Physical Geodesy
Physical Geodesy
Physical Geodesy
Bernhard Hofmann-Wellenhof
Helmut Moritz
Physical Geodesy
Second,
corrected edition
SpringerWienNewYork
Dr. Bernhard Hofmann-Wellenhof
Dr. Helmut Moritz
Institut für Navigation und Satellitengeodäsie
Technische Universität Graz, Graz, Austria
Almost the period of one generation has passed since 1967, the year of the
first release of Physical Geodesy by Weikko A. Heiskanen and Helmut Moritz.
Soon this book became a bestseller. Today, when studying publications deal-
ing with physical geodesy, not surprisingly the book is still frequently quoted.
Have the clocks been stopped since then? Not at all, time has flown as fast
as usual or maybe even faster – at least in someone’s imagination. However,
excellent quality is correlated with a long life expectation. This is the reason
why “the book” still plays an important role in geodetic science and beyond.
In the last decades, nevertheless, geodesy has certainly continually de-
veloped further – on the one hand by new computational methods and ideas
and on the other hand by modern measurement techniques. This is where
the story of this book starts.
Several years ago, I tried to convince Helmut Moritz on the necessity
of a new edition of Physical Geodesy. Even if I encountered some interest,
I did not manage to completely succeed. “Steter Tropfen höhlt den Stein”
(persistent drops hollow out the stone), I thought and started to repeat my
request regularly. The reason for my somehow obstinacy originated from
the past. In 1993, I got the chance to support Helmut Moritz in writing
the book entitled Geometry, Relativity, and Geodesy. For me, this was a
tremendously exciting time where we developed a great cooperation in any
respect. Immediately after this experience, I manifested my desire of another
chance for a cooperation. In these days, the idea of a new edition of Physical
Geodesy matured.
Finally, the persistent drops succeeded. I cannot tell you the Why and the
When; suddenly we had a contract with the Springer Publishing Company.
To me it seemed as if the wheel of time had been turned back – thank you,
Helmut!
Many persons deserve credit and thanks. Prof. Dr. Klaus-Peter Schwarz,
retired from the Department of Geomatics Engineering of the University
of Calgary, strongly influenced the balance between keeping, eliminating,
updating, and adding topics.
Prof. Dr. Herbert Lichtenegger, retired from the Institute of Navigation
and Satellite Geodesy of the Graz University of Technology, was a reviewer
of the book. He has critically read and corrected the full volume. His many
suggestions and improvements, critical remarks and proposals are gratefully
acknowledged.
Prof. Dr. Norbert Kühtreiber from the Institute of Navigation and Satel-
viii Foreword
lite Geodesy of the Graz University of Technology has helped with construc-
tive critique and valuable suggestions. Furthermore, he has strongly helped
to shape Chap. 11 by providing numerical examples and his valuable expe-
rience on the practical aspects of geoid computation.
In several fruitful discussions, Prof. Dr. Roland Pail from the Institute
of Navigation and Satellite Geodesy of the Graz University of Technology
has provided his rich experience on the space gravity missions. Parts of his
structured lecture notes are mirrored in the corresponding section. He also
deserves thanks for a careful proofreading of this section.
The cover illustration was designed and produced by Dipl.-Ing. Elmar
Wasle of TeleConsult Austria GmbH (www.teleconsult-austria.at). When
presenting this illustration to the Springer Publishing Company, the response
was extremely positive because of its eye-catcher quality.
The index of the book was produced using a computer program written
by Dr. Walter Klostius from the Institute of Geoinformation of the Graz
University of Technology. Also, his program helped in the detection of some
spelling errors.
The book is compiled with the text system LATEX2ε. One of the figures
included is also developed with LATEX2ε. The remaining figures are drawn
by using CorelDRAW 11. Primarily Dr. Klaus Legat from the Institute of
Navigation and Satellite Geodesy of the Graz University of Technology de-
serves the thanks for the figures. He was supported by Prof. Dr. Norbert
Kühtreiber. The highly academic level of the producers assures a seal of
quality. Many of these figures are redrawings of the originals in Heiskanen
and Moritz (1967).
