boundary surface delimiting
this inner space.
a = a1 i + a2 j + a3 k.
í Õ ! ¤. û
The Gauss divergence theorem 1.12
25
a
a+
3
+
a2
a+
∆z 1
a−
1
−
a2
a−
3
∆y
∆x
≈ (a+ − + − + −
1 − a1 ) ∆y ∆z + (a2 − a2 ) ∆x ∆z + (a3 − a3 ) ∆x ∆y.
Here, a+1 is the value of component a1 on the one face in the x direction,
and a1 its value on the other face, and so on. For example, a+
−
3 is the
value of a3 on the box’s upper and a− 3 on its lower face. A box has of
course six faces, in each of three co-ordinate directions both “up- and
downstream”.
Then
∂a1 ∂a2 ∂a3
a+ −
1 − a1 ≈ ∆x, a+ −
2 − a2 ≈ ∆y, a+ −
3 − a3 ≈ ∆z,
∂x ∂y ∂z
and by substitution we see that
x ⟨︁ ⟩︁
a · n dS ≈
∂V
∂a1 ∂a ∂a
≈ ∆x · ∆y ∆z + 2 ∆y · ∆x ∆z + 3 ∆z · ∆x ∆y =
∂x ∂y ∂z
í Õ ! ¤. û
26 1 Fundamentals of the theory of gravitation
(︃ )︃
∂a1 ∂a2 ∂a3
= + + ∆x ∆y ∆z,
∂x ∂y ∂z
the same expression as 1.21. So, in this simple case, the Gauss divergence
theorem applies.
Obviously the equation also works, if we build out of these “Lego™
bricks” a larger body, because the faces of the bricks touching each
other are oppositely oriented and cancel from the surface integral of the
whole body. It is a little harder to prove that the equation also applies
to bodies having inclined surfaces.
∆V = −4πGρ. (1.14)
∂V ∂ GM ⃓ GM
⃓
= r =− 2 ,
∂n ∂r R
⃓
r=R
í Õ ! ¤. û
The Gauss divergence theorem 1.12
27
z
+1
z
0
GM(0, 0, 0)
x
−1 y
0 −1
0
y +1
x
í Õ ! ¤. û
28 1 Fundamentals of the theory of gravitation
The surface integral is six times that over the top face
x w +1 w +1
(︄ )︄
⟨︁ ⟩︁ 1
a · n dS = −GM dx dy.
(x2 + y2 + 1) /2
top face −1 −1 3
2
= √︁ .
(y2 + 1) y2 + 2
Integrating this with respect to y yields
]︄+1
w +1
[︄
2 y
√︁ dy = 2 arctan √︁ =
−1 (y2 + 1) y2 + 2 y2 + 2 −1
1 π
= 4 arctan √ = 4 · = 23 π.
3 6
Adding the six faces together yields
w +1 w +1
(︄ )︄
1
− 6 · GM dx dy = −6 · GM · 23 π =
−1 −1
(x2 + y2 + 1)
3/2
= −4πGM,
agreeing with the volume result above.
í Õ ! ¤. û
Green’s theorems 1.13
29
and
x ⟨︁ ⟩︁ x ⟨︁ ⟩︁ x ⟨︁ ⟩︁
F · n dS = U ∇V · n dS = U ∇V · n dS =
∂V ∂V ∂V
x
∂V
= U dS.
∂V ∂n
The result is Green’s19 first theorem: 19
y y (︃ )︃
∂U ∂V ∂U ∂V ∂U ∂V
U ∆V dV + + + dV =
V V ∂x ∂x ∂y ∂y ∂z ∂z
x
∂V
= U dS.
∂V ∂n
This may be cleaned up, because the second term on the left is symmetric
for the interchange of the scalar fields U and V. Let us therefore
interchange U and V, and subtract the equations obtained from each
other. The result is Green’s second theorem:
y x (︂
∂V ∂U
)︂
(U ∆V − V ∆U) dV = U −V dS.
V ∂V ∂n ∂n
We assume in all operations that the functions U and V are “well-
behaved”: for example, all required derivatives exist everywhere in
body V.
A useful special case arises by choosing for the function U:
1
U= ,
ℓ
in which ℓ is the distance from the given point of evaluation P. This
function U is well-behaved everywhere except precisely at point P, where
it is not defined.
In the case when point P is outside the surface ∂V, the result, Green’s
third theorem, is now obtained by substitution (remember that ∆U = 0
inside ∂V):
y x (︃ (︂ )︂)︃
1 1 ∂V ∂ 1
∆V dV = −V dS.
V ℓ ∂V ℓ ∂n ∂n ℓ
í Õ ! ¤. û
30 1 Fundamentals of the theory of gravitation
Distance ℓ
Surface element dS
Surface
normal
n
Body V
Surface S = ∂V
Figure 1.8. Geometry for deriving Green’s third theorem if point P is outside
^ surface ∂V.
í Õ ! ¤. û
Green’s theorems 1.13
31
Surface ∂V, part 1
Surface ∂V, part 2
Point P Volume V
Figure 1.9. Geometry for deriving Green’s third theorem if point P is inside
^ surface ∂V.
í Õ ! ¤. û
32 1 Fundamentals of the theory of gravitation
Point P
Integration space V
Matter
(Limit)
^ Figure 1.10. Green’s third theorem for the space external to a body.
í Õ ! ¤. û
The Chasles theorem 1.14
33
^ 1.14 The Chasles theorem
We study the above-described case where the “body” is the space
outside the surface ∂V — in practice: the space outside the Earth.
From Green’s theorem 1.25 derived above, we may derive for a
harmonic function V (so, ∆V = 0) in the exterior space:
x x
1 1 ∂ 1 ∂ 1
(︂ )︂
VP = − V dS + V dS. (1.26)
4π ∂V ℓ ∂n 4π ∂V ∂n ℓ
Interpretation The exterior, harmonic potential of an arbitrarily shaped
body can be represented as the sum of a single and a double
mass-density layer on the body’s surface.
Explanation We obtain the surface density of a single mass layer using
equation 1.15,
1 ∂
κ=− V, (1.27)
4πG ∂n
and similarly the surface density of a double mass layer using
equation 1.18,
V
µ= .
4πG
If we plug these into equation 1.26, we obtain
x (︃ (︂ )︂)︃
κ ∂ 1
VP = G +µ dS.
∂V ℓ ∂n ℓ
point P, is harmonic inside the Earth’s body. The surface of the Earth is
∂V.
í Õ ! ¤. û
34 1 Fundamentals of the theory of gravitation
í Õ ! ¤. û
Boundary-value problems 1.15
35
potential V throughout the part of space exterior or interior to a given
boundary surface from given values relating to V on that boundary
surface, for example on the surface of the Earth. The simplest boundary-
value problem is Dirichlet’s23 problem: the potential V itself is given on 23
the boundary surface. More complicated boundary-value problems are
based on linear functionals of the potential: on the boundary, some linear
expression in V is given, for example a derivative or linear combination
of derivatives, generally
{︁ }︁
L V ,
{︁ }︁
with L · being a linear functional, see section 10.2.
The Dirichlet boundary-value problem in the form used in geodesy is:
determine the potential field V if its values are given on a closed surface
S, and V is harmonic (∆V = 0) outside surface S. In the vacuum of space,
the potential is always harmonic, as already noted earlier: the potential
of a point mass mP , V = GmP ℓ , is a harmonic function everywhere
/︁
23 Peter
Gustav Lejeune Dirichlet (1805–1859) was a German mathematician also
known for his contributions to number theory.
í Õ ! ¤. û
36 1 Fundamentals of the theory of gravitation
^ Self-test questions
1. Which instrument was used to determine the constant G? Why is
it difficult to obtain a precise value for this constant?
2. Why do all objects, irrespective of their mass, undergo the same
acceleration of free fall, although the gravitational attraction on a
more massive body is obviously stronger?
3. What is a conservative force field?
(a) A force field for which the force can be written as the gradient
of a unique potential.
(b) A force field in which an object carried along a closed loop
will not gain and not lose energy.
(c) An attractive force field from which no object can escape.
(d) A force field the curl of which vanishes everywhere.
4. On the surface of a homogeneous, spherical asteroid the accel-
í Õ ! ¤. û
Exercise 1 – 1: Core of the Earth
37
eration of free fall is 1 cm/s2 . What is the acceleration of free fall
on another asteroid that is otherwise similar, but has twice the
diameter?
(a) 0.25 cm/s2
(b) 1 cm/s2
(c) 2 cm/s2
(d) 4 cm/s2
5. What is a harmonic potential?
6. What is the order of the Laplace differential equation? kertaluku
7. Is a linear potential, V(x, y, z) = a + bx + cy + dz (a, b, c, d
constants), harmonic?
8. If the potential in the previous question is a gravitational potential,
calculate its acceleration vector.
9. Under what condition is it possible to describe the external grav-
itational field emanating from a body as produced by a single
mass-density layer on the surface of that body?
10. The dipole-layer density µ is mentioned in section 1.11. What is
the SI unit of this quantity?
í Õ ! ¤. û
38 1 Fundamentals of the theory of gravitation
a a0
Σ1
d
∞ ∞
Σ2
∂
4. Derive the equation for the radial gravitational gradient ∂r g on
the surface of a homogeneous-density sphere of density ρ.
^ Exercise 1 – 2: Atmosphere
1. The mean pressure of the atmosphere at sea level is 1013.25 hPa
(the unit for pressure, the pascal, is defined as Pa = N/m2 .) On
the Earth’s surface gravity is 9.81 m/s2 . Calculate the mean surface
density as a thin layer κ in units of kg/m2 .
2. Calculate the total mass of the atmosphere using the spherical-shell
approximation. You may take as its radius 6371 km.
3. Calculate the attraction generated by the atmosphere outside it,
both as acceleration and as a fraction of the total Earth attraction.
4. How much is the attraction from the atmosphere inside the
atmosphere?
í Õ ! ¤. û
Exercise 1 – 3: The Gauss divergence theorem
39
^ Exercise 1 – 3: The Gauss divergence theorem
There is a deposit (body) of iron ore inside the Earth, figure 1.11. The
deposit generates an attraction on the Earth’s surface, which has been
drawn as the a curve. We use the flat Earth approximation.
The true attraction curve is approximated by a simple function
⎧
⎨a if s ⩽ d
0
a=
⎩0 if s > d
(red dashed line), where s is the distance from the point on the Earth’s
surface straight above the ore deposit. So, the area where a ̸= 0 is a disk
of radius d on the surface of the Earth.
1. Compute, using the above approximation for a, the surface integral
x
a dS,
Σ1
í Õ ! ¤. û
40 1 Fundamentals of the theory of gravitation
í Õ ! ¤. û
^ The Laplace equation and its
solutions
2
^ 2.1 The nature of the Laplace equation
The Laplace equation is central to the study of the Earth’s gravitational
field: (︃ )︃
∂2 ∂2 ∂2
∆V = + + V = 0.
∂x2 ∂y2 ∂z2
We call the symbol ∆ the Laplace operator. Sometimes the alternative
notation ∇2 is used.
If we study gravitation as a field, the Laplace equation is more natural
than Newton’s approach. Newton’s equation is used when the mass
distribution is known: it yields directly the gravitational force caused
by the masses.
The Laplace equation, on the other hand, is a partial differential
equation. Its solution gives the potential V(x, y, z) of the gravitational
field throughout space or a part of space. From this potential one
may then calculate the effect of the field on a body moving in space
at the location where the body is. This is a two-phase process. It is
conceptually new here that a certain property, a field, is attached to
empty space, and we no longer talk about action at a distance directly
between two bodies.
Solving the Laplace equation in the general case may be difficult. The
approach generally taken is that we choose some co-ordinate frame
– 41 –
42 2 The Laplace equation and its solutions
∆V1 = 0, ∆V2 = 0,
V = αV1 + βV2 , α, β ∈ R
í Õ ! ¤. û
The Laplace equation in rectangular co-ordinates 2.2
43
values (“boundary values”) have to be given only on the boundary of
the part of space of interest.
For example, field values are given on the Earth’s surface. From this,
one calculates the values of the field in outer space, where the Laplace
equation applies — the behaviour of the field inside the Earth remains
outside the scope of our interest. From the perspective of the exterior
gravitational field, one does not even need to know the precise mass
distribution inside the Earth — and one cannot even determine it using
only measurement values obtained on and above the Earth’s surface.
∂2 ∂2 ∂2
YZ X + XZ Y + XY Z = 0.
∂x2 ∂y2 ∂z2
í Õ ! ¤. û
44 2 The Laplace equation and its solutions
1 ∂2 1 ∂2 1 ∂2
X(x) + Y(y) + Z(z) = 0.
X(x) ∂x2 Y(y) ∂y2 Z(z) ∂z2
Because this has to be true throughout the space, that is for all combi-
nations of values x, y, and z, it follows that each term must be a constant.
If we take for the first and second constants −ω2x and −ω2y , we get in
conclusion for the third constant ω2x + ω2y . By writing this definition
and result out and moving the denominator to the other side, we obtain
∂2 ∂2
X(x) = −ω2x X(x), Y(y) = −ω2y Y(y),
∂x2 ∂y2
(the reason for choosing a negative constant will become apparent), and
∂2 (︁ 2 2
)︁
Z(z) = ω x + ω y Z(z).
∂z2
Now, the solution is readily found at least to the first two equations:
2 they are harmonic oscillators, and their basis solutions2 are
harmoninen (︁ )︁ (︁ )︁
värähtelijä X(x) = exp ±iωx x , Y(y) = exp ±iωy y .
í Õ ! ¤. û
The Laplace equation in rectangular co-ordinates 2.2
45
Let us assume that in both the x and y directions the size of our
world is L (“shoebox world”3 ). Let us make things a little simpler by 3
assuming that on the boundary surfaces of our shoebox world we have
the boundary conditions
It then follows that the only pairs (ωx , ωy ) that yield a solution that fits
the box are
πj πk
ωx = , ωy = , j, k ∈ Z,
L L
and the only suitable functions are sine functions. Thus we obtain as a
solution
x y z
(︂ )︂ (︂ )︂ (︂ √︁ )︂
Vjk (x, y, z) = sin πj sin πk exp ±π (j2 + k2 ) .
L L L
This particular solution may now be generalised by multiplying it by
suitable coefficients, and summing it over different index values j = 0,
±1, ±2, . . . and k = 0, ±1, ±2, . . . .
We may however remark that the terms for which j = 0 or k = 0
will always vanish, and the terms that contain j = +n and j = −n,
or k = +n and k = −n, n ∈ N, are (apart from their algebraic signs)
identical. Therefore in practice we sum over the values j = 1, 2, . . . and
k = 1, 2, . . . .
Different boundary conditions will give slightly different general
solutions. Their general form is, however, always similar.
The zero-level z = 0 expansion resulting from the general solution is
the familiar Fourier4 sine expansion: 4
í Õ ! ¤. û
46 2 The Laplace equation and its solutions
Vjk (x,y)
⏟ (︂ )︂⏞⏞ (︂ )︂⏟
∑︂
∞ ∑︂
∞
jx ky
V(x, y, 0) = vjk sin π sin π , (2.1)
L L
j=1 k=1
í Õ ! ¤. û
The Laplace equation in rectangular co-ordinates 2.2
47
L
13πx 5πx
sin sin
Sea level L L
The takeaway from this is that the operation of vertically shifting the
field V from zero level to the level z, which is not straightforward in the
space domain, becomes simple — as in a straightforward multiplication
Figure 2.2. Vertically shifting the harmonic field V in the space and frequency
(wave-number) domains, commutative diagram. Rectangular
^ geometry.
í Õ ! ¤. û
48 2 The Laplace equation and its solutions
and then split the above equation into two equations, one for the right-
hand side function R(r) and one for the function A(α). Substitution
yields
r2 ∂2 R(r) 1 ∂2 A(α)
(︃ )︃
r ∂R(r)
+ + = 0.
R(r) ∂r2 R(r) ∂r A(α) ∂α2
Both terms must be constant:
(︃ 2 )︃
∂ R(r) ∂R(r)
r r + − k2 R(r) = 0,
∂r2 ∂r
∂2 A(α)
+ k2 A(α) = 0.
∂α2
Here, the algebraic sign of the constant k2 has been chosen so that A(α)
gets a periodic solution. Such a general solution would be
5 Thereason for this, as we shall later discuss more generally, is that the vertical shift
operation is a convolution.
í Õ ! ¤. û
The Laplace equation in polar co-ordinates 2.3
49
in which, because angle α has a period of 2π, k has to be a non-negative
integer: k = 0, 1, 2, . . . . Negative k does not give different solutions,
because
r rq (q − 1) rq−2 + qrq−1 − k2 rq = 0
(︁ )︁
=⇒ q 2 − k2 = 0
=⇒ q 2 = k2 .
í Õ ! ¤. û
50 2 The Laplace equation and its solutions
X = r cos ϕ cos λ,
Y = r cos ϕ sin λ, (2.4)
Z = r sin ϕ.
í Õ ! ¤. û
Spherical, geodetic, ellipsoidal co-ordinates 2.4
51
Pole Z
r cos ϕ P
r
Equator
r sin ϕ
ϕ Y
λ
X
Greenwich meridian
are much-used. On the Earth’s surface most often geodetic — also called
geographical — co-ordinates (φ, λ, h) are used:
X = (N + h) cos φ cos λ,
Y = (N + h) cos φ sin λ, (2.5)
Z = N + h − e2 N sin φ,
(︁ )︁
in which
a a2
N(φ) = √︁ = √︁ . (2.6)
1 − e2 sin2 φ a2 cos2 φ + b2 sin2 φ
The quantity N defined in equation 2.6 is the west-east direction, or
transversal, radius of curvature of the reference ellipsoid. In the equation, poikittainen
a is the equatorial radius of the reference ellipsoid used, b is the polar vertaus-
radius, ellipsoidi
2 2
def a − b
e2 = 2
(2.7)
a
is the square of the first eccentricity,7 and in equations 2.5, h is the height 7
í Õ ! ¤. û
52 2 The Laplace equation and its solutions
Z Ellipsoidal
P normal
x
(︂(︁ )︂ h
1 − e2 N + h sin φ
)︁
φ X, Y
O (N + h) cos φ
Reference
ellipsoid
isoakselin If the semimajor axis of the Earth ellipsoid is a and its semiminor axis
puolikas
pikkuakselin 7 The parameter is connected to the Earth’s flattening f through the equation e2 =
puolikas 2f − f2 .
í Õ ! ¤. û
The Laplace equation in spherical co-ordinates 2.5
53
b, it follows that E2 = a2 − b2 . Equation 2.7 tells us that E2 = a2 e2 .
The first eccentricity e is the eccentricity of the meridian ellipse, and
E = ae is the distance from the geocentre of the two focal points of this
ellipse, which lie on both sides of the geocentre in the equatorial plane.
Equations 2.8 tell us that all ellipses u = constant in the meridian plane
for different values of u have the same two focal points: they are confocal.
See figure 4.6.
We note still — see Heiskanen and Moritz (1967) figure 1-14 — that the
curves β = constant describe hyperbolas with these same focal points.
∂2 V 2 ∂V 1 ∂2 V tan ϕ ∂V 1 ∂2 V
∆V = 2
+r + 2 − 2 + 2 = 0, (2.9)
∂r ∂r r ∂ϕ 2 r ∂ϕ r cos ϕ ∂λ2
2
of which the first is nonphysical in outer space, because, unlike the true
geopotential, these expressions grow to infinity for r → ∞.
In the above equations, the functions Yn (ϕ, λ) are called surface pinta-
spherical harmonics, whereas the functions Vn (ϕ, λ, r) are solid spherical pallofunktio
harmonics. The latter are harmonic functions everywhere in space except avaruus-
pallofunktio
í Õ ! ¤. û
54 2 The Laplace equation and its solutions
The functions Pnm are Legendre functions, on which more later on. With
the help of expression 2.11, we obtain, by using the second, physically
pallofunktio- realistic alternative from equations 2.10, the following solution or series
kehitelmä expansion for the potential V in outer space:
∑︂
∞
1 ∑︂
n
V(ϕ, λ, r) = Pnm (sin ϕ) (anm cos mλ + bnm sin mλ) .
rn+1
n=0 m=0
(2.12)
The coefficients anm and bnm are called the coefficients of the spherical-
harmonic expansion, in short, spectral coefficients. Together they re-
present the function V, in somewhat the same way that the Fourier
coefficients vjk do in rectangular co-ordinates in equation 2.2. The
asteluku, subscripts n and m are called degree and order.
järjestysluku
Often we will be using a somewhat freer notation for the scaled
functions Yn Rn+1 . For example, if we expand the disturbing potential
/︁
häiriö-
potentiaali T into spherical harmonics, we shall use the notation Tn (ϕ, λ) for its
pallofunktio surface harmonics. Similarly, ∆gn (ϕ, λ) is the surface harmonic of the
gravity anomaly ∆g for degree n, and so on. Then, it holds on the
asteosuus- Earth’s surface r = R (degree constituent expansion) that
hajotelma
∑︂
∞ ∑︂
∞
T (ϕ, λ, R) = Tn (ϕ, λ), ∆g(ϕ, λ, R) = ∆gn (ϕ, λ),
n=0 n=0
and so on.
í Õ ! ¤. û
Self-test questions
55
the distance r from the Earth’s centre, or equivalently, on the height
H = r − R, if by R we denote the radius of the Earth sphere or sea level.
The dependence is
Y (ϕ, λ)
Vn (ϕ, λ, r) = n n+1 .
r
At sea level
Y (ϕ, λ) def
Vn (ϕ, λ, R) = n n+1 = Vn (ϕ, λ).
R
Therefore we may write
(︂ )︂n+1
R R + H −(n+1)
(︂ )︂
Vn (ϕ, λ, r) = r Vn (ϕ, λ) = Vn (ϕ, λ) =
R
H −(n+1) H
(︂ )︂ (︂ )︂
= 1+ Vn (ϕ, λ) ≈ exp − (n + 1) Vn (ϕ, λ).
R R
We see that the attenuation of the potential with height is exponential,
and the harmonic degree number n appears in the exponent, as also
did the wave number in rectangular geometry, see equation 2.2 and
figure 2.1. The analogy works.
^ Self-test questions
1. In what essential way does the approach of the Laplace equation
differ from Newton’s approach?
2. How does the linearity of the Laplace equation help in finding
solutions?
3. How does separation of variables work?
4. Why does solving the Laplace equation require boundary condi-
tions?
5. Show, using equations 2.8, that for the curves u = constant in the
meridian plane Y = 0, the sum of the distances from the curve point
[︂ √ ]︂T [︂ ]︂T
2 2
u + E cos β 0 sin β to the focal points ±E 0 0
is constant (and that these curves thus are confocal ellipses), and
that for the curves β = constant the difference of these distances is
í Õ ! ¤. û
56 2 The Laplace equation and its solutions
constant (and that these curves thus are confocal hyperbolas). See
figure 4.6.
í Õ ! ¤. û
^ Legendre functions and spherical
harmonics
3
^ 3.1 Legendre functions
In equations 2.11 and 2.12, the functions P are Legendre1 functions that 1
pop up whenever we solve a Laplace-like equation in spherical co-
ordinates. There exist various effective, so-called recursive algorithms
for their calculation, for example the following algorithm only for
ordinary Legendre polynomials Pn = Pn0 :
Similar equations also exist for the functions Pnm , m > 0. There are even
alternatives to choose from, although most equations are complicated.
One should be careful that, in their computation, the factorials do not kertoma
go overboard. Already 30! (factorial of 30) is a larger number than
computers can handle as 64-bit integers — not to mention 360!, the
factorial of 360, which overflows even the standard 64-bit floating-point
format. Heiskanen and Moritz’s (1967) equation 1-62, contrary to what
is stated there, is not directly suitable for computer use!
1 Adrien-Marie Legendre (1752–1833) was a French mathematician known for his work
– 57 –
58 3 Legendre functions and spherical harmonics
The first Legendre polynomials are listed in table 3.1. Higher polyno-
mials than this are rarely needed in manual computation.
kantafunktio For comparison, the Fourier basis functions (like, in a more complicated
P6
−0.5 P4 P2 P3
P1
−1
−1 −0.5 0 −→ t 0.5 1
í Õ ! ¤. û
Legendre functions 3.1
59
way, sines and cosines together as well!)
x
(︂ )︂
Fj (x) = exp 2πij ,
L
in which i2 = −1, can also be calculated recursively:
through zero.
◦ As the values in the end points t = ±1, ϕ = ±90◦ are ±1, it follows
that there are precisely n + 1 “algebraic-sign intervals”, meaning
open intervals of t or ϕ on which the polynomial assumes only
positive or only negative values.
í Õ ! ¤. û
60 3 Legendre functions and spherical harmonics
15
P32
P31
10 P21
P25,25
P33 5 · 1030 P33
5 P32
P22
P11
0
P21 P31
P32
−5
−1 −0.5 0 −→ t 0.5 1
Figure 3.2. Some associated Legendre functions. Note the extremely different
^ scale used for the function P25,25 , see equation 3.8.
í Õ ! ¤. û
Legendre functions 3.1
61
or ϕ ∈ −90◦ , 90 , precisely n − m times through zero. This is
◦
[︁ ]︁
Yn (ϕ, λ) =
∑︂
n
(︁ )︁
= anm Pnm (sin ϕ) cos mλ + bnm Pnm (sin ϕ) sin mλ =
m=0 ∑︂n
= vnm Ynm (ϕ, λ),
m=−n
These are the surface spherical harmonics of degree n and order m. pinta-
pallofunktio
Surface spherical harmonics, like also solid spherical harmonics,
come in three kinds:
Zonal harmonics m = 0. These functions depend only on latitude. vyöhyke-
funktiot
Sectorial harmonics m = n. the algebraic signs of these functions
sektorifunktiot
depend only on longitude and not on latitude. The functions
themselves however do depend on both latitude and longitude!
Tesseral harmonics 0 < m < n. These functions, the algebraic sign ruutufunktiot
of which changes with both latitude and longitude, form a
checkerboard pattern on the surface of the sphere, if the positive
í Õ ! ¤. û
62 3 Legendre functions and spherical harmonics
Figure 3.3. The algebraic signs of spherical harmonics on the Earth’s surface.
White means positive, grey negative. The functions “wave” in a
^ fashion similar to that of sine or cosine functions.
values are painted white and the negative ones grey (Latin tessera:
a tile, as used in a mosaic).
[︁ ]︁
Every function will, on the interval sin ϕ ∈ −1, +1 , go precisely n − m
times through zero. Every function is either symmetric or antisymmetric
through the origin as a function of ϕ or of t = sin ϕ.
Spherical harmonics thus represent a wave phenomenon of sorts.
They are however not proper wave functions (sines or cosines): the
connection to those is complicated at least. It nevertheless makes sense
to speak of their wavelength.
pallofunktio Figure 3.3 depicts how the algebraic signs of the different spherical
harmonics behave on the Earth’s surface — and above. This is a
perspective sketch and not all white and grey areas are visible!
In equation 2.11, the expressions cos mλ and sin mλ go around a full
circle, the equator, 0◦ ⩽ λ < 360◦ , or 0 ⩽ λ < 2π, precisely 2m times
puoli- through zero. The “semi-wavelength” is thus
aallonpituus
2πR R
= πm,
2m
í Õ ! ¤. û
Legendre functions 3.1
63
πR
n − m.
If we plug various values for m and n − m into this, we obtain table 3.3.
This table also gives the resolution that can be achieved with a spherical-
harmonic expansion, or in how detailed a fashion the expansion can
describe the Earth’s gravity field. The expansions available today, like
the model EGM2008, go to harmonic degree n = 2159; the “sharpness”
of a geopotential image based on them is thus 9 km. Models based on
satellite orbit perturbations often extend only to degree 20, meaning
í Õ ! ¤. û
64 3 Legendre functions and spherical harmonics
∑︂
∞
1 ∑︂
n
V(ϕ, λ, r) = Pnm (sin ϕ) (anm cos mλ + bnm sin mλ) .
rn+1
n=0 m=0
(2.12)
í Õ ! ¤. û
Symmetry properties of spherical harmonics 3.2
65
or (antisymmetric case)
(︁ )︁
Pnm (sin ϕ) = −Pnm sin(−ϕ) .
or (antisymmetric case)
í Õ ! ¤. û
66 3 Legendre functions and spherical harmonics
^ Tableau 3.4. Plotting a surface spherical-harmonic map. Note that the octave
code for associated Legendre functions used here contains an additional factor
(−1)m , so compared to equation 3.2 all odd orders have the opposite sign.
í Õ ! ¤. û
Orthogonality of Legendre functions 3.3
67
identically, so the coefficients b00 , b10 , b20 , . . . simply do not matter.
They may be any value, including zero. The coefficients a00 , a10 , a20 ,
. . . however do matter, as for m = 0, cos mλ = 1 identically. So we
obtain as the rotationally symmetric expansion
∑︂
∞
1
V(ϕ, λ, r) = V(ϕ, r) = an Pn (sin ϕ),
rn+1
n=0
def def
in which Pn = Pn0 are the familiar Legendre polynomials, and an =
an0 .
í Õ ! ¤. û
68 3 Legendre functions and spherical harmonics
In this case, the orthonormal functions are those of equation 3.3, but
3 Carl
Gustav Jacob Jacobi (1804–1851) was a German mathematician known for his
work on elliptic functions.
4 And also ⎧
⟨︁ ⟩︁ def 1 w +1 ⎨1 if n = n ′ ,
Pn · Pn ′ t = 2 Pn (t) Pn ′ (t) dt =
−1 ⎩0 if n ̸= n ′ ,
í Õ ! ¤. û
Low-degree spherical harmonics 3.4
69
normalized:
⎧
⎨P
nm (cos ψ) cos mα if m ⩾ 0,
Y nm (ψ, α) =
⎩Pn|m| (cos ψ) sin |m| α if m < 0.
yielding
1 ⟨︁ ⟩︁
V1 (r) = (a 11 i + b 11 j + a 10 k) · r .
r3
The potential field of a dipole is
G ⟨︁ ⟩︁
V(r) = 3
d·r ,
r
in which d is the dipole moment. Comparison yields
í Õ ! ¤. û
Splitting a function into degree constituents 3.5
71
d
Figure 3.5. Monopole, dipole, and quadrupole at the Earth’s centre and their
^ effects on the geoid.
∑︂
∞
f(ϕ, λ) = f(ϕ, λ, R) = fn (ϕ, λ)
n=0
í Õ ! ¤. û
72 3 Legendre functions and spherical harmonics
(︁ ′ ′ )︁ ∑︂
∞ ∑︂
n
Pnm cos ψ (anm cos mλ ′ + bnm sin mλ ′ ) ,
(︁ )︁
f ϕ ,λ =
n=0 m=0
substituting this into the degree constituent equation 3.9, and exploiting
the orthogonality of the Legendre functions, we obtain on the right-hand
side:
x
2n + 1
f ϕ ′ , λ ′ Pn (cos ψ) dσ ′ =
(︁ )︁
IR =
4π σ
x
2n + 1
= an0 2
Pn (cos ψ) dσ ′ .
4π σ
2n + 1 4π
IR = an0 = an0 = an .
4π 2n + 1
On the left-hand side of the degree constituent equation we obtain,
because on the assumed north pole ϕ = 90◦ and thus sin ϕ = 1:
IL = fn (ϕ, λ) = fn (90◦ , λ) =
∑︂
n
= Pnm (1) (anm cos mλ + bnm sin mλ) = Pn0 (1) an0 = an ,
m=0
Pn0 (1) = 1,
Pnm (1) = 0, m ̸= 0.
As this applies for every point (ϕ, λ), it follows that the degree con-
stituent equation 3.9 is universally true. Note that the values an depend
on the point choice!
í Õ ! ¤. û
Spectral representations of various quantities 3.6
73
^ 3.6 Spectral representations of various quantities
^ 3.6.1 Potential
Starting from equation 2.10, we write the spectral expansion of the
geopotential V in space:
∑︂
∞ (︂ )︂n+1
R
V(ϕ, λ, r) = r Vn (ϕ, λ), (3.10)
n=0
Yn (ϕ, λ)
Vn (ϕ, λ) = =
Rn+1
1 ∑︂
n
= n+1 Pnm (sin ϕ) (anm cos mλ + bnm sin mλ) =
R
1 ∑︂
m=0 n
= n+1 vnm Ynm (ϕ, λ).
R
m=−n
í Õ ! ¤. û
74 3 Legendre functions and spherical harmonics
Figure 3.6. Vertically or radially shifting the harmonic field V, in the space
^ and frequency (degree number) domains. Spherical geometry.
^ 3.6.2 Gravitation
5In the Neumann5 boundary-value problem we solve a function V of
reuna-arvo- which the normal derivative, ∂ V, is given on a closed surface in space,
∂n
tehtävä for example the surface of a body.
∂ ∂
If the body is a sphere, one may assume ∂n V = ∂r V and work with
spherical-harmonic expansions. By differentiating equation 3.10 we
obtain
∂V ∑︂
∞
n + 1 R n+1
(︂ )︂ ∑︂
∞
n + 1 R n+2
(︂ )︂
=− r r Vn (ϕ, λ) = − r Vn (ϕ, λ).
∂r R
n=0 n=0
∂V ⃓
⃓ ∑︂
∞
n+1
=− Vn (ϕ, λ).
∂r r=R R
⃓
n=0
∂V ⃓
⃓
def def
∑︂
∞
g(ϕ, λ, R) = = gn (ϕ, λ),
∂r r=R
⃓
n=0
í Õ ! ¤. û
Spectral representations of various quantities 3.6
75
it follows by analogy that
n+1
gn (ϕ, λ) = − Vn (ϕ, λ),
R
and conversely that
R
Vn (ϕ, λ) = − g (ϕ, λ).
n+1 n
As a result of this, we obtain the spectral representation of the solution to a
certain Neumann problem:
∑︂
∞ (︂ )︂n+1
R ∑︂
∞ (︂ )︂n+1
R gn (ϕ, λ)
V(ϕ, λ, r) = r Vn (ϕ, λ) = −R r . (3.12)
n+1
n=0 n=0
For the gravitation one may write, analogously to expression 3.11 for
the potential:
∑︂
∞
def
∑︂
∞
1 ∑︂
n
g(ϕ, λ, R) = gn (ϕ, λ) = gnm Ynm (ϕ, λ), (3.13)
Rn+1
n=0 n=0 m=−n
n+1
gnm = − v . (3.14)
R nm
This is an interesting result worth thinking about:
1. Firstly, note how simple the connection 3.14 between potential
vnm and gravitation gnm is in the frequency domain!
2. Secondly, if measurement values of gravitational acceleration
g(ϕ, λ) are available over the whole surface area of the Earth, we
may derive from these the degree constituent functions gn (ϕ, λ) asteosuus-
using the method presented earlier. In this way we can then funktio
obtain the solution by means of equation 3.12 for the whole
exterior geopotential field! This is the basic idea of geopotential —
or geoid — determination, from the spectral perspective.
í Õ ! ¤. û
76 3 Legendre functions and spherical harmonics
GM
(︃ ∑︂
360 (︂ )︂ ∑︂
a n
n )︃
V= r⊕
(︁ )︁
1+ r Pnm (sin ϕ) Cnm cos mλ + Snm sin mλ . (3.15)
n=2 m=0
6 Here a = a⊕ is used to signify the equatorial radius of the Earth’s reference ellipsoid,
not R, and ϕ signifies geocentric latitude. The co-ordinates (ϕ, λ, r) form a spherical
co-ordinate system.
í Õ ! ¤. û
Ellipsoidal harmonics 3.8
77
and satellite radar altimetric data (over the ocean). After the launches
of the gravimetric satellite missions CHAMP, GRACE, and GOCE, and as a
result of their measurements, nowadays at least the harmonic degree
interval 20 < n ⩽ 200 is the product of space geodesy. Only the still
higher-degree coefficients continue to come from terrestrial data. The
newer model EGM2008 (Pavlis et al., 2012) goes up to degree 2159.
In tableau 3.5 we give the first and last coefficients of the EGM96
model, the newest and best spherical-harmonics model from the time pallofunktio-
just before the satellite gravity missions. The values tabulated are n, m, malli
Cnm , Snm and the mean errors (standard deviations) of both coefficients
from their computation. Note that all Sn0 vanish!
Sometimes non-normalised coefficients are also used, and we write
GM⊕
(︃ ∑︂
∞ (︂ )︂ ∑︂
a n
n )︃
V= r 1− r Pnm (sin ϕ) (Jnm cos mλ + Knm sin mλ) . (3.16)
n=2 m=0
def
Then we use the notation Jn = Jn0 . The coefficient J2 is the most
important spherical-harmonic coefficient of the Earth’s gravity field,
expressing the flattening of the Earth. Based on equations 3.7 and 3.8,
the relationship with the parameters C, S is
{︄ }︄ {︄ }︄
Jn0 √ Cn0
= − 2n + 1 ,
Kn0 Sn0
{︄ }︄ √︄ {︄ }︄ (3.17)
Jnm (n − m)! Cnm
= − 2 (2n + 1) , m ̸= 0.
Knm (n + m)! Snm
í Õ ! ¤. û
78 3 Legendre functions and spherical harmonics
V(β, λ, u) =
í Õ ! ¤. û
Ellipsoidal harmonics 3.8
79
^ Table 3.6. Legendre functions of the second kind.
1 z+1
Q0 (z) = 2 ln
z−1 (n + 1)Qn+1 (z) − (2n + 1) zQn (z) + nQn−1 (z) = 0
1 z+1
Q1 (z) = 2 z ln −1
z−1
(︁ 3 2 1 )︁ z + 1 3
Q2 (z) = 4 z − 4 ln − z )︁ m/2 dm
z−1 2 Qnm (z) = 1 − z2
(︁
Qn (z)
(︁ 5 3 3 )︁ z + 1 5 2 2 dzm
Q3 (z) = 4 z − 4 z ln − z +3
z−1 2
∑︂
∞ ∑︂n
Qnm i u
(︁ )︁
= (︁ Eb )︁ Pnm (sin β) (Aenm cos mλ + Benm sin mλ) , (3.18)
n=0 m=0
Qnm i E
in which Qnm (z) are the Legendre functions of the second kind, sampled
in table 3.6. Although the general argument z is complex, equation 3.18
gives a real result for real-valued coefficients Aenm , Benm .
Those interested in the derivation of the above equation 3.18 can find
it in Heiskanen and Moritz (1967) section 1-20 or other textbooks on
potential theory.
GM⊕ E
Ae00 = Ae0 = arctan
E b
and the expansion specialised for a rotationally symmetric field becomes
∑︂
∞ ∑︂
∞
Qn i u
(︁ )︁
∗
V (β, u) = Vn∗ (β, u) = (︁ Eb )︁ Ae∗
n0 Pn (sin β). (3.19)
n=0 n=0
Qn i E
Also
Q0 i u
(︁ )︁
GM⊕ E
V0 (u) = V0∗ (u) = (︁ Eb )︁ arctan ,
Q0 i E E b
í Õ ! ¤. û
80 3 Legendre functions and spherical harmonics
u E b E
(︂ )︂ (︂ )︂
Q0 i = −i arctan u , Q0 i = −i arctan (3.20)
E E b
we obtain
GM⊕ E
V0 (u) = V0∗ (u) = arctan u . (3.21)
E
This corresponds to the “central field” of a spherical-harmonic expansion
GM⊕ r . Using this, we may scale equation 3.18 by substituting the
/︁
í Õ ! ¤. û
Ellipsoidal harmonics 3.8
81
and thus also b → a, β → ϕ, and u → r. We assume that Heiskanen
and Moritz (1967) equation 1-112,
Qnm i u
(︁ )︁ (︂ )︂
a n+1
lim (︁ Eb )︁ = r
E→0 Qnm i
E
V(u, β, λ) = V(r, ϕ, λ) =
∑︂∞ ∑︂n (︂ )︂
a n+1
= r Pnm (sin ϕ) (Aenm cos mλ + Benm sin mλ) , (3.22)
n=0 m=0
Substituting these into equation 3.22 affirms its equivalence with equa-
tions 3.15 and 3.16 for spherical harmonics.
í Õ ! ¤. û
82 3 Legendre functions and spherical harmonics
^ Self-test questions
1. What are the harmonic degree and harmonic order in a spherical-
harmonic expansion? How do they relate to the resolution of the
expansion on the Earth’s surface?
2. What types of spherical harmonics are there? Explain their
dependence on latitude and longitude.
3. How many times does a surface spherical harmonic Ynm (ϕ, λ)
change its algebraic sign travelling along a meridian from the
south pole to the north pole? How many times when travelling
around the Earth along the equator?
4. What does it mean if it is said that two functions are mutually
orthogonal? Give a possible definition of the scalar product of two
functions.
5. How does the attenuation of spherical harmonics with height
behave? Why does a gravimetric satellite that is trying to map the
gravity field of the Earth at a high resolution fly in as low an orbit
as possible?
í Õ ! ¤. û
Exercise 3 – 1: Attenuation with height of a spherical-harmonic expansion
83
6. What does the degree constituent equation express?
7. Which spherical-harmonic coefficients are associated with the
dipole moment of the Earth’s mass distribution? Why are they
missing from tableau 3.5?
í Õ ! ¤. û
84 3 Legendre functions and spherical harmonics
3. For what degree number n will they be less than 10−4 × of what
they are on the Earth’s surface?
1 dn (︁ 2 )︁n
Pn (t) = t −1 .
2n n! dtn
We can observe the following properties:
◦ Differentiating a symmetric function of t will produce an antisym-
metric function, and vice versa.
(︁ )︁
◦ The function t2 − 1 and its powers are symmetric.
◦ Thus: for even n values, Pn (t) = Pn (−t): Pn is symmetric between
the northern and southern hemispheres, and for odd n values
Pn (t) = −Pn (−t): Pn is antisymmetric between hemispheres.
(︁ )︁
◦ Similarly, for even n, Pn (sin ϕ) = Pn sin(−ϕ) , and for odd n,
(︁ )︁
Pn (sin ϕ) = −Pn sin(−ϕ) .
Questions
1. What is the corresponding rule for the functions Pnm , in
other words, for which values n and m is it symmetric and
for which values antisymmetric?
2. Fill in the diagram (n = 0, . . . , 5, m = 0, . . . , n) with either
‘S’ (symmetric) or ‘A’ (antisymmetric) in each framed cell:
í Õ ! ¤. û
Exercise 3 – 3: Algebraic-sign domains of spherical harmonics
85
n= 0 1 2 3 4 5
m=0
1
2
3
4
5 ×
3. What is the logic of symmetry?
4. If the field is mirror symmetric between the northern and
southern hemispheres, so V(ϕ, λ, r) = V(−ϕ, λ, r), which of
the spherical-harmonic coefficients anm and bnm drop out
of the series expansion? Why?
Hint: see the example formulas and graphs for Pnm (sin ϕ)
in this chapter and try to guess a general rule. Then, verify.
5. The same question if the potential is rotationally symmetric
about the Earth’s rotation axis: V(ϕ, λ, r) = V(ϕ, r).
í Õ ! ¤. û
86 3 Legendre functions and spherical harmonics
í Õ ! ¤. û
^ The normal gravity field
4
^ 4.1 The basic idea of a normal field
Just as the figure of the Earth can be approximated by an ellipsoid of pyörähdys-
revolution, the gravity field of the Earth can also be approximated by a ellipsoidi
field of which one equipotential surface, or level surface, is precisely this
ellipsoid of revolution, the reference ellipsoid. vertaus-
ellipsoidi
This brings a logical idea to mind: why not define intercompatibly
a reference ellipsoid and a model geopotential or normal potential, one
of the equipotential surfaces of which is the reference ellipsoid? After
that, a gravity formula is obtained by taking the gradient of this normal
potential.
After this we may define anomalous quantities, such as the disturbing häiriö-
potential and the gravity anomaly, which then again will be intercom- potentiaali
patible, while being numerically much smaller.
Let the normal potential be U(x, y, z). Then, normal gravity will be
⟨︁ ⟩︁ ∂U
γ (x, y, z) = ∥γ∥ = ∥∇U∥ = − γ · n = − ,
∂n
∂
in which ∂n denotes differentiation in the direction of the exterior
surface normal n to a level surface of the normal field, itself an ellipsoid
as well, see figure 4.1. This direction will differ from the direction of
the normal to the level surfaces of the gravity field, or plumb line, by an luotiviiva
– 87 –
88 4 The normal gravity field
X Field lines
of the normal
n gravity field
Normal
gravity
n
γ
γ
γ Equipotential surfaces of
the normal gravity field
X
Reference ellipsoid
(flattening exaggerated)
U = V ∗ + Φ,
í Õ ! ¤. û
The centrifugal force and its potential 4.2
89
Z
p Centrifugal
force
k X
Gravitation
Gravity
i j
Y
X
p = Xi + Yj.
{︁ }︁
The vectors i, j, k form an orthonormal basis along the (X, Y, Z) axes. ortonormaali
kanta
í Õ ! ¤. û
90 4 The normal gravity field
It follows that
√︂⟨︁ ⟩︁ √︁
p = ∥p∥ = p · p = X2 + Y 2 .
with ω⊕ the rotation rate of the Earth in radians per unit of time. If X
and Y are in metres and ω⊕ is in radians per second, then fω is obtained
in m/s2 .
Here on Earth, gravity measurements are generally made using an
instrument that is at rest with respect to the Earth’s surface: it follows
the rotation of the Earth. If the instrument moves, one must, in addition
to the centrifugal force, take into account another pseudo-force: the
1 Coriolis1 force. Fluids — water, air — on the Earth’s surface, if they
are at rest, only sense gravity, which includes the centrifugal force.
Currents in addition also sense the Coriolis force, which deflects them
sideways and causes the well-known vortex phenomena in the oceans
and atmosphere, such as hurricanes.
We may describe centrifugal force as the gradient of a potential. If
we write for this centrifugal potential
Φ = 12 ω2⊕ X2 + Y 2 ,
(︁ )︁
∂Φ ∂Φ ∂Φ
fω = ∇Φ = i+ j+ k=
∂X ∂Y ∂Z
= 12 ω2⊕ · 2X · i + 12 ω2⊕ · 2Y · j + 0 = ω2⊕ (Xi + Yj) ,
í Õ ! ¤. û
The centrifugal force and its potential 4.2
91
If we add to the gravitational potential V the centrifugal potential Φ,
we obtain the gravity potential or geopotential W:
W = V + Φ.
í Õ ! ¤. û
92 4 The normal gravity field
W(x, y, z) = constant.
{︁ }︁
Let i, j, k be an orthonormal basis along the (x, y, z) axes. Then, in
the direction of the unit vector
e = e1 i + e2 j + e3 k
∇W = g.
Level surfaces and gravity vectors, or plumb lines, are always perpen-
dicular to each other.
í Õ ! ¤. û
Level surfaces and plumb lines 4.3
93
Tangent plane
P W = WP + δW
ϵ Equipotential
surface W = WP
x0 X
x
x axis
Radius of
curvature ρx
∂2 ∂2 g
2
δW = 2
W = ∂xx W = − ρ ,
∂x ∂x x
from which
g
ρx = − .
∂xx W
By determining the curvature in the x and y directions,
def 1 ∂ W def 1 ∂yy W
Kx = ρ = − xxg , Ky = ρ = − g , (4.4)
x y
3 We use here compact Euler notation for partial derivatives, ∂xx , ∂yy , ∂zz , which is
often convenient.
í Õ ! ¤. û
94 4 The normal gravity field
Plumb line
Radius of curvature ρx
P
WP
W=
g
∆W g
/︁
g
∆W x −→
+
W = WP
^ Figure 4.4. The curvature of the plumb line.
we obtain
−2gJ + ∂zz W = −4πGρ + 2ω2⊕ .
By using
∂g ∂g
∂zz W = − =− ,
∂z ∂H
in which H is the height co-ordinate, we obtain for the vertical gradient
of gravity (Heiskanen and Moritz, 1967, equation 2-20):
∂g
= −2gJ + 4πGρ − 2ω2⊕ ,
∂H
an equation found by Ernst Heinrich Bruns (Bruns, 1878, page 13).
í Õ ! ¤. û
Level surfaces and plumb lines 4.3
95
parallel. This means that the plumb lines, being perpendicular to all
equipotential surfaces, must be curved in that direction.
Consider two equipotential surfaces, one for potential WP and one for
potential WP + ∆W. The distance separating them will be ∆H = ∆W g .
/︁
In the direction of co-ordinate x the relative tilt between the two surfaces
will be (︃ )︃
∂ ∂ ∆W ∆W ∂g
∆H(x) = =− 2 .
∂x ∂x g(x) g ∂x
If the starting distance between the surfaces is ∆H, it will take a distance
of
)︂/︄(︃
∆W
/︂ (︂ )︃ /︂
ρx = − ∆H ∂ ∆W ∂g g ∂g
∆H = − g − 2 =
∂x g ∂x ∂x
to bring the tangents together, see figure 4.4. The curvature of the
plumb line is the inverse of this, in both the x and the y co-ordinate
directions:
1 1 ∂g 1 1 ∂g
κx = ρ = g , κy = ρ = g .
x ∂x y ∂y
We can derive the curvature of the field lines of the normal gravity
field, or normal plumb lines, in the same way. The difference is, however,
that we can find a simple mathematical expression for gravity on the
surface of the reference ellipsoid, for example equation 4.8. A good
approximation is
γ(φ) ≈ γa cos2 φ + γb sin2 φ.
With the chain rule
∂γ ∂γ ∂φ 1 ∂γ 1
= = = (−2γa cos φ sin φ + 2γb sin φ cos φ) =
∂x ∂φ ∂x R ∂φ R
γ − γa
= b sin 2φ.
R
This means in the x or south-north and y or west-east direction:
1 ∂γ 1 γb − γa 1 ∂γ
κ∗x = γ ≈ sin 2φ, κ∗y = γ = 0.
∂x R γa ∂y
This also means that the direction of the normal plumb line at height h
is φ(h) = φ(0) + κ∗x h, in which numerically κ∗x = 0.171 ′′ km−1 · sin 2φ
(Heiskanen and Moritz, 1967, equation 5-34).
í Õ ! ¤. û
96 4 The normal gravity field
Astronomical
co-ordinates Φ, Λ
Plumb line
n
Greenwich
n
O Φ
í Õ ! ¤. û
The normal potential in ellipsoidal co-ordinates 4.5
97
natural co-ordinates.
Instead of the potential, orthometric height H may be used. Its definition
is easy to understand if one writes
w WP
∂W 1 1
= −g =⇒ dH = − g dW =⇒ HP = − dW, (4.5)
∂H W0 g(W)
∂
in which the integral is taken along the plumb line of point P. ∂H is
the derivative in the direction of the plumb line, the local normal to
the level surfaces. g is the acceleration of gravity along the plumb
line as a function of place — or of geopotential level. In this case
of orthometric heights, g is the true gravity inside the rock, which
is a non-linear function of place and will also depend on the rock
density. This trickiness of their determination is a problem specific to
orthometric heights. We will return to this later (Heiskanen and Moritz,
1967, chapter 4).
The co-ordinates Φ, Λ, and H also form a natural co-ordinate system.
= 21 ω2⊕ u2 + E2 1 − sin2 β =
(︁ )︁ (︁ )︁
í Õ ! ¤. û
98 4 The normal gravity field
Now
U(β, u) = V ∗ (β, u) + Φ(β, u).
On the reference ellipsoid u = b we have as a requirement U(β, b) = U0 ,
def
which is possible only if (Ae∗ e∗
n = An0 ):
1 2 1 2 2
U0 = Ae∗ = Ae∗
(︁ 2 2
)︁
0 + ω
3 ⊕
b + E 0 + 3 ω⊕ a ,
0 = Ae∗
1 ,
1 2 1 2 2
0 = Ae∗ = Ae∗
(︁ 2 2
)︁
2 − 3 ω⊕ b + E 2 − 3 ω⊕ a ,
0 = Ae∗
n, n = 3, 4, 5, . . . .
GM⊕ E
U0 = arctan + 13 ω2⊕ a2 . (4.6)
E b
From this follows
1 2 2 GM⊕ E
Ae∗
0 = U0 − 3 ω⊕ a = arctan .
E b
The normal gravity potential U is obtained as follows:
V ∗ (u)
⏟ ⏞⏞
0
⏟
∗ GM⊕ E
U(β, u) = V (β, u) + Φ(β, u) = arctan u +
E
A e∗
P2 (sin β) Φ(β,u)
⏟ ⏞⏞2 ⏟ Q (︁i u )︁ (︁⏟ ⏞⏞ ⏟ ⏟ ⏞⏞ )︁ ⏟
1 2 2 2 E 3 2 1
)︁ 1 2
(︁ 2 2 2
+ 3 ω⊕ a sin β − 2 + 2 ω⊕ u + E cos β =
Q2 i Eb 2
(︁ )︁
U(β, b) =
í Õ ! ¤. û
Normal gravity on the reference ellipsoid 4.6
99
V ∗ (b)
⏟ ⏞⏞
0
⏟ ⏟ Ae∗
2 P2 (sin β)
⏞⏞ ⏟ ⏟
Φ(β,b)
⏞⏞ ⏟
GM⊕ E
= arctan + 12 ω2⊕ a2 sin2 β − 61 ω2⊕ a2 + 12 ω2⊕ a2 cos2 β =
E b
GM⊕ E
= arctan + 13 ω2⊕ a2 ,
E b
the constant U0 (equation 4.6), as it had better be!
Z b
/︁
sin β a Z a
tan β = = √ 2 = √ = tan ϕ
cos β X +Y a2
/︁ b X +Y
2 2 b
and /︂(︁
Z 1 − e2 N
)︁
sin φ 1 Z a2
tan φ = cos φ = √ 2 = √ = tan ϕ,
X + Y2 N
/︁ 1 − e2 X2 + Y 2 b2
in which ϕ is the geocentric latitude, see equations 2.4. From this follows
directly
b
tan β = a tan φ,
in which the latitude angle φ is the geodetic or geographic latitude. β
is still the reduced latitude. Now it can be shown (exercise!) that redukoitu
leveysaste
aγ cos2 φ + bγb sin2 φ
γ(φ) = √︁ a . (4.8)
a2 cos2 φ + b2 sin2 φ
This is the famous Somigliana–Pizzetti5 equation. These geodesists de- 5
í Õ ! ¤. û
100 4 The normal gravity field
Z
Q
a
P
β
F2 E ϕ φ F1 X/Y
a O
Hyperbola β = constant
Figure 4.6. Geometry of the meridian ellipse and various types of latitude, as
^ well as the focal points F1 and F2 .
monstrated for the first time that an “ellipsoidal” normal gravity field
which has the reference ellipsoid as one of its equipotential or level
surfaces exists exactly, and that the gravity formula is also a closed
expression in geographic latitude.
í Õ ! ¤. û
Numerical values and calculation formulas 4.7
101
a the equatorial radius of the ellipsoid of revolution, its semimajor isoakselin
axis puolikas
f the flattening
defa−b
f= a ,
in which b is the polar radius or semiminor axis
ω⊕ the rotation rate of the Earth
GM⊕ the total mass of the Earth, including the atmosphere.
U ≈ 62 636 860.8500 +
(︄ )︄
− 9.780 326 77 − 0.051 630 75 sin2 φ −
+ h+
− 0.000 227 61 sin4 φ − 0.000 001 23 sin6 φ
(︄ )︄
+ 0.015 438 99 · 10−4 − 0.000 021 95 · 10−4 sin2 φ −
+ −4 4
h2 +
− 0.000 000 10 · 10 sin φ
+ − 0.000 024 22 · 10−8 + 0.000 000 07 · 10−8 sin2 φ h3 , (4.9)
(︁ )︁
and normal gravity (note the minus sign, U is positive and diminishes
going upwards):
∂U
γ=− ≈ + 9.780 326 77 + 0.051 630 75 sin2 φ +
∂h
í Õ ! ¤. û
102 4 The normal gravity field
Here, the unit of potential is m2/s2 , and the unit of gravity, m/s2 . φ
is geodetic latitude; h (in metres) is the height above the reference
ellipsoid. More precise equations can be found from Heikkinen (1981).
In these equations, the coefficient 9.780 32 . . . m/s2 is equatorial gravity,
and the value −0.030 87 . . . · 10−4 s−2 is the vertical gradient of gravity
on the equator.
Other gravity formulas and reference ellipsoids still in legacy use
(and slowly fading away) are Helmert’s 1906 ellipsoid, the Krasovsky
ellipsoid or SK-42 in Eastern Europe, the International or Hayford
ellipsoid (1924) and its gravity formula, and the Geodetic Reference
System 1967.
í Õ ! ¤. û
The normal potential as a spherical-harmonic expansion 4.8
103
X
Cubic
80 000 000 Quadratic
Linear
60 000 000 Realistic
40 000 000
20 000 000
Figure 4.7. The normal field’s potential over the equator. Heights in kilometres,
^ potential in m2/s2 .
Answers
1. See figure 4.7. The minimum of the quadratic function
is at height 3000 km. The cubic function does not have a
minimum.
2. Not very realistic: the stationary point for potential U (the
normal potential in a co-rotating system) should be located at
approximately 36 000 km height, at the geostationary orbit.
This tells us that polynomial approximation cannot be extrapo-
lated very far. In this case, the interval of extrapolation is of the
same order as the radius of the Earth, and that will no longer
work.
í Õ ! ¤. û
104 4 The normal gravity field
in the following form (Heiskanen and Moritz, 1967 equation 2-39, also
equation 3.16):
GM⊕
(︃ ∑︂
∞ (︂ )︂ ∑︂
a n
n )︃
V(ϕ, λ, r) = r 1− r Pnm (sin ϕ) (Jnm cos mλ + Knm sin mλ) ,
n=2 m=0
then we may also write the normal gravitational potential, V ∗ , into the
form
GM⊕
(︃ ∑︂
∞ (︂ )︂n
a
)︃
∗ ∗
V (ϕ, r) = r 1− Jn r Pn (sin ϕ) ,
n=2
even
def
which contains only even coefficients J∗n = J∗n0 , because the normal
field is symmetric about the equatorial plane.
The coefficients for the GRS80 normal gravitational potential are
6 found6 in table 4.1. Higher terms are usually not needed. The rela-
6 They can also be calculated from equation 2-92 given in Heiskanen and Moritz (1967):
(︁ )︁n
3 e2
(︃ )︃
n+1 J2
J∗2n = (−1) 1 − n + 5n 2 ,
(2n + 1) (2n + 3) e
starting from the values J2 and e2 . The results are the same as in the table’s left
column.
í Õ ! ¤. û
The disturbing potential 4.9
105
degree-two coefficients are non-zero! This is one reason why these
functions are used at all.
Instead of using an ellipsoidal model, we may also use as a normal
gravity potential formula the first two or three terms of the spherical-
harmonic expansion of the real geopotential. Then we obtain, taking
the centrifugal potential along:
Y Y (ϕ, λ) 1 2 (︁ 2
U = r0 + 2 3 + 2 ω⊕ X + Y 2 ,
)︁
r
with the corresponding equipotential surface U = U0 being the “Bruns
spheroid”, or
Y Y (ϕ, λ) Y4 (ϕ, λ) 1 2 (︁ 2
U = r0 + 2 3 + 2 ω⊕ X + Y 2 ,
)︁
+ 5
r r
def
the “Helmert spheroid”. Here, Y0 = GM⊕ while Y2 (ϕ, λ) and Y4 (ϕ, λ)
are taken from the true geopotential.
These equations are easy to compute, but their equipotential or level
surfaces are not ellipsoids of revolution, and in fact not even rotationally
symmetric. They are quite complicated surfaces (Heiskanen and Moritz,
1967, section 2-12)!
However, in geometric geodesy we always use a reference ellipsoid,
so this is also a wise thing to do in physical geodesy.
W = V + Φ,
í Õ ! ¤. û
106 4 The normal gravity field
def
T = W − U = V − V ∗.
W =V +Φ=Φ+
GM⊕
(︃ ∞ (︂ )︂ ∑︂
∑︂ a n
n )︃
+ r 1− r Pnm (sin ϕ) (Jnm cos mλ + Knm sin mλ) ,
n=2 m=0
GM
(︃ ∑︂
∞ (︂ )︂
a n ∗
)︃
U=Φ+ r ⊕ 1− r Jn Pn (sin ϕ) ,
n=2
even
GM⊕ ∑︂ (︂ a )︂n ∑︂
(︃ ∞ n )︃
T =W−U=− r r Pnm (sin ϕ) (δJnm cos mλ + Knm sin mλ) ,
n=2 m=0
(4.11)
in which
⎧
⎨δJ = Jn0 − J∗n if n even,
n0
⎩δJnm = Jnm otherwise.
asteosuus where, in every term, the degree constituent Tn has the same dimension
as T , and
GM⊕ ∑︂
n
Tn (ϕ, λ) = − a Pnm (sin ϕ) (δJnm cos mλ + Knm sin mλ) .
m=0
í Õ ! ¤. û
Self-test questions
107
∑︂
∞
T (ϕ, λ) = T (ϕ, λ, a) = Tn (ϕ, λ),
n=2
from which we see that on the reference level, the terms Tn (ϕ, λ) are
really the degree constituents of the disturbing potential T for a certain
degree number n.
The above expansions are all missing the terms n = 0, 1. Of these,
T0 (ϕ, λ) = T0 is a constant — the global average of the disturbing
potential — and T1 (ϕ, λ) has the form of a dipole field. Its value is
proportional to the cosine of the angle between the geocentric location
vector of the point of calculation and that of the dipole vector. Both
values vanish because it is assumed that
◦ The total mass of the Earth GM⊕ assumed by the normal field is
realistic.
◦ The origin of the co-ordinate reference system is assumed to be at
the centre of mass of the Earth.
See section 3.4 for more.
^ Self-test questions
1. What is the basic idea behind using a normal gravity field?
2. What is the difference between gravity and gravitation?
3. Given the centrifugal potential
Φ = 21 ω2⊕ X2 + Y 2 ,
(︁ )︁
7 Earlier
on we also used for this reference radius (in spherical approximation) the
symbol R.
í Õ ! ¤. û
108 4 The normal gravity field
í Õ ! ¤. û
Exercise 4 – 2: Centrifugal force
109
4. Compute for both a geodetic and a reduced latitude of 45◦ numer-
ical values of gravity for the case of the GRS80 reference ellipsoid.
By how much do they differ?
í Õ ! ¤. û
^ Anomalous quantities of the
gravity field
Here, (Φ, Λ) are astronomical latitude and longitude, that together make
up the direction of the local plumb line, and (φ, λ) are the geodetic luotiviiva
latitude and longitude that similarly make up the direction of the normal
gravity vector or “normal plumb line”.1 See figure 5.1. 1
– 111 –
112 5 Anomalous quantities of the gravity field
Plumb-line
deflections (ξ, η)
Topography
Reference ellipsoid
^ Figure 5.1. Geoid undulations N and deflections of the plumb line ξ and η.
í Õ ! ¤. û
Disturbing potential, geoid height, deflections of the plumb line 5.1
113
Figure 5.2. A geoid model for Finland from 1984. Deflections of the plumb
^ line computed from observations in red (Vermeer, 1984).
í Õ ! ¤. û
114 5 Anomalous quantities of the gravity field
T
N = γ. (5.2)
4 This is the famous Bruns4 equation (Heiskanen and Moritz, 1967, equa-
2 This
is not exactly true, due to the “normal plumb line” not being the same as the
normal on the reference ellipsoid. The error made is tiny.
3 This is not self-evident!
In a local vertical datum the potential of the zero point could
well differ by as much as the equivalent of a metre from the normal potential of a
global reference ellipsoid.
í Õ ! ¤. û
Gravity disturbances 5.2
115
−γ = − grad U
−g = − grad W
P
WP Geoid
UP
N Ellipsoid
UQ (= WP )
Q
Figure 5.3. Equipotential or level surfaces of the gravity field (W) and the
^ normal gravity field (U).
tion 2-144).
Figure 5.3 depicts the situation still better. In this figure, the normal
∂
gravity vector γ = grad U has a length of γ = ∥γ∥ = − ∂h U, from
which it follows, with equation T = W − U, that the separation between
“matching” surfaces W = WP and U = UQ , when WP = UQ , is
UQ − UP W −U T
N≈ γ = P γ P = γ.
í Õ ! ¤. û
116 5 Anomalous quantities of the gravity field
in which the differentiation is done along the plumb line for W and —
slightly imprecisely — along the normal on the reference ellipsoid for
U. The directions of the plumb line and surface normal on the ellipsoid
are actually very close to each other. Therefore, to good approximation
∂W ∂U ∂T
(︂ )︂
δg ≈ − − =− .
∂H ∂H ∂H
In spherical approximation we have
∂T
δg ≈ − . (5.3)
∂r
asteosuus We already expanded the disturbing potential T into constituents for
different spherical-harmonic degree numbers, equation 4.12, and now
we obtain by differentiating with respect to r:
∂ ∑︂ R n+1
(︃ ∞ (︂ )︂ )︃
∂T (ϕ, λ, r)
δg(ϕ, λ, r) = − =− r Tn (ϕ, λ) =
∂r ∂r
n=2
∑︂
∞
n + 1 R n+1
(︂ )︂ ∑︂
∞
n + 1 R n+2
(︂ )︂
= r r T n (ϕ, λ) = r Tn (ϕ, λ), (5.4)
R
n=2 n=2
í Õ ! ¤. û
Gravity anomalies 5.3
117
Topography Plumb line
Telluroid Q P (measurement point)
ζ
Mean sea
level (geoid)
HP
hQ
Ellipsoid
N
Figure 5.4. Reference ellipsoid, mean sea level (geoid), telluroid, and gravity
^ measurement.
í Õ ! ¤. û
118 5 Anomalous quantities of the gravity field
∂T 1 ∂γ
∆g = − +γ T, (5.5)
∂H ∂H
∂T 2
∆g = − − r T, (5.6)
∂r
in which r = R + H is the distance from the Earth’s centre.
By substituting into this equation 5.3 for δg we obtain
2
∆g = δg − r T.
í Õ ! ¤. û
Gravity anomalies 5.3
119
(but for T ) and 5.4 for δg:
∑︂
∞ (︂
n+1 2
)︂ (︂ )︂n+1
R
∆g(ϕ, λ, r) = r − r r Tn (ϕ, λ) =
n=2
∑︂
∞
n − 1 R n+1
(︂ )︂ ∑︂
∞
n − 1 R n+2
(︂ )︂
= r r Tn (ϕ, λ) = r Tn (ϕ, λ) =
R
n=2 n=2
∑︂
∞ (︂ )︂n+2
R
= r ∆gn (ϕ, λ), (5.7)
n=2
global average the same as the normal potential. Also the total mass
GM⊕ and geoid volume5 of the Earth are assumed to be the same as 5
the total mass and volume of the reference ellipsoid. The assumption is
largely justified because GM⊕ can be, and has been, determined very
precisely by satellites, and modern models for the normal potential, like
GRS80, are based on these determinations.6 6
í Õ ! ¤. û
120 5 Anomalous quantities of the gravity field
like EGM2008 give a smaller value of 6 378 136.3 m as the mean location of global mean
sea level. This is something to be well aware of when using the model in production
work. Uncertainty continues to be of decimetre order.
í Õ ! ¤. û
The boundary-value problem of physical geodesy 5.5
121
we do not precisely know. Even if we know the height of the measurement
location above sea level, that still does not give us the measurement
point’s location in space. This location depends additionally on the
location of sea level, the geoid, in space; more precisely its height above
or below the reference ellipsoid.
This is how we arrive at the third boundary-value problem.7 The 7
boundary-value problem of physical geodesy is to determine the potential V
outside a body if given on its surface is the linear combination
∂V
c1 V + c2 ,
∂n
with c1 and c2 suitable coefficients. The variable n represents here
differentiation in the direction of the normal to the boundary surface,
in practice the same as H or r.
In physical geodesy, the following linear combination is given as a
boundary condition:
∂T 1 ∂γ
∆g = − +γ T. (5.5)
∂H ∂H
∂
We see that c1 = −1 and c2 = γ−1 ∂H γ. This equation, the definition
5.5 of gravity anomalies, is known as the fundamental equation of physical
geodesy.
Again in spherical approximation, inverting equation 5.8 yields
R
Tn (ϕ, λ) = ∆gn (ϕ, λ).
n−1
Remember that the functions ∆gn (ϕ, λ) are computable using the degree asteosuus-
constituent equation 3.9 when ∆g(ϕ, λ) is known all over the Earth. yhtälö
7 The third or mixed boundary-value problem is associated with Victor Gustave Robin
(1855–1897), a French mathematician. Then, the Dirichlet problem could be called the
first and the Neumann problem the second boundary-value problem.
í Õ ! ¤. û
122 5 Anomalous quantities of the gravity field
exterior space:
∑︂
∞ (︂ )︂n+1
R ∑︂
∞ (︂ )︂n+1
R R
T (ϕ, λ, r) = r Tn (ϕ, λ) = r ∆gn (ϕ, λ) =
n−1
n=2 n=2
R ∑︂
∞
2n + 1 R n+1
(︂ )︂ x (︁ ′ ′ )︁
= ∆g ϕ , λ , R Pn (cos ψ) dσ ′ . (5.9)
4π n−1 r σ
n=2
In section 8.1 we will give closed form 8.3 for this function for the case
r = R, and a graph.
8 Sir
George Gabriel Stokes PRS (1819–1903) was an Irish-born, gifted mathematician
and physicist who made his career in Cambridge.
í Õ ! ¤. û
The telluroid mapping and the “quasi-geoid” 5.6
123
below the topography by an amount ζ if positive, or above it by an
amount −ζ otherwise. The quantity ζ is called a height anomaly.
Telluroid mapping is an important tool in Molodensky’s gravity field
theory. It is, however, a pretty abstract concept. One may say that the
telluroid is a model of the Earth’s surface, obtained by assuming that
◦ The true potential field of the Earth is the normal potential.
◦ The mathematical mean sea surface or geoid, the reference surface
for height measurement, coincides with the reference ellipsoid.
In other words, the telluroid is a model for the Earth’s topographic
surface that is obtained by taking levelled heights — more precisely,
geopotential numbers obtained from levelling — as if they represented
differences of the normal potential from that of the reference ellipsoid.
In practice, a map of values ζ is often called a “quasi-geoid” model.
The quasi-geoid is usually close to the geoid, except in the mountains,
where the differences can exceed a metre.
One should however remember that the height anomaly ζ is defined
on the topographic surface; a surface that is quite rough in many places.
This means that all variations in topographic height will also be reflected
as variations in this quasi-geoid, in such a way that the quasi-geoid
correlates strongly with the small details in the topography. One can
thus not say that the shape of the quasi-geoid only expresses the figure
of the Earth’s potential field. In it, variations in geopotential and in
topographic height are hopelessly mixed up.
This is why the quasi-geoid is an unfortunate compromise, a conces-
sion to “reference-surface thinking”, which only really works within the vertauspinta
classical geoid concept. Better stick — within Molodensky’s theory —
to the concept of height anomaly, which is a three-dimensional function
or field
ζ(X, Y, Z) = ζ(φ, λ, h).
í Õ ! ¤. û
124 5 Anomalous quantities of the gravity field
9 For the greatest precision, one should consider that the latitude Φ may also not
be a latitude on a geocentric reference ellipsoid. It could be astronomical latitude,
or latitude in some old national co-ordinate system computed on a non-geocentric
ellipsoid, like in Finland KKJ, the National Map Grid Co-ordinate System, which was
computed on the Hayford ellipsoid. The error caused by this is however two, three
orders of magnitude smaller than the effect caused by H − h.
í Õ ! ¤. û
Self-test questions
125
def
in which γ0 (φ) = γ(φ, 0), normal gravity at sea level, is only a function
of latitude. In a country like Finland, equation 5.11 is often sufficiently
precise, although the evaluation of the original equation 4.10 is also
easy.
Free-air anomalies are widely used. Generally, when one discusses
gravity anomalies, one means just this, free-air anomalies. They express
the Earth’s exterior gravity field, including mountains, valleys and
everything.
Questions
1. If gravity at sea level is 9.81 m/s2 , at what height will gravity
disappear, as computed according to the above-mentioned
vertical gravity gradient −0.3 mGal/m ?
2. How physically realistic is this?
Answers
(︂ /︂ )︂
5
1. At −0.3 mGal/m , it takes 9.81 · 10 0.3 m = 3270 km to go
to zero.
2. Not very. The gravity gradient itself drops quickly from the
value of −0.3 mGal/m going up, so this linear extrapolation is
simply wrong.
^ Self-test questions
1. How do deflections of the plumb line and geoid heights relate to
each other?
2. What is the fundamental equation of physical geodesy in spherical
approximation?
3. In what way is a gravity disturbance different from a gravity
anomaly?
4. What units are used for measuring gravity anomalies and gravity
í Õ ! ¤. û
126 5 Anomalous quantities of the gravity field
64◦ 64◦
62◦ 62◦
60◦ 60◦
Figure 5.5. Free-air gravity anomalies for Southern Finland, computed from
the EGM2008 spherical-harmonic expansion. Data © Bureau
Gravimétrique International (BGI) / International Association of
^ Geodesy. Web service BGI, EGM2008.
í Õ ! ¤. û
Exercise 5 – 1: The spectrum of gravity anomalies
127
^ Exercise 5 – 1: The spectrum of gravity anomalies
Use equation 5.8. If we assume that the quadratic mean of the degree
constituents ∆gn of gravity anomalies,
√︃ x
def 1
∥∆gn ∥σ = ∆g2 (ϕ, λ) dσ,
4π σ n
does not depend on the chosen degree number n, how then does the
similarly defined ∥Tn ∥σ depend on n?
In other words: which degree numbers of the gravity field are
relatively strongest in the disturbing potential, and which in the gravity
anomalies?
∂T 2
∆g = − − r T,
∂r
∂
T and compare it with the quantity 2T r . Assume
/︁
calculate ∂r
∂
T or 2T r , dominates?
/︁
r ≈ R. Which of the two, ∂r
í Õ ! ¤. û
128 5 Anomalous quantities of the gravity field
2. Assume that the point is close to sea level. Using the Bruns
equation
T
N = γ,
where γ is average gravity 9.81 m/s2 , compute the geoid height N
of the point.
í Õ ! ¤. û
^ Geophysical reductions
^ 6.1 General
6
We see that integral equations, like Green’s third theorem 1.25, offer a
possibility to calculate the whole exterior potential of the Earth — as
well as all the quantities that may be calculated from the potential, such
as, for example, the gravitational acceleration — using values of V or
∂
∂n
V — or a linear combination — observed on the boundary surface
only. A requirement is, that there are no masses outside the boundary
surface.
Green’s third theorem is but one example of many: every integral
theorem is the solution of some boundary-value problem. reuna-arvo-
tehtävä
There are three alternatives for choosing a boundary surface:
1. Choose the topographic surface of the Earth.
2. Choose mean sea level, more precisely, an equipotential surface
close to mean sea level called the geoid.
3. Choose the reference ellipsoid. vertaus-
ellipsoidi
◦ Alternative 1 has been developed mostly by the Molodensky
school (Molodensky et al., 1962) in the former Soviet Union. The
advantage of the method is that we need no gravity reduction, as all
masses are already inside the boundary surface. Its disadvantage
is that the, often complex, shape of the topography must be taken
– 129 –
130 6 Geophysical reductions
í Õ ! ¤. û
Bouguer anomalies 6.2
131
z
P
dz
β ds ⏟ ⏞⏞ ⏟
ℓ
dβ
cos β
H ℓ
d
s dV x
dV = ds · dz · s dα
in the public discussion on the figure of the Earth, and in 1735–1743 led an expedition
of the French Academy of Sciences doing a grade measurement in Peru, South America,
at the same time as De Maupertuis was carrying out a similar grade measurement in
Lapland. In addition to geodesy, he was also active in astronomy.
í Õ ! ¤. û
132 6 Geophysical reductions
right)
ℓ
ds dz = dβ dz,
cos β
in which the determinant of Jacobi needed, ℓ cos β , is seen.
/︁
í Õ ! ¤. û
Bouguer anomalies 6.2
133
Evaluation point P
II
I Bouguer plate
I
Topography d=H
where we assume for the density ρ of the plate an often-used value for
the average density of the Earth’s crust, ρ = 2670 kg/m3 . By substituting
í Õ ! ¤. û
134 6 Geophysical reductions
^ 6.2.2 Properties
Unlike free-air anomalies which vary on both sides of zero, Bouguer
anomalies are strongly negative, especially in the mountains. For example,
if the mean elevation of a mountain range is H = 1000 m, the Bouguer
systematiikka anomalies will, as a consequence of this, contain a bias of 1000 ×
(−0.1119 mGal) = −112 mGal, about −100 mGal for every kilometre of
elevation.
The advantage of Bouguer anomalies is their smaller variation with
place. For this reason they are suited especially for the interpolation and
prediction of gravity anomalies, in situations where the available gravi-
metric material is geographically sparse. This requires that topographic
heights are known to a better spatial density.
í Õ ! ¤. û
Terrain effect and terrain correction 6.3
135
Topography
Free-air anomaly
Bouguer anomaly
Both errors work in the same direction! Because volumes I are below
the point of evaluation, their attraction — which the simple Bouguer
reduction considers present, and removes — acts downwards. And
because volumes II are above the point of evaluation, their attraction
— which in the simple Bouguer reduction is not corrected for — acts
upwards. The error made is in the same direction as in the previous
case.
We write
∆gB′ = ∆gB + T C,
where T C — the “terrain correction” — is positive. ∆gB′ is called the
terrain-corrected Bouguer anomaly.
The terrain correction is calculated by numerical integration. Figure
6.5 shows the prism integration method, and how both prisms, I and II, lead
to a positive correction, because prism I is computationally added and
prism II removed. One needs a digital terrain model, DTM, which must
be, especially around the evaluation point, extremely dense: according
to experience, 500 m is the maximum inter-point separation in a country
like Finland, and in the mountains one needs even 50 m. The systematic
nature of the terrain correction makes a too-sparse terrain model cause,
í Õ ! ¤. û
136 6 Geophysical reductions
64◦ 64◦
62◦ 62◦
60◦ 60◦
í Õ ! ¤. û
Terrain effect and terrain correction 6.3
137
H x ′, y ′
(︁ )︁
Topography
θ II
H(x, y) P
I
H x ′, y ′
(︁ )︁
Geoid
x ′, y ′ x, y x ′, y ′
^ Figure 6.5. Calculating the classical terrain correction using the prism method.
To compute the terrain correction with the prism method, we use the
following equation, assuming a constant crustal density ρ and a flat
Earth, in rectangular map co-ordinates x, y:
w +Dw +D )︂2
1
(︂ (︁
′ ′
= 12 Gρ dx ′ dy ′ ,
)︁
T C(x, y) H x , y − H(x, y)
−D −D ℓ3
in which
√︃ (︂ (︁ )︁)︂2
ℓ= (x − x ′ )2 + (y − y ′ )2 + 1
2
H(x ′ , y ′ ) − H(x, y)
í Õ ! ¤. û
138 6 Geophysical reductions
Given: gravity g
on terrain Bouguer-
Terrain
plate
correction
correction
Free-air
reduction Subtraction of
to sea normal gravity
level at sea level, −γ0 (φ)
Figure 6.6. The steps in calculating the Bouguer anomaly. The reduction
to sea level uses the standard free-air vertical gravity gradient,
^ −0.3084 mGal/m , the vertical gradient of normal gravity.
í Õ ! ¤. û
Terrain effect and terrain correction 6.3
139
P
Q
300 m
200 m
Q′ Sea level
Figure 6.7. A special terrain shape. The vertical rock wall at PQ is also straight
^ on a map and extends to infinity in both directions.
1
TC = 2
· 2πGρ · H =
1
= 2
· 0.1119 mGal/m · 100 m = 5.595 mGal.
TC = 5.595 mGal,
í Õ ! ¤. û
140 6 Geophysical reductions
í Õ ! ¤. û
Spherical Bouguer anomalies 6.4
141
remote part of the shell contributes as much attraction as the
neighbourhood of the evaluation point!
2. The bathymetry of the oceans is accounted for2 by replacing the 2
water with standard-density crustal rock. This contribution to the
anomalies is positive.
3. The topography and bathymetry of remote parts of the globe
are also taken into account realistically. As most of the Earth is
covered by deep ocean, this causes a strong positive general bias,
also in low-lying areas, where the planar Bouguer reduction is
typically small.
4. As the terrain correction is now calculated over the whole globe in
spherical geometry — though only for the topography. Therefore
the rule that all its contributions are positive no longer applies:
Abrehdary et al. (2016) report that in places near mountain ranges
below the local horizon, the spherical terrain correction can be as
negative as −200 mGal.
There is a large systematic difference between the planar and spherical
Bouguer anomalies, which however is very long-wavelength in nature,
and even in an area the size of Australia almost a constant, −18.6 mGal
within a variation interval of a few milligals. The details in the Bouguer
maps look the same (Kuhn et al., 2009).
Just for fun, we compute the net mass effect of doing the complete
spherical Bouguer reduction globally. The mean height of the land
topography is 800 m, land occupying 29 % of the globe. The mean ocean
depth is 3700 m, corresponding to an equivalent rock depth to be “filled
in” of
2670 − 1030
3700 × m = 2272 m,
2670
assuming a density for crustal rock of 2670 kg/m3 , a sea-water density of
1030 kg/m3 , and ocean occupying 71 % of the globe. The sum weighted
2 One can also do so, and often does, in connection with the Bouguer-plate correction.
í Õ ! ¤. û
142 6 Geophysical reductions
by area is thus
κ = ρH,
where H is the height of the topography above sea level and ρ its mean
matter density. This mass surface density can be interpreted as a column
mass integral:
w R+H
κ=ρ dz.
R
3 Friedrich
Robert Helmert (1843–1917) was a German geodesist known for his work
on mathematical and statistical geodesy.
í Õ ! ¤. û
Helmert condensation 6.5
143
Equipotential surface
Topography
g′ g
Condensation layer
Figure 6.8. Helmert condensation and the changes it causes in the gravity
^ field.
í Õ ! ¤. û
144 6 Geophysical reductions
^ 6.6 Isostasy
í Õ ! ¤. û
Isostasy 6.6
145
Plumb-line deflections
Geoid
Mountain
Earth’s crust
“Root”
Earth’s mantle
^ Figure 6.10. Isostasy and the deflection of plumb lines towards the mountain.
4 John
Henry Pratt (1809–1871) was a British clergyman and mathematician who
worked as the archdeacon of Kolkata, India. Wikipedia, John Pratt.
5 John Fillmore Hayford (1868–1925) was a United States geodesist who studied isostasy
í Õ ! ¤. û
146 6 Geophysical reductions
Mountains
Sea
Compen-
sation Earth’s crust
depth
Compen-
sation
level
Mantle
lightest material, and the boundary between this light root material
and the denser material of the Earth’s mantle would be at a fixed depth.
This model, which nowadays finds little acceptance, is illustrated in
figure 6.11.
6 Another classical isostatic hypothesis is due to G. B. Airy.6 Because V.
7 A. Heiskanen7 used it extensively and developed its mathematical form,
í Õ ! ¤. û
Isostasy 6.6
147
Mountains
Sea ρw
Earth’s crust
Compensation depth t0
ρc
Anti-root Compen-
sation
level
Mountain root
ρm
Mantle
t r
t0 t
r
Anti-root
Root
í Õ ! ¤. û
148 6 Geophysical reductions
ρc , the density of the mantle ρm , the density of sea water ρw , sea depth
d, crustal thickness t, and topographic height H. We obtain
d (ρm − ρw ) + c
tρc + dρw − (t + d) ρm = c =⇒ t=− ρm − ρc
on the sea, and
Hρ − c
tρc − (t − H) ρm = c =⇒ t = ρ m− ρ
m c
In the equations, the constant c is, at least from the viewpoint of isostatic
equilibrium, arbitrary and expresses the fact that the level from which
one computes the depth of the root — less precisely, the “average
thickness of the crust” — can be chosen arbitrarily.
Another approach: instead of c, use the “zero topography compen-
sation level”, for short compensation depth, t0 , to be computed from the
above equations by setting H = d = 0:
t0 (ρc − ρm ) = c.
Hρc − t0 (ρc − ρm ) ρ
r= ρm − ρc = t0 + H ρ −c ρ , (6.5)
m c
8 Its
dimension, after multiplication with ambient gravity g, is pressure: according
to Archimedes’ law, the pressure of the crustal (plus sea-water) column minus the
pressure of the column of displaced mantle material.
í Õ ! ¤. û
Isostasy 6.6
149
and under the sea
d (ρc − ρw ) + t0 (ρc − ρm ) ρ −ρ
r=− ρm − ρc = t0 − d ρc − ρw , (6.6)
m c
In other words,
∑︂
(deviation × density contrast) = constant.
interfaces
í Õ ! ¤. û
150 6 Geophysical reductions
Answers
1. We use equation 6.5, finding
ρ
r − t0 = H ρ −c ρ =
m c
2670 kg/m3
= 1100 m × = 4196 m.
(3370 − 2670) kg/m3
Here we have used standard densities for crustal and mantle
rock, respectively.
2. We use equation 6.6, finding
ρ −ρ
r − t0 = −d ρc − ρw =
m c
(2670 − 1030) kg/m3
= −2000 m × = −4686 m,
(3370 − 2670) kg/m3
using the standard density value for sea water.
3. The depth contrast between root and anti-root is 4196 −
(−4686) m = 8882 m. For perspective, Mount Everest is
8848 m above sea level.
í Õ ! ¤. û
Isostasy 6.6
151
Mid-oceanic ridge
Plate motion
Deep-sea trench Conrad
discontinuity
Sea X
Earth’s crust
Mohorovičić
Lithosphere discontinuity
Figure 6.14. The modern understanding of isostasy and plate tectonics. Deep-
^ sea trenches are known to be in isostatic disequilibrium.
í Õ ! ¤. û
152 6 Geophysical reductions
í Õ ! ¤. û
The “isostatic geoid” 6.8
153
areas, section 6.2. This of course is because isostatic reduction is only
the shifting of masses from one place to another — from mountains to
roots beneath the same mountains, the mass deficit of which is pretty
precisely the same as the mass of the mountains themselves sticking out
above sea level — rather than removal of masses, which is what Bouguer
reduction does.