I am also grateful to the Springer Publishing Company for their advice
and cooperation.
The inclusion by name of a commercial company or product does not
constitute an endorsement by the authors. In principle, such inclusions were
avoided whenever possible.
Finally, your ideas for a future edition of this book and your advice are
appreciated and encouraged.
The selection of topics is certainly different from the original book written
by Heiskanen and Moritz. However, basically we tried not only to keep the
overall structure wherever possible but also to leave the text unchanged.
The primary selection criteria of the topics were relevancy, tutorial content,
and the interest and expertise of the authors. A detailed description of the
contents is given in the Preface.
Internet citations within the text omit the part “http://” if the address con-
tains “www”; therefore, “www.esa.int” means “http://www.esa.int”. Usu-
ally, internet addresses given in the text are not repeated in the list of refer-
ences. Therefore, the list of references does not yield a complete picture of
the references of which we have been benefiting.
The use of the internet sources caused some troubles for the following
reason. When looking for a proper and concise explanation or definition,
quite often identical descriptions were found at different locations. So the
unsolvable problem arose to figure out the earlier and original source. In
these cases, sometimes the decision was made, to avoid a possible conflict of
interests, by omitting the citation of the source at all. This means that some
phrases or sentences may have been adapted from internet sources. On the
other side, as soon as this book is released, it may and will also serve as an
input source for several homepages.
For bibliographical references, the most readily accessible or most com-
prehensive publication of an author on a particular topic is given rather than
his first. The list of references does not aim at completeness; some important
publications may have been omitted but never on purpose.
The (American) spelling of a word is adopted from Webster’s Dictionary
of the English Language (third edition, unabridged). Apart from typical
differences like the American “leveling” in contrast to the British “level-
ling”, this may lead to other divergences when comparing dictionaries. Web-
ster’s Dictionary always combines the negation “non” and the following word
without hyphen unless a capital letter follows. Therefore “nongravitational”,
“nonpropulsed”, “nonsimultaneity” and “non-European” are corresponding
spellings.
Symbols representing a vector or a matrix are in boldface. The inner or
scalar product of two vectors is indicated by a dot “·”. The norm of a vector,
i.e., its length, is indicated by two double-bars “”. Vectors not related to
matrices are written either as column or as row vectors, whatever is more
convenient.
Motivation 1
4 Heights 157
4.1 Spirit leveling . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
4.2 Geopotential numbers and dynamic heights . . . . . . . . . . 159
4.3 Orthometric heights . . . . . . . . . . . . . . . . . . . . . . . 161
4.4 Normal heights . . . . . . . . . . . . . . . . . . . . . . . . . . 166
4.5 Comparison of different height systems . . . . . . . . . . . . . 168
4.6 GPS leveling . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
References 389
m
F =G (1–3)
l2
l F
¯ Y || y
m y–´
z
P (x,y,z)
|| z
l
F
° z–³
m (»,´,³) ¯ Z
|| y
® X x–»
|| x y
x
Fig. 1.1. The components of the gravitational force; upper figure
shows y-component
1.1 Attraction and potential 5
Gm x − ξ x−ξ
X = −F cos α = − 2
= −G m 3 ,
l l l
Gm y − η y−η
Y = −F cos β = − = −G m 3 , (1–4)
l2 l l
Gm z − ζ z−ζ
Z = −F cos γ = − 2
= −G m 3 ,
l l l
where
l= (x − ξ)2 + (y − η)2 + (z − ζ)2 . (1–5)
We next introduce a scalar function
Gm
V = , (1–6)
l
called the potential of gravitation. The components X, Y, Z of the gravita-
tional force F are then given by
∂V ∂V ∂V
X= , Y = , Z= , (1–7)
∂x ∂y ∂z
that is, the force vector is the gradient vector of the scalar function V .