The reduction methods used in isostatic calculations are similar
to those in other reductions. We will discuss them later: numerical
integration in the space domain — grid integration, spherical-cap
integration, least-squares collocation (LSC), finite elements, etc. — or in pienimmän
the spectral domain, for example FFT and “Fast Collocation”. neliösumman
kollokaatio
The question of the hypothesis assumed to apply is a more interest-
ing one. Traditionally, the Pratt or Airy hypotheses have been used,
developed into quantitative methodologies by Hayford or Heiskanen or
Vening Meinesz.9 A newer approach has been to use real measurement 9
data from seismic tomography in order to model the interior structure
of the Earth. With real measurement data, if reliable, one should get
better results.
í Õ ! ¤. û
154 6 Geophysical reductions
64◦ 64◦
62◦ 62◦
60◦ 60◦
í Õ ! ¤. û
The “isostatic geoid” 6.8
155
small indirect effect, it follows that isostatic methods are not perhaps
the best possible if the intent is to calculate a model of the geoid or
quasi-geoid representing the exterior potential.10 Heiskanen and Moritz 10
(1967) call on their page 152 the indirect effect ”moderate”.
However, isostatic methods are very suitable for elucidating the
interior structure of the Earth, because both the topography and the
imprint it makes on the Earth’s mantle, the isostatic compensation, are
computationally removed.
Research has shown that the great topographic features of the Earth
are some 85–90 % isostatically compensated (Heiskanen, 1960). This is
valuable information if no other knowledge is available.
This is the second reason why the isostatic geoid is of interest: the
gravity field of an Earth from which the effect of mountains has been
removed completely — mountain roots and all — can uncover physical
unbalances existing in deeper layers, and processes causing these. Such
processes include especially convection currents in the Earth’s mantle
as well as the possible effect of the liquid outer core of the Earth
on these currents. Interesting correlations have been found between
mantle convection patterns, the global map of the geoid, and the electric
current patterns in the core causing the Earth’s magnetic field (Wen and
Anderson, 1997; Prutkin, 2008; Kogan et al., 1985).
Isostatic reduction consists of two parts:
◦ computational removal of the topography
◦ computational removal of the isostatic compensation of the topog-
raphy.
It is possible to calculate both parts exactly using prism integration, see
section 6.3. Here however we shall gain understanding by a qualitative
approach. We approximate both parts with a single mass-density layer,
with density for example κ = ρH for the topography. We place the first
10 Of
course Bouguer reduction is even worse! The indirect effect can be hundreds of
metres.
í Õ ! ¤. û
156 6 Geophysical reductions
1 ∑︂ R n+1
∞ (︂ )︂
1
= r Pn (cos ψ),
ℓ R
n=0
We obtain for the potential field of the mass-density layer at sea level,
when the evaluation point is also placed at sea level, H = 0 =⇒ r = R:
x ∑︂
∞
Vtop = GR κ Pn (cos ψ) dσ
σ
n=0
Vcomp =
x (︃ (︂ )︂2 )︃ ∑︂
∞ (︂ )︂n+1
R R−D
= G (R − D) −κ Pn (cos ψ) dσ =
σ R−D R
n=0
x ∑︂
∞ (︂
R−D n
)︂
= −GR κ Pn (cos ψ) dσ,
σ R
n=0
í Õ ! ¤. û
The “isostatic geoid” 6.8
157
Sea level
Compensation depth
11 This works on dry land and on the ocean. Lakes, glaciers and areas like the Dead Sea
are more complicated.
í Õ ! ¤. û
158 6 Geophysical reductions
∑︂
∞
4πGR
(︃ (︂
R−D n
)︂ )︃
− 1− κn (ϕ, λ) =
2n + 1 R
n=1
x ∑︂
∞ (︃ (︂
R−D n
)︂ )︃
′ ′
= − GR κ(ϕ , λ ) 1− Pn (cos ψ) dσ ′ = δViso .
σ R
n=1
It follows that
∑︂
∞
4πGR
(︃ (︂
R−D n
)︂ )︃
δViso =− 1− κn (ϕ, λ) =
2n + 1 R
n=1
∑︂
∞
2
(︃ (︂
R−D n
)︂ )︃
=− R 1− 2πGκn =
2n + 1 R
n=1
∑︂
∞
2
(︃ (︂ )︂ )︃
R−D n
=− R 1− (AB )n .
2n + 1 R
n=1
∑︂
N
2nD ∑︂
N
δViso ≈− (A ) ≈ − D (AB )n ≈ −DA
˜︁ B ,
2n + 1 B n
n=1 n=1
∞
∑︂ 2R
δViso ≈ − (AB )n ,
2n + 1
n=N+1
where the terms are small and rapidly falling to zero. In this degree range, the mass-
density layer approximation for the topography and its compensation breaks down,
but it hardly matters as these short wavelengths aren’t even isostatically compensated.
í Õ ! ¤. û
The “isostatic geoid” 6.8
159
and
δViso DA˜︁ B DAB
δNiso = γ ≈ − γ ≈− γ . (6.8)
This is the indirect effect of isostatic reduction.
Let us substitute realistic values. Let the depth of the Mohorovičić13 13
discontinuity be on average ∼ 20 km.14 14
í Õ ! ¤. û
160 6 Geophysical reductions
These are values for typical, extended continental or ocean areas, and
only indicative. Precise calculation needs to be numerical.
Equation 6.8 is, through the Bouguer-plate attraction AB , linear in the
height H. This means that every added kilometre of topography causes
about −2.2 m in the quantity δNiso, land , and every added kilometre of
bathymetry similarly +1.4 m in the quantity δNiso, sea . We may also
conclude that in the isostatic reduction’s effect on the geoid – at least
for longer wavelengths 2πR n , longer than the compensation depth D
/︁
^ Self-test questions
1. Which effects are computationally removed in
(a) the simple Bouguer reduction?
(b) the terrain-corrected Bouguer reduction?
(c) the isostatic reduction?
2. Why is the terrain correction always positive?
3. Why do Bouguer anomalies have good interpolation properties,
and on what condition — in other words, which additional infor-
mation must be available at the interpolation stage?
4. How was it discovered that mountains have roots?
5. Explain the isostatic hypotheses of Pratt–Hayford and Airy–
Heiskanen.
í Õ ! ¤. û
Exercise 6 – 1: Gravity anomaly
161
P
Q
600 m
300 m
Q′ Sea level
í Õ ! ¤. û
162 6 Geophysical reductions
^ Exercise 6 – 4: Isostasy
Assume Airy–Heiskanen isostatic compensation (figure 6.12). The
density of the Earth’s crust ρc = 2670 kg/m3 , density of the mantle
ρm = 3370 kg/m3 , so the density contrast at the crust-mantle interface is
vertaustaso 700 kg/m3 . Let the reference level for the interface corresponding to zero
topography be −25 km, so t0 = 25 km.
1. Calculate the depth of the “root” of an 8 km high mountain
below the reference level −25 km, assuming it is isostatically
compensated.
2. The volcano Mauna Kea, Hawaii, is 4 km above sea level, however
the surrounding sea is 5 km deep. How deep is the root of Mauna
Kea below the reference level?
3. How much is the “anti-root” of the surrounding sea above the
reference level? Let the density of sea water be 1030 kg/m3 .
4. So, how deep is the root of Mauna Kea relative to its surroundings?
í Õ ! ¤. û
^ Vertical reference systems
7
^ 7.1 Levelling, orthometric heights and the geoid
Heights have traditionally been determined by levelling. Levelling is
a technique for determining height differences using a level (levelling
instrument) and two rods or staffs. The level comprises a telescope and
a spirit level, and in the measurement situation the telescope’s optical
axis, the sight axis, is pointing along the local horizon. Levelling staffs
are placed on two measurement points, and through the measuring
telescope, measurement values are read off them. The difference
between the two values gives the height difference between the two
points in metres.
The distance between level and staffs is 40–70 m, as longer distances
would cause too large errors due to atmospheric refraction. Longer dis-
tances are measured by repeat measurements using several instrument
stations and intermediate points.
The height differences ∆H thus obtained are not, however, directly
useable. The “height difference” between two points P and Q, obtained
by directly summing the height differences ∆H, depends namely also
on the path chosen when levelling from P to Q. Also the sum of height
∑︁
differences ⃝ ∆H around a closed path is (generally) not zero.
– 163 –
164 7 Vertical reference systems
00
10
Level
back fore
20
View
∆H = back − fore
∑︂
P
WP = W0 − (∆H · g) ,
sea level
∑︂
P
CP = − (WP − W0 ) = (∆H · g) ,
sea level
í Õ ! ¤. û
Orthometric heights 7.2
165
system is called N60. However, the precise realisation is a special pillar
in the garden of the Helsinki astronomical observatory in Kaivopuisto.1 1
The new Finnish height system is called N2000, and the realisation of its
reference level is a pillar at the Metsähovi research station. In practice
N2000 heights are, at the decimetric precision level, heights over the
Amsterdam NAP datum.
Other countries have their own, similar height reference or datum
points: Russia has Kronstadt, Western Europe the widely used Ams-
terdam datum NAP, southern Europe has the old Austro-Hungarian
harbour city of Trieste, North America the North American Vertical Da-
tum 1988 (NAVD88) with the datum point Father Point (Pointe-au-Père)2 2
in Rimouski, Quebec, Canada, etc.
1 However, the value engraved on the pillar is the reference height of the still older
NN system, not N60. The correct reference value for N60 for this pillar, 30.513 76 m, is
given in the publication Kääriäinen (1966), page 49.
2 The district Pointe-au-Père of the city of Rimouski was named after the Jesuit priest
Father Henri Nouvel (1621?–1701?), who served forty years with the native population
of New France. Pointe-au-Père is also notorious as the location of the RMS Empress of
Ireland shipwreck in 1914, in which over a thousand passengers perished.
í Õ ! ¤. û
166 7 Vertical reference systems
í Õ ! ¤. û
Orthometric heights 7.2
167
P
WP
∆H3 g
∆H3′
g ∆H2 H ∆H2′
∆H1 ∆H1′
O W0
Geoid
Figure 7.3. Levelled heights and geopotential numbers. The height obtained
∑︁
by summing the levelled height differences, 3i=1 ∆Hi , is not the
∑︁3 ′
correct height above the geoid: i=1 ∆Hi computed along the
plumb line.
The equipotential or level surfaces of the geopotential are not parallel:
because of this, a journey along the Earth’s surface may well go
“upwards”, to increasing heights above the geoid, although the
geopotential number decreases. Thus, water may flow “upwards”.
The gravity vector g is everywhere perpendicular to the level
surfaces, and its length is inversely proportional to the distance
^ separating the surfaces.
“The level surface of the Earth’s gravity field that fits on average best to
the mean sea level.”
í Õ ! ¤. û
168 7 Vertical reference systems
and z is the measured distance from the geoid along the plumb line.
Because the equation for g already itself contains H, we obtain the
solution iteratively, starting from a crude initial estimate for H. The
suppeneminen iteration converges fast.
We shall see that determining precise orthometric heights is challeng-
ing, especially in the mountains.
í Õ ! ¤. û
Normal heights 7.3
169
South
North
Lake Päijänne: C = − (W − W0 ) = 76.9 GPU gS
gN HS
Lake Päijänne
HN
Geoid: W = W0
Figure 7.4. In terms of orthometric heights, water may sometimes flow “up-
wards”. Although the north and south ends of Lake Päijänne are
on the same geopotential level — 76.9 geopotential units below
that of mean sea level — the orthometric height of the south end
HS is greater than that of the north end HN , because local gravity
g is stronger in the north than in the south. The height difference
in the case of Lake Päijänne is 8 mm (Jaakko Mäkinen, personal
communication). Calculation using the normal gravity field yields
6 mm. The balance of 2 mm comes from the difference between
^ the gravity anomalies at the northern and southern ends.
in which γ0H is the average normal gravity computed between the zero
level (reference ellipsoid) and H∗ along the ellipsoidal normal. So, the vertaus-
method of computing is the same as in the case of orthometric heights, ellipsoidi
but using the normal gravity field instead of the true gravity field.
Heights “above sea level” are for practical reasons given in metres.
For large, continental networks we want to give heights above a compu-
tational reference ellipsoid in metres, and thus heights above “sea level”
also have to be in metres.
Molodensky also proposed that instead of the geoid, height anomalies
would be used, the definition of which is
def T
ζ= , (7.1)
γHh
in which now γHh is the average normal gravity at terrain level. More
precisely, the average of normal gravity along the ellipsoidal normal
over the interval z ∈ H∗ , h , in which H∗ is the normal height of the
[︁ ]︁
í Õ ! ¤. û
170 7 Vertical reference systems
point and h its height from the reference ellipsoid. The parameter z is
the distance from the reference ellipsoid reckoned along the ellipsoidal
häiriö- normal. T is the disturbing potential at the point.
potentiaali
Based on these assumptions, Molodensky showed that
H∗ + ζ = h.
H + N = h.
í Õ ! ¤. û
Normal heights 7.3
171
height type that can be computed directly from geopotential numbers,
and that also would be compatible with similarly defined, so-called
height anomalies, and with geometric heights h reckoned from the
reference ellipsoid.
The geometric height h from the reference ellipsoid may be connected
to the potential U of the normal gravity field through the following
integral equation:
wh
U = U0 − γ(z) dz.
0
Here, U is the normal potential and γ normal gravity. One level surface
of U, U = U0 , is also the reference ellipsoid. The variable z is the
distance from the ellipsoid along its local normal.3 3
By defining
wh
def 1
γ0h = γ(z) dz (7.2)
h 0
we obtain
U − U0
h=− .
γ0h
By using W = U + T and dividing by γ0h we obtain
W − W0 T
= −h
γ0h γ0h
assuming W0 = U0 , the normal potential on the reference ellipsoid.
Next, one could define
? W − W0
H+ = −
γ0h
as a new height type, and
? T
N+ = h − H+ =
γ0h
as the corresponding new geoid height type. It has however the aesthetic
flaw that we divide here by the average normal gravity computed
3 Here we ignore that the normal gravity vector γ(z) is for z ̸= 0 not precisely parallel
with the ellipsoidal normal: the curvature of the field lines of the normal gravity field
or normal plumb lines, section 4.3.2.
í Õ ! ¤. û
172 7 Vertical reference systems
and with equations 7.3, 7.4, and the definitions above of H+ and N+ ,
(︃ )︃
def T T γ0h + H+ N+ H+
ζ= = · ≈N 1+ = N+ + ,
γHh γ0h γHh R R
(︃ )︃
∗ def W − W0 W − W0 γ0h + N+
H =− =− · ≈H 1− =
γ0H γ0h γ0H R
N+ H+
= H+ − .
R
+ +/︁
Because the, already small, correction terms N H R cancel, we finally
obtain
H∗ + ζ = H+ + N+ = h. (7.5)
The quantity γ0H , and thus also normal height H∗ , can be, unlike γ0h ,
computed using only information obtained by (spirit or trigonometric)
í Õ ! ¤. û
Normal heights 7.3
173
+ +
− N RH
N+
ζ
h
H∗
H+
N+ H+
R
γ(z)
í Õ ! ¤. û
174 7 Vertical reference systems
Topography
ζ
Telluroid
H∗ H h
H∗ N ζ
Geoid Reference ellipsoid Quasi-geoid
Figure 7.7. Geoid, quasi-geoid, telluroid and topography. Note the correlation
between the quasi-geoid and topography. Depicted is an area
where N > 0. The separation between geoid and quasi-geoid is
^ exaggerated.
C W − W0
H∗ = =− , (7.6)
γ γ
Height anomaly
W−U T
ζ= = ,
γHh γHh
in which w
1 h
γHh = γ(z) dz.
ζ H∗
The height anomaly ζ, which otherwise is a quantity similar to the
geoid height N, is however located at the level of the topography,
not at sea level. The surface formed by points which are a distance
H∗ above the reference ellipsoid (and thus a distance ζ below or
−ζ above the topography), is called the telluroid. It is a mapping
í Õ ! ¤. û
Difference between geoid height and height anomaly 7.4
175
of sorts of the topographic surface: the set of points Q whose
normal potential UQ is the same as the true potential WP of the
corresponding point P on the topography. See figure 5.4.
Often, as a concession to old habits, we construct a surface that
is at a distance ζ above or a distance −ζ below the reference
ellipsoid. This surface is called the quasi-geoid. It lacks physical
meaning: it is not a level surface, although out at sea it coincides
with the geoid. Its short-wave features, unlike those of the geoid,
correlate with the short-wavelength features of the topography.
Height above the ellipsoid (assumed U0 = W0 )
U − U0
h= ,
γ0h
where wh
1
γ0h = γ(z) dz.
h 0
h = H∗ + ζ.
í Õ ! ¤. û
176 7 Vertical reference systems
both, H and N, one needs the topographic mass density ρ, for which
a standard constant value is often assumed (2670 kg/m3 ), and the local
vertical gradient of gravity, for which generally the vertical normal
gravity gradient (−0.3084 mGal/m ) is assumed.
The difference between height anomaly and geoid height is calculated
as follows.
1. First, calculate the separation between the quasi-geoid and the
“free-air geoid”. The free-air geoid is an equipotential surface of
the harmonically downwards continued exterior potential. If TFA is
the disturbing potential of the exterior, harmonically downwards
continued field, then its difference between topography level and
sea level is:
w H dT (z)
FA
TFA (H) − TFA (0) = dz ≈ −∆gFA H, (7.7)
0 dz
and by using the Bruns equation twice, ζ = T (H) γ = TFA (H) γ
/︁ /︁
4 Here we made the approximation that γ is the same on the topography level as at
sea level.
í Õ ! ¤. û
Difference between geoid height and height anomaly 7.4
177
Because the surface mass density of the plate is Hρ, its assumed
attraction is everywhere on the plumb line of point P:
2πGρ H. (7.9)
See also Heiskanen and Moritz (1967), pages 327–328. As the Bouguer
anomaly ∆gB is strongly negative in the mountains, it follows that the
quasi-geoid is there always above the geoid: approximately, using
equation 6.2:
0.1119 mGal/m 2
ζ−N≈ H ≈ 10−7 m−1 · H2 .
9.81 m/s2
Or, if H is in units of km and ζ − N in units of m:
ζ − N ≈ 0.1 m/km2 · H2 .
í Õ ! ¤. û
178 7 Vertical reference systems
h = H∗ + ζ ,
in which g is the true gravity inside the topographic masses. From this
we obtain
C − (W − W0 )
H= = ,
g g
in which the mean gravity along the plumb line is
w
1 H
g= g(z) dz.
H 0
í Õ ! ¤. û
Calculating orthometric heights precisely 7.6
179
The method is recursive: H appears on both the left and right sides. This
is not a problem: both H and g are obtained iteratively. Convergence is
fast.
In practice one calculates orthometric height using an approximate
formula. In Finland, Helmert orthometric heights have long been used, for
(︁ )︁
which gravity measured on the Earth’s surface, g H , is extrapolated
downwards by using the estimated vertical gravity gradient interior
to the rock. It is assumed that its standard value outside the rock, the
value −0.3084 mGal/m (the free-air gradient), changes to a value that is painovoiman
0.2238 mGal/m greater (twice the standard-density 2670 kg/m3 Bouguer- ilmagradientti
plate effect): the end result is the total inside-rock gravity gradient,
−0.0846 mGal/m .
This is called the Prey5 reduction. As the end result we obtain the 5
following equations (the coefficient is half the gravity gradient, so the
mean gravity along the plumb line is the same as gravity at the midpoint
of the plumb line):
thus
C C
H= = , (7.13)
g g(H) + 0.0423 mGal/m · H
í Õ ! ¤. û
180 7 Vertical reference systems
6 Theodor Niethammer (1876–1947) was a Swiss astronomer and geodesist who created
í Õ ! ¤. û
Calculating normal heights precisely 7.7
181
^ 7.7 Calculating normal heights precisely
For this we use equation 7.6:
C W − W0
H∗ = =− , (7.6)
γ γ
where the average value of normal gravity along the plumb line is
w ∗
1 H
γ = γ0H = ∗ γ(z) dz.
H 0
Because normal gravity is in good approximation a linear function of z,
we may write
∂γ
γ = γ0 + 12 H∗ ,
∂z
∂ def
in which ∂z γ = −0.3084 mGal/m and γ0 (φ) = γ(φ, 0) is normal gravity
computed at height zero. We obtain
γ = γ0 − 0.1542 mGal/m · H∗ .
í Õ ! ¤. û
182 7 Vertical reference systems
Questions
1. Calculate the orthometric height of point P.
ilma-anomalia 2. Calculate the free-air gravity anomaly ∆gFA of point P.
3. Calculate the Bouguer anomaly (without terrain correction)
∆gB of point P.
4. Calculate the normal height of point P.
5. If the geoid height at point P is N = 25.000 m, how much is
the height anomaly (“quasi-geoid height”) ζ?
Answers
1. First attempt:
C 5000
H(0) = g = m = 519.165 m.
9.82
Second attempt (equation 7.13):
= 7.023 mGal.
C
H∗(0) = γ = 509.087 m.
0
í Õ ! ¤. û
Orthometric and normal corrections 7.9
183
The second, equation 7.14:
5000 m2/s2
H∗(1) = =
9.821 500 m/s2 − 0.1542 · 10−5 s−2 · 509.087 m
= 509.128 m,
ζ = N + 0.026 m = 25.026 m.
∑︂
B
HB = HA + ∆H + OCAB .
A
The fact that the difference in orthometric heights between two points
A and B is not equal to the sum of the levelled height differences is due
to gravity not being the same everywhere.
With CA , CB and ∆C the geopotential numbers at A and B, and the
geopotential differences along the levelling line, it holds that CB − CA −
í Õ ! ¤. û
184 7 Vertical reference systems
∑︁B
∆C = 0 because of the conservative nature of the geopotential.
A
Dividing by a constant γ0 yields
CB CA ∑︂ ∆C
B
γ0 − γ0 − γ0 = 0.
A
with gA and gB average gravity values along the plumb lines of A and
B and g gravity along the levelling line. This expression compares
∑︁B
A ∆H, the naively calculated sum of levelled height differences, with
the difference between the orthometric heights of the end points A and
B, calculated according to the definition.
Subtraction yields
(︃
CB CB
)︃ (︃
CA CA
)︃ ∑︂
B (︂
∆C ∆C
)︂
OCAB −0= − γ − − γ − g − γ0 ,
gB 0 gA 0
A
in which
(︃ )︃ (︃ )︃
CB CB γ0 − gB CB γ0 − gB
− γ = γ0 = γ0 HB ,
gB 0 gB
(︃ )︃
CA CA γ0 − gA
− γ = γ0 HA ,
gA 0
∆C ∆C γ0 − g
(︂ )︂
g − γ0 = γ0 ∆H,
í Õ ! ¤. û
Orthometric and normal corrections 7.9
185
Similarly we may also calculate the normal correction (NC) in calculating
normal heights. Start from the equation
∑︂
B
CB CA ∑︂ ∆C
B
NCAB = H∗B − H∗A − ∆H = − − g , (7.16)
γB γA
A A
The identical first term in both equation 7.15 and equation 7.17 can be
traced back to the term
∑︂
B
∆C ∑︂
B
g = ∆H,
A A
í Õ ! ¤. û
186 7 Vertical reference systems
metriikka
c2 dτ2 =
2GM 2 2 2GM −1 2
(︂ )︂ (︂ )︂
dr − r2 dϕ2 + cos2 ϕ dλ2 =
(︁ )︁
= 1− 2 c dt − 1 − 2
c r c r
2W 2 2 2W −1 2
(︂ )︂ (︂ )︂
dr − r2 dϕ2 + cos2 ϕ dλ2 ,
(︁ )︁
= 1 − 2 c dt − 1 − 2
c c
in spherical co-ordinates plus time (ϕ, λ, r, t). Here we see how the rate
ominaisaika of proper time τ is slowed down compared to stationary co-ordinate
time t (time at infinity r → ∞), when the geopotential W increases
closer to the mass. The slowing-down ratio is
√︃
∂τ 2W W
= 1− 2 ≈ 1− 2.
∂t c c
Now c2 , the speed of light squared, is, in the units of daily life, a huge
number: 1017 m2/s2 . This means that measuring a potential difference
of 1 m2/s2 — corresponding to a height difference of 10 cm — using this
method, requires a precision of 1 : 1017 . More traditional, microwave-
based atomic clocks can do precisions of 10−12 –10−14 (Vermeer, 1983a).
With the new optical clocks, the objective should be achievable and
relativistic levelling may become a reality.
The clock works in this way: extreme cooling produces a so-called
Bose–Einstein condensate of atoms, which is trapped inside an optical
valohila lattice formed by six laser beams, an electromagnetic pattern of standing
waves. The clock oscillation is on a different frequency. A Bose–Einstein
7 Karl Schwarzschild (1873–1916) was a German physicist who was the first to derive, in
1915 while serving on the Russian front, a closed spherically symmetric, non-rotating
solution to the field equation of Albert Einstein’s general theory of relativity, the
Schwarzschild metric.
í Õ ! ¤. û
A vision for the future: relativistic levelling 7.10
187
Braunschweig
100 km
Garching
Figure 7.8. An optical lattice clock: the ultra-precise atomic clock of the future
operates at optical wavelengths. To the right, the trajectory of the
^ Predehl et al. (2012) experiment.
condensate has the property that all atoms are in precisely the same
quantum state — like the photons in an operating laser: their matter
waves are coherent. In a way, all the atoms together act as one virtual
atom. The condensate may consist of millions of atoms.
Unfortunately it is not enough that just one laboratory measures time
to extreme precision. One has also to be able to compare the ticking
rates of different clocks over geographical distances. For this, a solution
has also been found: existing fibreoptic cables already in global use for
Internet and telephony are useable for this with small modifications.
The modifications concern the amplifiers in the cables at distances of
some 100 km, which must be replaced by modified ones (Predehl et al.,
2012). In this way, both the traditional precise levelling networks and
the height systems based on GNSS technology and geoid determination
may be replaced by this hi-tech (and hi-science!) solution.
í Õ ! ¤. û
188 7 Vertical reference systems
^ Self-test questions
1. Why are heights calculated directly from levelled height differ-
ences not good enough as a height system?
2. What is a geopotential number?
3. What are orthometric heights?
4. What are normal heights?
5. What is the classical definition of the geoid?
6. What is a height anomaly?
7. What is the quasi-geoid?
8. Why might water sometimes flow in the “wrong” direction, to a
greater height?
9. What is the telluroid?
10. What are the orthometric correction and the normal correction?
− (W − W0 ) = 5000 m2/s2 .
Below the point at sea level, normal gravity is γ0 = 9.821 500 m/s2 .
Calculate the normal height of the point.
í Õ ! ¤. û
Exercise 7 – 3: Difference between orthometric and normal height
189
^ Exercise 7 – 3: Difference between orthometric and
normal height
At point P, the Bouguer anomaly is ∆gB = −120 mGal. The orthometric
height of the point is 1150 m.
1. Calculate the normal height of point P.
2. If the geoid height in point P is N = 21.75 m, calculate the height
anomaly ζ of the point.
í Õ ! ¤. û
^ The Stokes equation and other
integral equations
∑︂
∞ ∑︂
∞
∆gn
T= Tn = R ,
n−1
n=2 n=2
R ∑︂ 2n + 1
∞ x
T= ∆g(ϕ ′ , λ ′ ) Pn (cos ψ) dσ ′ =
4π n−1 σ
n=2
– 191 –
192 8 The Stokes equation and other integral equations
Mass excess
Mass
deficit
−N
x ∑︂
∞
(︄ )︄
R 2n + 1
= P (cos ψ) ∆g(ϕ ′ , λ ′ ) dσ ′ =
4π σ n−1 n
n=2
x
R
= S(ψ) ∆g(ϕ ′ , λ ′ ) dσ ′ , (8.1)
4π σ
in which
∑︂
∞
2n + 1
S(ψ) = P (cos ψ),
n−1 n
n=2
the Stokes kernel function. The angle ψ is the geocentric angular
distance between the evaluation point and moving observation point,
see figure 8.2. The equation above allows the calculation, from global
gravimetric data and for every point on the surface of the Earth sphere,
í Õ ! ¤. û
The Stokes equation and the Stokes integral kernel 8.1
193
N(ϕ, λ)
Evaluation S(ψ)
point
∆g ϕ ′ , λ ′ dσ ′
(︁ )︁
Moving data or
ψ integration point
Earth’s centre
of the disturbing potential T , and from that, the geoid height N using
Bruns equation 5.2, N = T γ . The result is
/︁
T (ϕ, λ) x
R
S(ψ) ∆g ϕ ′ , λ ′ dσ ′ ,
(︁ )︁
N(ϕ, λ) = γ = (8.2)
4πγ σ
in which (ϕ, λ) and (ϕ ′ , λ ′ ) are the evaluation point and the moving
point (“data point”), respectively, and the angular distance between
them is ψ. Equation 8.2 is the classical Stokes integral equation of
gravimetric geoid determination.
The above illustrates the correspondence between integral equations
and spectral expansions. There are other examples of this, like the
spectral representation of the function 1 ℓ , equation 8.7, Heiskanen and
/︁
í Õ ! ¤. û
194 8 The Stokes equation and other integral equations
25
S(ψ)
20 1
sin 12 ψ
15 − 6 sin 12 ψ + 1 − 5 cos ψ
0
30◦ 60◦ 90◦ 180◦
ψ −→
−5
0 0.5 1 1.5 2 2.5 3 3.5
Figure 8.3. The Stokes kernel function S(ψ). The argument ψ is in radians
[︁ )︁
0, π . Also plotted are the three parts of analytical expression 8.3
^ with their different asymptotic behaviours.
equation 5.10:
∑︂
∞ (︂ )︂n+1
R 2n + 1
S(ψ, r, R) = P (cos ψ). (5.10)
r n−1 n
n=2
í Õ ! ¤. û
Example: The Stokes equation in polar co-ordinates 8.2
195
infinity when ψ → 0. The next three terms, − 6 sin 12 ψ + 1 − 5 cos ψ,
[︁ )︁
are all bounded on the whole interval 0, π and
(︂ the value for ψ)︂ = 0
is −4. The last, complicated term − 3 cos ψ ln sin 12 ψ + sin2 21 ψ also
goes to infinity — positive infinity! — for ψ → 0, but much more slowly
because of the logarithm.
◦ Normal gravity:
∂U b
γ(r) = − = − r0 .
∂r
◦ Normal gravity gradient:
∂γ ∂2 U b
= − 2 = 20 .
∂r ∂r r
◦ Gravity anomaly, equation 5.5:
∂T 1 ∂γ
∆g(α, r) = − + T=
∂r γ ∂r
í Õ ! ¤. û
196 8 The Stokes equation and other integral equations
∑︂
∞
k −k
= r r (ak cos kα + bk sin kα) −
k=1
1 ∑︂ −k
∞
−r r (ak cos kα + bk sin kα) =
k=1
∑︂
∞
k − 1 −k
= r r (ak cos kα + bk sin kα) .
k=2
it follows that
∑︂
∞ (︂ )︂k+1
R
∆g(α, r) = r ∆gk (α),
k=2
def
∆gk (α) = (k − 1) R−(k+1) (ak cos kα + bk sin kα) ,
k−1
∆gk (α) = T (α). (8.4)
R k
kantafunktio According to Fourier theory, the basis functions cos kα and sin kα are
orthonormal on the circle r = R when choosing the following integral as
the scalar product:
⎧
w 2π w 2π ⎨0 if k ̸= m,
1 1
cos kα cos mα dα = sin kα sin mα dα =
π 0 π 0 ⎩1 if k = m,
w
1 2π
π cos kα sin mα dα = 0 always.
0
∑︂
∞
∆g(α, R) = ∆gk (α)
k=2
í Õ ! ¤. û
Example: The Stokes equation in polar co-ordinates 8.2
197
into its Fourier terms as follows:
def
∆gk (α) = (k − 1) R−(k+1) (ak cos kα + bk sin kα) =
A B
⏟ ⏞⏞k ⏟ ⏟ ⏞⏞k ⏟
−(k+1) −(k+1)
= (k − 1) R ak cos kα + (k − 1) R bk sin kα.
The substitutions
{︄ }︄ {︄ }︄
ak Rk+1 Ak
=
bk k−1 Bk
yield
∑︂
∞ ∑︂
∞
T (α, R) = Tk (α) = R−k (ak cos kα + bk sin kα) =
k=2 k=2
∑︂
∞ (︃
Rk+1 Rk+1
)︃
−k
= R A cos kα + B sin kα =
k−1 k k−1 k
k=2
∑︂
∞
R
= (A cos kα + Bk sin kα) .
k−1 k
k=2
í Õ ! ¤. û
198 8 The Stokes equation and other integral equations
1
T (α, R) = π ·
∑︂
∞
R
(︃ w 2π w 2π )︃
′ ′ ′ ′ ′ ′
· cos kα ∆g(α , R) cos kα dα + sin kα ∆g(α , R) sin kα dα =
k−1 0 0
k=2
1 ∑︂ R
∞ w 2π
∆g(α ′ , R) · cos k (α − α ′ ) dα ′ .
(︁ )︁
=π
k−1 0
k=2
∑︂∞
cos (k ′ + 1) (α − α ′ ) ∑︂∞
cos k ′ (α − α ′ )
(︁ )︁ (︁ )︁
′
S(α − α ) = ≈ =
k′ k′
k ′ =1 k ′ =1
(︃ )︃
1 1 )︁ ≈ − ln(α − α ′ ).
= ln
2
(︁
2 1 − cos(α − α ′ )
í Õ ! ¤. û
Example: The Stokes equation in polar co-ordinates 8.2
199
^ Tableau 8.1. Stokes equation in two dimensions, octave code.
í Õ ! ¤. û
200 8 The Stokes equation and other integral equations
−2π −π π 2π
−1
geoid undulations on the same circle. Both curves display fairly realistic
statistical behaviour. The code used is given in tableau 8.1.
200
−→ ∆g (mGal)
−→ N (m)
100
−100
−200 −→ α
0◦ 90◦ 180◦ 270◦ 360◦
í Õ ! ¤. û
Plumb-line deflections and Vening Meinesz equations 8.3
201
^ 8.3 Plumb-line deflections and Vening Meinesz
equations
By differentiating the Stokes equation with respect to place, we obtain
integral equations for the components of the deflection of the plumb luotiviivan
line (Heiskanen and Moritz, 1967, equation 2-210’): poikkeama
{︄ }︄ {︄ }︄
ξ(ϕ, λ) x dS(ψ) cos α
1
= ∆g(ϕ ′ , λ ′ ) dσ ′ =
η(ϕ, λ) 4πγ σ dψ sin α
{︄ }︄
x dS(ψ) cos α
1
= ∆g(ϕ ′ , λ ′ ) sin ψ dα dψ, (8.5)
4πγ σ dψ sin α
í Õ ! ¤. û
202 8 The Stokes equation and other integral equations
z P
ℓ r
Q ψ y
R
O
Figure 8.6. The geometry of the generating function of the Legendre polyno-
^ mials.
def def
With the definitions R = ∥R∥ and r = ∥r∥, we may write (cosine
rule):
ℓ2 = r2 + R2 − 2rR cos ψ. (8.6)
We may also write the function 1 ℓ as the following expansion (Heiska-
/︁
in which r and R are the distances of points P and Q from the origin O,
usually the centre of the Earth. The function 1 ℓ is called the generating
/︁
í Õ ! ¤. û
The Poisson integral equation 8.4
203
Now add together this equation and equation 8.7:
1 ∑︂
∞ (︂ )︂n+1
−2r2 + 2rR cos ψ + ℓ2 R
3
= − (2n + 1) r Pn (cos ψ).
ℓ R
n=0
∑︂
∞
(︁ )︁
R r2 − R2 (︂ )︂n+1
R
3
= (2n + 1) r Pn (cos ψ). (8.8)
ℓ
n=0
∑︂
∞ (︂ )︂n+1
R
V(ϕ, λ, r) = r Vn (ϕ, λ),
n=0
we obtain
V(ϕ, λ, r) =
1 ∑︂ R n+1
∞ (︂ )︂ x (︁
V ϕ ′ , λ ′ , R Pn (cos ψ) dσ ′ =
)︁
= r (2n + 1)
4π σ
n=0
∑︂
[︄ ∞
x (︁
]︄
(︂ )︂n+1
1 R
V ϕ ′, λ ′, R Pn (cos ψ) dσ ′ =
)︁
= (2n + 1) r
4π σ
n=0
x R (︁r2 − R2 )︁ (︁
1 ′ ′
)︁ ′
= V ϕ , λ , R dσ
4π σ ℓ3
by replacing the expression in square brackets by equation 8.8.
í Õ ! ¤. û
204 8 The Stokes equation and other integral equations
í Õ ! ¤. û
Gravity anomalies in the exterior space 8.5
205
we can express the gravity anomaly in the external space ∆g(ϕ, λ, r) as
a function of gravity anomalies ∆g(ϕ ′ , λ ′ , R) on a reference sphere of vertauspallo
radius R. The function r∆g is harmonic, because according to equation
5.7
1 ∑︂
∞ (︂ )︂n+1
R
∆g = r (n − 1) r Tn ,
n=2
so
∑︂
∞ (︂ )︂n+1
R ∑︂
∞ (︂ )︂n+1
R
r∆g = r (n − 1) Tn = r Tn′ ,
n=2 n=2
An alternative notation is
x (︁ 2 2
)︁
1 RR r −R
∆g = r ∆g∗ dσ,
4π σ ℓ3
in which ∆g∗ denotes the gravity anomaly at sea level, again calculated
by harmonic downwards continuation of the exterior field, in this case the
expression r∆g.
From equation 8.10 we may lift the closed form of the kernel, which
is dimensionless: (︁ 2 )︁
2
RR r −R
K(ℓ, r, R) = r ,
ℓ3
with which
x
1
K(r, ψ, R) ∆g ϕ ′ , λ ′ , R dσ ′ .
(︁ )︁
∆g(ϕ, λ, r) =
4π σ
í Õ ! ¤. û
206 8 The Stokes equation and other integral equations
∆g(ϕ, λ, r) =
1 ∑︂ R n+1 ∑︂
∞ (︂ )︂ ∞ (︂ )︂n+2
R
= r r (n − 1) Tn (ϕ, λ) = r ∆gn (ϕ, λ).
n=2 n=2
∆g(ϕ, λ, r) =
1 ∑︂ R n+2
∞ (︂ )︂ x
= r (2n + 1) ∆g(ϕ ′ , λ ′ , R) Pn (cos ψ) dσ ′ =
4π σ
n=2
x ∑︂
(︄ ∞ )︄
(︂ )︂n+2
1 R
= r (2n + 1) Pn (cos ψ) ∆g(ϕ ′ , λ ′ , R) dσ ′ =
4π σ
n=2
x
1
= K ∆g(ϕ ′ , λ ′ , R) dσ ′ ,
4π σ mod
in which
∑︂
∞ (︂ )︂n+2
def R
Kmod (ψ, r, R) = r (2n + 1) Pn (cos ψ)
n=2
is the modified spectral version of the Poisson kernel for gravity anoma-
lies. From this kernel, the constituents of degree number 0 and 1 have
been removed, see Heiskanen and Moritz (1967) equation 2-159.
Compared to the Stokes kernel, the Poisson kernel drops off fast to
zero for growing values of ℓ. In other words, the evaluation of the
kalotti integral equation may be restricted to a very local area, like a cap of
í Õ ! ¤. û
The vertical gradient of the gravity anomaly 8.6
207
2
Poisson kernel K
1 1 km K 1 r2 − R2
2
=
2 km R r ℓ3
0
−4
Distance (km) −→
0 1 2 3 4 5 6 7
Figure 8.7. The Poisson kernel function for gravity anomalies as well as the
kernel for the anomalous vertical gravity gradient, both for various
height differences r − R. These kernels are used when evaluating
^ the surface integral in map co-ordinates (x, y) in kilometres.
radius 1◦ . See figure 8.7. The main use of Poisson’s kernel is the harmonic
continuation, upwards or downwards, of gravity anomalies measured
and computed at various levels, shifting them to the same reference vertaustaso
level.
In the limit r → R (sea level becomes the level of evaluation), this
kernel function goes asymptotically to the two-dimensional Dirac δ
function. This is inevitable for a kernel that computes gravity anomalies
from gravity anomalies.
í Õ ! ¤. û
208 8 The Stokes equation and other integral equations
so
1 ∑︂ R n+3
∞ (︂ )︂ x
∂ ∆g(ϕ, λ, r) (︁ ′ ′ )︁
=− r (2n + 1) (n + 2) ∆g ϕ , λ , R Pn (cos ψ) dσ ′ =
∂r 4πR σ
n=2
x
1
K ′ (ψ, r, R) ∆g ϕ ′ , λ ′ , R dσ ′ , (8.11)
(︁ )︁
=
4πR σ
in which the (dimensionless) kernel function is now
∑︂
∞ (︂ )︂n+3
R
′
K (ψ, r, R) = − r (2n + 1) (n + 2) Pn (cos ψ).
n=2
í Õ ! ¤. û
The vertical gradient of the gravity anomaly 8.6
209
^ Tableau 8.2. Derivation of the kernel K ′ for the vertical gradient of gravity
anomaly. Used is the definition of ℓ, equation 8.6, as well as Poisson’s integral
equation 8.10.
x (︃ R
(︄ )︃ )︄
∂∆g(ϕ, λ, r) 1 ∂ (︁ 2 2
)︁ −3 (︁ ′ ′ )︁ ′
= ·R r −R ·ℓ ∆g ϕ , λ , R dσ =
∂r 4π ∂r σ r
R2 x ∂ 1 (︁ 2
(︃ )︃
2
∆g ϕ ′ , λ ′ , R dσ ′ =
)︁ −3 (︁ )︁
= · r −R ·ℓ
4π σ ∂r r
(︁ 2 )︁− 3/2
R2 x
(︄(︃ )︄
r2 − R2 2r 2
)︃
1 1 (︁ 2 d ℓ ∂ℓ
· 3 + · r − R2 · ∆g ϕ ′ , λ ′ , R dσ ′ =
)︁ (︁ )︁
= − 2
+ 2
·
4π σ r r ℓ r d (ℓ ) ∂r
R2 x
(︄ (︃ )︄
r2 − R2 r2 − R2 (︁ 3 −5 )︁
)︃
1
· (2r − 2R cos ψ) ∆g ϕ ′ , λ ′ , R dσ ′ =
(︁ )︁
= 3
− 2
+2 + · −2ℓ
4π σ ℓ r r
R2 x 1 2 2 2 2 2
(︃ )︃
3r −R 1 ℓ +r −R
∆g ϕ ′ , λ ′ , R dσ ′ −
(︁ )︁
= 3
2− 2 2
4π σ ℓ r ℓ r
1 1 x R R r2 − R2
(︁ )︁
− · 3
∆g(ϕ ′ , λ ′ , R) dσ ′ =
r 4π σ r ℓ
R2 x 1
(︄ (︁ 2 )︁ )︄
2
)︁ (︁ 2
2
3r −R
2
3 r −R r − R2 (︁ ′ ′ )︁ ′ 1
= 2 − 2 − 2 ∆g ϕ , λ , R dσ − ∆g(ϕ, λ, r) =
4π σ ℓ3 r2 r2 ℓ2 r
R2 x 1 2 2
(︄ (︁ 2 )︁ )︄ (︃ )︃
3 r −R 1 3
(︁ ′ ′ )︁ ′
= 2− 2 ∆g ϕ , λ , R dσ − + ∆g(ϕ, λ, r) =
4π σ ℓ3 r2 ℓ2 r 2r
R2 x 1 2 2
[︄ (︁ 2 )︁ ]︄
3 r −R 5
∆g ϕ ′ , λ ′ , R dσ ′ − ∆g(ϕ, λ, r). (8.12)
(︁ )︁
= 3
2− 2 2 2
4π σ ℓ r ℓ 2r
Looking at figure 8.7, it is seen that the Poisson kernel K gets narrower
in proportion to r − R and taller in proportion to (r − R)−2 . As the
integral over the Poisson kernel is two-dimensional and scales with the
square of the width, it remains constant when r → R, and in fact the
kernel converges to the two-dimensional Dirac δ function.
For the kernel K ′ of the vertical gradient of the gravity anomaly, the
behaviour is more unpleasant: it narrows in the same way, but, as figure
í Õ ! ¤. û
210 8 The Stokes equation and other integral equations
8.7 shows, gets taller in proportion to (r − R)−3 . This makes its integral
over the sphere diverge in proportion to (r − R)−1 .
Regularisation can be done by observing that a globally constant
gravity anomaly field
(︂ )︂2
∆g ˜︂ 0 (r) = R ∆g0
˜︂ 0 (ϕ, λ, r) = ∆g
r
has a gradient of
∂ ∆g
˜︂ 0 (ϕ, λ, r) 2 ˜︂
= − r ∆g 0 (ϕ, λ, r), (8.15)
∂r
but also, like equation 8.13:
x ∆g
˜︂ (ϕ ′ , λ ′ , R)
∂ ∆g
˜︂ 0 (ϕ, λ, r) R2 5 ˜︂
= κ 0 3 dσ ′ − ∆g (ϕ, λ, r). (8.16)
∂r 4π σ ℓ 2r 0
Subtract equation 8.16 from equation 8.13 and substitute equation 8.15,
yielding
(︁ )︁
∂∆g(ϕ, λ, r) ∂ ∆g(ϕ, λ, r) − ∆g ˜︂ 0 (ϕ, λ, r) ∂∆g
˜︂ 0 (ϕ, λ, r)
= + =
∂r ∂r ∂r
x ∆g ϕ ′ , λ ′ , R − ∆g
(︁ )︁
˜︂ 0 (ϕ ′ , λ ′ , R)
R2
= κ dσ ′ −
4π σ ℓ3
5 ˜︂ 0 (ϕ, λ, r) − 2 ∆g
(︂ )︂
− ∆g(ϕ, λ, r) − ∆g r 0 (ϕ, λ, r) =
˜︂
2r
x ∆g ϕ ′ , λ ′ , R − ∆g
(︁ )︁
R2 0
= κ dσ ′ −
4π σ 3
ℓ (︃ )︃
(︂ )︂2
5 R 2 R 2
(︂ )︂
− ∆g(ϕ, λ, r) − r ∆g0 − r r ∆g0 .
2r
def
Choose the constant ∆g0 = ∆g(ϕ, λ, R), the anomaly at sea level of the
evaluation point:
í Õ ! ¤. û
The vertical gradient of the gravity anomaly 8.6
211
x ∆g ϕ ′ , λ ′ , R − ∆g(ϕ, λ, R) ′ 2 (︂ R )︂2
(︁ )︁
R2
≈ κ dσ − r r ∆g(ϕ, λ, R).
4π σ ℓ3
(8.17)
and the integral 8.17 will converge for r → R. We posit without proof
that for r → R, convergence will be to the same limit as the Heiskanen
and Moritz equation, in other words, the second term in expression
8.14 fades away and, effectively, κ → 2.
If we are integrating over the surface of a spherical Earth of radius
R rather than the unit sphere σ of radius 1 — or, equivalently, in local
metric co-ordinates (x, y) — we can make the substitution dS = R2 dσ,
with dS a surface element on a sphere of radius R. This removes the
factor R2 from integral equations such as 8.10, 8.12, and 8.17.
In Molodensky’s method this or similar equations can be rapidly
evaluated from very local gravimetric data.
The closed expression given in Heiskanen and Moritz (1967, expres-
sion 2-217), is the anomalous vertical gravity gradient evaluated at
sea level (on the reference sphere). In our equations 8.17 and 8.11
we also need gravity anomalies at sea level. However, anomalies at
the topographic surface level are available. In practice, we may proceed
iteratively, by initially assuming that the anomaly values observed at
topography level are at sea level:
∆g(0) ϕ, λ, R ≈ ∆g ϕ, λ, r = ∆g ϕ, λ, R + H ,
(︁ )︁ (︁ )︁ (︁ )︁
í Õ ! ¤. û
212 8 The Stokes equation and other integral equations
approximation:
(0)
(︁ )︁ ⃓⃓
∂∆g ϕ, λ, z
∆g(1) ϕ, λ, R ≈ ∆g ϕ, λ, r −
(︁ )︁ (︁ )︁
H.
⃓
∂z
⃓
⃓
z=r
í Õ ! ¤. û
Gravity reductions in geoid determination 8.7
213
^ 8.7.2 Downwards continuation in linear approximation
The approach described above can, following Molodensky, be linearised:
⎛ (︁ )︁ ⎞
∆g∗ ϕ ′ ,λ ′
⎜⏟ ⏞⏞ ⃓ ⏟⎟
x ⎜
R ∂∆g ⃓⃓ ∂T ⃓
⎟ ⃓
′⎟ ′
T= ⎜∆g − H ⎟ S(ψ) dσ + H. (8.18)
⎜
4π ∂z ⃓z=H ′ ⎟ ∂z z=H
⃓
σ⎜
⎝ ⎠
⏞ ⏟⏟ ⏞
T ∗ (ϕ,λ)
After this, we apply, at sea level, the Stokes equation, and obtain the
disturbing potential at sea level T ∗ . After this, the disturbing potential
is “unreduced” back to terrain level, to the evaluation point, with the
equation
(︁ )︁ ⃓⃓
∂T ϕ, λ, z
T (ϕ, λ, H) = T ∗ (ϕ, λ) + H.
⃓
∂z
⃓
⃓
z=H
∂
In these equations T , its vertical derivative T,
∆g, and its vertical
∂H
∂
derivative ∂H ∆g always belong to the exterior harmonic gravity field.
The connection between them is the fundamental equation of physical fysikaalisen
geodesy, equation 5.5, in spherical geometry geodesian
perusyhtälö
∂T 2
∆g = − − r T, (5.6)
∂r
in which r = R + H. Here, we need firstly the vertical derivative of the
disturbing potential. This is easy:
∂T ∂T 2
= = −∆g − r T,
∂H ∂r
í Õ ! ¤. û
214 8 The Stokes equation and other integral equations
where the first term on the right is directly measured, and the second
term’s T is obtained iteratively from the main product of the solution
process.
Calculating the vertical gradient of gravity anomalies, that is the
anomalous vertical gradient of gravity, is harder. For this task, section
8.6 offers calculation options. Luckily for practical calculations, the
kernels of the integral equations are very localised and one does not
need gravity anomalies from a very large area.
In this case, the reduction takes place from the height of the ∆g mea-
surement point to the height of the T evaluation point. This is likely to
be a shorter distance than from sea level to evaluation height, especially
in the immediate surroundings of the evaluation point. This means
1 that the linearisation error will remain smaller.1 What is bad, on the other
1 The linearisation error could be even further tuned down by choosing as the evaluation
í Õ ! ¤. û
Gravity reductions in geoid determination 8.7
215
Here, we were all the time discussing the determination of the
disturbing potential T (ϕ, λ, H); this is in practice the same as determining
the height anomaly
T (ϕ, λ, H) T (ϕ, λ, H)
ζ(ϕ, λ, H) = ≈ (︁ 1
γHh
)︁,
γ ϕ, 2 (H + h)
equation 7.1. Here, γ is normal gravity calculated for point latitude2 ϕ 2
and topographic height 12 (H + h) ≈ H + 12 ζ.
í Õ ! ¤. û
216 8 The Stokes equation and other integral equations
2πGρ (H − HRTM ) ,
í Õ ! ¤. û
Gravity reductions in geoid determination 8.7
217
P −
− − P′ +
+ + −
Bouguer plate, down- Inverse
Terrain correction wards continuation terrain correction
Figure 8.8. Residual terrain modelling (RTM). One removes the short wave-
lengths, the deviations from the red dashed line, from the terrain
computationally: the masses rising above it are removed, the
valleys below it are filled. After reduction, the red dashed line,
smoother than the original terrain, is the new terrain surface. The
exterior potential of the new mass distribution will differ only
little from the original one, but may be harmonically downwards
continued to sea level.
Left, terrain correction for point P, middle, Bouguer-plate and
gradient reduction to the level of smoothed terrain point P ′ , and
^ right, the inverse terrain correction for point P ′ .
to section 5.4, −0.3 mGal/m — and this operation will leave the
gravity anomaly unchanged.
4. Rigorously speaking, an inverse terrain correction for the shapes
of the smoothed terrain should be applied, to arrive at gravity
anomalies realistic for this new replacement topography. Often
also this step is left out as the effect is small.
5. After that, harmonic downwards continuation of the exterior field
succeeds: almost only long wavelengths are left in the exterior
field.
Because the mass shifts in the RTM method are so small, take place
í Õ ! ¤. û
218 8 The Stokes equation and other integral equations
over such small distances, and are of such a short wavelength in nature,
the indirect effect or “restore” step — the change in geopotential due to
the mass shifts that has to be applied in reverse to arrive at the final
geopotential or geoid solution — is so small as to often be negligible.
For the same reason, the effect of unknown topographic density will
also remain small.
Finally we note that, because the RTM method removes the effect
of the short-wavelength topography, it is also a suitable method for
interpolating gravity anomalies. See Märdla (2017).
í Õ ! ¤. û
Kernel modification 8.9
219
“Remove” “Restore”
Brute force
∆g −−−−−−−−−−−−−−−−−−→ N
⏐ ↑
⏐ Global gravity Global gravity ⏐
−↓ ⏐+
field model field model
∆gloc Nloc
⏐ ↑
⏐ Exterior masses Exterior masses ⏐
−↓ (topography) (topography) ⏐
+
Stokes
∆gred −−−−−−−−−−−−−−−−−−→ Nred
í Õ ! ¤. û
220 8 The Stokes equation and other integral equations
Stokes equation, we will evaluate this integral, after removing the effect
of the global model from the given gravity data, only over a limited
kalotti area or cap: evaluate the equation
x
R
Nred = S(ψ) ∆gred (ϕ ′ , λ ′ ) dσ ′ , (8.19)
4πγ σ0
in which σ0 is a cap on the unit sphere the radius of which is, say, ψ0 .
The (possibly dangerous) assumption behind this is that, outside the
cap, ∆gred is both small and rapidly varying, because the longer wave-
lengths have been removed from it with the global-model reduction.
Write both parts of the integrand in equation 8.19 into spectral form:
∑︂
∞
2n + 1
S(ψ) = Pn (cos ψ)
n−1
n=2
and ∑︂
∞
′ ′
∆gred (ϕ , λ ) = ∆gn (ϕ ′ , λ ′ ),
n=L+1
and also
í Õ ! ¤. û
Kernel modification 8.9
221
it follows from the orthogonality of the Y functions that
x
Pn (cos ψ) ∆gn ′ (ϕ ′ , λ ′ ) dσ ′ = 0 if n ̸= n ′ .
σ
in which
∑︂
∞
2n + 1
L
S (ψ) = P (cos ψ)
n−1 n
n=L+1
í Õ ! ¤. û
222 8 The Stokes equation and other integral equations
25
S(ψ)
20 S2 (ψ)
S3 (ψ)
15 S4 (ψ)
S5 (ψ)
10 S6 (ψ)
S2−5 (ψ)
5
S(ψ)
S4 (ψ) S2 (ψ) S(ψ)
0
S(ψ)
S6 (ψ) Angular distance ψ (rad) −→
−5
0 0.5 1 1.5 2 2.5 3 3.5
Figure 8.10. Modified Stokes kernel functions. Note how the kernel values
for higher modification degree numbers L approach zero outside
the local area. The red curve has been ”soft modified” over a
^ modification degree range of 2–5 using a cosine taper.
4 Josiah
Willard Gibbs (1839–1903) was an American physicist, chemist, thermody-
namicist, mathematician and engineer.
í Õ ! ¤. û
Advanced kernel modifications 8.10
223
^ 8.10 Advanced kernel modifications
Other kernel modification methods are found in the literature. Their
general form is
∑︂
∞
2n + 1 ∑︂
L
2n + 1
L
S (ψ) = Pn (cos ψ) + (1 − sn ) P (cos ψ) =
n−1 n−1 n
n=L+1 n=2
∑︂
L
2n + 1
= S(ψ) − sn P (cos ψ), (8.20)
n−1 n
n=2
∑︂
L
2n + 1
L
S (ψ) = S(ψ) − sn P (cos ψ)
n−1 n
n=2
over the area outside a local cap, σ − σ0 . Let us multiply this expression
with each of the Legendre polynomials Pn (cos ψ), n = 2, . . . , L in turn,
integrate over the area σ − σ0 outside the local cap, and require the
result to vanish:\{
w
S(ψ) Pn (cos ψ) dσ −
σ−σ0
∑︂
L
2n ′ + 1
w
− sn ′ ′ Pn ′ (cos ψ) Pn (cos ψ) dσ = 0,
n −1 σ−σ0
n ′ =2
n = 2, . . . , L,
5 The
choice sn = 1 again gives the simply (Wong–Gore) modified Stokes kernel from
which the low degrees have been completely removed.
í Õ ! ¤. û
224 8 The Stokes equation and other integral equations
∑︂
L
2n ′ + 1
e ′ s ′ = Qn ,
′
n ′ − 1 nn n
n =2
in which
w wπ
1
Qn = S(ψ) Pn (cos ψ) dσ = S(ψ) Pn (cos ψ) sin ψ dψ
2π σ−σ0 ψ0
and
w
1
enn ′ = Pn (cos ψ) Pn ′ (cos ψ) dσ =
2π σ−σ0
wπ
= Pn (cos ψ) Pn ′ (cos ψ) sin ψ dψ.
ψ0
í Õ ! ¤. û
Block integration 8.11
225
cap radius ψ0 and modification degree L, it may be so smooth that
it does not contain any significant contribution from degree numbers
higher than L. If this applies for S, it will also apply for SL . This means
that SL will be a linear combination of the Legendre polynomials: an
element of the vector space spanned by the polynomials Pn , n = 2, . . . ,
L. But if this is so, and the scalar products 8.21 with each of the basis kantavektori
vectors vanish, then SL must be the zero function on σ − σ0 .
See also Featherstone (2003).
Appendix A section A.1 explains more about linear vector spaces and
the scalar product of vectors.
R ∑︂
N(ϕ, λ) ≈ Si (ϕ, λ) ∆gi , (8.22)
4πγ
i
in which σi is the area of block i and its size on the unit sphere is
x x
def
ωi = dσ = cos ϕ dϕ dλ.
σi σi
í Õ ! ¤. û
226 8 The Stokes equation and other integral equations
4 j=
1 36 1
36 36
1
16
4 36 4
36 36 0
1 1 -1
36 4 36
36
k = -1 0 1
w λi + ∆λ/2 w ϕi + ∆ϕ/2 (︁
′ ′
cos ϕ ′ dϕ ′ dλ ′ ≈
)︁
Si (ϕ, λ) = S ψ(ϕ, λ; ϕ , λ )
λi − ∆λ/2 ϕi − ∆ϕ/2
∑︂ 1 ∑︂
1
≈ ∆ϕ ∆λ wj wk Sjk
i ,
j=−1 k=−1
in which ∆λ and ∆φ are the block sizes in the latitude and longitude
directions, and w−1 = w1 = 16 , w0 = 46 are the weights.
(︂ (︁ )︁)︂
Sjk 1 1
cos ϕi + 21 j ∆ϕ ,
(︁ )︁
i (ϕ, λ) = S ψ ϕ, λ; ϕ i + 2
j ∆ϕ, λ i + 2
k ∆λ
j, k = −1, 0, 1
used in the evaluation, 3×3 of them. See figure 8.11. More complicated
formulas (repeated Simpson or Romberg) can also be employed.
6 Thomas Simpson FRS (1710–1761) was an English mathematician and textbook writer.
Actually Simpson’s rule was already being used a century earlier by Johannes Kepler.
í Õ ! ¤. û
Effect of the local zone 8.12
227
^ 8.12 Effect of the local zone
One can show that the effect of the local (inner) zone on the geoid height
at the evaluation point (ϕ, λ) is proportional to the gravity anomaly
in the point itself, ∆g(ϕ, λ). Starting from Stokes equation 8.2 with
S(ψ) ≈ 1 sin 12 ψ ≈ 2 ψ , we find, for a circular inner zone of radius ψ0 :
/︁ /︁
w 2π w ψ0
R 2 dψ dα ≈
δN0 = sin
∆g(ψ, α) ψ
4πγ 0 0 ψ
w ψ0 (︃ w 2π )︃
R 1 R s
≈γ ∆g(ψ, α) dα dψ ≈ γ · ψ0 · ∆g0 = γ0 ∆g0 .
0 2π 0
Here s0 = Rψ0 is the radius of the local block or cap in units of length.
The quantity
w ψ0 (︃ w 2π )︃
def 1 1
∆g0 = ∆g(s, α) dα dψ =
ψ0 0 2π 0
w (︃ w )︃
1 s0 1 2π
=s ∆g(s, α) dα ds
0 0 2π 0
2 d 2
S(ψ) ≈ =⇒ S(ψ) = − 2 :
ψ dψ ψ
{︄ }︄ {︄ }︄
δξ0 w ψ0 w 2π cos α
1 2
(︂ )︂
≈ − 2 ∆g(ϕ ′ , λ ′ ) sin ψ dα dψ.
δη0 4πγ 0 0 ψ sin α
∂∆g ∂∆g
∆g ≈ ∆g0 + x +y ≈
∂x ∂y
í Õ ! ¤. û
228 8 The Stokes equation and other integral equations
(︃ )︃
∂∆g ∂∆g
≈ ∆g0 + R ψ cos α + sin α ,
∂x ∂y
and substitute:
{︄ }︄
δξ0 1
≈ ·
δη0 4πγ
{︄ }︄
w ψ0 w 2π (︃ )︂ )︃ cos α
2 ∂∆g ∂∆g
(︂
· − 2 ∆g0 + Rψ cos α + sin α sin ψ dα dψ.
0 0 ψ ∂x ∂y sin α
r 2π
Here, the terms in ∆g0 drop out in α integration, because 0 sin α dα =
r 2π
0 cos α dα =r0. So do the mixed
r 2π
terms in sin α cos α. The only nonzero
2π 2
terms contain 0 sin α dα = 0 cos2 α dα = π :
w ψ0 w 2π
1 2 ∂∆g
δξ0 ≈ − Rψ cos α cos α sin
ψ dα dψ ≈
4πγ 0 0
ψ 2 ∂x
w ψ0 w 2π
R ∂∆g Rψ ∂∆g
≈− cos2 α dα dψ ≈ − 0 ,
2πγ 0 0 ∂x 2γ ∂x
w ψ0 w 2π
1 2 ∂∆g
δη0 ≈ − R sin α
ψ sin α
sin
ψ dα dψ ≈
4πγ 0 0 ψ 2 ∂y
w ψ0 w 2π
R ∂∆g Rψ ∂∆g
≈− sin2 α dα dψ ≈ − 0 .
2πγ 0 0 ∂y 2γ ∂y
Evaluating these integrals assumes the partial derivatives to be constant
within the cap. Using Rψ0 = s0 yields now
s0 ∂∆g s0 ∂∆g
δξ0 ≈ − , δη0 ≈ − .
2γ ∂x 2γ ∂y
These equations might be useful as standard block integration, equation
8.22, behaves numerically poorly in the immediate surroundings of the
evaluation point if the kernel function is singular at the origin ψ = 0.
Both the Stokes 8.2 and Vening Meinesz 8.5 kernels are of this kind.
^ Self-test questions
1. What do the Stokes equation and its spectral form look like?
í Õ ! ¤. û
Exercise 8 – 1: The Stokes equation in the near zone
229
(︁ )︁
2. What does the Stokes kernel function S ψ look like when ex-
panded in Legendre polynomials?
3. What is a suitable approximation of the Stokes kernel when ψ is
small?
4. What is an isotropic, what an anisotropic quantity on the Earth’s
surface? Give an example of the latter.
5. What does the Poisson integral equation describe?
6. Why are gravity reductions necessary when using the Stokes
equation for computing a geoid model?
7. Which gravity reduction methods are available?
8. Explain the residual terrain modelling (RTM) method.
9. Explain the remove–restore approach.
10. Why, in geoid determination, is the Stokes kernel function often
modified? What does such a modification look like?
11. What is the Gibbs phenomenon?
í Õ ! ¤. û
230 8 The Stokes equation and other integral equations
Hint: you need to consider Jacobi’s determinant for the polar co-
ordinates (s, α).
3. Compute N (as an equation) if ∆g = ∆g0 only within a circular
area s ⩽ s0 , and outside it ∆g = 0. Assume that s0 is small.
í Õ ! ¤. û
^ Spectral techniques, FFT
9
^ 9.1 The Stokes equation as a convolution
We start from the Stokes equation 8.1,
x
R
S(ψ) ∆g ϕ ′ , λ ′ dσ ′ ,
(︁ )︁
T (ϕ, λ) =
4π σ
in which (ϕ ′ , λ ′ ) is the location of the moving integration or observation
point, and (ϕ, λ) is the location of the evaluation point, both at sea level,
on the surface of a spherical Earth. So, the locations of both points are
given in spherical co-ordinates. The integration is carried out over the
surface of the unit sphere σ: a surface element is dσ = cos ϕ dϕ dλ, in
which cos ϕ is the determinant of Jacobi, for the spherical co-ordinates
(ϕ, λ).
However locally, in a sufficiently small area, one may also write
the point co-ordinates in rectangular form and express the integral
in rectangular co-ordinates. Suitable rectangular co-ordinates are, for
example, map projection co-ordinates, see figure 9.1.
A simple example of rectangular co-ordinates in the tangent plane
would be
– 231 –
232 9 Spectral techniques, FFT
x
Data point
α
y
Evaluation point
R ψ
Earth’s centre
tangent plane touches the sphere. The locations of other points are
measured by the angle ψ at the Earth’s centre, the geocentric angular
distance, and by the direction angle in the tangent plane or azimuth α.
A more realistic example uses a popular conformal map projection
called the stereographic projection:
ψ ψ
x = 2 tan R cos α, y = 2 tan R sin α.
2 2
In the limit for small values of ψ this agrees with equations 9.1.
Taking the squares of equations 9.1, summing them, and dividing
the result by R2 yields
x2 + y2
ψ2 ≈ .
R2
More generally ψ is the angular distance between the points (x, y)
(evaluation point) and (x ′ , y ′ ) (data, integration or moving point) seen
from the Earth’s centre, approximately
(︃ )︃2 (︃ )︃2
2 x − x′ y − y′
ψ ≈ + .
R R
í Õ ! ¤. û
The Stokes equation as a convolution 9.1
233
Furthermore, we must account for Jacobi’s determinant R2 of the
projection:
dσ = R−2 dx dy ⇐⇒ dx dy = R2 dσ,
and the Stokes equation now becomes
x ∞ (︁
1
S x − x ′ , y − y ′ ∆g x ′ , y ′ dx ′ dy ′ ,
)︁ (︁ )︁
T (x, y) ≈ (9.2)
4πR −∞
a two-dimensional convolution.1 1
í Õ ! ¤. û
234 9 Spectral techniques, FFT
ing potential Tloc , and the corresponding geoid undulation Nloc , have
been obtained, we must remember to add to them again the effect of
the global spherical-harmonic expansion on the disturbing potential T
entistämisvaihe and geoid undulation N to be calculated separately. This is the “restore”
kommutoiva step of the computation; see the commutative diagram 8.9.
kaavio
(︁ )︁
∆gij = ∆g xi , yj ,
xi = i δx, yj = j δy, i, j = 0, 1, . . . , N − 1,
for suitably chosen grid spacings (δx, δy) . The integer N is the grid size,
assumed for simplicity to be the same in both directions.
Next, we do the same for the kernel function
2 Alternatively,
one could for example calculate for every grid point the average over a
square cell surrounding the point.
í Õ ! ¤. û
Integration by FFT 9.2
235
so we write
(︁ )︁
Sij = S ∆xi , ∆yj ,
where again
landing the origin in the centre of the grid. This is the correct way to
compute the values of the true, non-periodic kernel, with both positive
and negative values ∆x and ∆y from an area symmetrically surrounding
the origin.
Next:
1. The grid representations ∆gij and Sij thus obtained of the func-
tions ∆g and S are transformed to the frequency domain — they
{︁ }︁ {︁ }︁
become functions Suv = F Sij and Guv = F ∆gij of the two
“frequencies”, the wave indices u and v in the x and y direc-
tions. The spatial frequencies or wave numbers3 ν ˜︁ and spatial 3
u L, ν v L , in which
/︁ /︁
wavelengths λ are ν˜︁x = λ−1
x = ˜︁y = λ−1
y =
L = Nδx = Nδy is the size of the area, assumed to be square.
3 This is the so-called linear wave number, whole waves per unit of length. The angular
wave number is k = 2π˜︁
ν, radians of phase angle per unit of length.
í Õ ! ¤. û
236 9 Spectral techniques, FFT
í Õ ! ¤. û
Solution in latitude and longitude 9.3
237
Observation
Interpolation Regular
points in their −−−−−−−−−−−−−−→
point grid
own places
⏐ ⏐
⏐ ⏐
↓Direct solution FFT↓
grid by interpolating from the given (φ, λ) one through a map projection
calculation is avoided. However, working in geographical co-ordinates
causes errors due to meridian convergence — as a latitude and longitude
co-ordinate system is not actually rectangular. The co-ordinate pair
(φ, λ cos φ) would be slightly more suitable.
The problem has also been addressed on a more conceptual level.
Substitute
λ − λ′
cos λ − λ ′ = 1 − 2 sin2
(︁ )︁
,
2
ψ
cos ψ = 1 − 2 sin2 ,
2
ϕ − ϕ′
cos ϕ − ϕ ′ = 1 − 2 sin2
(︁ )︁
,
2
í Õ ! ¤. û
238 9 Spectral techniques, FFT
λ − λ′
cos ψ = cos ϕ − ϕ ′ − 2 cos ϕ cos ϕ ′ sin2
(︁ )︁
2
′
ψ ϕ − ϕ λ − λ′
=⇒ sin2 = sin2 + cos ϕ cos ϕ ′ sin2 .
2 2 2
Here we may use the following approximation:
ψ ϕ − ϕ′ λ − λ′
sin2 ≈ sin2 + cos2 ϕ0 sin2 , (9.4)
2 2 2
def def
which depends only on the differences ∆ϕ = ϕ − ϕ’ and ∆λ = λ − λ ′ .
After this, the FFT method may be applied by using co-ordinates
4 (ϕ, λ)4 and the Stokes kernel written as
(︄ √︃ )︄
∆ϕ ∆λ
S(ψ) = S(∆ϕ, ∆λ) = S 2 arcsin sin2 + cos2 ϕ0 sin2 ,
2 2
which is now a function of only the differences ∆ϕ and ∆λ, as the con-
volution theorem requires. This clever way of using FFT in geographical
5 co-ordinates was invented by the Dutchman G. Strang van Hees5 in
1990.
í Õ ! ¤. û
Solution in latitude and longitude 9.3
239
Write the Stokes equation as follows:
x (︁
R )︁[︂ (︁ ]︂
S ∆ϕ, ∆λ; ϕ ∆g ϕ ′ , λ ′ cos ϕ ′ dϕ ′ dλ ′ ,
)︁
N(ϕ, λ) = (9.5)
4πγ
where we have expressed S as a function of latitude difference, longitude
difference and evaluation latitude. Now, choose two support latitudes,
ϕi and ϕi+1 . Assume furthermore that between these S is a sufficiently
linear function of ϕ. In that case we may write
(ϕ − ϕi ) Si+1 (∆ϕ, ∆λ) + (ϕi+1 − ϕ) Si (∆ϕ, ∆λ)
S(∆ϕ, ∆λ; ϕ) = ,
ϕi+1 − ϕi
where ∆ϕ = ϕ − ϕ ′ , ∆λ = λ − λ ′ , and
Si (∆ϕ, ∆λ) = S ϕ − ϕ ′ , λ − λ ′ ; ϕi ,
(︁ )︁
ψ ϕ − ϕ′ λ − λ′
sin2 = sin2 + cos ϕ cos ϕ ′ sin2 =
2 2 2
∆ϕ ∆λ
= sin2 + cos ϕ cos (ϕ − ∆ϕ) sin2 .
2 2
We calculate Si and Si+1 for the support latitude values ϕi and ϕi+1 ,
we evaluate the integrals with the aid of the convolution theorem, and
í Õ ! ¤. û
240 9 Spectral techniques, FFT
6 Inthe literature the method has been generalised by also expanding the kernel with
respect to height.
í Õ ! ¤. û
Solution in latitude and longitude 9.3
241
Linearise:
This expansion into two terms will be accurate only for a limited range
in ∆φ, and the kernel function C is assumed to be of bounded support.
In this case, the integrals may be calculated within a limited area instead
of over the whole Earth.
Substitution yields
x
C(∆ϕ, ∆λ; ϕ) · m ϕ ′ , λ ′ cos ϕ ′ dϕ ′ dϕ ′ =
(︁ )︁
ℓ(ϕ, λ) =
x (︁
C0 + (ϕ − ϕ0 ) Cϕ · m cos ϕ ′ dϕ ′ dϕ ′ =
)︁
=
x x
= C0 · m cos ϕ ′ dϕ ′ dλ ′ + (ϕ − ϕ0 ) Cϕ · m cos ϕ ′ dϕ ′ dλ ′ .
(9.7)
It is important here now that the integrals in the first and second terms,
x [︂ (︁ ]︂
C0 (∆ϕ, ∆λ) m ϕ ′ , λ ′ cos ϕ ′ dϕ ′ dλ ′ = C0 ⊗ m cos ϕ ,
)︁ [︁ ]︁
x [︂ (︁ ]︂
Cϕ (∆ϕ, ∆λ) m ϕ , λ cos ϕ dϕ ′ dλ ′ = Cϕ ⊗ m cos ϕ ,
′ ′ ′
)︁ [︁ ]︁
í Õ ! ¤. û
242 9 Spectral techniques, FFT
Let us choose, say, a 256 × 256 size grid (so N = 256) and fill in
the missing values by extrapolation.
The values of the kernel functions C0 and Cϕ are calculated on
a 256 × 256 grid (∆ϕ, ∆λ) as well. The number of these is thus
[︁ ]︁
also 65 536. Calculating the convolutions C0 ⊗ m cos ϕ and
7
[︁ ]︁
Cϕ ⊗ m cos ϕ by means of FFT — like7
x
C0 (∆ϕ, ∆λ) m ϕ ′ , λ ′ cos ϕ ′ dϕ ′ dλ ′ =
(︁ )︁
{︂ {︁ }︁ {︁ }︁}︂
= C0 ⊗ m cos ϕ = F −1
F C0 · F m cos ϕ ,
[︁ ]︁
x
Cϕ (∆ϕ, ∆λ) m ϕ , λ cos ϕ ′ dϕ ′ dλ ′ =
(︁ ′ ′ )︁
{︂ {︁ }︁ {︁ }︁}︂
= Cϕ ⊗ m cos ϕ = F−1 F Cϕ · F m cos ϕ ,
[︁ ]︁
(︁ )︁
requires N2 × 2 log N2 = 65 536×16 = more than a million “stan-
8 dard operations”,8 multiplication with the coefficients (ϕ − ϕ0 )
and adding together, again 65 536 standard operations.
The grid matrices corresponding to kernel functions C0 and Cϕ
are obtained as follows: for three reference latitudes ϕ−1 , ϕ0 , ϕ+1
we compute numerically the grids
(︁ )︁
C−1 = C ∆ϕ, ∆λ; ϕ−1 ,
(︁ )︁
C0 = C ∆ϕ, ∆λ; ϕ0 ,
(︁ )︁
C+1 = C ∆ϕ, ∆λ; ϕ+1 ,
in which C0 is directly available, and
C+1 − C−1
Cϕ ≈ .
ϕ+1 − ϕ−1
í Õ ! ¤. û
Bordering and tapering of the data area 9.4
{︁ }︁ }︃
243
{︃
{︁ }︁ {︁ }︁ {︁ }︁ ]︁(0) F ℓ
F C0 · F m cos ϕ = F ℓ = F−1 {︁ }︁ .
[︁
=⇒ m cos ϕ
F C0
The second approximation is obtained by first calculating
]︁(0) ]︁(0)
ℓ(0) = C0 ⊗ m cos ϕ
[︁ [︁
+ (ϕ − ϕ0 ) · Cϕ ⊗ m cos ϕ ,
on the Earth’s surface can be described exactly in geodetic co-ordinates, the method
may be used with geodetic latitude φ instead of geocentric latitude ϕ. Thus, errors
caused by ignoring the Earth’s flattening are avoided.
í Õ ! ¤. û
244 9 Spectral techniques, FFT
eastern edge of the data area to the western edge, and the northern
edge to the southern edge, the data has to be continuous across these
11 edges.11 In practice, this is not the case. We are faced with two different
requirements:
◦ The data on the other side of an edge must be so far away as to
have no noticeable influence across the edge on the result of the
calculation.
◦ The data must be continuous across the edges.
Therefore, always when using FFT with the convolution theorem, two
measures need to be taken.
1. We continue the data by adding a border area to the data area,
so-called bordering. Often, the width of the border area is 25 % of
the size of the data area, making the surface area of the whole
calculation area four times that of the data area itself. The border
is filled with measured values where those exist, otherwise with
predicted (inter- or extrapolated) values.
The calculation area for the kernel function is also made similarly
four times larger. In this case the whole grid including the border
area is filled with real (computed) values.
The grid of the kernel function must be filled in such a way,
that index values i, j > N 2 are interpreted as negative values
/︁
11 Topologically the area with the edges thus connected is equivalent to a torus, and
the data is presupposed to be continuous on the surface of the torus.
í Õ ! ¤. û
Bordering and tapering of the data area 9.4
245
25 % 50 % 25 %
Data area
1
If the values at the edges are not zero, they may be forced to
zero by multiplying the whole calculation area by a so-called
tapering function, which goes smoothly to zero towards the edges.
Such a function can easily be built: examples are a cubic spline
polynomial or a Tukey or cosine taper. See figure 9.3, showing
a 25 % tapering function, as well as example images 9.4, which
show how non-continuity — sharply differing left and right, and
upper and lower, edges — causes horizontal and vertical artefacts
in the Fourier transform. These artefacts are related to the Gibbs Gibbsin ilmiö
phenomenon, already mentioned in section 8.9: a sharp cut-off or
edge in the space domain will generate signal on all frequencies,
up to the highest ones.
Many journal articles have appeared on these technicalities. Groups that
were already involved in early development of FFT geoid determination
in the 1980s include Forsberg’s group in Copenhagen, The group
of Klaus-Peter Schwarz and Michael Sideris in Calgary, Canada, the
Delft group (Strang van Hees, Haagmans, De Min, Van Gelderen),
the Milanese group (Sansò, Barzaghi, Brovelli), Heiner Denker at the
Hannover “Institut für Erdmessung”, and many others.
í Õ ! ¤. û
246 9 Spectral techniques, FFT
Figure 9.4. Example images for the FFT transform without (above) and with
(below) tapering. The online FFT service from Watts (2004) was
used. The images are greytone amplitude spectra |Fuv | plotted
^ with the origin u = v = 0 in the centre, see appendix C.
í Õ ! ¤. û
Computing a geoid model with FFT 9.5
247
^ 9.5.1 GRAVSOFT software
The GRAVSOFT geoid determination software has been mainly produced
in Denmark. Authors include Carl Christian Tscherning,12 René Fors- 12
berg, Per Knudsen, the Norwegian Dag Solheim, and the Greek Dimitris
Arabelos. The manual for the software is Forsberg and Tscherning
(2008).
This package is in widespread use and also provides, in addition
to variants of FFT geoid determination, for example least-squares col- pienimmän
location, as well as routines for evaluating various terrain effects. Its neliösumman
kollokaatio
popularity can be partly explained by it being free for scientific use, and
being distributed as source code. It is also well-documented. There-
fore it has also found commercial use, for example in the petroleum
extraction industry.
GRAVSOFT has also been used a great deal for teaching, for example at
many research schools organised by the IAG (International Association
of Geodesy) in various countries. ISG, Geoid Schools.
í Õ ! ¤. û
248 9 Spectral techniques, FFT
20˚ ◦
20 24◦
24˚ 28◦
28˚ 32˚ ◦
32
20
◦
0 ◦˚
770 7700˚
19
8 ◦˚
668 6688˚◦
24 25
22
21
23
31 28
29
30
20
26
27
6 ◦˚
666 19 6666˚◦
17
18
18
664
4 ◦˚ 6644˚◦
18
5
23 24 2
20 212
2
19
2 ◦˚
662 6622˚◦
18
17
19 16
660
0 ◦˚ 6600˚◦
15
16
20˚ ◦
20
28◦ 32˚◦
32
24◦
24˚ 28˚
^ Figure 9.5. The Finnish FIN2000 geoid. Data © Finnish Geodetic Institute.
í Õ ! ¤. û
Use of FFT in other contexts 9.6
249
^ 9.6 Use of FFT computation in other contexts
í Õ ! ¤. û
250 9 Spectral techniques, FFT
í Õ ! ¤. û
Computing terrain corrections with FFT 9.7
251
From equation 9.8 we obtain by expansion into terms:
x +∞ x +∞ ′
1 ′ ′ H
T C(x, y) = 12 GρH2 dx dy − GρH dx ′ dy ′ +
−∞ ℓ3 −∞ ℓ3
x +∞ (H ′ )2
+ 12 Gρ dx ′ dy ′ , (9.9)
−∞ ℓ3
in which every integral is a convolution with kernel ℓ−3 , and the functions
to be integrated are 1, H ′ , and (H ′ )2 .
Unfortunately the function ℓ−3 as implicitly defined above has no
Fourier transform. Therefore, we change the above definition a tiny bit
by adding a small term:
ℓ2 = (x − x ′ ) 2 + (y − y ′ ) 2 + δ2 . (9.10)
The terms in the above equation 9.9 are large numbers that almost
cancel each other, giving a nearly correct result. Numerically this is an
unpleasant situation. There is a solution for this which we present next.
If ℓ is defined according to equation 9.10, then the Fourier transform
of kernel ℓ−3 is (Harrison and Dickinson, 1989; Forsberg, 1984):
{︁ −3 }︁ 2π
(︃ )︃
2π 4π2 δ2 q2
F ℓ = exp(−2πδq) = 1 − 2πδq + − ··· ,
δ δ 1·2
def
√︂ √ /︂
2 2 u 2 + v2
in which q = ν ˜︁x + ν
˜︁y = L , u and v are wave indices, and
˜︁x = u L and ν ˜︁y = v L are (linear) “spatial frequencies” or wave
/︁ /︁
ν
numbers in the x and y directions in the (x, y) plane. If we substitute
this into equation 9.9, we notice that the terms containing 1 δ sum to
/︁
í Õ ! ¤. û
252 9 Spectral techniques, FFT
í Õ ! ¤. û
Self-test questions
253
^ Self-test questions
1. What is the definition of a convolution?
2. Explain the convolution theorem.
3. Check that the dimensions of the quantities on both sides of
equation 9.2 match.
4. What is spatial frequency? What is the difference between linear
and angular spatial frequency?
5. Explain the basic idea of the Strang van Hees method.
6. What other approaches are there to applying the FFT method on a
curved (spherical or ellipsoidal) surface?
7. Why are bordering of the data area and tapering of the calculation
area necessary?
8. In addition to geoid determination, where in physical geodesy is
the FFT method also used?
9. When computing the terrain correction on the Earth’s surface,
explain the “δ trick” used in the derivation. Why is it necessary,
and how does one make the δ vanish again?
í Õ ! ¤. û
^ Statistical methods
10
^ 10.1 The role of uncertainty in geophysics
In geophysics, we often obtain results based on uncertain, incomplete,
or otherwise deficient observational data. This also applies in the
study of the Earth’s gravity field: the density of gravity observations on
the Earth’s surface, for example, varies greatly, and large areas of the
oceans and polar regions are covered only by a very sparse network of
measurements. We speak of spatial undersampling.
Measurement technologies that work from space, on the other hand,
usually provide coverage of the whole globe, oceans, poles and all.
They, however, do not always measure at a very high resolution. Either
the resolution of the method is limited — this holds for example for the
gravity-field parameters calculated from satellite orbit perturbations ratahäiriöt
— or the instruments measure only directly underneath the satellite’s
path, like satellite altimetry.
Another often relevant uncertainty factor is that one can do precise
measurements on the Earth’s surface, but inside the Earth the uncer-
tainty is much larger and the data is obtained much more indirectly.
In previous chapters we described techniques by which we could
calculate desired values or parameters for the Earth’s gravity field,
assuming that, for example, gravity anomalies are available everywhere
on the Earth’s surface, and with arbitrarily high resolution. In this
– 255 –
256 10 Statistical methods
d
⃓
f ↦→ f(x)⃓ .
⃓
dx x=x0
and so on.
We may write symbolically
d⃓
⃓ {︁ }︁ d
⃓
L= , meaning L f = f(x)⃓ .
⃓
dx x=x0 dx
⃓
x=x0
Remember that all partial derivatives, as also the Laplace operator ∆, are
linear.
In physical geodesy, all interesting functionals are functionals of the
häiriö- function T (ϕ, λ, R) = T (ϕ, λ, r)|r=R , that is, of the disturbing potential
potentiaali at the surface of a spherical Earth. The theory thus uses the spherical
1 approximation,1 and the surface of the sphere of radius R corresponds
í Õ ! ¤. û
Statistics on the Earth’s surface 10.3
257
def
to mean sea level. For example, the disturbing potential TP = T (ϕ, λ, R)
at a point P at sea-level location (ϕ, λ) is such a functional:
If the quantity is not the disturbing potential, but, say, the gravity
anomaly or the deflection of the plumb line: luotiviivan
poikkeama
T (·, ·, R) ↦→ ξ(ϕ, λ, r),
T (·, ·, R) ↦→ ∆g(ϕ, λ, r),
T (·, ·, R) ↦→ η(ϕ, λ, r).
even the spherical-harmonic coefficients anm , bnm are all linear func- pallofunktio-
tionals of the disturbing potential T : kerroin
ϕ ∈ − π 2 ,+ π 2 ,
[︁ /︁ /︁ ]︁ [︁ )︁
T (ϕ, λ, R), λ ∈ 0, 2π .
í Õ ! ¤. û
258 10 Statistical methods
í Õ ! ¤. û
Statistics on the Earth’s surface 10.3
259
realisations x1 , x2 , x3 , . . . are obtained, which together have certain
statistical properties.