It is of basic importance that according to (1–7) the three components
of the vector F can be replaced by a single function V . Especially when we
consider the attraction of systems of point masses or of solid bodies, as we
do in geodesy, it is much easier to deal with the potential than with the three
components of the force. Even in these complicated cases the relations (1–7)
are applied; the function V is then simply the sum of the contributions of
the respective particles.
Thus, if we have a system of several point masses m1 , m2 , . . . , mn , the
potential of the system is the sum of the individual contributions (1–6):
mi n
G m1 G m2 G mn
V = + + ··· + =G . (1–10)
l1 l2 ln li
i=1
6 1 Fundamentals of potential theory
dm
= , (1–11)
dv
where dv is an element of volume and dm is an element of mass. Then the
sum (1–10) becomes an integral (Newton’s integral),
dm
V =G =G dv , (1–12)
l l
v v
where l is the distance between the mass element dm = dv and the at-
tracted point P . Denoting the coordinates of the attracted point P by x, y, z
and of the mass element m by ξ, η, ζ, we see that l is again given by (1–5),
and we can write explicitly
(ξ, η, ζ)
V (x, y, z) = G dξ dη dζ , (1–13)
(x − ξ)2 + (y − η)2 + (z − ζ)2
v
dv = dξ dη dζ . (1–14)
z
P (x,y,z)
l
v
d³
dm (»,´,³)d ´ d »
y
The components of the force of attraction are given by (1–7). For in-
stance,
∂V ∂ (ξ, η, ζ)
X= =G dξ dη dζ
∂x ∂x l
v
(1–15)
∂ 1
=G (ξ, η, ζ) dξ dη dζ .
∂x l
v
Note that we have interchanged the order of differentiation and integration.
Substituting (1–8) into the above expression, we finally obtain
x−ξ
X = −G dv . (1–16)
l3
v
between two points P (ξ, η, ζ) and P (x, y, z). It is the potential of a point
mass m = 1/G, situated at the point P (ξ, η, ζ); compare (1–5) and (1–6).
It is easy to show that 1/l is harmonic. We form the following partial
derivatives with respect to x, y, z in the fashion of (1–8):
∂ 1 x−ξ ∂ 1 y−η ∂ 1 z−ζ
=− 3 , =− 3 , =− 3 ;
∂x l l ∂y l l ∂z l l
∂2 1 −l2 + 3(x − ξ)2 ∂2 1 −l2 + 3(y − η)2
= , = , (1–23)
∂x2 l l3 ∂y 2 l l3
∂2 1 −l2 + 3(z − ζ)2
= .
∂z 2 l l3
Adding the last three equations and recalling the definition of ∆, we find
1
∆ = 0; (1–24)
l
Not only the potential of a point mass but also any other gravitational
potential is harmonic outside the attracting masses. Consider the potential
(1–12) of an extended body. Interchanging the order of differentiation and
integration, we find from (1–12)
1
∆V = G ∆ dv = G ∆ dv = 0 ; (1–25)
l l
v v
that is, the potential of a solid body is also harmonic at any point P (x, y, z)
outside the attracting masses.
If P lies inside the attracting body, the above derivation breaks down,
since 1/l becomes infinite for that mass element dm(ξ, η, ζ) which coincides
with P (x, y, z), and (1–24) does not apply. This is the reason why the po-
tential of a solid body is not harmonic in its interior but instead satisfies
Poisson’s differential equation (1–17).