The classical example is the dice throw. A die can be thrown again
and again, and one can practise the art of statistics on the results of the
throws. Another classic example is measurement. Measurement of the
same quantity can be repeated, and is repeated, in order to improve
precision.
For a stochastic process defined on the Earth’s surface, the situation
is different.
{︁ }︁ def 1 x w w
1 2π + π/2
M x = x(ϕ, λ) dσ = x(ϕ, λ) cos ϕ dϕ dλ.
4π σ 4π 0 − π/2
(10.1)
Here, x(ϕ, λ) is the one and only realisation of process x that is available
on this Earth.
Clearly this definition makes sense only in the case where the statis-
tical behaviour of the process x(ϕ, λ) is the same everywhere on Earth,
independently of location (ϕ, λ). This is called the assumption of homo-
geneity. It is in fact the assumption that the spherical symmetry of the
Earth extends to the statistical behaviour of field x.
Similarly to the statistical variance based on expectancy, we may
define the geographic variance:
{︃(︂ }︃
{︁ }︁ def {︁ }︁)︂2
Cxx (ϕ, λ) = Var x(ϕ, λ) = M x − M x . (10.2)
í Õ ! ¤. û
260 10 Statistical methods
their definition:
{︁ }︁
M ∆g = 0.
In that case, equation 10.2 is simplified as follows:
{︁ }︁ {︁ }︁ x (︁
1 )︁2
C∆g∆g (ϕ, λ) = Var ∆g(ϕ, λ) = M ∆g2 = ∆g(ϕ, λ) dσ.
4π σ
{︁ }︁
The definition given here of the geographic mean M · is based on
integration of the one and only realisation over the surface of the Earth.
As has been seen, in statistics the mean is defined slightly differently, as
{︁ }︁
the expectancy of a stochastic process. For gravity anomalies it is E ∆g ,
in which ∆g is the anomaly considered as a stochastic process, the series
of values of ∆g that results if we look at an endless series of randomly
formed Earths. Not very practical!
If the expectancy of a stochastic process is the same as the mean of one
realisation computed by integration — and other statistical properties
are similarly the same — we speak of an ergodic process. Establishing
empirically in geophysics that a process is ergodic is typically difficult
to impossible.
2 Thisis not exactly valid if, for example, the normal gravity field used in calculating
the anomalies contains the mass of the atmosphere, but gravity values measured close
to sea level do not contain the attraction of the atmosphere.
í Õ ! ¤. û
The covariance function of the gravity field 10.4
261
α
P
Q
gravity field, the covariance function will not depend on the absolute
location of the points, but only on the difference in location between
points P and Q.
Write
(︁ )︁ (︁ )︁
ϕQ = ϕQ ϕP , λP , ψPQ , αPQ , λQ = λQ ϕP , λP , ψPQ , αPQ .
í Õ ! ¤. û
262 10 Statistical methods
í Õ ! ¤. û
Least-squares collocation 10.5
263
Remark The true gravity field of the Earth is not terribly homogeneous
or isotropic, but in spite of this, both hypotheses are widely used.
í Õ ! ¤. û
264 10 Statistical methods
matrix and all of its elements can be calculated, provided all argument
values (observation times) ti are also known.
Let the shape of the problem now be that one should estimate, or
predict, the value of process s at the moment in time T , that is s(T ), based
on our knowledge of the above-described observations s(ti ), i = 1, . . . ,
N.
In the same way as we calculated above the covariances between
s(ti ) and s(tj ) — elements of the signal variance matrix Cij — we also
compute the covariances between s(T ) and all s(ti ), i = 1, . . . , N. We
obtain
{︁ }︁ [︂ (︁ )︁ (︁ )︁ (︁ )︁ ]︂
Cov s(T ), s(ti ) = C T, t1 C T, t2 · · · C T, tN .
For this we may use the notation CT j . It is assumed here that there is
only one point in time T for which estimation is done. Generalisation to
the case where there are several Tp , p = 1, . . . , M, is straightforward.
In that case, the signal covariance matrix will be of size M × N:
⎡ (︁ )︁ (︁ )︁ (︁ )︁ ⎤
C T1 , t 1 C T1 , t 2 · · · C T1 , t N
{︂ (︁ )︁ }︂ ⎢
(︁ )︁ (︁ )︁ (︁ )︁ ⎥
⎢ C T2 , t 1 C T2 , t 2 · · · C T2 , t N ⎥
Cov s Tp , s(ti ) = ⎢ .. .. .. ⎥.
. . .
⎢ ⎥
⎣ ⎦
(︁ )︁ (︁ )︁ (︁ )︁
C TM , t 1 C TM , t 2 · · · C TM , t N
For this we may use the more general notation Cpj .
í Õ ! ¤. û
Least-squares collocation 10.5
265
Here, ni is a stochastic quantity: observational error or noise. Let its
variance — or more precisely, the joint noise variance matrix of multiple
observations — be Dij . This is a similar matrix to the above Cij , and
also symmetric and positive definite. The only difference is that Dij
designates noise, which we are not interested in. Often, it may be
assumed that the errors ni and nj of two different observations ℓi and
ℓj do not correlate, in which case Dij is a diagonal matrix.
Here, for the sake of writing convenience, we left the summation sign
∑︁
off (Einstein summation convention): We always sum over adjacent,
identical indices, in this case i.
Study the variance of this difference, the so-called variance of prediction: ennustus-
{︂ (︁ )︁ (︁ )︁}︂ varianssi
def
Σpp = Var ˆ︁ s Tp − s Tp .
and5 5
í Õ ! ¤. û
266 10 Statistical methods
{︃(︂ }︃
(︁ )︁ (︁ )︁)︂ (︂ (︁ )︁ (︁ )︁ )︂
Σpq = Cov ˆ︁ s Tp − s Tp , ˆ︁ s Tq − s Tq =
{︂(︁ )︁ (︁ )︁}︂ {︂ (︁ )︁ (︁ )︁}︂
= Λpi Cov s(ti ) + ni , s(tj ) + nj Λjq + Cov s Tp , s Tq −
{︂ (︁ )︁}︂ {︂ (︁ )︁ }︂
− Λpi Cov s(ti ), s Tq − Cov s Tp , s(tj ) Λjq =
= Λpi (Cij + Dij ) Λjq + Cpq − Λpi Ciq − Cpj Λjq . (10.6)
The variances, or diagonal elements, Σpp of the matrix are now obtained
by setting q = p.
Then, from equation 10.6 and exploiting the symmetry of the C and D
matrices, we obtain
in which
5 The matrix Ciq is the transpose of Cpj , the matrix Λjq the transpose of Λpi .
í Õ ! ¤. û
Least-squares collocation 10.5
267
I = Λpi (Cij + Dij ) Λjp =
(︂ )︂ (︂ )︂
−1 −1
= Cpi (Cij + Dij ) + δΛpj (Cij + Dij ) (Cjk + Djk ) Ckp + δΛkp =
hhhh ˂
˂˂˂˂
−1
=˂
Cpi
˂˂ (C˂ijh
˂+h
˂Dij ) hh
˂h
h Ch
jp
h
+ Cpi δΛ
hhhh
hip
h˂+˂δΛ
˂˂pi Cip + δΛpi (Cij + Dij ) δΛjp ,
˂˂
(︂ )︂
II = − Λpj Cjp = − Cpi (Cij + Dij )−1 + δΛpj Cjp =
hhhh ˂
˂˂˂˂
−1
= −˂C˂ ˂(C ij )hhC
h+ Dh −˂δΛ pi Cip
h h˂
˂ ˂˂
˂˂
˂ pi
˂ij
hjp
˂ h
h˂
and
(︂ )︂
III = − Cpi Λip = − Cpi (Cij + Dij )−1 Cjp + δΛip =
= − Cpi (Cij + Dij )−1 Cjph
−hCh δΛ
pih h,
hip
Here, the last term — the only difference from result 10.7 — is positive,
′
because the matrices Cij and Dij are positive definite: Σpp > Σpp , except
when δΛpi = 0. In other words, the solution given above,
í Õ ! ¤. û
268 10 Statistical methods
í Õ ! ¤. û
Least-squares collocation 10.5
269
This equation was derived from gravimetric data for Ohio state, USA,
but it has broader validity. C(0) = C0 , the signal variance. The variable
d = R ψ0 is also called the correlation length. It is the distance d for
which C(d) = 12 C0 , as seen from the equation.
The quantity C0 varies considerably between areas, from hundreds to
thousands of mGal2 , and tends to be largest in mountainous areas. The
quantity d is generally of the order of magnitude of tens of kilometres.
Alternative functions that are also often used in local applications
include the covariance function of the stationary Gauss–Markov process,
and a quadratic variant:
(︃ (︂ )︂ )︃
ψ ψ 2
(︂ )︂
C(ψ) = C0 exp − , C(ψ) = C0 exp − .
ψ0 ψ0
7 The correlation is
{︂ }︂ C0
{︂ }︂ Cov ∆gP , ∆gQ
/︁ )︁2
1 + ψ ψ0
(︁
1
Corr ∆gP , ∆gQ = √︃ {︂ }︂ {︂ }︂ = √ = (︁ /︁ )︁2 ,
C0 C0 1+ ψ ψ0
Var ∆gP Var ∆gQ
í Õ ! ¤. û
270 10 Statistical methods
0.8
0.6
C(ψ)
0.4
0.2
4
2
−4 −2 0
0 2 −2 y
x 4 −4
20
16
12
∆g
8
30 40
y 20 30
20
10 10 x
Figure 10.3. An example of least-squares collocation. Here are given two data
points ∆g1 and ∆g2 (stars); the surface plotted gives the estimated
value ∆g
ˆ︂ P for each point P in the area. We use least-squares
^ collocation for inter- and extrapolating gravimetric data.
í Õ ! ¤. û
Least-squares collocation 10.5
271
x
1 (15 mGal)
30
P′
2 (20 mGal)
20
10 P
y
10 20 30
í Õ ! ¤. û
272 10 Statistical methods
Cij + Dij ≈ [︄ ]︄ [︄ ]︄
C11 C12 1000 666.66
≈ Cij = = mGal2 ,
C21 C22 666.66 1000
í Õ ! ¤. û
Least-squares collocation 10.5
273
and its inverse matrix
[︄ ]︄
0.0018 −0.0012
(Cij + Dij )−1 = mGal−2 .
−0.0012 0.0018
Furthermore
[︂ ]︂ [︂ ]︂
CPi = CP1 CP2 = 444.44 444.44 mGal2 .
∆g
ˆ︂ P = [︄ ]︄ [︄ ]︄
[︂ ]︂ 0.0018 −0.0012 15
= 444.44 444.44 mGal =
−0.0012 0.0018 20
= 9.333 mGal.
í Õ ! ¤. û
274 10 Statistical methods
zero. In fact, not using the observational data at all would leave
us with the a priori estimate
√
ˆ︂ P = 0 ±
∆g 1000 mGal = 0 ± 31.623 mGal,
almost as good.
If, instead, we had used point P ′ in between points 1 and 2, at
location (25 km, 25 km), then
/︂
2 (︁
CP ′ 1 = CP ′ 2 = 1000 mGal 1 + 50 400 = 888.89 mGal
2
/︁ )︁
í Õ ! ¤. û
Least-squares collocation 10.5
275
geoid heights Np and g of gravity anomalies ∆gi . In the latter case, the
Stokes equation is “covertly” along in the structure of the C matrices.
These matrices are built from covariance functions. Their elements
can be expressed as follows:8 8
{︁ }︁ {︁ }︁ {︁ }︁
= M ′ fp gi , = M ′ g i gj ,
[︁ ]︁ [︁ ]︁ [︁ ]︁
Cfg pi
Cgg ij
Dgg ij
= E ni nj ,
ℓi = gi + ni , or equivalently ℓ = g + n.
8 Here,
{︁ }︁
we use the geographic mean M ′ · for evaluating the signal covariances. In
doing so, f and g are no longer considered stochastic. It is assumed that their global
{︁ }︁ {︁ }︁
geographic mean vanishes: M f = M g = 0.
í Õ ! ¤. û
276 10 Statistical methods
Here are several points j where gravity has been measured: let us say,
N observations ℓj = ∆gj + nj , j = 1, . . . , N. The number of points to be
predicted may be one: point P, or many. The matrices Cij and Dij are
square, and the inverse of their sum exists. CPi is a rectangular matrix.
If there is only one point P, CPi is a size 1 × N row matrix.
9 The prediction error is now the difference quantity9 ∆gˆ︂ P − ∆g , and
P
its variance (“variance of prediction”) is
{︂ }︂
def
ΣPP = Var ∆g ˆ︂ P − ∆g =
P
{︁ }︁ {︂ }︂ {︂ }︂ {︂ }︂
= Var ∆g ˆ︂ P + Var ∆g − Cov ∆g
P
ˆ︂ P , ∆g − Cov ∆g , ∆g
P P
ˆ︂ P .
and
{︂ }︂ {︃ (︂ )︂ }︃
−1
Cov ∆gP , ∆gP = Cov CPi (Cij + Dij )
ˆ︂ ∆gj + nj , ∆gP =
(︃ {︂ }︂ )︃
−1
= CPi (Cij + Dij ) Cov ∆gj , ∆gP + 0 =
9 Beaware that here, ∆gP is the true value of the gravity anomaly at point P, which we
do not know empirically. The measured value would be ℓP = ∆gP + nP , in which nP is
the random error or “noise” of the gravimetric observation.
í Õ ! ¤. û
Prediction of gravity anomalies 10.6
277
and also
{︂ }︂
ˆ︂ P = CPi (Cij + Dij )−1 CjP
Cov ∆gP , ∆g
{︂ }︂
and finally, the signal variance Var ∆gP = CPP .
Here, CiP (also called CjP , or even CℓP ) is the transpose of CPi . The
matrix (Cij + Dij )−1 is symmetric and its own transpose.
The end result is
Borderline cases
◦ Point P is far from all points i. Then CPi ≈ 0 and ΣPP ≈ CPP ,
so prediction is impossible in practice, and the prediction
equation 10.10 will yield the value zero. The mean error of
√ √
prediction σ∆gP = ΣPP is the same as the variability CPP
of the gravity anomaly signal, the square root of the signal
variance.
◦ Point P is identical with one of the points i. Then, if we use
only that point i, we obtain
í Õ ! ¤. û
278 10 Statistical methods
∑︂
∞ ∑︂
n
K(ψ) = knm Ynm (ψ, α)
n=2 m=−n
∑︂
∞ ∑︂
∞
K(ψ) = kn0 Yn0 (ψ) = kn Pn (cos ψ). (10.14)
n=2 n=2
í Õ ! ¤. û
Covariance function and degree variances 10.7
279
The coefficients kn are called the degree variances (of the disturbing astevarianssi
potential). For isotropic covariance functions K(ψ), the information
content of the degree variances kn , n = 2, 3, . . . is the same as that of
the function itself, and is in fact its spectral representation.
í Õ ! ¤. û
280 10 Statistical methods
The degree variances are the geographic variances of the degree con-
stituents of the disturbing potential.
T (ϕ, λ, r) =
GM⊕ ∑︂ (︂ R )︂n+1 ∑︂
∞ n
(︁ )︁
= r P nm (sin ϕ) δC nm cos mλ + Snm sin mλ ,
R
n=2 m=0
∗
in which the normal field, coefficients Cn , has been subtracted out:
⎧
⎨δC = C − C∗ if n even,
n0 n0 n
⎩δCnm = Cnm otherwise.
We see that
GM⊕ ∑︂
n
(︁ )︁
Tn (ϕ, λ) = Pnm (sin ϕ) δCnm cos mλ + Snm sin mλ .
R
m=0
We obtain
x (︃ )︃2 ∑︂
n (︂
1 GM⊕ 2 2
)︂
Tn2
⟨︁ ⟩︁
kn = dσ = Tn · Tn σ = δCnm + Snm .
4π σ R
m=0
í Õ ! ¤. û
Propagation of covariances between various quantities 10.8
281
The degree variances kn of the disturbing potential can be calculated
directly from the spherical-harmonic coefficients.
kn = σ2n = σTi T .
def
∑︂
∞
T (ϕ, λ, R) = Tn (ϕ, λ),
n=2
it holds that ∑︂
∞ (︂ )︂n+1
R
T (ϕ, λ, r) = r Tn (ϕ, λ).
n=2
Symbolically
{︁ }︁
T (ϕ, λ, r) = L T (·, ·, R) .
Here, L is the linear functional
{︁ }︁ ∑︂
∞ (︂ )︂n+1
R
L f = r fn ,
n=2
í Õ ! ¤. û
282 10 Statistical methods
∑︂
∞
f= fn .
n=2
Symbolically
{︁ }︁ ∑︂
∞
L f = Ln fn ,
n=2
in which (︂ )︂n+1
n R
L = r
is the spectral representation of the functional L.
We may write at a certain point P, location (ϕP , λP , rP ) in space:
{︁ }︁ ∑︂
∞
LP f = Ln
P fn,P ,
n=2
in which (︂ )︂n+1
R
LnP = rP .
(︁ )︁
Concretely, for the disturbing potential T ϕP , λP , rP in point P, this
means
(︁ )︁ {︁ }︁ ∑︂
∞
n
∑︂
∞ (︂
R n+1 (︁
)︂ )︁
T ϕP , λP , rP = LP T (·, ·, R) = LP Tn,P = r T n ϕ P , λ P .
P
n=2 n=2
í Õ ! ¤. û
Propagation of covariances between various quantities 10.8
{︁ }︁
283
Based on orthogonality,10 M ′ Tn,P Tn ′ ,Q(P) = 0 if n ̸= n ′ . So 10
)︁ ∑︂
∞
′
{︁ }︁
Ln n
(︁
K rP , rQ , ψPQ = P LQ M Tn,P Tn,Q(P) . (10.16)
n=2
)︁ ∑︂
∞
{︁ }︁
M ′ Tn,P Tn,Q(P) .
(︁
K ψPQ =
n=2
(︁ )︁ ∑︂
∞
K ψPQ = kn Pn (cos ψ), (10.14)
n=2
we see that
{︁ }︁
M ′ Tn,P Tn,Q(P) = kn Pn (cos ψPQ ).
10 Likein the proof of the harmonicity of r ∆g in section 8.5, one must take along the
third dimension. Write
{︁ }︁ 1 w 2π {︁ }︁
M ′ Tn,P Tn ′ ,Q(P) = M Tn,P Tn ′ ,Q(P) dαPQ =
2π 0 {︃ }︃
1 w 2π {︂ }︂
= M Tn,P · Tn ′ ,Q(P) dαPQ = M Tn,P Tn⃝′ ,P ,
2π 0
1 w 2π 1 w 2π (︂ r )︂n ′ +1
Tn⃝′ ,P (r) = Tn ′ ,Q(P) (r) dαPQ = Tn ′ ,Q(P) (R) dαPQ =
2π 0 2π 0 R
(︂ r )︂n ′ +1 1 w 2π (︂ r )︂n ′ +1
= · Tn ′ ,Q(P) (R) dαPQ = Tn⃝′ ,P (R).
R 2π 0 R
This shows that Tn⃝′ ,P (r) is a perfectly legal solid spherical harmonic of degree
n ′ , Tn⃝′ ,P = Tn⃝′ ,P (R) a legal surface harmonic, and the orthogonality of spherical
harmonics applies:
{︁ }︁ {︂ }︂
M ′ Tn,P Tn ′ ,Q(P) = M Tn,P Tn⃝′ ,P = 0 if n ̸= n ′ .
í Õ ! ¤. û
284 10 Statistical methods
)︁ ∑︂
∞
Krn rP , rQ Kψ
(︁ (︁ )︁ (︁ )︁
K rP , rQ , ψPQ = n ψPQ ,
n=2
(︁ )︁
and Kψn ψPQ must have the same form as K(ψ) in equation 10.14, and
for the same reason:
Kψ
(︁ )︁
n ψPQ = kn Pn (cos ψPQ ).
(︁ )︁
At sea level Krn R, R = 1 and the coefficients kn are those given by
equation 10.15.
11 Equation 10.16 now becomes11
)︁ ∑︂
∞
Ln n
(︁ (︁ )︁
K rP , rQ , ψPQ = L
P Q k n Pn cos ψPQ =
n=2
∑︂ (︂
∞
R
)︂n+1 (︂
R n+1
)︂ (︁ )︁
= rP rQ k n Pn cos ψ PQ =
n=2
∑︂
∞ (︃
R2
)︃n+1
(︁ )︁
= rP rQ k n Pn cos ψPQ . (10.17)
n=2
11 Thisonly works this cleanly because in this case, the operator Ln is of multiplier
(︁ /︁ )︁n+1
type, R r .
í Õ ! ¤. û
Propagation of covariances between various quantities 10.8
285
^ 10.8.2 Example: the covariance function of gravity anomalies
We know from equation 5.7 that there exists the following relationship
between gravity anomalies and the disturbing potential:
∑︂
∞
n − 1 R n+1
(︂ )︂
∆g = r r Tn ,
n=2
{︁ }︁
symbolically: ∆g = L∆g T for a suitable operator L∆g :
{︁ }︁ ∑︂
∞
n − 1 R n+1
(︂ )︂
L∆g f = Ln
∆g fn , Ln
∆g = r r .
n=2
Often, we write
)︁ def {︁ }︁ {︁ }︁
C rP , rQ , ψPQ = Cov ∆gP , ∆gQ = M ′ ∆gP ∆gQ(P) =
(︁
∑︂∞ (︃
R2
)︃n+2
(︁ )︁
= rP rQ cn P n cos ψPQ ,
n=2
í Õ ! ¤. û
286 10 Statistical methods
í Õ ! ¤. û
Global covariance functions 10.9
287
^ 10.9 Global covariance functions
Empirical covariance functions have been calculated often. There have
been only a few empirical covariance functions for the whole Earth.
They are commonly given in the form of a degree variance formula. The
best known is the rule observed by William Kaula (Rapp, 1989):12 12
2n + 1
kn = α .
n4
By writing
n−1 2
(︂)︂
cn = kn ,
R
in which cn are the degree variances of gravity anomalies, we obtain
2n + 1 n − 1 2 2α
(︂ )︂
cn = α ≈ .
n4 R nR2
Here, α is a planet specific constant, according to Kaula’s estimate
/︁ )︁2
α = 10−10 GM⊕ a⊕ .
(︁
The Kaula rule does not hold very precisely. It applies, by the way,
fairly well for the gravity field of Mars, of course with a different
constant (Yuan et al., 2001).
Another well-known rule is the Tscherning–Rapp equation (Tschern-
ing and Rapp, 1974):
A (n − 1) n−1 2
(︂ )︂
cn = = kn .
(n − 2) (n + B) R
12 William
M. Kaula (1926–2000) was an American geophysicist and space geodesist
who studied the determination of the Earth’s gravity field by means of satellite
geodesy.
13 Arne Bjerhammar (1917–2011) was a Swedish geodesist.
í Õ ! ¤. û
288 10 Statistical methods
Kaula EGM96
108 EGM2008
Tscherning–Rapp
GOCE, Gatti et al. (2014)
106 EGM96 error variances
EGM2008 error variances
104
)︁
102
Degree variance kn
100
10−2
10−4
10−6
Figure 10.5. Global covariance functions as degree variances. The GOCE model
^ cuts off at degree 280.
í Õ ! ¤. û
Collocation and the spectral viewpoint 10.10
289
that also the results of the calculation, estimates fˆ︁i of the result function
f(ψ), are desired at the same points. Then we have equation 10.9:
with
[︁ ]︁ (︁ )︁ (︁ )︁
Cfg = Cfg f(ψi ), g(ψj ) = Cfg ψi , ψj ,
ij
[︁ ]︁ (︁ )︁ (︁ )︁
Cgg ij = Cgg g(ψi ), g(ψj ) = Cgg ψi , ψj ,
[︁ ]︁ (︁ )︁ (︁ )︁
Dgg ij = Dgg g(ψi ), g(ψj ) = Dgg ψi , ψj .
{︂ )︁}︂ 1 ∑︂
N−1
[︁ ]︁ (︁ (︁ )︁ def [︁ ]︁
Cfg = M⃝ f(ψi ) g ψj(i) = f(ψi ) g ψj(i) = Cfg k ,
i,j(i) N
i=0
which, like the geographic average in section 10.4, replaces the statistical
average.
In the same way we obtain
[︁ ]︁ {︂ (︁ )︁}︂
Cgg i,j(i)
= M⃝ g(ψi ) g ψj(i) =
1 ∑︂
N−1
(︁ )︁ def [︁ ]︁
= g(ψi ) g ψj(i) = Cgg k .
N
i=0
Now Cfg , Cgg are functions of only k, and we may write them
[︁ ]︁ (︁ )︁ [︁ ]︁
Cfg
ij
= C fg ψ i , ψj = C fg (∆ψk ) = Cfg k ,
[︁ ]︁ (︁ )︁ [︁ ]︁
Cgg ij = Cgg ψi , ψj = Cgg (∆ψk ) = Cgg k ,
í Õ ! ¤. û
290 10 Statistical methods
2 1
i
0
N−1
N−2
j
∆ψk
ψj
ψi
def
in which ∆ψk = (ψj − ψi ) mod 2π and k = (j − i) mod N.
Furthermore
[︁ ]︁ (︁ )︁ [︁ ]︁ {︁ }︁
Dgg ij
= Dgg ψi , ψj = Dgg (∆ψk ) = Dgg k = E ni nj(i) ,
Dgg = σ2 IN ,
í Õ ! ¤. û
Self-test questions
291
Without proof we present that the spectral version of equation 10.9
looks like this:
{︁ }︁ {︂ }︂ {︁ }︁ {︂ }︂
{︁ }︁ F C fg F Cfg
f = {︁ }︁
F ˆ︁ {︁ }︁ F g + n = {︁ }︁ F g + n .
F Cgg + F Dgg F Cgg + σ N
2/︁
This is an easy and fast way to calculate the solution using FFT. If for a
{︁ }︁
suitable operator L we have f = L g , the equation becomes
{︁ }︁ {︁ }︁ {︂ }︂
{︁ }︁ F L · F Cgg
F ˆ︁f = {︁ }︁ 2/︁
· F g + n .
F Cgg + σ N
For example, if g are gravity anomalies and f are values of the disturbing
potential, then16 16
{︁ }︁ R
F L = .
n−1
The approach is called Fast Collocation, for example Bottoni and Barzaghi
(1993). Of course it is used in two dimensions on the Earth’s surface,
although our example is one-dimensional. As always, it requires that
the observations are given on a grid, and in this case also that the
precision of the material is homogeneous — the same everywhere — over
the area. This requirement is hardly ever precisely fulfilled.
^ Self-test questions
1. What is the difference between signal and noise?
2. What is a functional?
16 In
real computation though, the equation must be changed to use, instead of the
harmonic degree number n, which refers to global spherical geometry, the Fourier
wave number expressed on the computation grid used.
í Õ ! ¤. û
292 10 Statistical methods
ΣPP ≈ DPP .
í Õ ! ¤. û
Exercise 10 –2: Hirvonen’s covariance equation and prediction
293
^ Exercise 10 – 2: Hirvonen’s covariance equation and
prediction
Hirvonen’s covariance equation is
C0
C(s) = (︁ /︁ )︁2 , (10.18)
1+ s d
with the Ohio parameters C0 = 337 mGal2 and d = 40 km (Heiskanen
and Moritz, 1967, equation 7-9). The equation gives the covariance
between the gravity anomalies at two points P and Q:
(︁ )︁ {︂ }︂
C sPQ = Cov ∆gP , ∆gQ .
í Õ ! ¤. û
294 10 Statistical methods
is the signal covariance matrix between ∆gP and ∆gi . Dij is the variance
matrix of the observation random uncertainty or noise ni , i = 1, 2:
[︄ {︁ }︁ {︁ }︁ ]︄
Var n1 Cov n1 , n2
Dij = {︁ }︁ {︁ }︁ .
Cov n1 , n2 Var n2
í Õ ! ¤. û
Exercise 10 –5: Propagation of covariances
295
triangular configuration, with a right angle at point P, and the distances
from P to points 1 and 2 still 40 km. The distance between points 1 and
√
2 is now only 40 2 km.
1. Compute Cij , CPi , ∆g
ˆ︂ P and ΣPP .
2. Compare the result with the previous one. Conclusion?
{︁ }︁ ∑︂
∞ (︃ )︃n+1
R2 (︁ )︁
Cov T P , T Q = r r k n Pn cos ψPQ .
P Q
n=2
∑︂
∞
δg = Ln
δg Tn
n=2
{︂ }︂ ∑︂
∞
Ln n
(︁ )︁
Cov δgP , δgQ = δg,P Lδg,Q kn Pn cos ψPQ .
n=2
∂2 ∂
2
T = − δg,
∂r ∂r
in other words, the vertical gradient of the gravity disturbance.
∑︂
∞ (︂ )︂n+1
R
T (ϕ, λ, r) = r Tn (ϕ, λ)
n=2
í Õ ! ¤. û
296 10 Statistical methods
2n + 1
kn = α .
n4
From these one can derive, using propagation of variances, the degree
variances of gravity anomalies
∑︂
∞ ∑︂
∞ (︂
n−1
)︂
∆g(ϕ, λ, R) = Ln
∆g (R) Tn (ϕ, λ) = Tn (ϕ, λ)
R
n=2 n=2
as follows:
n−1 2 2α
)︁2 (︂ )︂
cn = Ln
(︁
∆g (R) k n = kn ≈ .
R nR2
Analogously, differentiate expansion 5.7 for the gravity anomaly
∑︂
∞
n − 1 R n+1
(︂ )︂
∆g(r, φ, λ) = r r Tn (ϕ, λ),
n=2
yielding
∂ ∆g ∑︂ (n − 1) (n + 2) (︂ R )︂n+1
∞
=− r Tn ,
∂r r2
n=2
with
(n − 1) (n + 2)
Ln
∆g ′ (R) = − .
R2
í Õ ! ¤. û
Exercise 10 –7: Underground mass points
297
1. Derive an approximate equation for the degree variances for the
anomalous gravity gradient. Designate them with the symbol cn′ ,
in an analogue fashion as above for the gravity anomaly degree
variances cn :
cn′ = x(n) · kn ≈ y · nz .
Find the expression x(n) and the constants y and z for the case of
the Earth.
2. Conclusion?
í Õ ! ¤. û
^ Gravimetric measurement devices
11
^ 11.1 History
The first measurement device ever built based on a pendulum was a
clock. The pendulum equation,
√︃
ℓ
P = 2π g ,
1 Christiaan Huygens FRS (1629–1695) was a leading Dutch natural scientist and
mathematician. Besides inventing the pendulum clock, he also was the first to realise
(in 1655) that the planet Saturn has a ring.
2 JeanRicher (1630?–1696) was a French astronomer. He is really only remembered for
his pendulum finding.
– 299 –
300 11 Gravimetric measurement devices
also used in Finland in the 1920s and 1930s (Pesonen, 1930; Hirvonen,
1937). We must also mention the submarine measurements, including
in the Java Sea by the Dutch Vening Meinesz, in which it was observed
syvänmeren that above the trenches in the ocean floor there is a notable shortage of
hauta gravity, and that they thus are in a state of strong isostatic disequilibrium
(Vening Meinesz, 1928).
Pendulum gravimeters are however too hard to operate and too slow
for high-productivity gravimetric observations. For that purpose, the
spring gravimeter was developed, see section 11.2.
3 Henry Kater FRS FRAS (1777–1835) was a British physicist who made contributions to
scientific instruments and metrology.
4 RobertFreiherr (baron) Daublebsky von Sterneck (1839–1910) was a major general in
the Austro-Hungarian army and a geophysicist, astronomer and geodesist.
í Õ ! ¤. û
The relative or spring gravimeter 11.2
301
Figure 11.2. Autograv™ CG-5 spring gravimeter from Scintrex. Image Mon-
^ niaux (2011).
í Õ ! ¤. û
302 11 Gravimetric measurement devices
d2 ℓ (︁ )︁
=0 =⇒ mg = k (ℓ − ℓ0 ) = k ℓ − ℓ0 , (11.2)
dt2
in which ℓ is the mean length of the spring during the oscillation, and
also the equilibrium length in the absence of oscillations.
When the test mass is disturbed, it starts oscillating about its equilib-
värähtely- rium position. The oscillation equation, obtained by summing equations
yhtälö 11.1 and 11.2, is
d2 (︁ )︁ k (︁ )︁
ℓ − ℓ = − m ℓ − ℓ .
dt2
The period is
√︃ √︃
m ℓ − ℓ0 δℓ
√︂
P = 2π = 2π g = 2π g , (11.3)
k
í Õ ! ¤. û
The relative or spring gravimeter 11.2
303
according to equation 11.4, a lengthening of only 5 · 10−8 m = 50 nm
(check!), one-twelfth the wavelength of a helium-neon laser. Clearly
then, the sensor observing or compensating this displacement must be
extremely sensitive!
^ 11.2.1 Astatisation
An astatised gravimeter uses a different measurement geometry. The
LaCoste-Romberg gravimeter, which long enjoyed great popularity,
serves as an example. Inside it, the test mass is at the end of a lever
beam, see figure 11.3. Two torques operate on the beam, which are in
equilibrium. The torque caused by the spring is
(︁ )︁
τs = k ℓ − ℓ0 b sin β,
The force of gravity pulling at the mass is mg, and the corresponding
torque
τg = mgp cos ϵ.
Between these there has to be equilibrium:
(︁ )︁ bc
τg − τs = mgp cos ϵ − k ℓ − ℓ0 cos ϵ = 0
ℓ
or
(︁ )︁
mgpℓ − kbc ℓ − ℓ0 = 0. (11.5)
í Õ ! ¤. û
304 11 Gravimetric measurement devices
Reality
Working range
Force
Spring, length ℓ Hooke’s law
c
(︁ )︁
k ℓ − ℓ0 sin β
Length
β
ϵ
b Test mass beam
p
mg
By differentiation
dℓ mpℓ mpℓ ℓ ℓ − ℓ0
=− =− )︁ = g ℓ0 .
dg mgp − kbc
/︂(︁
mgp − mgp ℓ ℓ − ℓ0
í Õ ! ¤. û
The relative or spring gravimeter 11.2
305
For example, assuming ℓ = 5 cm, ℓ0 = 0.1 cm, g = 10 m/s2 gives
dℓ
= 2.5 · 10−6 m/mGal ,
dg
a 50 times5 better(︁ result)︁ than earlier! The improvement or astatisation 5
ratio is precisely ℓ − ℓ0 ℓ0 .
/︁
but for a state of disequilibrium. Then, the test mass will be undergoing
an acceleration a, positive downwards, and we have
m (g − a) pℓ − kbc (ℓ − ℓ0 ) = 0,
5 For
comparability we should still multiply by p b sin β , if the position of the test
/︁
mass is measured.
6 LucienLaCoste (1908–1995) was an American physicist and metrologist, who, as
an undergraduate, together with his physics professor Arnold Romberg (1882–1974)
discovered the principle of the astatised gravimeter and zero-length spring.
í Õ ! ¤. û
306 11 Gravimetric measurement devices
ℓ0 (︁ )︁
mapℓ = mgp ℓ−ℓ
ℓ − ℓ0
or
g ℓ0 (︁ )︁
a=− ℓ−ℓ .
ℓ ℓ − ℓ0
(︁ )︁/︁
Here we see again the “astatisation ratio” ℓ − ℓ0 ℓ0 appear, which for
a zero-length spring (ℓ0 ≈ 0) is very large.
Now the string length disequilibrium ℓ − ℓ is connected with the
vertical displacement z (reckoned downwards) of the test mass, as
follows:
(︁ )︁ p
z= ℓ−ℓ .
b sin β
With this, we obtain
d2 g ℓ0 b sin β
a= z=− z.
dt 2 ℓ ℓ − ℓ0 p
P = 4.4 s.
What this long oscillation period also means is that the instrument is
less sensitive to high-frequency vibrations, for example from passing
traffic or microseismicity. This is a significant operational advantage.
í Õ ! ¤. û
The relative or spring gravimeter 11.2
307
δ(ε)
ε)
α F(ε) cos(α + δ + ε)
F(
−ε
mg
mg
Figure 11.4. The idea of astatisation. The elastic force of an ordinary spring
grows steeply with extension (left), whereas the weight of the
test mass is constant. The lever beam and diagonal arrangement
(right) causes the part of the force of the spring in the direction
of motion of the lever (red) to diminish with extension, while
the spring force itself grows almost similarly with extension.
This near-cancellation boosts sensitivity. The spring used is a
^ zero-length spring.
í Õ ! ¤. û
308 11 Gravimetric measurement devices
í Õ ! ¤. û
The absolute or ballistic gravimeter 11.3
309
Vacuum pump system
Cage transporter
Falling prism
“Superspring”
g
Reference prism
Semi-transparent mirror
Laser
Mirror
purpose of the cage is to prevent the last remaining traces of air from
affecting the motion of the prism. Approaching the bottom, the cage,
which moves along a rail under computer control, decelerates, and the
prism lands relatively softly on its base. After that, the cage moves back
to the top of the tube and a new measurement cycle starts.
A laser interferometer measures the locations of the prism during its
fall. The measurements are repeated thousands of times to get good
precision through averaging. Another prism, the reference prism, is sus- vertausprisma
pended in another tube from a very soft spring (actually an electronically
simulated “superspring”) to protect it from microseismicity.
The instrument is designed to achieve the greatest precision possible;
for example, the vibration caused by the drop is controlled by a well-
designed mount. Precisions are of the order of several microgals, similar
to what LaCoste-Romberg relative gravimeters are capable of.
í Õ ! ¤. û
310 11 Gravimetric measurement devices
d2
z = g(z),
dt2
where it is assumed — realistically — that gravity g depends on the
location z within the drop tube, reckoned downwards. If we nevertheless
take g to be constant, we obtain by integration
d
z = v0 + gt, z = z0 + v0 t + 12 gt2 ,
dt
from which we obtain the observation equations of the measurement
í Õ ! ¤. û
The absolute or ballistic gravimeter 11.3
311
process ⎡ ⎤
[︂ ]︂ z 0
1 2
zi = 1 ti t · ⎣ v0 ⎦ + ni .
⎢ ⎥
2 i
g
Here, the unknowns9 to be estimated are z0 , v0 and g. The quantities 9
zi are the interferometrically measured vertical locations of the falling
prism, and ni are the measurement errors or “noise”. Determining
precisely the corresponding measurement time or epoch ti reckoned
from a moment close to that of release of the prism is of course essential.
The number of measurement values obtained from each individual
drop is substantial.
We write the observation equations in matric form:
ℓ = A x + n,
in which
⎡ ⎤ ⎡ ⎤ ⎡ 1 2
⎤
z1 n1 1 t1 t
2 1
⎡ ⎤
1 2 z0
⎢ z2 n2 1 t2 t
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
2 2
⎥ ⎢ ⎥ ⎢ ⎥
ℓ=⎢ ⎥, n = ⎢ ⎥, A = ⎢ ⎥, x = ⎢
⎣ v0 ⎦ .
⎥
⎢ .. .. .. .. ..
⎣ . . . . .
⎥ ⎢ ⎥ ⎢ ⎥
⎦ ⎣ ⎦ ⎣
1 2
⎦ g
zn nn 1 tn t
2 n
The solution follows from this according to the method of least-squares pienimmän
adjustment, from the normal equations neliösumman
tasoitus
T T
x = A ℓ,
A A ˆ︁
9 It
would be easy (exercise!) to add an unknown representing the vertical gradient of
gravity to this. Whether a useable value for this unknown can be obtained from the
measurements is a good question.
í Õ ! ¤. û
312 11 Gravimetric measurement devices
fall fast at first and then slower. To bring this about, a second laser
pulse pair is used that acts like a mirror, or perhaps a tennis racket. The
third and last laser pulse recombines the beams. Then, constructive or
destructive interference is observed using a fluorescence detector. The
í Õ ! ¤. û
Network hierarchy in gravimetry 11.4
313
phase difference between the two arms of the interferometer is inferred
from these observations.
As the atoms travel through space-time along two different paths on
which the gravity potentials are different,11 a phase difference is formed 11
between these, which in principle can be measured. See figure 11.7, in
which the horizontal axis is time. Without gravity (dashed lines) this
phase difference would be zero.
Like in all (non-kinematic) interferometric methods, the ambiguity
problem — the circumstance that a measured phase is always in the
[︁ )︁
interval 0, 2π , although a phase change or difference may contain
any number of whole cycles — poses its own challenge. Ambiguity
resolution is possible by measuring at several different inter-pulse time
intervals T , figure 11.7.
11 In
fact, the spinning of the phase angle of the atom’s wave function acts like a clock,
and the speed at which time elapses depends on the local geopotential (Vermeer,
1983a).
í Õ ! ¤. û
314 11 Gravimetric measurement devices
t
T T
Recombination
beam
z
Splitting
beam
Mirroring
beam
g Without gravity
In gravity field
í Õ ! ¤. û
The superconducting gravimeter 11.5
315
the Meissner effect. Of course the field itself must be constant. It is gen-
erated by superconducting solenoids inside a vessel made of mu-metal käämi
(Wikipedia, Mu-metal), which keeps out the Earth’s magnetic field.
Superconduction applied for this still demands working at the tem-
perature of liquid helium (He). For this reason the device is not only
expensive, but also requires an expensive laboratory in an environment
where the societal infrastructure allows.
There are more than thirty superconducting gravimeters in the world.
The work is co-ordinated by the IAG service IGETS, the International
Geodynamics and Earth Tide Service. One GWR20-type instrument has
been working since 1994 in Kirkkonummi at the Metsähovi research
station of the then Finnish Geodetic Institute, now the National Land
Survey, Virtanen and Kääriäinen (1995), Virtanen (1998). The instrument
was upgraded in 2014.
í Õ ! ¤. û
316 11 Gravimetric measurement devices
Middle capac-
itor plate
Levitation coil
Feedback
coil
Levitation coil
— and they are of considerable geophysical interest, Wikipedia, Earth normal modes.
í Õ ! ¤. û
Gravity measurement and the atmosphere 11.6
317
remotely controllable superconducting gravimeters, for example the
GWR iGrav® , weighing 30 kg and not expending any liquid helium at
all. On the other hand it needs over a kilowatt in grid power for its
refrigeration system (GWR Instruments, Inc., iGRAV® Gravity Sensors).
Perhaps this will lead to improvement over the current situation where
the bulk of the instruments are located in Europe and North America.
κ= p γ,
/︁
14 Soyes, the force acting on a standard 14-inch laptop screen (aspect ratio 16 : 9) is
547 kg. . . but it doesn’t matter as it is not an old-fashioned vacuum cathode-ray tube.
í Õ ! ¤. û
318 11 Gravimetric measurement devices
í Õ ! ¤. û
Airborne gravimetry and GNSS 11.7
319
twice the Bouguer-plate atmospheric reduction 11.7 calculated
above. This is the value that should be added to the measured
gravity values.
One may also think of this value as the change in gravity if
the local atmospheric Bouguer plate were condensed, Helmert
condensation style, to below the measurement location, producing
a double planar Bouguer correction.
At sea level, the correction is 0.87 mGal. At height, the correction is
p(H)
0.87 p mGal,
0
in which p(H) and p0 are the air pressures at height H and at sea
level, respectively.
í Õ ! ¤. û
320 11 Gravimetric measurement devices
Attached to the aircraft are a number of GNSS antennas. With these and
a geodetic GNSS instrument, the motions of the aircraft can be monitored
with centimetre accuracy. From these motions, we may then calculate
the pseudo-forces mentioned above under item 2.
If we measure the position of the plane (or instrument) xi at moments
ti , ∆t = ti+1 − ti , we obtain estimated acceleration values as follows
(in an inertial frame):
x∗i+1 + x∗i−1 − 2x∗i
a∗i ≈ . (11.8)
∆t2
When the acceleration measured by the gravimeter is g˜︁ and the direction
luotiviiva of the local plumb line (upwards) n, local gravity g follows:
⟨︂(︁ )︁ ⟩︂
∗
˜︁ − a∗ · n − ω2⊕ N(φ) cos φ,
⟨︁ ⟩︁
g=g ˜︁ − a + fω · n = g
and we obtain
⟨︂(︁ )︁ ⟩︂
⊕ ′
˜︁ − a⊕ · n + 2ω⊕ veast cos φ .
⟨︁ ⟩︁
˜︁ − a − fω · n = g
g=g
í Õ ! ¤. û
Airborne gravimetry and GNSS 11.7
321
Often the high-frequency part removed from the signal is 10 000 times
stronger than the gravity signal we are after! See for example Lu et al.
(2017) figure 2.
If the uncertainty (mean error) of one GNSS vertical position co-ordinate
measurement is σz (and the different co-ordinates do not correlate with
each other), then according to equation 11.8 the uncertainty of the
vertical acceleration is √
σz 6
σa = .
∆t2
Making the time interval ∆t as long as possible without resolution
suffering requires a low flight speed. Generally a propeller aircraft or
even a helicopter is used. Of course the price of the measurement grows
with the duration of the flight — a helicopter rotor hour is expensive!
The flight height H is chosen in accordance with resolution ∆x:
H ∼ ∆x = v ∆t,
where v is the flight speed. The separation between adjacent flight lines
is chosen similarly.
The first major airborne gravimetry project was probably the
Greenland Aerogeophysics Project (Brozena, 1992). In this ambitious
American-Danish project in the summers of 1991 and 1992, over
200 000 km was flown, all the time measuring gravity and the magnetic
field, and the height of the ice surface using altimetry.
Since then, other large uninhabited areas in the Arctic and Antarctic
regions have also been mapped, see Brozena et al. (1996), Brozena and
Peters (1994). Already in subsection 9.6.2 we made mention of other
large surveys. Activity continues, see Coakley et al. (2013), Kenyon et al.
(2012). The method is very suitable for large, uninhabited areas, but
also, for example, for sea areas near the coast or inside archipelagos,
where ship gravimeters would have difficulty navigating long straight
tracks. In 1999, an airborne gravimetry campaign was undertaken over
the Baltic Sea, including the Gulf of Finland (Jussi Kääriäinen, personal
communication).
í Õ ! ¤. û
322 11 Gravimetric measurement devices
í Õ ! ¤. û
Measuring the gravity gradient 11.8
323
up, gravity diminishes, about 0.3 mGal for every metre of height.
In topocentric co-ordinates (x, y, z), where z points to the zenith, this
matrix is approximately
⎡ ⎤
−0.15 0 0
M≈⎣ 0 −0.15 0 ⎦ mGal/m ,
⎢ ⎥
0 0 0.3
∂ GM ∂ (R + z) 2gz
g =2 · =− ≈
∂z z (R + z) 3 ∂z (R + z)
≈ 3 · 10−6 m/s2 m = 0.3 mGal/m .
/︁
The quantities ∂xx W and ∂yy W again represent the curvatures of the
equipotential or level surfaces in the x and y directions, equations 4.4:
∂2 W g ∂2 W g
∂xx W = = −ρ , ∂yy W = = −ρ ,
∂x2 x ∂y2 y
í Õ ! ¤. û
324 11 Gravimetric measurement devices
0 0 3000
Note that
∂2 W ∂2 W ∂2 W
+ + = ∂xx W + ∂yy W + ∂zz W ≈ 0,
∂x2 ∂y2 ∂z2
^ Self-test questions
1. For the spring gravimeter described in section 11.2, one milligal
of change in gravity g produces according to equation 11.4 a
lengthening of 5 · 10−8 m. Do a calculational check.
2. Why is a pendulum gravimeter, although theoretically absolute,
not very accurate as an absolute gravimeter?
3. By which method choices do we, in practical measurements, take
the drift of a relative gravimeter into account?
4. Why were, before the advent of absolute gravimeters, the reference
points of international fundamental gravimetric networks often
on airports?
5. In an absolute or ballistic gravimeter, what is the role of:
í Õ ! ¤. û
Exercise 11 –1: Absolute gravimeter
325
(a) the “cage” surrounding the falling prism
(b) the “superspring”?
6. According to Google
◦ The Gulf War from 1990 to 1991 was the first conflict in which
the military used GPS widely.
◦ By December 1993, GPS achieved initial operational capabil-
ity (IOC), indicating a full constellation (24 satellites) being
available.
◦ The Greenland Aerogeophysics Project, the first ever large-
scale airborne gravimetric mission, mapped the gravity field
of Greenland during the summers of 1991 and 1992.
Why are these three dates so close together?
í Õ ! ¤. û
326 11 Gravimetric measurement devices
í Õ ! ¤. û
Exercise 11 –3: Air pressure and gravity
327
2. How much does sea water rise due to the “inverted barometer ylösalainen
effect” under a low-pressure zone? ilmapuntari
3. To how much does the effect from point 2 amount in local gravity
measured on a ship? Assume that you are on the open sea, that
the free-air vertical gravity gradient is −0.3 mGal/m, and that the
density of sea water is 1030 kg/m3 . Analyse the situation carefully.15 15
í Õ ! ¤. û
^ The geoid, mean sea level, and
sea-surface topography
– 329 –
330 12 The geoid, mean sea level, and sea-surface topography
“The level surface of the Earth’s gravity field that agrees on average best
with the mean sea level.”
The practical problem with this definition is that determining the correct
level of the geoid requires knowledge of the mean sea level everywhere
on the world ocean. This is why many “geoid” models in practice do
not coincide with global mean sea level, but with some locally defined
mean sea level — and often only approximately.
Mean sea level in its turn is also a problematic concept. It is sea level
from which all periodic effects have been computationally removed
— but who can know if a so-called secular effect in reality is perhaps
long-period? The measure of permanency are the time series that are
mareografi available, as tide gauges have been widely operating already for about
í Õ ! ¤. û
Geoid models and national height datums 12.2
331
a century, when again modern satellite time series — TOPEX/Poseidon
and its successors — are just about a quarter of a century long.
A sensible compromise may be the average sea level over 18 years, an
important periodicity, saros (Wikipedia, Saros), in the orbital motion of
the Moon.
í Õ ! ¤. û
332 12 The geoid, mean sea level, and sea-surface topography
δN = a + b (λ − λ0 ) + c (φ − φ0 ) + · · ·
í Õ ! ¤. û
334 12 The geoid, mean sea level, and sea-surface topography
the geoid must also change. The geoid rise is, however, small
compared to the land uplift, only a few percent of it.
Equation (the dot above a quantity denotes the derivative with respect
1 to time1 ):
̇ = ḣ − N
H ̇ =H ̇r+H ̇e+H ̇ t,
in which
Ḣ relative land uplift from the geoid
ḣ absolute land uplift from the reference ellipsoid
̇r
H relative land uplift from the local mean sea level
̇e
H eustatic (global mean sea level) rise
̇t
H change over time of the sea-surface topography (likely small)
̇
N geoid rise from the reference ellipsoid.
The rise of the geoid as a result of land uplift can be simply calculated
with the Stokes integral equation:
x
dN R d
(︂ )︂
= S(ψ) ∆g dσ.
dt 4πγ σ dt
d
Here, dt ∆g is the change of gravity anomalies over time due to land
uplift. Unfortunately we do not precisely know the mechanism by which
mass flows in the Earth’s mantle to underneath the land-uplift area. We
may posit
d dH ̇,
∆g = c = cH
dt dt
in which the constant c may range from −0.16 to −0.31 mGal/m .
◦ The value −0.16 mGal/m is called the “Bouguer hypothesis”: it
corresponds to the situation in which upper mantle matter flows
into the space freed up underneath the rising Earth’s crust, filling
it. This matter may be roughly modelled as a Bouguer plate.
í Õ ! ¤. û
The geoid and post-glacial land uplift 12.3
335
Earth’s crust
Asthenosphere
(a)
Bouguer hypothesis. . .
Earth’s crust
Upper mantle
(b)
. . . and free-air hypothesis.
◦ The value −0.31 mGal/m is the opposite extreme, the “free-air hy-
pothesis”. By this hypothesis, the ice load during the last ice
age has only compressed the Earth’s mantle, and now it is slowly
expanding again to its former volume (the “rising dough model”). pullataikina-
malli
Up until fairly recently, the most likely value was about −0.2 mGal/m ,
with substantial uncertainty. The latest results (Mäkinen et al., 2010;
Olsson et al., 2019) may be summarized as −0.16 ± 0.02 mGal/m (one
standard deviation), which would seem to settle the issue. It looks like
the Bouguer hypothesis is closer to physical reality. The flow of mass is
í Õ ! ¤. û
336 12 The geoid, mean sea level, and sea-surface topography
5◦ ◦
65 ◦ ◦
35 ◦
10 ◦ ◦ 30 65
15◦ 20◦ 25
Föllinge
Meldal
Kopperå Stugun Vaasa Joensuu
Vågstranda Kramfors
Äänekoski
◦
60 ◦ 60
◦
5◦ 35
10 ◦ ◦
30
15◦ 20◦ 25◦
^ Figure 12.2. The Fennoscandian gravity line on the 63rd parallel north.
í Õ ! ¤. û
Global sea-surface topography and heat transport 12.5
337
◦ positioning of tide gauges along the coast using GNSS, together
with gravimetric geoid determination
◦ precise levelling along the coast connecting tide gauges.
In addition to this, we still have the oceanographic method: physical
modelling. The method is termed steric levelling if temperature and
salinity measurements along vertical profiles are used on the open
ocean, and geostrophic levelling if ocean current measurements are used
to determine the Coriolis effect, generally close to the coast.
All methods should give the same results. The Baltic Sea is a textbook
example where all three geodetic methods have been used. It has been
found that the whole Baltic Sea surface is tilted: relative to a level
surface, the sea surface goes up from the Danish straits to the bottoms pohjukka
of the Gulf of Finland and the Bothnian Bay by 25–30 cm.
Oceanographic model calculations show that this tilt is mainly due
to a salinity gradient: in the Atlantic Ocean, the salinity is 30–35 o/oo,
when in the Baltic it drops to 5–10 o/oo, due to the massive production
of fresh water by the rivers (Ekman, 1992). Of course on top of this
come temporal variations, like oscillations caused by storms resembling
those in a bathtub, the amplitude of which can be over a metre.
In Ekman (1992) more is said about the sea-surface topography of the
Baltic and its determination.
í Õ ! ¤. û
338 12 The geoid, mean sea level, and sea-surface topography
′
⟨︁ ⟩︁
fω = −2 ω⊕ × v , (12.1)
def ⟨︁ ⟩︁
ω⊕ = ω⊕ · n n = (ω⊕ sin φ) n.
′ def
⟨︁ ⟩︁ ⟨︁ ⟩︁
fω = −2 ω⊕ × v = −2ω⊕ sin φ n × v ,
′
fω ′
= ∥fω ∥ = 2v ω⊕ sin |φ| .
í Õ ! ¤. û
Global sea-surface topography and heat transport 12.5
339
For example, in the case of the Gulf Stream, the height changes
caused by this effect are several decimetres. If we define a local (x, y) co-
ordinate system in which x(φ, λ) is pointing north and y(φ, λ) east, we
may write for the sea-surface topography H the geostrophic equations
∂H ω ∂H ω
= −2vy γ⊕ sin φ, = +2vx γ⊕ sin φ. (12.2)
∂x ∂y
As we will see in chapter 13, we can measure the location in space of
the sea surface at a precision of a few centimetres using satellite radar
altimetry. If we furthermore have a precise geoid map, we may calculate
the sea-surface topography, and with the aid of equations 12.2 solve for
the flow velocity vector field2 2
[︂ ]︂T [︂ ]︂T
vx (x, y) vy (x, y) = vx (φ, λ) vy (φ, λ) .
2A popular, though unofficial, unit for ocean current is the sverdrup (Wikipedia,
Sverdrup), a million cubic metres per second. All the rivers of the world together
make about one sverdrup, while the Gulf Stream is 30–150 Sv. “There is a river in the
ocean” – Matthew Fontaine Maury (1806–1873), American polymath and pioneer of
oceanography.
í Õ ! ¤. û
340 12 The geoid, mean sea level, and sea-surface topography
North Atlantic Ocean, mesoscale eddies have been moving alongside the
Gulf Stream; eddies of size 10–100 km which show up in altimetric
imagery. It is interesting that the eddies also show up in maps of the
ocean surface temperature, and biologists have observed that life inside
the eddies differs from that outside them (Godø et al., 2012). The life
span of the eddies can be weeks, even months.
A good, though somewhat dated, introduction to “geodetic oceanog-
raphy” is given by Rummel and Sansó (1992).
í Õ ! ¤. û
The global behaviour of the sea level 12.6
341
í Õ ! ¤. û
342 12 The geoid, mean sea level, and sea-surface topography
í Õ ! ¤. û
The sea-level equation 12.7
343
Sea level
drops Sea level Sea level
rises drops
Greenland
Antarctica
Figure 12.5. The sea-level equation. Sea level reacts in a complicated way
^ when continental ice sheets melt.
2015):
G (︁
(︂ )︁ (︁ )︁)︂
S = SE + ρi Gs ⊗i I − Gs ⊗i I + ρw Gs ⊗o S − Gs ⊗o S , (12.3)
R
in which
◦ S = S(ω, t) = S(ϕ, λ, t) means the variations of sea level as a
function of place ω = (ϕ, λ) and time t. These variations are
relative to the solid Earth’s surface: they are changes in sea depth.
S is also what tide gauges measure.
◦ I = I(ω, t) is similarly a function of place and time describing the
variations in thickness of ice sheets and glaciers.
◦ SE is the eustatic term, the variation in ice volume converted into
“equivalent global sea-level variation”, in an equation
mi (t)
SE (t) = − ,
ρw A o
in which mi (t) is the variation in total ice mass as a function of
time, ρw the density of sea water, and Ao the total surface area of
the oceans.
◦ R is the mean radius of the Earth, G Newton’s universal gravita-
tional constant, section 1.2.
í Õ ! ¤. û
344 12 The geoid, mean sea level, and sea-surface topography
def
in which ∆t = t − t ′ ⩾ 0.
This equation expresses simply that the sea depth S is the distance
between sea surface and sea floor, and that a change in it is the
difference between the vertical displacements of those two: that
of the sea surface caused by a change in potential V, and that of
the sea floor, meaning a change in local radius r.
Here the Green’s function of the geopotential is
GV (ψ, ∆t) = GrV (ψ, ∆t) + GeV (ψ, ∆t) + GvV (ψ, ∆t),
/︂
r
The function GV (ψ, ∆t) = δ(∆t) 2 sin 12 ψ is the rigid partial
(︁ )︁
í Õ ! ¤. û
The sea-level equation 12.7
345
^ Figure 12.6. Sea-level rise after the last ice age (Rohde, 2005).
The functions GeV and GvV are the elastic and viscous deformation
partial Green’s functions of the geopotential. These thus charac-
terise the rheological behaviour of the Earth, and their theoretical
calculation requires the internal density and viscosity distribu-
tions ρ(r) and η(r) of the Earth — assuming they are isotropic, only
dependent upon r.
í Õ ! ¤. û
346 12 The geoid, mean sea level, and sea-surface topography
in location between sea level and the Earth’s solid body or Earth’s
crust. It is a function of place: one may not assume that it would be the
same everywhere. Mitrovica et al. (2001) show how, for example, the
meltwater from Greenland flees to the southern hemisphere, when the
meltwater from Antarctica again comes similarly to the north. This is
a consequence of the change in the Earth’s gravity field and the geoid,
when large volumes of ice melt. And also the physical shape of the Earth
changes when the ice load changes: glacial isostatic adjustment, GIA.
This also complicates the monitoring of the global mean sea level
from local measurements: the problem is familiar in Fennoscandia,
where the Earth’s crust, for now, is rising faster than the global sea level.
Green’s functions in the sea-level equation are functions of both ψ
and time difference ∆t. This tells us that GIA is a function of both place
and time. On a spherically symmetric Earth, the functions may be
written as expansions. See Wieczerkowski et al. (1999).
The elastic response of the Earth to loading is instantaneous on the
geological time-scale. It is described by similar elastic Love numbers as
those that appear in the theory of tidal deformation, for long (though
geologically short) periods P. See section 14.2. Like this:
GeV (ψ,∆t) Ge (ψ,∆t)
⏟ ⏞⏞ ⏟
⏟ r
⏞⏞ ⏟
1 ∑︂
∞
1 ∑︂
∞
Ges (ψ, ∆t) = γ · δ(∆t) kn Pn (cos ψ) − γ · δ(∆t) hn Pn (cos ψ),
n=0 n=0
í Õ ! ¤. û
Self-test questions
347
with the viscous Love numbers for potential and vertical displacement:
∑︂
I ∑︂
I
kvn (∆t) = rkni exp(−sni ∆t), hvn (∆t) = rh
ni exp(−sni ∆t).
i=1 i=1
^ Self-test questions
1. List all the causes of sea-level variations that you are aware of.
2. What is the sea-surface topography?
3. What is eustatic sea-level rise?
4. What is the origin of the name “El Niño”?
5. What is absolute, and what is relative land uplift? What does the
difference between the two consist of?
í Õ ! ¤. û
348 12 The geoid, mean sea level, and sea-surface topography
6. Which two main models are on offer for the mechanism of land
uplift?
7. Which three geodetic techniques are available for determining the
sea-surface topography?
8. What is the shape of the sea-surface topography of the Baltic Sea,
and what is its cause?
9. What is the Coriolis force, and how does it affect ocean currents?
10. What is the geostrophic balance?
11. In whose honour is the unit sverdrup named?
12. How can one invert a map of the sea-surface topography into a
map of ocean currents? Where on Earth does this not work?
13. What is the Peltier effect? What is the mid-Holocene highstand?
14. What does the sea-level equation describe?
15. How would sea-level equation 12.3 change if the convolution inte-
grals like equation 12.4 were over the unit sphere dσ = cos ϕ dϕ dλ
instead of over dω = R2 cos ϕ dϕ dλ?
16. Why does the mean sea level in the Baltic Sea not rise when the
Greenland continental ice sheet melts? What will happen in the
Baltic Sea when the West Antarctic ice sheet melts?
í Õ ! ¤. û
Exercise 12 –2: Land subsidence and the mechanism of land uplift
349
height difference between the left and the right edges.
3. (For fun) if the depth of the current is 1 km, what is the water
transport in sverdrup?
í Õ ! ¤. û
^ Satellite altimetry and satellite
gravity missions
– 351 –
352 13 Satellite altimetry and satellite gravity missions
í Õ ! ¤. û
Satellite altimetry 13.1
353
of the sea surface geocentrically. Together with its successors
Jason-1, 2 and 3 (2001-055A, 2008-032A, 2016-002A), this satellite
mission has also produced, and continues to produce, valuable
information on the global rise of the sea level over the last 25 years,
of about 3 mm per year. See figure 13.1.
The famous oceanographer Walter Munk2 characterised 2
TOPEX/Poseidon in 2002 as “the most successful ocean experiment
of all time” (Munk, 2002).
◦ Haiyang-2A (2011-043A) is a Chinese satellite also launched by
China.
◦ SARAL (2013-009A) is a satellite launched by India. The altimeter
AltiKa and DORIS are French contributions.
◦ CryoSat-2 (2010-013A) is a satellite launched by the ESA to study
polar sea ice. Of special interest is the freeboard, the amount by varalaita
which the ice sticks out of the water. From this, the thickness, and,
with the surface area, the total volume may be calculated. In-orbit
positioning is done using the French DORIS system.
The launch of CryoSat-1 failed.
◦ Sentinel-3A (2016-011A) is a versatile ESA remote-sensing satellite,
the first of a planned constellation. It carries several instruments,
among them the SRAL, or Synthetic Aperture Radar Altimeter. synteettinen
aukko
The measurement method of satellite radar altimetry is presented
in figure 13.2. The figure shows all the quantities playing a role in
altimetry: the measured range s is the height h of the satellite above
the reference ellipsoid corrected for the geoid height N, the sea-surface vertaus-
topography H, and variations of the sea surface, like tides, eddies, ellipsoidi
annual variation, and so on.
Furthermore, if the satellite does not contain a precise positioning
device, the true orbit of the satellite will differ from the calculated orbit
í Õ ! ¤. û
354 13 Satellite altimetry and satellite gravity missions
60
TOPEX/Poseidon, Jason satellites, global mean sea level, mm
20
Jason-3
TOPEX/Poseidon
Jason-2
0
Jason-1
−20
1998, 2016 Super El Niño
−40
5
−60
1990 1995 2000 2005 2010 2015 2020
Figure 13.1. Results from the TOPEX/Poseidon and Jason satellites. Annual
cycle removed. Data © Colorado University at Boulder’s Sea Level
Research Group; Nerem et al. (2010). Comparison with ENSO (“El
Niño”), SOI = Southern Oscillation Index, Climate Research Unit,
^ Climate Research Unit; Ropelewski and Jones (1987).
h = h0 + ∆h,
í Õ ! ¤. û
Satellite altimetry 13.1
355
True orbit
Calculated orbit
h s
Sea-surface
topography H Footprint
Geoid
Reference ellipsoid
Geoid height N
Sea surface Mean sea surface
í Õ ! ¤. û
356 13 Satellite altimetry and satellite gravity missions
the SWH is large, the altimeter footprint — the area on the sea from
which radio energy returns to the receiver — will also be larger, and the
distance travelled by the radio waves will on average be a little longer.
The newest satellites use an interferometric technique that differs
somewhat from the description above.
Of all the corrections related to instrumentation, atmosphere, ocean,
and solid Earth, we mention
1. the height of sea waves (SWH)
2. solid-Earth tides
3. ocean tides
4. the “wet” tropospheric propagation delay, best derived from
measurements with a downlooking water vapour radiometer on
the satellite, otherwise from meteorological modelling
5. the “dry” tropospheric propagation delay
6. the ionospheric delay, only for the part of the ionosphere below
the satellite, depending on flight height
7. the altimeter’s own calibration correction — nowadays “in-flight”
calibration is always strived for, using an ensemble of GNSS-
mareografi positioned tide gauges, see section 13.4.
The measurements and all corrections to be made to them are collected
into a “geophysical data record” (GDR), one per observation epoch. The
files built this way are distributed to researchers. This allows all kind of
experimentation; for example, the replacement of a correction by one
calculated from improved models.
í Õ ! ¤. û
Crossover adjustment 13.2
357
The observation equation is
s = h − N − H − ϵ + n = h0 + ∆h − N − H − ϵ + n,
in which s is the altimetric measurement of the height of the sea surface
(including the known corrections 1–7 in the previous section), h the
actual, and h0 the calculated height of the satellite above the reference
ellipsoid. N is the geoid height, H is the sea-surface topography: the
permanent deviation of the sea surface from an equipotential surface,
∆h is the orbit-error correction, ϵ is the residual variation of the sea
surface, the variation remaining after correcting for the tides and other
effects that can be modelled, and n is the random uncertainty, or noise,
in the radar altimetry observations.
From this we obtain in the crossing point of tracks i and j:
def (︁ i i
)︁ (︂ j j
)︂ (︁ )︁
ℓk = s − h0 − s − h0 = (∆hi − ∆hj ) − (ϵi − ϵj ) + ni − nj .
This is the observation equation of crossover adjustment. Here we see risteyskohta-
the complication that sea-surface residual variation and orbit corrections tasoitus
appear in the equation in the same way. They cannot be separately
determined by crossover adjustment.
If we forget for now the sea-surface residual variation — or assume
that it behaves randomly, in other words it is part of the noise n — we
may write more simply
def (︁ )︁
ℓk = ∆hi − ∆hj + nk , in which nk = ni − nj − (ϵi − ϵj ) .
The index k counts crossover points, the indices i and j count tracks.
Next, we choose a suitable model for the satellite orbit error. The
simplest choice, sufficient for a small area, is the assumption that the
orbit correction is a constant for each track. See a simple example,
figure 13.3.
í Õ ! ¤. û
358 13 Satellite altimetry and satellite gravity missions
∆h2 ∆h3
∆h2 − ∆h3
∆h1
∆h1 − ∆h3
Crossover 1
Crossover 2
2
1
ℓ1 = ∆h2 − ∆h3 + n1 ,
ℓ2 = ∆h1 − ∆h3 + n2 ,
3 or in matric form3
x
ℓ A ⏟ ⏞⏞ ⎤⏟ n
⏟[︄ ⏞⏞ ]︄⏟ ⏟[︄ ⏞⏞ ]︄⏟ ∆h1
⎡ ⏟[︄ ⏞⏞ ]︄⏟
ℓ1 0 1 −1 ⎢ n1
= ⎣ ∆h2 ⎦+ , (13.1)
⎥
ℓ2 1 0 −1 n2
∆h3
í Õ ! ¤. û
Crossover adjustment 13.2
359
symbolically
ℓ = Ax + n .
If one now tries to calculate the solution using ordinary least-squares,
)︁−1 T
x = AT A
(︁
ˆ︁ A ℓ,
this will not work. The normal matrix AT A is singular (check!). This
makes sense, as one can move the whole track network up or down
without the observations ℓk changing. No unique solution can be found
for such a system.
Finding a solution requires that something must be fixed: for example,
one track — or, more democratically, the mean level of all tracks. This
fixing is achieved by adding the following “observation equation”:
[︂ ]︂
def
ℓ3 = 0 = c c c · x , (13.2)
c c c
∆h
ˆ︂ 3 0
3 Notethe similarity with the observation equations for levelling! Instead of bench-
marks, we have tracks, instead of levelling lines, crossover points.
í Õ ! ¤. û
360 13 Satellite altimetry and satellite gravity missions
1 1 1 c
⎡ ⎤⎡ ⎤ ⎡ /︁ ⎤
−1 2 1 1 −1 2 1 c
1 ⎢ ⎥ 1⎢
−1 1 c ⎦ ,
/︁ ⎥
= 3 ⎣ 2 −1 1 ⎦ ⎣ 1 ⎦= 3⎣ 2
⎥⎢
1 c −1 1 c
/︁ /︁
−1 −1 1 −1
and the solution is
⎡ ⎤
∆h
ˆ︂ 1
⎣ ∆h2 ⎦ = A−1 ℓ =
⎢ ˆ︂ ⎥
∆h
⎡ /︁ ⎤ ⎡ ⎤ ⎡ ⎤
−1 2 1 c ℓ1 −1 2
ˆ︂ 3 [︄ ]︄
ℓ1
= 13 ⎣ 2 −1 1 c ⎦ ⎣ ℓ2 ⎦ = 1 ⎢
/︁ ⎥ ⎢
3 ⎣ 2 −1 ⎦ ,
⎢ ⎥ ⎥
ℓ2
−1 −1 1 c
/︁
0 −1 −1
from which c has vanished.
Another way to look at this is to first write the observation equations
13.1 and 13.2 together as
ℓ A x n
⏟⎡ ⏞⏞ ⎤⏟ ⏟⎡ ⏞⏞ ⎤⏟⎡
⏟ ⏞⏞ ⎤⏟ ⎡⏟ ⏞⏞ ⎤⏟
ℓ1 0 1 −1 ∆h1 n1
⎣ ℓ2 ⎦ = ⎣ 1 0 −1 ⎦⎣ ∆h2 ⎦ + ⎣ n2 ⎦,
⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥
0 c c c ∆h3 0
and then multiply the left-hand side and both terms on the right with
the diagonal matrix ⎡ ⎤
1 0 0
def ⎢
D = ⎣ 0 1 0 ⎦.
⎥
0 0 1 c
/︁
í Õ ! ¤. û
Crossover adjustment 13.2
361
The result is
Dℓ DA Dn
⏟⎡ ⏞⏞ ⎤⏟ ⏟⎡ ⏞⏞ ⎤⏟ ⎡ ⎤ ⏟⎡ ⏞⏞ ⎤⏟
ℓ1 0 1 −1 ∆h1 n1
⎣ ℓ2 ⎦ = ⎣ 1 0 −1 ⎦ ⎣ ∆h2 ⎦ + ⎣ n2 ⎦,
⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥
0 1 1 1 ∆h3 0
∆h = a + bτ,
where the parameter τ is the location along the track reckoned from its
starting point. The dimension of this location can be time (seconds) or
distance (degrees or kilometres). Now, the set of observation equations
for the situation described above is
x
⏟ ⏞⏞ ⏟
⎡ ⎤
a1
ℓ A n
⏟[︄ ⏞⏞ ]︄⏟ ⏟[︄ ⏞⏞ ]︄⏟⎢ ⏟[︄ ⏞⏞ ]︄⏟
⎢ b ⎥
1 ⎥
1 τ21 −1 −τ31 ⎢
⎢ ⎥
ℓ1 0 0 ⎢ a2 ⎥ + n1 .
⎥
=
ℓ2 1 τ12 3
0 0 −1 −τ2 ⎢ b2 ⎥⎢
⎥ n2
⎢ ⎥
⎣ a3 ⎦
b3
The design matrix A contains, besides the values 1 and −1, also values rakennematriisi
±τik , in which i is the number of the track and k that of the crossover
point. These values are computable when the geometry of the tracks is
known.
í Õ ! ¤. û
362 13 Satellite altimetry and satellite gravity missions
Now there are two unknowns for every track, a and b, a constant and
a trend. Of course also this system will prove to be singular. Removing
the singularity can be done by fixing all three parameters b and one
4 parameter a.4
4 Inorder to understand this, build, say, a three-track “wire-frame model” from pieces
of iron wire, tied together by pieces of string at the crossover points. The crossover
conditions do not in any way fix the values of the trends b, and the whole absolute
level of the frame continues to be unconstrained.
í Õ ! ¤. û
Crossover adjustment 13.2
363
crossover observation.
6 Andrey Nikolayevich Tikhonov (1906–1993) was a Russian mathematician and
geophysicist.
í Õ ! ¤. û
364 13 Satellite altimetry and satellite gravity missions
í Õ ! ¤. û
Choice of satellite orbit 13.3
365
^ 13.2.4 Global crossover adjustment
In a global crossover adjustment, often a still more sophisticated model
is used,
∆h = a + b sin τ + c cos τ, (13.3)
in which now τ is an angular measure, for example the place along the
track reckoned from the last south-north equator crossing or ascending
node. See Schrama (1989), where this problem is treated more extensively.
In this model, a represents the size of the orbit, while b and c denote
the offset of the centre of the orbit from the geocentre. This model
is three-dimensional: the orbital arcs with their crossovers form a
spherical network surrounding the Earth. The degrees of freedom left
by the crossover conditions are now the size of this sphere and the offset
of its centre from the geocentre:
7 One could argue that, in equation 13.3, the parameter a should be zero, as Kepler’s
third law allows a very precise determination of the orbital size, see section 13.3. Then,
also a0 = 0 in equation 13.4.
í Õ ! ¤. û
366 13 Satellite altimetry and satellite gravity missions
Z
Earth’s
rotation E
Perigee Nodal line
axis
Y
Nodal line
ν ω
ν Peri-
Satellite Apogee
i a gee
Ω ̇ b
Ω ω
X
θ
line
dal
Earth ea
X
si
rotation
Ap
X X ′, (Greenwich)
vernal equinox
Apogee Ascending node
8 Thisis why it is said that Henry Cavendish was the first to “weigh the Earth”. . . .
Determining GM⊕ was already straightforward back then using the orbital motion of
the Moon, or even gravity on the Earth’s surface. The challenge was separating G and
the mass of the Earth M⊕ from each other, obtaining the latter in ordinary units of
mass.
í Õ ! ¤. û
Choice of satellite orbit 13.3
367
using on-board thrusters, so that the satellite passes over the same rakettimoottori
place, for example once a day, after 14 orbital periods. Alternatively one
chooses an orbit that flies over the same place every third, seventeenth,
168th day. . . . This is called the repeat period.
The choice of the repeat period depends on the mission objective:
◦ If one wishes to study the precise shape of the mean sea surface,
one chooses a long repeat period, in order to get the tracks as close
together as possible on the Earth’s surface.
◦ If one wishes to study the variability of the sea surface, one chooses
an orbit that returns to the same location after a short time interval.
Then, the grid of tracks on the Earth’s surface will be sparser.
Parameters describing the figure of the Earth also affect satellite motion,
for example the quantity J2 , the dynamic flattening, having a value of
J2 = 1082.63 · 10−6 . It is the largest of the many spherical-harmonic pallofunktio-
coefficients that together represent the figure of the Earth and that affect kerroin
satellite orbits. In the case of J2 , the effect is that the plane of the satellite
orbit rotates at a certain angular rate around the Earth’s rotation axis:
orbital or nodal precession. This typically makes the satellite, if it flies
over the same location the next day, do so several minutes earlier. For a
circular orbit of radius a, the equation is
√︃
dΩ GM⊕ (︂ a⊕ )︂2
= − 32 a J2 cos i,
dt a3
in which a⊕ is the equatorial radius of the Earth reference ellipsoid, M⊕
the mass of the Earth, and i the inclination of the orbital plane relative
to the equator.
Substituting numerical values into this yields
dΩ cos i
= −1.318 95 · 1018 m3.5 s−1 · ,
dt (a⊕ + h)3.5
í Õ ! ¤. û
368 13 Satellite altimetry and satellite gravity missions
Attraction Solar
caused by Satellite
apparent
Earth’s orbital
daily motion
flattening motion
X
X
dΩ (︁ ◦
= − 1.331 03 · 10−6 rad/s · cos i = −6 . 589 day · cos i.
/︁ )︁
dt
For practical reasons — solar panels! — we often choose the satellite
orbit such that the orbital plane turns along with the annual apparent
◦/︁ ◦
motion of the Sun, 360 365.25 days = 0 . 9856 day . See figure 13.6.
/︁
9 If
the height of the orbit is less than 1400 km, it cannot be completely no-shadow. In
midwinter or in midsummer the satellite will then fly through the Earth’s shadow.
í Õ ! ¤. û
Choice of satellite orbit 13.3
369
Spring
Summer Winter
Autumn
Autumn
Winter
Summer
Morning hemisphere
Evening hemisphere
Spring
Figure 13.7. The geometry of a “no-shadow” orbit. In this figure, the satellite
flies north over places where it is morning and south over places
^ where it is evening.
unless the inclination is precisely 90◦ , there will be areas around both
poles that the satellite will never overfly: the “polar holes”.
A drawback of a Sun-stationary orbit is that the altimetric observations
are always made at the same local time of day. For example, the diurnal
and semidiurnal tides caused by the Sun will always have the same
phase angle (“resonance”), and thus they cannot be observed with a
satellite in this type of orbit. Therefore, the oceanographic satellite
TOPEX/Poseidon, and the follow-up Jason satellites, were placed in
non-Sun-stationary orbits.
^ 13.3.1 Example
A satellite moves in a Sun-stationary orbit, in other words, always, day
after day, flies over the same latitude at the same local mean solar time.
í Õ ! ¤. û
370 13 Satellite altimetry and satellite gravity missions
θ̇
i
3 2 1
Figure 13.8. A satellite in a retrograde orbit around the rotating Earth, crossing
the equator south to north three successive times. The angle
between the orbit and the equator, the inclination i, or for a
retrograde orbit, its supplement 180◦ − i, is also the highest
northern or southern latitude that the satellite can fly over. The
^ unreachable “polar holes” are indicated by blue dashed lines.
Questions
1. What is the period of the satellite if it always flies again over
the same spot after 14 revolutions?
2. The same question if the satellite always flies over the same
spot after 43 revolutions (3 days)?
3. And after 502 revolutions (35 days)?
4. What is the height of the satellite in a “three-day orbit”? Use
Kepler’s third law, equation 13.5. GM⊕ = 3 986 005·108 m3/s2 ,
and the height of the satellite is h = a − a⊕ , with a⊕ =
í Õ ! ¤. û
Choice of satellite orbit 13.3
371
6 378 137 m.
5. What is the satellite height in a “35-day orbit”? And the
height difference from the previous question?
6. What is, for the three-day orbit, the mean separation between
north-going orbital tracks (so, at what level of detail is the
altimeter able to image the sea surface!)?
7. The same question for a 35-day orbit.
8. Questions for reflection:
(a) For what purpose would you use a 35-day orbit, for
what purpose a three-day orbit?
(b) Would it be possible, or easy, to fly both orbits with the
same satellite (see question 5)?
Answers
1. The satellite completes 14 orbits per day, a day being 1440
minutes: P = 1440 min 14 = 102.857 min.
/︁
latitudes.
◦/︁
7. 360 502 = 0◦. 717, or 40 000 km 502 = 80 km.
/︁
í Õ ! ¤. û
372 13 Satellite altimetry and satellite gravity missions
format long
GM=3986005e8;
ae=6378137;
P=100.465*60;
fac=4*pi*pi;
a=(GM*P*P/fac)^0.33333333;
h = a - ae;
printf(’\n\nOrbital height: %8.3f km.\n’, h/1000);
8.
(a) The 35-day orbit would be excellent for detailed map-
ping. The three-day orbit would be able to see, for
example, tides or weather-related phenomena, albeit at
poorer spatial resolution.
(b) The difference in height being only 3 km and in period
4 s, the change in orbit between the two repeat periods
should be easily within reach of even small on-board
10 thrusters.10 So, yes.
10 Disclosure: the total velocity change needed is ∆v = 1.6 m/s , a brisk walking pace.
í Õ ! ¤. û
Retracking 13.5
373
Travel time
Half-
height
rule
Figure 13.9. Analysing the altimeter return pulse. The classical return pulse
^ time measurement uses the “half-height point”.
^ 13.5 Retracking
The results of a satellite altimetry mission are published already during
flight in the form of a geophysical data record (GDR) file, containing
everything related to the measurement, such as atmospheric correction
terms, tidal corrections, and sea-state parameters.
It is common practice today to re-process older altimetry measure-
ments, applying improved methodologies in order to extract additional
useful information. The complete return pulse is analysed again in an
approach called retracking (Altimetry, Retracking).
The method of analysis uses the point on the leading edge of the
return pulse which is at half the height of the maximum value of the
pulse. This is according to experience a good way to get the travel
time associated with the point at the centre of the footprint, directly
underneath the satellite. In the back part of the pulse are reflections
from the further-away peripheral areas of the footprint.
There are three situations where the automatic analysis technique
applied during flight does not work properly, and a more careful a
í Õ ! ¤. û
374 13 Satellite altimetry and satellite gravity missions
11 Thenewest satellites such as Sentinel-3 use a digital terrain model for steering the
tracking window when not over the open ocean.
12 Assuming that there is no snow on the ice. Also, ice density varies, and differs
between one-year and multi-year ice.
í Õ ! ¤. û
Oceanographic research using satellite altimetry 13.6
375
40
30
Ice volume (1000 km3 )
20
10
Year
0
1970 1980 1990 2000 2010 2020
Figure 13.10. Ice volume on the Arctic Ocean. PIOMAS; Schweiger et al.
(2011).
^
í Õ ! ¤. û
376 13 Satellite altimetry and satellite gravity missions
í Õ ! ¤. û
Satellite gravity missions 13.7
377
or geopotential; in other words, to determine a global high-resolution
model of the geoid.
^ 13.7.1 CHAMP
CHAMP (Challenging Minisatellite Payload for Geophysical Research
and Applications, 2000-039B) was a German satellite project under
the auspices of the German Research Centre for Geosciences GFZ. The
satellite was launched into orbit from Plesetsk, Russia, in 2000. The orbit
height was initially 454 km, coming down over the time of the mission
to some 300 km due to atmospheric drag. The orbital inclination was
87◦ . On 19th September, 2010, the satellite returned into the atmosphere.
Project description: CHAMP Mission.
CHAMP contained a GPS receiver in order to determine the satellite
location in space x(t) over time t. From successive satellite locations
one may calculate the geometric acceleration a(t) by differentiation:
d2
a(t) = x(t).
dt2
The differentiation is done numerically as presented in the part on
airborne gravimetry, equation 11.8.
The satellite also contained an accelerometer, which eliminated the kiihtyvyys-
satellite’s accelerations caused by the atmosphere’s aerodynamic forces, mittari
the deviations from free-fall motion. Then, only the accelerations caused
by the Earth’s gravitational field remain, from which a precise global
geopotential or geoid model may be calculated using the techniques
described earlier.
A number of global geopotential models based on CHAMP data have
been calculated and published.
^ 13.7.2 GRACE
GRACE (Gravity Recovery And Climate Experiment Mission, 2002-
012 A and B) measured temporal changes in the Earth’s gravity field
í Õ ! ¤. û
378 13 Satellite altimetry and satellite gravity missions
Solar
Earth internal mass
cells
density variations
(example)
Figure 13.11. Determining the Earth’s gravity field from GPS orbital tracking
^ of a low flying satellite.
í Õ ! ¤. û
Satellite gravity missions 13.7
379
Difference between line-of-sight accelerations
1 Satellite 2
Satellite Distance 220 km
Precise
ranging,
wavelength
45 0 km 1.5 cm
e ight
H
Figure 13.12. The principle of the GRACE satellites: measuring the minute
variations in time of the gravity field using SST, satellite-to-
satellite tracking. The changes are due to mass shifts in the
“blue film” — the atmosphere and hydrosphere — and expressed
^ as variations in “total sea-floor pressure” (↓).
^ 13.7.3 GOCE
GOCE (2009-013A, Geopotential and Steady-state Ocean Circulation
Explorer) was the most ambitious of all the satellites. Built by the
European Space Agency ESA, the satellite was launched successfully
í Õ ! ¤. û
380 13 Satellite altimetry and satellite gravity missions
from Plesetsk in March 2009. The orbital height was only 270–235 km
during the mission and the satellite contained an ionic rocket engine with
ajoaine a stock of propellant in order to maintain the orbit against atmospheric
13 drag. The orbital inclination was 96◦. 7, so the orbit was Sun-stationary.13
13 Because of this inclination angle, there was a cap of radius 6◦. 7 at each pole within
which no measurements were obtained. Over recent years these “poles of ignorance”
have been gradually filled in by airborne gravimetry campaigns, for example Forsberg
et al. (2017).
í Õ ! ¤. û
Satellite gravity missions 13.7
381
Measuring
acceleration
GOCE satellite
differences
1
5
4 X
3
6
2
Gradiometer
Accelerometer
Unknown
density variations
Figure 13.14. Determining the Earth’s gravity field with the gravitational
^ gradiometer on the GOCE satellite.
í Õ ! ¤. û
382 13 Satellite altimetry and satellite gravity missions
12.5 and figure 12.4. This is the background for the name of the GOCE
satellite.
^ Self-test questions
1. What is the footprint of a radar altimeter? How does it depend on
wave height?
2. What is the freeboard of ocean ice? How can it be used to determine
the volume of the ice?
3. What three alternative models for the satellite orbit-error correc-
tion exist?
4. What is, in satellite altimetry crossover adjustment, a datum defect,
and how can it be corrected?
5. How can Kepler’s third law be used to determine the mean height
of a satellite orbit if the satellite’s period is given?
6. What is the repeat period of a satellite orbit?
7. What is J2 , and how does it affect the motion of a satellite?
8. What is a Sun-synchronous orbit, and why is it useful?
9. What is a retrograde orbit?
10. Why are the orbits of the TOPEX/Poseidon and Jason satellites not
Sun-synchronous?
11. In table 13.1 some satellites have a repeat period that is an integer
number of days, some satellites do not. What do satellites with
non-integer repeat periods seem to have in common?
12. With which three satellite altimetric methods can one study sea-
surface variability?
13. Three satellite missions have been launched so far to study the fine
structure of the Earth’s gravity field and its temporal variability.
Present them and the methods used by them.
í Õ ! ¤. û
Exercise 13 –1: Altimetry, crossover adjustment
383
∆h = a + bτ,
í Õ ! ¤. û
384 13 Satellite altimetry and satellite gravity missions
GM⊕ P2 = 4π2 a3 ,
GM⊕ = 3 986 005 · 108 m3/s2 , and the height of the satellite is
h = a − a⊕ , in which a⊕ = 6 378 137 m.
2. What is the orbital inclination i of the satellite, if it is given that
the orbit is circular and Sun-synchronous? See section 13.3.
í Õ ! ¤. û
^ Tides, the atmosphere, and Earth
crustal movements
– 385 –
386 14 Tides, the atmosphere, and Earth crustal movements
Lo
Hi w
gh tid
t id e
e
R
X ζ ζ′
z
Hi
gh
Lo tid
w e ℓ
tid
e
d
Parallax
Figure 14.1. Theoretical tide. ζ ′ is the local zenith angle of the Moon (or Sun),
^ ζ the corresponding geocentric angle.
where z is the co-ordinate defined along the line connecting the Earth’s
centre and the attracting body, see figure 14.1. R is the radius of the
Earth, assumed spherical.
The net tidal potential as felt by mass elements on the Earth’s surface
is now
GM GM R 2
(︂ )︂
′ ′′
V =V −V = − cos ζ =
ℓ R d
GM ∑︂ R n+1
∞ (︂ )︂
GM R 2
(︂ )︂
= Pn (cos ζ) − cos ζ
R d R d
n=0
using expansion 8.7. Here, the term for n = 0 is constant and thus
arbitrary for a potential, and we delete it. The term n = 1 produces a
í Õ ! ¤. û
The theoretical tide 14.1
387
precise cancellation. Remains
GM ∑︂ R n+1
∞ (︂ )︂
V= Pn (cos ζ),
R d
n=2
or for n = 2,
1 Thedeclination of a celestial body is its latitude on the celestial sphere, its angular
distance from the celestial equator (Wikipedia, Declination), in this case as seen from
the geocentre.
2 Thehour angle is the angle, or difference in longitude, between the meridian of the
Moon and the local meridian, measured along the celestial equator (Wikipedia, Hour
angle), in this case as seen from the geocentre. It vanishes when the Moon is in upper
culmination, at its greatest elevation in the local meridian.
í Õ ! ¤. û
388 14 Tides, the atmosphere, and Earth crustal movements
and we obtain
3
+ 4
cos2 ϕ cos2 δ cos 2h.
From this
⎛(︁ )︁ ⎞
3 sin2 ϕ − 1 3 sin2 δ − 1 +
)︁ (︁
GMR2 ⎜⎜ + 3 sin 2ϕ sin 2δ cos h + ⎟ .
⎟
V= 3
4d ⎝ ⎠
2 2
+ 3 cos ϕ cos δ cos 2h
GMR2 (︁
3 sin2 ϕ − 1 3 sin2 δ − 1 ,
)︁ (︁ )︁
V1 = 3
4d
that still depends on the lunar declination δ and is therefore
periodic with a 14-day (half-month) period. Using spherical
trigonometry:
í Õ ! ¤. û
The theoretical tide 14.1
389
plane with respect to the equator, on average 23◦. 5 but varying
between 18◦. 3 and 28◦. 6. Thus we obtain
GMR2 (︁ )︁ (︂ (︁ 1 1 )︂
2 2
)︁
V1 = 3 sin ϕ − 1 3 sin ϵ 2 − 2 cos 2ℓ − 1 ,
4d3
where we have used result 14.1. We split V1 = V1a + V1b into two
parts, a constant3 and a periodic, semi-monthly (“fortnightly”) 3
part:
GMR2 (︁
3 sin2 ϕ − 1 23 sin2 ϵ − 1 ,
)︁ (︁ )︁
V1a = 3
(14.2)
4d
GMR2 (︁
3 sin2 ϕ − 1 23 sin2 ϵ cos 2ℓ .
)︁ (︁ )︁
V1b =− 3
4d
í Õ ! ¤. û
390 14 Tides, the atmosphere, and Earth crustal movements
^ Table 14.1. The various periods in the theoretical tide. The widely used
symbols were standardised by George Darwin.
a Lunar fortnightly
b Solar semi-annual
The value for the Moon equals D$ = 26.8 cm × γ and for the Sun
D⊙ = 12.3 cm × γ, with γ ≈ 9.81 m/s2 . See figure 14.2.
6 The periods are listed in table 14.1 with their Darwin6 symbols.
In practice, the diurnal and semidiurnal tides can be divided further
into many “spectral lines” close to each other, also because the lunar
orbit, like the Earth’s orbit, is significantly eccentric.
í Õ ! ¤. û
Deformation caused by the tidal potential 14.2
391
V1a , permanent V1b , fortnightly, ϵ = 23◦
80 0.2 80 0.1
40 0 40 0.05
Latitude ( ◦ )
−0.2 0
0 0
−0.05
−40 −0.4 −40
−0.1
−80 −0.6 −80
−0.15
Obliquity ϵ, 0◦ − 90◦ Lunar longitude, 0◦ − 360◦
X
V2 , diurnal, δ = 23◦ V3 , semidiurnal, δ = 23◦
80 80 0.6
0.4
0.4
40 0.2 40
Latitude ( ◦ )
0.2
0 0
0 0
−0.2 −0.2
−40 −0.4 −40 −0.4
−0.6
−80 −0.6 −80 −0.8
0 5 10 15 20 0 5 10 15 20
Hour angle Hour angle
Figure 14.2. The main components of the theoretical tide. These values must
^ still be multiplied by Doodson’s constant D.
7 Augustus Edward Hough Love FRS (1863–1940) was a British mathematician and
student of Earth elasticity.
í Õ ! ¤. û
392 14 Tides, the atmosphere, and Earth crustal movements
1 ∑︂ ∑︂
∞ ∞
int
ur (ϕ, λ, r) = γ Hn (r) Vn (ϕ, λ, r) = Hn (r) ζn (ϕ, λ, r),
n=2 n=2
1 ∑︂
∞
∂Vnint (ϕ, λ, r) ∑︂
∞
uϕ (ϕ, λ, r) = γ Ln (r) =r Ln (r) ξn (ϕ, λ, r),
∂ϕ
n=2 n=2
1 ∑︂
∞
∂Vnint (ϕ, λ, r) ∑︂
∞
uλ (ϕ, λ, r) = γ Ln (r) =r Ln (r) ηn (ϕ, λ, r).
cos ϕ ∂λ
n=2 n=2
Here, r is the distance from the geocentre. It is assumed here that the
Love numbers Hn and Ln depend only on r, so that the elastic properties
of the Earth are spherically symmetric. The symbols ζn , ξn and ηn
represent the effect of the tidal potential of harmonic degree n on the
level of an equipotential surface and on the components of the direction
luotiviiva of the plumb line.
epäsuora The deformation of the Earth also causes a change, the “indirect effect”
vaikutus
í Õ ! ¤. û
Deformation caused by the tidal potential 14.2
393
in addition to the Moon’s original tidal potential V, in the geopotential.
We write
∑︂
∞
δV(ϕ, λ, r) = Kn (r) Vnint (ϕ, λ, r),
n=2
Because of the large distances to Sun and Moon, the only non-negligible
part of the tidal potential V is the part for the degree number n = 2, the
“rugby-ball part” V2int .
The Love numbers will still depend on the frequency, or on the tidal
period P:
(︁ )︁ (︁ )︁ (︁ )︁
hn = hn P , ℓn = ℓn P , k n = kn P .
The tides offer an excellent means of determining all these Love numbers
(︁ )︁ (︁ )︁ (︁ )︁
h2 P , ℓ2 P , and k2 P empirically, because, being periodic variations,
they cause Earth deformations at the same periods, but with different
amplitudes and phase angles.9 In this way we may determine at 9
least those Love numbers that correspond to periods occurring in the
theoretical tide.
The numbers h and ℓ are nowadays obtained for example by GNSS
positioning. The GNSS processing software contains a built-in reduction
for this phenomenon. From gravity measurements one obtains infor-
mation on a certain linear combination of h and k, δ = 1 + h − 32 k: the
lunar tidal force changes gravity directly, vertical displacement changes
gravity though its gradient, and deformation of the Earth, the shifting
of masses, also changes gravity directly.
The long water-tube tilt meter is also a useful research instrument,
like the instrument of the Finnish Geodetic Institute that has long been
9 The phase angles may be represented by making the Love numbers complex.
í Õ ! ¤. û
394 14 Tides, the atmosphere, and Earth crustal movements
10 Alsocalled the Shida number. Toshi Shida (1876–1936) was a Japanese Earth tide
researcher.
í Õ ! ¤. û
Tidal corrections between height systems 14.4
395
practically impossible to know. See Poutanen et al. (1996).
More generally, the reduction of a geodetic quantity, for example the
height of the geoid, for the permanent part of the tide can be carried
out in three different ways:
◦ No reduction whatever is made for the permanent part. The
quantity thus obtained is called the “mean geoid”. The surface
obtained is in the hydrodynamic sense an equilibrium surface,
directly suitable for use in oceanography.
◦ The direct effect of the tidal field of Sun and Moon is removed in its
entirety from the quantity, but the effect of the Earth’s deformation
caused by it is left uncorrected. The quantity thus obtained is
called the “zero geoid”.
◦ Both the effect of the tidal field of a celestial body, and the effect of
the deformation it causes, can be calculated according to a certain
deformation model (Love numbers), and removed. The result
obtained is called the “tide-free geoid”. Its problem is, as explained,
the empirical indeterminacy of the elasticity model used.
See figure 14.3. It is good to be critical and precisely analyse the way
in which the data reduction has been done!
GMR2 (︁ 2
)︁ (︁ 3 2
)︁
Vperm = 3 sin ϕ − 1 sin ϵ − 1 ≈
4d3 2
3GMR2 (︁ 2 1
)︁
≈ −0.7615 · sin ϕ − .
4d3 3
With the combined Doodson’s constant 14.3 for Sun and Moon equal to
3GM⊙ R2 3GM$ R2
D= + =
4d3⊙ 4d3$
í Õ ! ¤. û
396 14 Tides, the atmosphere, and Earth crustal movements
we obtain
(︁ 1
− sin2 ϕ × γ.
)︁
Vperm = 29.77 cm × 3
We can express this, with Bruns equation 5.2, as a permanent tidal geoid
effect:
Nperm = 29.77 cm × 13 − sin2 ϕ .
(︁ )︁
From this, Nperm (0◦ ) = 9.92 cm on the equator, and Nperm (±90◦ ) =
−19.85 cm on the poles.
This, the geoid effect of the permanent part of the external potential
of the Sun and Moon, is also equal to the difference between the mean
geoid and the zero geoid as defined above:
def (︁ 1
∆mean − sin2 ϕ .
)︁
zero N = Nmean − Nzero = 29.77 cm × 3
í Õ ! ¤. û
Loading of the Earth’s crust by sea and atmosphere 14.5
397
and for two different latitudes ϕ1 and ϕ2 we have for the effect on the
height difference
∆mean mean
(︁ 2 2
)︁
zero H(ϕ2 ) − ∆zero H(ϕ1 ) = 29.77 cm × sin ϕ2 − sin ϕ1 .
∆mean mean
(︁ 2 2
)︁
tidefree H(ϕ2 ) − ∆tidefree H(ϕ1 ) = 20.84 cm × sin ϕ2 − sin ϕ1 .
∆zero zero
(︁ 2 2
)︁
tidefree H(ϕ2 ) − ∆tidefree H(ϕ1 ) = −8.93 cm × sin ϕ2 − sin ϕ1 .
í Õ ! ¤. û
398 14 Tides, the atmosphere, and Earth crustal movements
^ Self-test questions
1. Present in words the three components of the theoretical tide
produced by the Laplace decomposition method.
2. How may the slowly varying part of the theoretical tide be further
decomposed into two parts? Present the parts in words.
3. What are the declination and hour angle of a celestial body, for
example the Moon?
í Õ ! ¤. û
Exercise 14 –1: The permanent tide
399
4. What is Doodson’s constant?
5. What do Love numbers express?
6. Why is it not possible to empirically determine the deformation
caused by the permanent part of the tide?
7. Present the three different ways to take the permanent part of the
tide into account when defining the geoid.
GMR2 (︁
3 sin2 ϕ − 1 32 sin2 ϵ − 1 ,
)︁ (︁ )︁
V1a = 3
4d
in which ϕ is latitude and ϵ is the obliquity of the Earth’s axis of rotation,
currently about 23◦. 5.
1. For what value ϕ does the permanent part of the tide vanish?
What is your interpretation?
2. For what value ϵ does the permanent part of the tide vanish?
What is your interpretation?
í Õ ! ¤. û
^ Earth gravity field research
15
^ 15.1 Internationally
In the framework of the IAG, the International Association of Geodesy,
research into the Earth’s gravity field is currently the responsibility of
the International Gravity Field Service (IGFS). The IGFS was created in
2003 at the IUGG General Assembly in Sapporo, Japan, and it operates
under the IAG’s new Commission 2 “Gravity Field”. The United States
National Geospatial-Intelligence Agency (NGA) serves as its technical
centre.
An important and well-reputed IAG service is the International Gravity
Bureau, the BGI, Bureau Gravimétrique International located in Toulouse,
France (http://bgi.obs-mip.fr/). The bureau works as an international
broker to which countries can submit their gravimetric materials. If a
researcher needs gravimetric material from another country, for example
in order to do a geoid computation, they can request it from the BGI, who
will provide it with the permission of the country of origin, provided the
country of the researcher has in its turn submitted its own gravimetric
materials for BGI use.
The French state has invested significant funds into this vital interna-
tional activity.
Another important IAG service in this field is the ISG, the International
Service for the Geoid. It has in fact been operating since as early as
– 401 –
402 15 Earth gravity field research
1992 under the name International Geoid Service (IGeS), the executive
arm of the International Geoid Commission (IGeC). The ISG office is
located in Milan, Italy (http://www.isgeoid.polimi.it/). The task of this
service is to support geoid determination in different countries. Existing
geoid solutions are collected into a common database, and international
research schools are organised to develop awareness about and skills in
the art of geoid computation, especially in developing countries. The
Italian state has provided significant funding for these activities.
Both services, BGI and ISG, are under the auspices of the IGFS, as two
of the many official services of the IAG. Other IGFS services include the
International Center for Earth Tides (ICET), the International Center for
Global Earth Models (ICGEM), and the International Digital Elevation
Model Service (IDEMS).
^ 15.2 Europe
The EGU, the European Geosciences Union, operates in Europe, co-
ordinating many publication and meeting activities relating to the
gravity field and geoid. The EGU organises annual symposia, where
sessions are always also included on subjects related to the gravity
field and geoid. American scientists also participate. Conversely the
1 American Geophysical Union’s (AGU) fall and spring meetings1 are also
1 Fall
(autumn) meetings are in San Francisco, spring meetings somewhere in the
world. The AGU, while American, is a very cosmopolitan player.
í Õ ! ¤. û
The Nordic countries 15.3
403
^ 15.3 The Nordic countries
In the Nordic countries, important work is being co-ordinated by the
NKG, the Nordiska Kommissionen för Geodesi, and its Working Group for
Geoid and Height Systems. Its activities include geoid determination,
studying the preconditions for still more precise geoid models, new
levelling technologies, and the study of post-glacial land uplift.
The group has for a long time computed high-quality geoid models at
its computing centre in Copenhagen, the next to last one being NKG2004
(Forsberg and Kaminskis, 1996; Forsberg and Strykowski, 2010). The
newest model, NKG2015, is the result of calculations by the computing
centres of several countries, including Sweden and Estonia. It was
published in October 2016.
^ 15.4 Finland
In Finland the study of the Earth’s gravity field has mainly been in
the hands of the Finnish Geodetic Institute, founded in 1918, one year
after Finnish independence. The institute has been responsible for
the national fundamental levelling and gravimetric networks and their
international connections. In 2001 the Finnish Geodetic Institute’s
gravity and geodesy departments were joined into a new department
of geodesy and geodynamics, to which gravity research also belongs.
Among topics studied are solid-Earth tides, the free oscillations of
the solid Earth, post-glacial land uplift, and vertical reference or height korkeus-
systems. järjestelmä
Geoid models have been computed all the time, starting with Hir-
vonen’s global model (Hirvonen, 1934) and ending, for now, with the
Finnish model FIN2005N00 (Bilker-Koivula, 2010). These geoid models
are actually based on the Nordic NKG2004 gravimetric geoid, and are
fitted to a Finnish set of GNSS levelling control points, serving as a
transformation surface for heights.
In 2015, the Finnish Geodetic Institute was merged into the National
í Õ ! ¤. û
404 Earth gravity field research
Land Survey as its geospatial data centre and research facility. The
English-language acronym continues as FGI, the Finnish Geospatial
Research Institute (https://www.maanmittauslaitos.fi/en/research).
Helsinki University of Technology (today part of Aalto University)
has also been active in research on the Earth’s gravity field. Heiskanen,
a professor at HUT in 1928–1949, acted in 1936–1949 as the director
of the International Isostatic Institute. After moving to Ohio State
University, he worked with many other, including Finnish and Finnish-
born, geodesists on calculating the first major global geoid model, the
“Columbus geoid” (Kakkuri, 2008).
^ 15.5 Textbooks
There are many good textbooks on the study of the Earth’s gravity field.
In addition to the already mentioned classic, Heiskanen and Moritz
(1967), which is in large part obsolete, we may mention Wolfgang Torge’s
book (1989). Moritz (1980) is difficult but good. Similarly difficult is
Molodensky et al. (1962). Worth reading also from the perspective of
physical geodesy is Vaníček and Krakiwsky (1987). A newer book in
the field is Hofmann-Wellenhof and Moritz (2006).
í ¤. û
^ Field theory and vector calculus
— core knowledge
– 405 –
406 A Field theory and vector calculus — core knowledge
í Õ ! ¤. û
Vector calculus A.1
407
^ A.1.3 The exterior or vectorial product
The exterior product, or cross product, of two vectors is itself a vector
called the vectorial product, at least in three-dimensional Euclidean space.
For example, the angular momentum L:
⟨︁ ⟩︁
L= r×p ,
í Õ ! ¤. û
408 A Field theory and vector calculus — core knowledge
⟨︁ ⟩︁
x= a×b
α
∥x∥
b
a
í Õ ! ¤. û
Scalar and vector fields A.2
409
Angular momentum
⟨︁ ⟩︁
r × ṙ
Planet
Sun
ṙ
1⃦
⃦⟨︁ ⟩︁⃦
r × ṙ ⃦
2 Velocity
Radius vector vector
r
Figure A.2. Kepler’s second law. In the same amount of time, the radius vector
of a planet will “sweep over” a same-sized area — conservation
^ of angular momentum.
So: the quantity on the left-hand side, angular momentum L per unit
of mass m, equation A.2, is conserved:
⟨︁ ⟩︁ L
r × ṙ = m .
Like, for example, the total amount of energy, electric charge and many
other quantities, the amount of angular momentum in a closed system
is also constant.
^ A.2.1 Definitions
In the Euclidean space we may define functions or fields.
í Õ ! ¤. û
410 A Field theory and vector calculus — core knowledge
i ⊥ j, i ⊥ k, j⊥k
means that
⟨︁ ⟩︁ ⟨︁ ⟩︁ ⟨︁ ⟩︁
i · j = i · k = j · k = 0. (A.6)
Orthonormality means in addition that
a = a1 i + a2 j + a3 k,
and scalar and vectorial products can now also be calculated with the
aid of their components:
⟨︁ ⟩︁ ⟨︁ ⟩︁
s = a · b = (a1 i + a2 j + a3 k) · (b1 i + b2 j + b3 k) =
∑︂
3
= a1 b1 + a2 b2 + a3 b3 = ai bi ,
i=1
using the identities stated above for the basis vectors A.6 and A.7.
í Õ ! ¤. û
Scalar and vector fields A.2
411
For the vectorial product, the calculation is more involved. For
orthogonal vectors, the angle α in equation A.3 is 90◦ , so
⃦⟨︁ ⟩︁⃦ ⃦⟨︁ ⟩︁⃦ ⃦⟨︁ ⟩︁⃦
⃦ i × j ⃦ = ⃦ i × k ⃦ = ⃦ j × k ⃦ = 1.
b1 b2 b3
= (a2 b3 − a3 b2 ) i + (a3 b1 − a1 b3 ) j + (a1 b2 − a2 b1 ) k.
So
c1 = a2 b3 − a3 b2 , c2 = a3 b1 − a1 b3 , c3 = a1 b 2 − a2 b 1 .
r = xi + yj + zk,
í Õ ! ¤. û
412 A Field theory and vector calculus — core knowledge
∂V ∂V ∂V
g = grad V = ∇V = i +j +k .
∂x ∂y ∂z
í Õ ! ¤. û
Scalar and vector fields A.2
413
^ Figure A.3. The gradient. The level curves of the scalar field are dashed blue.
a1 a2 a3
⎡ ⎤ ⎡ ⎤
∂ ∂ [︄
∂ ∂
]︄ ∂ ∂
= det ⎣ ∂y ∂z ⎦ i − det ∂x ∂z j + det ⎣ ∂x ∂y ⎦ k =
a2 a3 a1 a3 a1 a2
(︃ )︃ (︃ )︃
∂a3 ∂a2 ∂a1 ∂a3 ∂a2 ∂a1
(︂ )︂
= − i+ − j+ − k,
∂y ∂z ∂z ∂x ∂x ∂y
í Õ ! ¤. û
414 A Field theory and vector calculus — core knowledge
í Õ ! ¤. û
Scalar and vector fields A.2
415
^ Figure A.5. The curl. Positive (anticlockwise) and negative (clockwise) eddies.
∂ ∂ ∂ ∂
b1 = V− V = 0,
∂y ∂z ∂z ∂y
∂ ∂ ∂ ∂
b2 = V− V = 0,
∂z ∂x ∂x ∂z
∂ ∂ ∂ ∂
b3 = V− V = 0,
∂x ∂y ∂y ∂x
thus
b = curl a = 0 !
In other words, if the vector field a(x, y, z) is the gradient of the scalar
field V(x, y, z), its curl will vanish:
⟨︁ ⟩︁ ⟨︁ ⟩︁
curl grad V = ∇ × ∇V = ∇ × ∇ V = 0,
í Õ ! ¤. û
416 A Field theory and vector calculus — core knowledge
then also
a(x, y, z) = grad (V(x, y, z) + V0 ) ,
with V0 an arbitrary constant, because
∂V0 ∂V ∂V
grad V0 = i + j 0 + k 0 = 0.
∂x ∂y ∂z
So the potential is not uniquely defined.
a = grad V = ∇V,
∂2 ∂2 ∂2
+ 2 + 2 = ∇ · ∇ = ∇2 .
⟨︁ ⟩︁
∆= 2
∂x ∂y ∂z
í Õ ! ¤. û
Integrals A.3
417
When operating on the potential of a “source free” field — for example
the gravitational potential in a vacuum or the electrostatic potential in
an area of space free of electric charges — the result of this Delta, or
Laplace, operator vanishes.
^ A.3 Integrals
from which one obtains the integral form, the work integral
w B ⟨︁ ⟩︁
∆EAB = F · dr .
A
Here, the amount of work done by a body moving from point A to point
⟨︁ ⟩︁
B is computed by integrating F · dr along the path AB.
If we parametrise the path according to arc length s, and the tangent
vector to the path is called
def dx dy dz
t= i+ j + k,
ds ds ds
we may also write
w B ⟨︁ ⟩︁
∆EAB = F · t ds,
A
the parametrised version of the integral.
í Õ ! ¤. û
418 A Field theory and vector calculus — core knowledge
vector field, the projection of a onto the normal vector of the surface,
the vector perpendicular to the surface in the surface element dS.
Let the normal vector on the surface be n. Then we must integrate
x ⟨︁ ⟩︁
a · n dS,
S
symbolically written x ⟨︁ ⟩︁
a · dS ,
S
í Õ ! ¤. û
Integrals A.3
419
Tangent vector t
curl a
curl a
z Integral Closed
s ⟨︁ Integral ⟩︁
⟨︁ ⟩︁
a · t ds path ∂S
∂S
S curl a · n dS
curl a
with r the location vector of the edge curve. The parametrised form of
the theorem is
x ⟨︁ ⟩︁ z ⟨︁ ⟩︁
curl a · n dS = a · t ds,
S ∂S
with n is the normal to surface S and t the tangent vector of edge curve
∂S.
In words The surface integral of the curl of a vector field over a surface
is the same as the closed path integral of the field around the
edge of the surface.
Special case For a conservative vector field a it holds that curl a = 0
everywhere. Then z ⟨︁ ⟩︁
a · dr = 0,
∂S
1 William Thomson, Lord Kelvin PRS FRSE (1824–1907) was a British physicist, mathe-
matician, engineer and inventor, and of course best known for the absolute temperature
scale proposed by him in 1848.
í Õ ! ¤. û
420 A Field theory and vector calculus — core knowledge
so also w B ⟨︁ ⟩︁ w B ⟨︁ ⟩︁
a · dr = a · dr .
A A
path 1 path 2
The work integral from point A to point B does not depend on the path
chosen. And the work done by a body transported around a closed path
is zero.
í Õ ! ¤. û
The continuity of matter A.4
421
n
∂V
div a V
Figure A.7. The Gauss divergence theorem. n is the normal vector to the
exterior surface. The Gauss divergence theorem can also be
presented with the aid of (Michael Faraday’s) field lines: a field line
starts or terminates on an electric charge (a place where div a ̸= 0)
^ or runs to infinity (through the surface ∂V).
í Õ ! ¤. û
422 A Field theory and vector calculus — core knowledge
over time. The two terms must balance for the “matter accounting” to
close.
If the moving fluid is incompressible, then ρ is constant:
d
ρ=0 =⇒ div(ρv) = ρ div v = 0 =⇒ div v = 0.
dt
í Õ ! ¤. û
^ Function spaces
∑︂
3
r = r1 e1 + r2 e2 + r3 e3 = ri ei .
i=1
Precisely because three basis vectors (not in the same plane) are always
enough, we call the ordinary (Euclidean) space three-dimensional.
In a vector space one can define a scalar product, which is a linear
mapping from two vectors to one number (“bilinear form”):
⟨︁ ⟩︁
r·s .
– 423 –
424 B Function spaces
⟨︁ ⟩︁
kantavektori If the basis vectors are orthogonal to each other, in other words, ei ·ej =
0 if i ̸= j, we may calculate the coefficients ri in a simple way:
∑︂
3 ⟨︁ ⟩︁ ⟨︁ ⟩︁
r · ei r · ei
r= ri ei , ri = ⟨︁ ⟩︁ = 2
. (B.1)
i=1
e i · e i ∥e i ∥
If, in addition,
{︁ }︁
ei · ei = ∥ei ∥2 = 1,
⟨︁ ⟩︁
i ∈ 1, 2, 3 ,
in other words, the basis vectors are orthonormal, equation B.1 becomes
simpler still:
∑︂
3
⟨︁ ⟩︁
r= ri ei , r i = r · ei . (B.2)
i=1
^ B.2.1 Description
Functions can also be considered elements in a vector space. If we define
1 the scalar product of two functions f and g as the following integral1
⟩︁ ⟨︂→− → w 2π
− def 1
⟨︁ ⟩︂
f·g = f · g = π f(x) g(x) dx, (B.3)
0
it is easily verified that the above requirements for a scalar product are
met.
kanta One basis in this vector space (a function space) is formed by the Fourier
1 The arrows over the function designators f and g here are just psychology: the
functions are really “vectors”. These arrows are not normally used.
í Õ ! ¤. û
The Fourier function space B.2
425
basis functions,
1√
e0 = 2,
2
ek = cos kx, k = 1, 2, 3, . . . , (B.4)
e−k = sin kx, k = 1, 2, 3, . . . .
1 √ ∑︂ ∞
f(x) = a0 2 + (ak cos kx + bk sin kx) ,
2
k=1
This is the familiar way in which the coefficients of a Fourier series are
calculated.
^ B.2.2 Example
As an example of Fourier analysis, we may take a step function on the
[︁ )︁
interval 0, 2π : ⎧
⎨0 x ∈ [︁0, π)︁,
f(x) =
⎩1 x ∈ [︁π, 2π)︁.
í Õ ! ¤. û
426 B Function spaces
def √ ∑︂
K
2 ∑︂ 1
K
f(K) (x) = 12 a0 2 + bk sin kx = 1
−π sin kx, (B.5)
2 k
k=1 k=1
odd
^ B.2.3 Convergence
suppeneminen The Fourier expansion converges in the square integral sense: if we
define the truncated expansion
def 1 √ ∑︂
K
(K)
f (x) = 2 a0 2 + (ak cos kx + bk sin kx) ,
k=1
í Õ ! ¤. û
The Fourier function space B.2
427
K=1
K = 25 K=3 K=5
1.0 f(x)
0.8 a0
0.6
X
0.4 a b
0.2 b5
b3
0 f(x)
b1
−0.2
0 1 2 3 4 5 6
Figure B.1. Fourier analysis on a step function. Plotted are the truncated
Fourier expansions f(K) (x), equation B.5, for values of K of 1, 3, 5,
^ and 25. The inset gives the spectrum of the function.
then w 2π (︁
1 )︁2
lim π f(K) (x) − f(x) dx = 0.
K→∞ 0
This does not mean that for an arbitrarily small value ε it holds that
⃓f (x) − f(x)⃓ < ε for every x ∈ 0, 2π , when K → ∞. Looking at
⃓ (K) ⃓ [︁ )︁
Also, note the “shoulder” of the expansion, even for K = 25. This
í Õ ! ¤. û
428 B Function spaces
shoulder will get narrower for higher K, but not any lower, remaining
Gibbsin ilmiö at approximately 0.09. This is known as the Gibbs phenomenon.
Lx − λx = 0,
∑︂
n
x= xi ei ,
i=1
On the other hand, we may write n different vectors Lei on the basis
{︁ }︁
ej in the following way:
∑︂
n
Lei = aij ej , i = 1, . . . , n.
j=1
This defines the coefficients aij , which may be collected into a size n × n
matrix A.
Now substitution yields
∑︂
n ∑︂
n ∑︂
n (︃∑︂
n )︃
Lx = xi · aij ej = aij xi ej . (B.6)
i=1 j=1 j=1 i=1
í Õ ! ¤. û
Sturm–Liouville differential equations B.3
429
Also
∑︂
n ∑︂
n
λx = λ xi ei = (λxj ) ej . (B.7)
i=1 j=1
Ax − λx = 0, (B.8)
í Õ ! ¤. û
430 B Function spaces
or
∑︂
n (︃∑︂
n )︃ ∑︂
n (︃∑︂
n )︃
xi aij yj = aij xj yi ,
i=1 j=1 i=1 j=1
aij = aji , i, j ∈ 1, . . . , n, or A = AT .
In other words,
Lxp = λp xp ,
If L is self-adjoint, then
⟨︁ ⟩︁ ⟨︁ ⟩︁ ⟨︁ ⟩︁ ⟨︁ ⟩︁ ⟨︁ ⟩︁
xq · Lxp = Lxq · xp = xp · Lxq =⇒ λ p xq · xp = λ q xp · xq .
í Õ ! ¤. û
Sturm–Liouville differential equations B.3
431
It follows that
⟨︁ ⟩︁
(λp − λq ) xp · xq = 0.
Remember that the scalar product is symmetric. If λp ̸= λq , we thus
⟨︁ ⟩︁
must have xp · xq = 0, or xp ⊥ xq , what was to be proven.
Example The variance matrix of location in the plane. The variance matrix
of the co-ordinates of point P in the plane is
{︄[︄
{︁ }︁
]︄}︄ [︄ ]︄
xP σ2x σxy
Var xP = Var = ΣPP = ,
yP σxy σ2y
a symmetric matrix. Here, σ2x and σ2y are the variances, or squares
of the mean errors, of the x and y co-ordinates, and σxy is the
covariance between the co-ordinates.
The eigenvalues of this matrix ΣPP are the solutions of the charac-
teristic equation
[︄ ]︄
(︁ )︁ σ2x − λ σxy
det ΣPP − λI = det = 0,
σxy σ2y − λ
or
(︁ 2
σx − λ σ2y − λ − σ2xy = 0.
)︁ (︁ )︁
This yields
1
(︁ 2 2
)︁ 1 √︂(︁ )︁2 (︁ )︁
λ1,2 = 2
σx + σy ± 2 σ2x + σ2y − 4 σ2x σ2y − σ2xy =
1
(︁ 2 2
)︁ 1 √︂(︁ )︁2
= 2 σx + σy ± 2 σ2x − σ2y + 4σ2xy .
í Õ ! ¤. û
432 B Function spaces
d2
x(t) + ω2 x(t) = 0. (B.9)
dt2
The solution has the general form (α amplitude, ϕ phase constant)
d ⃓ d ⃓
(︁ )︁ ⃓ ⃓
x(0) = x T , x⃓ = x⃓ .
dt x=0 dt x=T
These boundary conditions are an essential part of being self-adjoint.
Then, a solution is found only for certain values of ω — quantisation.
Equation B.9 is an eigenvalue problem, form-wise:
Lx + ω2 x = 0,
í Õ ! ¤. û
Sturm–Liouville differential equations B.3
433
it holds that (integration by parts):
⟩︁ w T
]︃T w
d2 y(t)
[︃
⟨︁→
− →
− dy(t) T dx(t) dy(t)
x ·Ly = x(t) dt = x(t) − dt,
0 dt2 dt 0 0 dt dt
⟩︁ w T d2 x(t)
[︃ ]︃T w
⟨︁ →
− →
− dx(t) T dx(t) dy(t)
Lx · y = 2
y(t)dt = y(t) − dt.
0 dt dt 0
0 dt dt
As, on the right-hand side, the first terms vanish and the second terms
are identical, it follows that
⟨︁→
−
x · L→
−
y = L→
⟩︁ ⟨︁ − →
x ·−
⟩︁
y ,
2πkt 2πkt
sin ωk t = sin , cos ωk t = cos .
T T
Any linear combination of these is a valid solution as well, and is of the general form
B.10.
5 JacquesCharles François Sturm FRS FAS (1803–1855) was a French mathematician,
one of the 72 names engraved on the Eiffel Tower. Eiffel Tower, 72 names.
í Õ ! ¤. û
434 B Function spaces
together form a complete basis for this vector space in such a way that
every function can be written as an — if necessary infinite — linear
combination of these basis functions. The situation is analogous to three-
dimensional space, where a complete basis consists of three vectors not
in the same plane.
í Õ ! ¤. û
Spherical harmonics B.5
435
An alternative, more compact way of writing this is
⎧
⎨P (sin ϕ) cos mλ if m ⩾ 0,
nm
Ynm (ϕ, λ) =
⎩Pn|m| (sin ϕ) sin |m| λ if m < 0,
see Heiskanen and Moritz (1967, equation 1-69). Proving this orthogo-
nality is not straightforward.
If we now divide the functions Ynm (or, equivalently, Rnm , Snm ) by
the square roots of the above factors, we obtain the fully normalised
surface spherical harmonics Y nm , for which it holds that pinta-
x 2 pallofunktio
⃦Y nm ⃦2 = 1
⃦ ⃦
Y (ϕ, λ) dσ = 1.
4π σ nm
í Õ ! ¤. û
436 B Function spaces
∑︂
∞
1 ∑︂
n
(︁ )︁
V(ϕ, λ, r) = Pnm (sin ϕ) anm cos mλ + bnm sin mλ .
rn+1
n=0 m=0
∑︂
∞
1 ∑︂
n
V(ϕ, λ, r) = vnm Y nm (ϕ, λ),
rn+1
n=0 m=−n
in which ⎧
⎨a if m ⩾ 0,
nm
vnm =
⎩bn|m| if m < 0.
í Õ ! ¤. û
Self-test questions
437
On the sphere r = R this becomes
∑︂
∞
1 ∑︂
n
V(ϕ, λ, R) = vnm Y nm (ϕ, λ),
Rn+1
n=0 m=−n
or
x
Rn+1
anm = V(ϕ, λ, R) Pnm (ϕ, λ) cos mλ dσ,
4π σ
x
Rn+1
bnm = V(ϕ, λ, R) Pnm (ϕ, λ) sin mλ dσ.
4π σ
^ Self-test questions
⟨︁ ⟩︁ ⟨︁ ⟩︁
1. The identity r · s = s · r , for two elements r and s of a vector
space, expresses the property of linearity | commutativity |
associativity.
í Õ ! ¤. û
^ Why does FFT work?
FFT
C
is a factorisation method for computing the discrete Fourier trans-
form that spectacularly reduces the number of calculations needed and
speeds up the calculation. It requires the number of data grid points to
be a factorisable number.
There are alternatives in choosing precisely which FFT method to use.
The fastest FFT requires a grid the number of points of which is a power
of 2. The size of the grid is then 2n × 2m . Alternative, “mixed-radix”
methods may also be considered and perform well if the grid size is
something like 360 × 480, for example N = 360 = 2 × 2 × 2 × 3 × 3 × 5.
If the grid size is a prime number, FFT is no better than the ordinary
discrete Fourier transform.
[︁ )︁
If the function f(x) is given on the interval x ∈ 0, L , on an equi-
spaced grid, xk = kL N , as values fk = f(xk ), k = 0, . . . , N − 1, the
/︁
in which
1 ∑︂
N−1
jk
(︂ )︂
F(˜︁
νj ) = f(xk ) exp −2πi , j = 0, . . . , N − 1. (C.1)
N N
k=0
˜︁j = j L ,
/︁
The frequency argument, spatial frequency or[︂wave number, /︂ ]︂ ν
j = 0, . . . , N − 1 is defined on the interval1 0, (N − 1) L . i is the 1
– 439 –
440 C Why does FFT work?
[︂ (︁ 1 )︁/︂ ]︂
1 Alternatively,the interval of definition can be chosen as − 12 N L , 2 N − 1 L .
/︁
í Õ ! ¤. û
441
value k = 0, 1, . . . , 12 N − 1, is either a summation, for even values of j, or
a subtraction, for odd values of j. In total, 12 N sums and 12 N differences
are pre-calculated. Also the exp expressions are pre-calculated into a
lookup table.
Altogether some 12 N2 standard operations are needed, half the origi-
nal number.
Equation C.3 is itself recognised as a Fourier series, but the number
of support points is only 12 N instead of N. If 12 N is also even, we may
repeat the above trick, resulting in an expression requiring only an
order of 14 N2 operations. Lather, rinse, repeat, and the number of
operations becomes 18 N2 , 16
1 1
N2 , 32 N2 , etc. . . . A more precise analysis
shows that if N is a power of 2, the whole discrete Fourier transform
may be computed in order N · 2 log N operations!
In the literature, smart algorithms are found implementing the
method described, for example fftw (“Fastest Fourier Transform in
the West”, FFTW Home Page; Frigo and Johnson, 2005).
í Õ ! ¤. û
^ Helmert condensation
D
In order to derive the equation for Helmert condensation, we first derive
the equation for the potential of the topography:
y ρ(ϕ ′ , λ ′ , r ′ ) ′ y
1
Vtop (ϕ, λ, r) = G dV ≈ Gρ dV ′ ,
top ℓ(ψ, r, r ′ ) top ℓ(ψ, r, r ′ )
– 443 –
444 D Helmert condensation
1 ∑︂ 1 r ′ ∑︂
∞ (︃ )︃n+1 ∞ (︃ )︃n
1 r′
= Pn (cos ψ) = Pn (cos ψ).
ℓ r′ r r r
n=0 n=0
∑︂∞ y (︃ )︃n
1 r′
ext
Vtop (ϕ, λ, r)
= Gρ r r Pn (cos ψ) dV ′ =
top
n=0
x w R+H(ϕ ′ ,λ ′ ) ∑︂
∞
(︄ (︃ )︃n )︄
1 r′ 2
= Gρ r r (r ′ ) dr ′ Pn (cos ψ) dσ ′ =
σ R
n=0
]︄R+H
x ∑︂
∞
[︄
1 1 n+3
= Gρ (r ′ ) Pn (cos ψ) dσ ′ =
σ rn+1 n + 3
n=0 r ′ =R
x ∑︂
∞
1 1
(︂ )︂
= Gρ (R + H)n+3 − Rn+3 Pn (cos ψ) dσ ′ .
σ rn+1 n + 3
n=0
1 Uniform convergence means that, given r and r ′ , for every ϵ > 0 there is an Nmin for
which
⃓ 1 1 ∑︂
⃓ ⃓
N (︃ ′ )︃n
r ⃓
⃓ − Pn (cos ψ)⃓ < ϵ
⃓ ⃓
⃓ℓ r r ⃓
n=0
for all N > Nmin , and for all values of ψ. This is a stronger property than mere
convergence.
í Õ ! ¤. û
The interior potential of the topography D.2
445
(R + H)n+3 =
(︃ )︃
n+3 H (n + 3) (n + 2) H2 (n + 3) (n + 2) (n + 1) H3
=R 1 + (n + 3) + + + ··· .
R 2 R2 2·3 R3
(D.2)
Substitution yields
ext
Vtop (ϕ, λ, r) = GρR2 ·
x ∑︂ ∞ (︂ )︂n+1 (︃ )︃
R H H2 H3
· r + 1
(n + 2) 2 + 1
(n + 2) (n + 1) 3 + · · · Pn (cos ψ) dσ ′ .
σ R 2 R 6 R
n=0
(D.3)
This is thus the exterior potential of the topography — or, inside the
topographic masses, the harmonic downwards continuation of the exterior
potential, assuming that this is mathematically possible and does not
diverge. In mountainous topography, this may be a problem.
1 ∑︂ r n+1
∞ (︂ )︂
1
= r Pn (cos ψ).
ℓ r′
n=0
Substitute:
1 ∑︂ r n+1
y ∞ (︂ )︂
int
Vtop (ϕ, λ, r) = Gρ Pn (cos ψ) dV ′ =
top r r′
n=0
I
⏟ ⏞⏞ ⏟
x w R+H(ϕ ′ ,λ ′ )
1 ∑︂
∞ (︂ )︂
r n+1 ′ 2 ′
= Gρ (r ) dr Pn (cos ψ) dσ ′ .
σ R r r′
n=0
í Õ ! ¤. û
446 D Helmert condensation
1 ∑︂ r n+1 ′ 2 ′
w R+H(ϕ ′ ,λ ′ ) ∞ (︂ )︂
I= (r ) dr =
R r r′
n=0
⎡ ⎤R+H(ϕ ′ ,λ ′ )
⎢ ∑︂
∞
(︄ )︄
(r ′ )−(n−2)
+ r2 ln r ′ ⎥
⎥
=⎣
⎢ rn − =
n−2 ⎦
n=0
n̸=2
r ′ =R
∑︂
∞
rn
(︂ )︂
R+H
= R−(n−2) − (R + H)−(n−2) + r2 ln ,
n−2 R
n=0
n̸=2
yielding
int
Vtop (ϕ, λ, r) =
⎛ ⎞
x ⎜ ∑︂
∞
rn R + H⎟
(︂ )︂
−(n−2) −(n−2) 2 ⎟ P (cos ψ) dσ ′ .
= Gρ R − (R + H) + r ln
R ⎠ n
⎜
σ⎝ n−2
n=0
n̸=2
(R + H)−(n−2) =
(︃ )︃
−(n−2) H (n − 2) (n − 1) H2 (n − 2) (n − 1) n H3
=R 1 − (n − 2) + − + ··· .
R 2 R2 2·3 R3
Also, the special case n = 2,
(︃ )︃
2 R+H 2 H 1 H2 1 H3 1 H4
r ln =r − + − + ... =
R R 2 R2 3 R3 4 R4
(︃ )︃
rn H n − 1 H2 (n − 1) n H3 (n − 1)n(n + 1) H4
= n−2 − + − + ··· ,
R R 2 R2 2 · 3 R3 2·3·4 R4
is cleanly included into the following expression obtained by substitu-
tion:
int
Vtop (ϕ, λ, r) =
x ∑︂ ∞ (︃ )︃
rn H H2 H3
= Gρ n−2
− 1
(n − 1) 2 + 1
(n − 1) n 3 − · · · Pn (cos ψ) dσ ′ . (D.4)
σ R R 2 R 6 R
n=0
í Õ ! ¤. û
The exterior potential of the condensation layer D.3
447
^ D.3 The exterior potential of the condensation layer
This is derived by specialising equation D.3 to the case H → 0, but
nevertheless ρ → ∞, so that κ = ρH remains finite. in this limit, all
terms containing H2 , H3 and higher powers go to zero. The result is
then
x ∑︂
∞ (︂ )︂n+1
R H
ext
Vcond (ϕ, λ, r) = GρR2
P (cos ψ) dσ ′ =
σ r R n
n=0
x ∑︂
∞ (︂ )︂n+1
R
= GR r κPn (cos ψ) dσ ′ .
σ
n=0
· Pn (cos ψ) dσ ′ =
x ∑︂
∞ (︂ )︂n+1 (︃ )︃
R H3
= −Gρ r
1 2 1
nH + 6 n (n + 3) + · · · Pn (cos ψ) dσ ′ .
σ 2 R
n=0
2 Theoretically speaking, the exterior space is the space outside a sphere that encloses
all of the Earth’s topography, a so-called Brillouin sphere. Practice is less restrictive.
í Õ ! ¤. û
448 D Helmert condensation
Then
ext
δVHelmert =
∑︂
∞ (︂ )︂n+1
R 1
(︃
H3n
)︃
1 2 1
= −4πGρ r nHn + 6 n (n + 3) + ··· .
2n + 1 2 R
n=0
∂ 2
∆gext
Helmert = − δV ext − δV ext ≈
∂r Helmert r Helmert
∑︂ 1 (︂ − (n + 1) 2 )︂ (︂ R )︂n+1 (︃
∞
H3n
)︃
1 2 1
≈ 4πGρ r +r r nHn + 6 n (n + 3) + ··· =
2n + 1 2 R
n=0
1 ∑︂ n − 1 R n+1 1
∞
H3n
(︂ )︂ (︃ )︃
2 1
= −4πGρ · r nHn + 6 n (n + 3) + · · · . (D.7)
2n + 1 r 2 R
n=0
í Õ ! ¤. û
Total potential of Helmert condensation D.4
449
Now, n = 1 also gives a zero result, expected as gravity anomalies do
not contain any constituents of degree number 1.
Result D.7 is approximate and not to be used on or close to the
topography. Note the strong dependence upon n: the gravity effect of
Helmert condensation is dominated by short wavelengths, the local
features of the topography.
from which one obtains with Bruns equation 5.2 the indirect effect of
Helmert condensation:
int
δVHelmert
δNHelmert = γ =
4πGρ ∑︂ n + 1 1 2
∞
H3
(︃ )︃
1
= γ H − (n − 2) n + · · · . (D.8)
2n + 1 2 n 6 R
n=0
í Õ ! ¤. û
450 D Helmert condensation
2πGρ 2
δNHelmert, const ≈ γ H ,
µ = 21 ρH2 . (D.9)
1 ∑︂ rQ n+1
∞ (︂
1
)︂
=r rP Pn (cos ψPQ ),
ℓPQ Q
n=0
3 In fact, a better place for this replacement layer would be the 41 H level.
í Õ ! ¤. û
The dipole method D.5
451
By substituting into this equation D.9 for the double mass-density layer
µQ we obtain, by taking the limit rP , rQ ↓ R:
1 ∑︂
∞ x
V= n (2πGρH) HPn (cos ψ) dσ ′ =
4π σ
n=0
1 ∑︂
∞ x
= n AB HPn (cos ψ) dσ ′ .
4π σ
n=0
Here, we have left off the designations P and Q again as they are no
longer needed for clarity.
The symbol AB denotes the attraction of a Bouguer plate of thickness
H and matter density ρ.
Let us develop the quantity (AB H) into a spherical-harmonic expan- pallofunktio-
sion. According to degree constituent equation 3.9: kehitelmä
x
2n + 1
(AB H)n = (AB H) Pn (cos ψ) dσ ′ ,
4π σ
yielding
∑︂
∞
n
V= (A H) ≈ 1 (A H) ,
2n + 1 B n 2 B
n=0
at least for the higher n values: regionally though not globally.
Thus we obtain again an estimate for the indirect effect of Helmert
condensation. In geoid computation by means of this method this
represents the shift in geoid surface caused by the condensation, which
must be undone, that is, accounted for with the opposite algebraic
sign. In other words, when looked upon as a remove–restore method, it poistamis-
constitutes its “restore” step: entistämis-
menetelmä
V AB H πGρH2
δNHelmert = γ ≈ 12 γ = γ .
For comparison, the more precise expansion D.8 yields in approxima-
tion for larger n values
4πGρ ∑︂
∞
n+1 2 πGρ ∑︂ 2
∞
πGρH2
1
δNHelmert ≈ γ · Hn ≈ γ Hn = γ ,
2 2n + 1
n=0 n=0
í Õ ! ¤. û
^ The Laplace equation in spherical
co-ordinates
^ E.1 Derivation
E
Consider a small volume element with sizes in co-ordinate directions of
def
∆ϕ, ∆λ, and ∆r. Look at the difference in flux of vector field a = ∇V vuo
between what comes in and what goes out through opposite faces.
We do the analogue of what was shown in subsection 1.12.4, using a
body or volume element with surfaces aligned along co-ordinate lines,
allowing the size of the element to go to zero in the limit, and exploiting
the divergence theorem 1.19 of Gauss. The quantity div a = ∆V is a
source density in space, and its average value multiplied by the volume
of an element must equal the total flux through the surfaces of the
element.
{︁ }︁
Define at the location of the body an orthonormal basis e1 , e2 , e3 ortonormaali
of type “north-east-up”. The vector e1 points to the local north, the kanta
vector e2 to the east, and the vector e3 “up”, in the radial direction. We
may write
a = a1 e1 + a2 e2 + a3 e3 .
Part of the difference in flux f between opposing faces is due to a
change in the normal component of a between the faces, part is due to
a difference in face surface area ω:
I II
⏟ ⏞⏞ ⏟ ⏟ ⏞⏞ ⏟
f+ − f− ≈ ω (a+ − a− ) + a (ω+ − ω− ).
– 453 –
454 E The Laplace equation in spherical co-ordinates
e1 e3
a
e2
r ∆ϕ
r cos ϕ ∆r
r r ∆λ cos ϕ
∆ϕ
∆λ
λ
Equatorial plane
ω−
ϕ = r cos ϕ ∆r ∆λ, ω+
ϕ = r cos(ϕ + ∆ϕ) ∆r ∆λ,
difference
ω+ −
ϕ − ωϕ ≈ −r sin ϕ ∆ϕ · ∆r ∆λ.
Multiply by
∂V 1 ∂V
a1 = = r
∂ (rϕ) ∂ϕ
í Õ ! ¤. û
Derivation E.1
455
and divide by element volume r2 cos ϕ ∆r ∆ϕ ∆λ, yielding
tan ϕ ∂V
∆IIϕ V = − .
r2 ∂ϕ
This of course in addition to the first contribution
a+
1 − a1
−
∆Iϕ V = ⟨∇a1 · e1 ⟩ = ,
r · ∆ϕ
with [︃ ]︃+ [︃ ]︃+
∂V 1 ∂V
a+
1 − a−
1 = = r ,
∂ (rϕ) −
∂ϕ −
yielding
[︂ ]︂+
∂
V
1 1 ∂ϕ − 1 ∂2 V
∆Iϕ V = r · r · ≈ .
∆ϕ r2 ∂ϕ2
◦ Longitudinal direction, λ, “west–east”: no change in surface area
ωλ = r ∆r ∆ϕ because of rotational symmetry:
∆IIλ V = 0.
We only have
a+
2 − a2
−
∆Iλ V = ⟨∇a2 · e2 ⟩ = ,
r cos ϕ · ∆λ
with
[︃ ]︃+ [︂ ]︂+
∂V 1 ∂V
a+
2 − a−
2 = = .
∂ (λr cos ϕ) −
r cos ϕ ∂λ −
Substitution yields
[︁ ∂
]︁+
1 1 ∂λ
V − 1 ∂2 V
∆Iλ V = · · ≈ .
r cos ϕ r cos ϕ ∆λ r2 cos ϕ ∂λ2
2
í Õ ! ¤. û
456 E The Laplace equation in spherical co-ordinates
ω+ −
r − ωr ≈ 2r ∆r · cos ϕ ∆ϕ ∆λ.
Multiply by
∂V
a3 =
∂r
and divide by the volume of the element r2 cos ϕ ∆r ∆ϕ ∆λ, yield-
ing for the second contribution to the Laplace operator
2 ∂V
∆IIr V = r .
∂r
This in addition to the first contribution
[︁ ∂ ]︁+
a+ −
3 − a3 ∂r
V − ∂2 V
∆Ir V = ⟨∇a3 · e3 ⟩ = = ≈ .
∆r ∆r ∂r2
All of this gives us the end result
^ E.2 Solution
í Õ ! ¤. û
Solution E.2
457
This must again apply for all values r and ϕ and thus both expressions
can only be equal to a constant, p. This yields two equations:
(︃ )︃
2
2∂ R ∂R
r + 2r − pR = 0,
∂r2 ∂r
(︃ )︃
1 ∂2 Y ∂2 Y ∂Y
+ − tan ϕ + pY = 0.
cos2 ϕ ∂λ2 ∂ϕ2 ∂ϕ
R(r) = rq ,
yielding
q (q − 1) rq + 2qrq − prq = 0 =⇒ (q (q + 1) − p) rq = 0
í Õ ! ¤. û
458 E The Laplace equation in spherical co-ordinates
pinta- With this, we find for the surface spherical harmonics the linear combi-
pallofunktio nations
∑︂
n
Yn (ϕ, λ) = Pnm (sin ϕ) (anm cos mλ + bnm sin mλ) .
m=0
pallofunktio- Here, anm and bnm are the spherical-harmonic coefficients specifying
kerroin the linear combination of special solutions. Only the second solution
is physically realistic for representing the Earth’s exterior gravitational
field, going to zero at infinity r → ∞.
1 This also explains why m must be an integer: the longitude λ is circular with a period
of 2π.
í Õ ! ¤. û
Bibliography
ABCDEFGHIKLMNOPRSTVWY
A
Altimetry, Retracking. Radar altimetry tutorial & toolbox. ESA and CNES.
URL http://www.altimetry.info/radar-altimetry-tutorial/data-flow/data-
processing/retracking/. Accessed 1st March, 2022. 373
– 459 –
460 ABCDEFGHIKLMNOPRSTVWY Bibliography
í ¤. û
Bibliography ABCDEFGHIKLMNOPRSTVWY
461
http://www.helmut-moritz.at/SciencePage/Molodensky.pdf. Helmut
Moritz and Maria I. Yurkina, editors. Accessed 1st March, 2020. 170
H. Bruns. Die Figur der Erde: Ein Beitrag zur europäischen Gradmessung.
Stankievicz, Berlin, 1878. URL
https://play.google.com/books/reader?id=DP0-
AAAAYAAJ&hl=en&pg=GBS.PP5. Accessed 31st January, 2020. 94
í ¤. û
462 ABCDEFGHIKLMNOPRSTVWY Bibliography
Eiffel Tower, 72 names. List of the 72 names on the Eiffel Tower. URL https:
//en.wikipedia.org/wiki/List_of_the_72_names_on_the_Eiffel_Tower.
Accessed 7th April, 2019. 16, 17, 34, 46, 57, 90, 433
í ¤. û
Bibliography ABCDEFGHIKLMNOPRSTVWY
463
L. Eötvös. Three Fundamental Papers of Loránd Eötvös. Loránd Eötvös
Geophysical Institute of Hungary, 1998. ISBN 963-7135-02-2. Editor Zoltán
Szabó. 323
G. Farmelo. The Strangest Man. Basic Books, reprint edition, 2011. ISBN
978-0-4650-2210-6. 27
FFTW Home Page. URL https://www.fftw.org. Accessed 15th May, 2019. 441
R. Forsberg and J. Kaminskis. Geoid of the Nordic and Baltic region from
gravimetry and satellite altimetry. In Segawa et al. (1996), pages 540–547.
URL https://doi.org/10.1007/978-3-662-03482-8_72. 403
R. Forsberg and G. Strykowski. NKG Gravity Data Base and NKG Geoid.
Technical University of Denmark, DTU Space, National Space Institute,
2010. URL
https://www.nordicgeodeticcommission.com/wp-content/uploads/20
í ¤. û
464 ABCDEFGHIKLMNOPRSTVWY Bibliography
14/10/8-WG_geoid_2010_March_presentation_ForsbergStrykowski.pdf.
Accessed 11th May, 2019. 403
J.-P. Friedelmeyer. Du côté des lettres (2) : une lettre de Sophie Germain à Carl
Friedrich Gauss (20 février 1807), et la réponse de celui-ci (30 avril 1807),
2014. URL https://images.math.cnrs.fr/Du-cote-des-lettres-une-lettre-de-
Sophie-Germain-a-Carl-Friedrich-Gauss-20. Accessed 22nd September,
2021. 94
í ¤. û
Bibliography ABCDEFGHIKLMNOPRSTVWY
465
https://www.researchgate.net/publication/275029640_SPACE-
WISE_GRIDS_OF_GRAVITY_GRADIENTS_FROM_GOCE_DATA_AT_N
OMINAL_SATELLITE_ALTITUDE. Accessed 11th May, 2019. 288
GRACE Mission. Measuring Earth’s Surface Mass and Water Changes. Jet
Propulsion Laboratory. URL https://grace.jpl.nasa.gov/. Accessed 1st
March, 2020. 378
í ¤. û
466 ABCDEFGHIKLMNOPRSTVWY Bibliography
M. Heikkinen. Solving the shape of the Earth by using digital density models.
Report 81:2, Finnish Geodetic Institute, Helsinki, 1981. 101, 102, 104, 318
í ¤. û
Bibliography ABCDEFGHIKLMNOPRSTVWY
467
Humboldt University Berlin. Friedrich Robert Helmert with a relative
pendulum, 2017. URL https://www.researchgate.net/publication/3189949
32_Friedrich_Robert_Helmert_founder_of_modern_geodesy_on_the_occ
asion_of_the_centenary_of_his_death. © 2017 Humboldt-Universität zu
Berlin, Universitätsbibliothek (CC BY 3.0). Accessed 19th May, 2019. 144
E. Hytönen. Absolute gravity measurement with long wire pendulum. PhD thesis,
University of Helsinki, 1972. Finnish Geodetic Institute publication 75. 301
J. Kakkuri, editor. Geodesy and Geophysics, lecture notes, NKG Autumn School
1992, Finnish Geodetic Institute publication 115, 1993. 462, 473, 474
í ¤. û
468 ABCDEFGHIKLMNOPRSTVWY Bibliography
í ¤. û
Bibliography ABCDEFGHIKLMNOPRSTVWY
469
ent/uploads/2014/10/1-Makinen_et_al_land_uplift_gravity_lines.pdf.
Accessed 22nd September, 2021. 335
S. Märdla. Regional Geoid Modelling by the Least Squares Modified Hotine Formula
Using Gridded Gravity Disturbances. PhD thesis, Tallinn University of
Technology, 2017. URL https://digikogu.taltech.ee/en/Item/baba5f3f-
22ce-43b4-8f81-7b1fe015ac19. Accessed 17th September, 2021. 130, 218
P. J. Melchior. The Tides of the Planet Earth. Pergamon Press, Oxford, 1978.
ISBN 978-0-0802-6248-2. 389, 392
NOAA, Ocean currents. How does the ocean affect climate and weather on
land? NOAA Ocean Exploration and Research. URL
í ¤. û
470 ABCDEFGHIKLMNOPRSTVWY Bibliography
W. R. Peltier. Closure of the budget of global sea level rise over the GRACE
era: the importance and magnitudes of the required corrections for global
glacial isostatic adjustment. Quaternary Science Reviews, 28(17–18):
1658–1674, 2009. URL https://doi.org/10.1016/j.quascirev.2009.04.004.
Special issue: Quaternary Ice Sheet-Ocean Interactions and Landscape
Responses. 342
í ¤. û
Bibliography ABCDEFGHIKLMNOPRSTVWY
471
PIOMAS. Polar Science Center, PIOMAS Arctic Sea Ice Volume Reanalysis.
URL http://psc.apl.washington.edu/research/projects/arctic-sea-ice-
volume-anomaly/. Accessed 11th May, 2019. 375
í ¤. û
472 ABCDEFGHIKLMNOPRSTVWY Bibliography
R. H. Rapp. The decay of the spectrum of the gravitational potential and the
topography for the Earth. Geophysical Journal International, 99(3):449–455,
1989. URL https://doi.org/10.1111/j.1365-246X.1989.tb02031.x. 287
í ¤. û
Bibliography ABCDEFGHIKLMNOPRSTVWY
473
A. R. Schweiger, R. Lindsay, L. Zhang, M. Steele, H. Stern, and R. Kwok.
Uncertainty in modeled Arctic sea ice volume. Journal of Geophysical Research,
116(C00D06), 2011. URL http://dx.doi.org/10.1029/2011JC007084. 375
D. Scuka. GOCE burning: Last orbital view. ESA blog, 2013. URL
https://blogs.esa.int/rocketscience/2013/11/11/goce-burning-last-
orbital-view/. Accessed 22nd September, 2021. 380
G. Spada and D. Melini. SELEN: a program for solving the “Sea Level
Equation”. User manual for version 2.9, Computational Infrastructure for
Geodynamics (CIG), 2015. URL
http://geodynamics.org/cig/software/selen/selen-manual.pdf. Accessed
1st March, 2020. 342
í ¤. û
474 ABCDEFGHIKLMNOPRSTVWY Bibliography
M. Vermeer. A new Seasat altimetric geoid for the Baltic. Report 83:4, Finnish
Geodetic Institute, 1983b. 351
M. Vermeer. Geoid studies on Finland and the Baltic. PhD thesis, University of
Helsinki, 1984. Report 84:3, Finnish Geodetic Institute. 34, 113
í ¤. û
Bibliography ABCDEFGHIKLMNOPRSTVWY
475
H. Virtanen and J. Kääriäinen. The installation and first results from the
superconducting gravimeter GWR20 at the Metsähovi station, Finland.
Report 95:1, Finnish Geodetic Institute, Helsinki, 1995. 315
í ¤. û
476 ABCDEFGHIKLMNOPRSTVWY Bibliography
í ¤. û
Bibliography ABCDEFGHIKLMNOPRSTVWY
477
Wikipedia, Strengthening mechanisms of materials. URL https:
//en.wikipedia.org/wiki/Strengthening_mechanisms_of_materials.
Accessed 10th May, 2019. 307
L. Wong and R. Gore. Accuracy of geoid heights from modified Stokes kernels.
Geophysical Journal of the Royal Astronomical Society, 18(1):81–91, 1969. URL
https://doi.org/10.1111/j.1365-246X.1969.tb00264.x. 221
í ¤. û
478 ABCDEFGHIKLMNOPRSTVWY Bibliography
í ¤. û
Index
ABCDEFGHIJKLMNOPQRSTUVWZ
– 479 –
480 ABCDEFGHIJKLMNOPQRSTUVWZ Index
í ¤. û
Index ABCDEFGHIJKLMNOPQRSTUVWZ
481
figure, 89 calculation by FFT, 242
centrifugal potential, 88 notation, 233
expression, 90 convolution theorem, 233
CHAMP (satellite), 377 co-ordinate time (relativity), 186
accelerometer, 377 co-ordinates
figure, 378 ellipsoidal, 52
GPS receiver, 377 geodetic, 51
characteristic equation, 431 definition, 52
Chasles theorem, 34 natural, 97
equipotential surface as boundary, figure, 96
33, 34 polar, 418
non-uniqueness of mass distribution, rectangular, 4, 50
36 Copenhagen (Denmark),
Chasles, Michel, 34 geoid determination, 403
circle mean, M⃝ , 289 Coriolis acceleration
circular disk, attraction, 132 direction, 338
closing error (gravimetry), 307 in airborne gravimetry, 320
co-geoid, 130 of an ocean current, 338
collocation, least-squares (LSC), 247 Coriolis force, 90
description, 263 Coriolis, Gaspard-Gustave, 90
FFT, 288 corkscrew rule, of the vectorial product,
figure, 270 408, 411
flexibility, 275 correlation length, 268, 269, 297
solution, 271 correlation, quasi-geoid & topography,
theory, 274 174, 175
Columbus geoid (model), 146, 404 correspondence, integral & spectral
commutative diagram equations, 193
FFT, 237 cosine rule on the sphere, 237, 387
radial shift, 74 half-angle, 238
remove--restore, 219 covariance function
vertical shift, 47 definition, 261
compensation depth, 148, 154 empirical, 287
components, of a vector, 410 Gauss–Markov, 269
condensation layer, exterior potential, 447 global, 288
confocality, 53 isotropy, 279
conservation law, 23 of gravity anomalies, 285
conservative field of Hirvonen, 268, 272
definition, 5, 416 figure, 270
as the gradient of a potential, 420 of the disturbing potential, 278, 295
divergence, 416 in space, 282, 284
potential, 416 spectral representation, 279
continental ice sheets and sea level, 340 cross product, see vectorial product
continuity equation, 421 cross-covariance, 258
convection, in the Earth’s mantle, 155 crossover adjustment
convergence, uniform, 427, 444 a priori uncertainties, 362
convolution allowed datum transformations, 364,
í ¤. û
482 ABCDEFGHIJKLMNOPQRSTUVWZ Index
í ¤. û
Index ABCDEFGHIJKLMNOPQRSTUVWZ
483
equatorial radius, 51, 101 Euler’s identity, 440
first eccentricity, 51 European Geosciences Union (EGU), 402
polar radius, 101 European Space Agency (ESA)
semimajor axis, 101 CryoSat-2, 353
semiminor axis, 101 ERS-1/2, 352
earthquake, 347 GOCE, 379
eccentricity, orbital, 366 sea-surface topography map, 341
EGM96 (geopotential model), 76 Sentinel-3A, 353
coefficients, mean errors, 78 eustatic rise, of mean sea level, 333, 343
EGM2008 (geopotential model), 63, 77, evaluation functional, 256
126, 136 Everest, Mount, 150
Eiffel Tower, 94 exterior product, see vectorial product
72 names, 16, 17, 34, 46, 57, 90, 433
eigenvalue problem F
linear operator, 428 factorial, 57
matrix, 429 Falkland Islands, 380
self-adjoint operator, 430 Faller, James E., 308
variance matrix of location, 431 falling atoms, gravity potential, 313
Einstein summation convention, 265 Faraday, Michael, 23, 421
Einstein, Albert, 4 Fast Collocation, 291
El Niño Southern Oscillation (ENSO), 329 Fast Fourier Transform (FFT)
elastic properties algorithms, 441
of the Earth, 391 and convolution, 234
of the Earth’s crust, 398 collocation, 288, 291
electric currents in the Earth’s core, 155 commutative diagram, 236
ellipsoidal-harmonic expansion mixed-radix, 439
definition, 77 radix 2, 439
centrifugal potential, 97 terrain correction, 250, 252
computation, 82 fast Fourier transform (FFT)
convergence, 81 geoid determination, 236
normal potential, 81 Fastest Fourier Transform in the West
standard form, 80 (fftw, software), 441
RMS Empress of Ireland, 165 Father Point / Pointe-au-Père (Rimouski,
Envisat (satellite), 352 Quebec, Canada), 165
eötvös (unit), 120, 324 Fennoscandia
Eötvös tensor, 322 continental ice sheet, 151, 341
Eötvös, Loránd, 91 gravity line, 336
equivalence principle, 4, 91 longevity of heights, 332
ERS-1/2 (satellites), 352 Fennoscandian Shield,
escape velocity, 86 isostatic anomalies, 154
estimator, 265 Fermat’s last theorem, 94
mean error, 271 field equations
optimal, 266 of electromagnetism, 17
Eterna (software), 398 of gravitation, 1, 17
Euclidean space, 10 field line, 23, 421
Euler notation, 93 field theory of gravitation, 1
field, the concept, 41
í ¤. û
484 ABCDEFGHIJKLMNOPQRSTUVWZ Index
í ¤. û
Index ABCDEFGHIJKLMNOPQRSTUVWZ
485
geoid computation education in geostrophic equations, 339
developing countries, 402 geostrophic equilibrium, 338
geoid computation research school, Germain curvature, 93
international, 402 Germain, Marie-Sophie, 93
geoid determination German Research Centre for Geosciences
1-D FFT, 243 (GFZ), 377
classical, 130 Gibbs phenomenon
comparison point, 332 Fourier transform, 245
FFT, 247 kernel modification, 222
research groups, 245 step function, 428
gravimetric, 34, 193 Gibbs, Josiah Willard, 222
principle, 192 glacial isostatic adjustment (GIA), 333,
2-D computation framework, 195 346
NKG, 403 glacier retreat, 150
satellite altimetry, 375 GM⊕ , best value, 6
software, 247 GNSS
spectral viewpoint, 75 height of gravimetric stations, 130
spherical cap, 220 in airborne gravimetry, 320
spherical FFT in height determination, 246
multi-band, 238 measuring ocean tidal loading, 398
Taylor expansion, 240 on an altimetric satellite, 376
standard crustal density, 180 positioning of tide gauges, 333, 337
geoid height or undulation GNSS levelling, 332
definition, 112 GOCE (satellite)
from satellite altimetry, 249 description, 379
global, 112 figure, 381
in Finland, 112 name, 382
geoid model ocean currents and heat transport,
Columbus, 146 339
computation, 246 precision, 381
global high resolution, 377 resolution, 322, 380
of Finland, 113 sea-surface topography, 341
geoid rise, 333, 334 GPS
geological map, density values, 180 on CHAMP satellite, 377
geophysical data record (GDR), 356, 373 reference system, 101
geophysical reduction, 130 GRACE (satellite pair)
geopotential description, 377, 379
image sharpness, 63 accelerometer, 378
level surface, 91 microwave link, 378
spectral expansion, 73 results, video, 380
geopotential number GRACE follow-on mission, 379
definition, 164 gradient
and levelled height, 167 of Earth attraction, 380
as energy level, 165 of gravity disturbance, 295
GEOS-3 (satellite), 351 of potential, normal direction, 34
Geosat (satellite), 351 gradient (operator), 9
í ¤. û
486 ABCDEFGHIJKLMNOPQRSTUVWZ Index
í ¤. û
Index ABCDEFGHIJKLMNOPQRSTUVWZ
487
CHAMP, 378 grid (Fourier)
GOCE, 381 formation, 234
fine structure, 376 interpolation, 236
observation density, 255 of gravity anomalies, 234
of mountains, 144 of the disturbing potential, 236
research of the Stokes kernel, 235
in Europe, 402 GRS80
in Finland, 403 definition, 101
in HUT, 404 atmospheric mass, 318
internationally, 401 GM⊕ , 6
residual, 218 spherical-harmonic coefficients, 104
statistical behaviour, 268 Gulf of Finland, airborne gravimetry, 321
temporal change, 377 Gulf Stream, 340
textbooks, 404 Guyana, French, 299
gravity formula, 87, 100, 124 GWR iGrav (gravimeter), 317
legacy, 102 GWR20 (gravimeter), 315
gravity gradient, 324
eötvös unit, 120 H
of Sun and Moon, 324 Haiyang-2A (satellite), 353
gravity mapping survey, 313 Hardanger plateau (Norway), 149
gravity measurement, reference surfaces harmonic downwards continuation
(figure), 117 existence, 204, 215
gravity potential, 91 exterior field, 217
gravity versus gravitation, 89, 91 Helmert condensation, 445
gravity-gradient tensor, 322 of gravity anomaly, 216
measurement, 323 of r ∆g, 205
GRAVSOFT (software), 247 of the exterior field, 217
Green equivalent-layer theorem, 34 harmonic field
Green, George, 29 definition, 16
Greenland attenuation with height, 55
continental ice sheet, 340 figure, 47
meltwater, 346 radial shift, 74
Greenland Aerogeophysics Project (GAP), r ∆g, 204
321 vertical shift, 47
Green’s first theorem, 29 harmonic oscillator, 44, 458
Green’s function Hayford, John Fillmore, 145
of sea level, 344 height
of the geopotential, 344 above mean sea level, 167
of vertical displacement, 345 above the reference ellipsoid, 52, 113
Green’s second theorem, 29 and geopotential number, 167
Green’s third theorem, 29 height anomaly
boundary point, 31 definition, 169
exterior point, 30 telluroid mapping, 123
exterior space, 32 three-dimensional, 123
interior point, 30, 31 height system, national, 331
Greenwich meridian, 50 height transformation surface, 332, 403
Heiskanen, Veikko Aleksanteri, 146, 404
í ¤. û
488 ABCDEFGHIJKLMNOPQRSTUVWZ Index
í ¤. û
Index ABCDEFGHIJKLMNOPQRSTUVWZ
489
isotropic process, 268 Wong–Gore, 221
isotropy and spectral representation, 201 KKJ, 124
isotropy assumption, 262 Knudsen, Per, 247, 249
Italy (ISG, funding), 402 Kolkata (India), 145
iteration Kronstadt (Russia) datum, 165
calculation of normal height, 181
calculation of orthometric height, L
179 LaCoste, Lucien, 305
LAGEOS (satellite), 366
J land uplift
J2 (dynamic flattening), 77, 101, 367, 368 absolute, 333
Jacobi, Carl Gustav Jacob, 68 effect on height system, 331
Jacobi’s determinant post-glacial
definition, 418 mechanism, 335
Bouguer-plate transformation, 132 relative, 333
map projection co-ordinates, 233 Laplace equation, 16
polar co-ordinates, 418 definition, 41
spherical co-ordinates (ϕ, λ), 231 basis solutions, 53
spherical co-ordinates (ψ, α), 68 co-ordinate transformation, 42
Jason (satellites) in ellipsoidal co-ordinates, 77
description, 353 in polar co-ordinates, 48
orbit choice, 369 in rectangular co-ordinates, 43
Java Sea (Dutch Indies, Indonesia), 300 in spherical co-ordinates, 53, 453
Jerry (GRACE satellite), 378 linearity, 42
local field behaviour, 42
K solving, 41
Kääriäinen, Jussi, 321 Laplace operator (∆), 416, 417
Kaivopuisto (Helsinki, Finland), definition, 16, 41
reference benchmark, 165, 166, linearity, 256
331 Laplace, Pierre-Simon, 16, 416
Kater, Henry, 300 Lapland, grade measurement, 131
Kaula, William, 287 latitude
Kepler, Johannes, 365 geocentric, 50
Simpson’s rule, 226 geodetic, 108
Kepler’s orbital elements, 366 reduced, 99, 108
Kepler’s second law, 408 definition, 52
figure, 409 types
Kepler’s third law, 365 figure, 100
octave script, 372 relationships, 99
kernel function grid matrix (FFT), 242 Laurentide ice sheet, 341
kernel modification law of motion, Newton’s, 4
advanced, 223 Legendre function, 54, 57
coefficients, 223 associated
degree, 221 algebraic-sign domains, 85
figure, 222 algebraic-sign intervals, 61
sharp cut-off, 222 figure, 60
soft cut-off, 222 fully normalised, 68
í ¤. û
490 ABCDEFGHIJKLMNOPQRSTUVWZ Index
í ¤. û
Index ABCDEFGHIJKLMNOPQRSTUVWZ
491
mesoscale eddy, 329, 340 National Map Grid Co-ordinate System
metal fatigue, 307 (KKJ), 124
metallurgy, 307 NAVD88, North American
Metsähovi research station (Finland), 316 Vertical Datum 1988, 165
N2000 reference benchmark, 165, 331 network hierarchy in gravimetry, 313
superconducting gravimeter, 315 Neumann, Carl Gottfried, 74
microgal (µGal), 120 Newton, Isaac, 3
microseismicity, effect on gravimeter, 306, Newton’s law of gravitation, 3
309 Newton’s law of motion, 4
Milan (Italy), ISG office, 402 Newton’s theory of gravitation, 1
milligal (mGal), 120 Niethammer, Theodor, 180
mixed covariance, 286 Niethammer’s method, 180
modal relaxation time, 347 NKG2004 (geoid model), 403
modal strength, 347 NKG2015 (geoid model), 403
Mohorovičić, Andrija, 159 NN (height system), 165
Mohorovičić discontinuity (“Moho”), 159, noise (definition), 264
252 noise variance matrix, 265, 275
Molodensky theory, 34, 123, 129, 168 Nordiska Kommissionen för Geodesi (NKG),
Molodensky, Mikhail Sergeevich, 34 403
book, 404 norm, of a vector, 5, 406, 424
photo, 170 Normaal Amsterdams Peil (NAP), 165
Molodensky’s method normal correction (NC)
evaluation point as reference level, equation, 185
214 for benchmark interval, 185
height anomaly, 215 normal gravitational potential,
linearisation, 213 spherical-harmonic expansion,
vertical gravity gradient, 211 104
Molodensky’s realisation, 170, 173 normal gravity
graphic cartoon, 173 definition, 87
Molodensky’s truncation coefficients, 224 at sea level, 125
monopole at Earth’s centre, 71 GRS80, 101
monsoon, 379 in a known location, 116, 117
Moritz, Helmut, 404 linearity along the plumb line, 170,
mu-metal, 315 173, 181
Munk, Walter, 353 on the reference ellipsoid, 99, 100
normal gravity field
N atmospheric mass, 318
N60 (height system), 164, 331 figure, 88
land uplift, 331 normal gravity vector, 171
reference surface, 247 normal height
N2000 (height system), 165, 247, 331 definition, 174
nabla (∇, operator), 10, 411 calculation, 181
National Geospatial-Intelligence Agency operationality, 175
(NGA), US, 76, 401 practical calculation, 185
National Imagery and Mapping Agency precise calculation, 181
(NIMA), US, 76 normal plumb line
í ¤. û
492 ABCDEFGHIJKLMNOPQRSTUVWZ Index
í ¤. û
Index ABCDEFGHIJKLMNOPQRSTUVWZ
493
definition, 17 calculation of Fourier
for the geopotential, 91 basis functions, 59
Poisson integral calculation of Legendre polynomials,
for computing a harmonic field 57
from surface values, 204 definition of normal height, 174
for r ∆g, 205 definition of orthometric height, 179
spectral form, 206 reference benchmark
Poisson kernel for gravity anomalies, N60, 166, 331
205–207 N2000, 331
Poisson, Siméon Denis, 17 reference ellipsoid
polynomial fit, of geoid surface, 332 as a level surface, 87, 100, 112
potential legacy, 102
definition, 5 normal potential at surface, 171
harmonic upwards continuation, 281 reference surfaces in gravity
origin of word, 29 measurement (figure), 117
uniqueness, 416 reference-surface thinking, 123
potential energy, 11 regularisation of the exterior field, 215
Potsdam system, 301, 313 relativity theory, 1
powers of height, degree constituents, 448 relativity, general, 17, 186
PRARE (positioning instrument), 352 “remove” step, 216
Pratt, John Henry, 145 remove–restore method, 130, 218, 219
Pratt–Hayford hypothesis, 145, 146 research school, international, 247
precession, nodal or orbital, 367 residual gravity field, 218
precise levelling residual terrain modelling
between tide gauges, 337 indirect effect, 218
height system creation, 331 residual terrain modelling (RTM), 215,
Prey reduction, 179 217
Prey, Adalbert, 179 as an interpolation method, 218
Principia (book), 3 “restore” step, 218
propagation of variances, 276 reversion pendulum, Kater’s, 300
propeller aircraft (airborne gravimetry), Richer, Jean, 299
321 rising dough model, of land uplift, 335
proper time (relativity), 186 river basin, tropical, 379
Robin, Victor Gustave, 121
Q Romberg, Arnold, 305
quadrature, block average, 226 Romberg, Werner, 226
quadrupole at Earth’s centre, 71 root of mountain, 144
quasi-geoid depth, 146, 148
concept, 123, 175 matter density, 145
figure, 174 rotational potential, 88
Royal Society of Edinburgh, 17
R Royal Society of London, 3, 17, 122
radar-altimeter calibration, 372
in-flight, 372 S
Raman effect, 312 Sacks, Oliver, 4
Rapp, Richard H., 76 sampling density, spatial, 255
recursion San Francisco (USA), 402
í ¤. û
494 ABCDEFGHIJKLMNOPQRSTUVWZ Index
í ¤. û
Index ABCDEFGHIJKLMNOPQRSTUVWZ
495
Simpson, Thomas, 226 figure, 7
Simpson’s rule potential, 6, 7
block average, 226 potential and attraction, figure, 9
nodal weights, 226 spherical-harmonic coefficient
sink (vector field), 22, 23, 413 as a functional, 257
figure, 414 fully normalised, 436
Skylab (space station), 351 spherical-harmonic expansion
snow clearing, 316 coefficients, 54
solar time, 369 degree-one part, 69
Solheim, Dag, 247 degree-zero part, 69
solid body, 11 first terms, 105
attraction, 12 global, 76
field at infinity, 12 model, 76
potential, 11 resolution, 63
total mass, 13 rotational symmetry, 66, 67
solid spherical harmonic, 53 spheroid
Somigliana, Carlo, 99 Bruns, 105
Somigliana–Pizzetti equation, 99 Helmert, 105
source (vector field), 22, 23, 413 spring balance, linear, 302
figure, 414 stabilised platform (gravimeter), 308, 319
space domain (Fourier), 46, 236 staff-reading difference, 183
spatial frequency (Fourier), 235, 251 steel manufacture, 307
spatial wavelength (Fourier), 235 stereographic map projection, 232
spectral coefficients, 54 Sterneck, Robert von, 300
spectral constituent function, 71 stochastic process
sphere, coated definition, 257
exterior attraction, 19 ergodicity, 260
exterior potential, 21 on the Earth’s surface, 258, 259
interior attraction, 19 stationarity, 263
interior potential, 21 variance function, 258
spherical co-ordinates, 50 Stokes curl theorem, 418
figure, 51 figure, 419
spherical harmonic Stokes equation
algebraic sign, 62 2-D simulation, 200
sectorial, 61, 62 and harmonicity, 212
semi-wavelength, 62 convolution, 231
table, 64 differentiation, 201
symmetries, 62 disturbing potential, 122
tesseral, 61, 62 exterior space, 193
wavelength, 62 geoid height, 193
zero points, 62 geoid rise, 334
zonal, 61, 62 in collocation, 275
spherical harmonics in plane co-ordinates, 233
of Laplace, 54 inner zone, 227
spherical shell integration geometry, 193
attraction, 8 spectral form, 191
í ¤. û
496 ABCDEFGHIJKLMNOPQRSTUVWZ Index
í ¤. û
Index ABCDEFGHIJKLMNOPQRSTUVWZ
497
Tikhonov regularisation, 363 normal component, 417
Tikhonov, Andrey Nikolayevich, 363 vector space
tilt meter, long water-tube, 393 abstract, 423
time slowing-down ratio (relativity), 186 basis, 423
Toeplitz circulant matrix, 290 bilinear form, 423
Toeplitz, Otto, 290 orthogonal basis, 424
Tom (GRACE satellite), 378 orthonormal basis, 424
tomography, seismic, 153 scalar product, 5, 423
TOPEX/Poseidon (satellite) vector, informal definition, 405
description, 352 vectorial product, 407
mean sea level, 331 figure, 408
orbit choice, 369 properties, 407
results, 354 Vening Meinesz integral equations, 201
topography Vening Meinesz, Felix Andries, 153, 201
exterior potential, 444 submarine measurement, 300
exterior Taylor-series expansion, 444 Verbaandert–Melchior pendulum, 394
interior potential, 445 vertical gravity gradient, 102
Taylor-series expansion, 446 anomalous
potential, 15 kernel, 207, 208
spherical-harmonic expansion, 64 reduction to sea level, 211
topography shift to inside geoid, 212 free-air, 323
Torge, Wolfgang, 404 inside-rock, 179
torsion balance viscous relaxation mode, 347
Cavendish, 4 Von Sterneck device, 300
Eötvös, 323 vortex phenomenon, 90
Toulouse (France), BGI office, 401 vorticity
trace, of a matrix, 432 in a vector field, 414
trench, ocean, 300 of flow, 422
Trieste (Italy) datum, 165
true anomaly, 366 W
Tscherning, Carl Christian, 247 Walferdange (Luxembourg), 315
Tscherning-Rapp formula, 287 water flowing upwards, 169
Tuvalu (Pacific Ocean), 342 water vapour radiometer, 356
twiddle factor (FFT), 441 water, phases, 340
Tytyri limestone mine (Lohja, Finland), wave equation
394 of matter, 42
relativistic, for the electron, 27
U weighing the Earth, 366
unit sphere, 435 weighing visitors, 316
upper culmination, of the Moon, 387 Wenzel, Hans-Georg, 398
wind pile-up, variation, 329
V wire pendulum, very long, 301
variance of prediction, 271, 276 work as a scalar product, 406
definition, 265 work integral, 406, 417
minimisation, 267 is path independent, 420
vector field parametrised, 417
definition, 410
í ¤. û
498 ABCDEFGHIJKLMNOPQRSTUVWZ Index
Z
zenith tube, 394
zero geoid, 395
zero potential, convention
at infinity, 18
at mean sea surface, 18
zero-length spring, 304, 306
how to build, 304
invention, 305
í ¤. û