P
#
r
z
#
y
¸ r si
x n#
equations
x = r sin ϑ cos λ ,
y = r sin ϑ sin λ , (1–26)
z = r cos ϑ ;
or inversely by
r= x2 + y 2 + z 2 ,
x2 + y 2
ϑ = tan−1 , (1–27)
z
y
λ = tan−1 .
x
To get Laplace’s equation in spherical coordinates, we first determine the
element of arc (element of distance) ds in these coordinates. For this purpose
we form
∂x ∂x ∂x
dx = dr + dϑ + dλ ,
∂r ∂ϑ ∂λ
∂y ∂y ∂y
dy = dr + dϑ + dλ , (1–28)
∂r ∂ϑ ∂λ
∂z ∂z ∂z
dz = dr + dϑ + dλ .
∂r ∂ϑ ∂λ
By differentiating (1–26) and substituting it into the elementary formula
we obtain
ds2 = dr 2 + r 2 dϑ2 + r 2 sin2 ϑ dλ2 . (1–30)
h1 = 1 , h2 = r , h3 = r sin ϑ . (1–33)
where the primes denote differentiation with respect to the argument (r, in
this case). Since the left-hand side depends only on r and the right-hand side
12 1 Fundamentals of potential theory
only on ϑ and λ, both sides must be constant. We can therefore separate the
equation into two equations:
and
∂2Y ∂Y 1 ∂2Y
+ cot ϑ + + n(n + 1) Y = 0 , (1–40)
∂ϑ2 ∂ϑ sin2 ϑ ∂λ2
where we have denoted the constant by n(n + 1).
Solutions of (1–39) are given by the functions
where the functions g and h individually depend on one variable only. Per-
forming this substitution in (1–40) and multiplying by sin2 ϑ/g h, we find
sin ϑ h
sin ϑ g + cos ϑ g + n(n + 1) sin ϑ g = − , (1–45)
g h
Yn (ϑ, λ) = Pnm (cos ϑ) cos mλ and Yn (ϑ, λ) = Pnm (cos ϑ) sin mλ (1–50)
are solutions of the differential equation (1–40) for Laplace’s surface spherical
harmonics.
Since these solutions are linear, any linear combination of the solutions
(1–50) is also a solution. Such a linear combination has the general form
n
Yn (ϑ, λ) = [anm Pnm (cos ϑ) cos mλ + bnm Pnm (cos ϑ) sin mλ] , (1–51)
m=0
where anm and bnm are arbitrary constants. This is the general expression
for the surface spherical harmonics Yn (ϑ, λ).
14 1 Fundamentals of potential theory
satisfies (1–56). Apart from the factor (1 − t2 )m/2 = sinm ϑ and from a
constant, the function Pnm is the (n + m)th derivative of the polynomial
(t2 − 1)n . It can, thus, be evaluated. For instance,
(1 − t2 )1/2 d2 2 1
P11 (t) = (t − 1) = 1 − t 2·2= 1 − t2 = sin ϑ . (1–58)
2·1 dt2 2
The case m = 0 is of particular importance. The functions Pn0 (t) are often
simply denoted by Pn (t). Then (1–57) gives
1 dn 2
Pn (t) = Pn0 (t) = (t − 1)n . (1–59)
2n n! dtn
Because m = 0, there is no square root, that is, no sin ϑ. Therefore, the
Pn (t) are simply polynomials in t. They are called Legendre’s polynomials.
We give the Legendre polynomials for n = 0 through n = 5:
P0 (t) = 1 , P3 (t) = 52 t3 − 3
2 t,
P1 (t) = t , P4 (t) = 35 4
8 t − 15 2
4 t + 38 , (1–60)
P2 (t) = 32 t2 − 1
2 , P5 (t) = 63 5
8 t − 35 3
4 t + 15
8 t.
Remember that
t = cos ϑ . (1–61)
The polynomials may be obtained by means of (1–59) or more simply by the
recursion formula
n−1 2n − 1
Pn (t) = − Pn−2 (t) + t Pn−1 (t) , (1–62)
n n
by which P2 can be calculated from P0 and P1 , P3 from P1 and P2 , etc.
Graphs of the Legendre polynomials are shown in Fig. 1.4.
The powers of cos ϑ can be expressed in terms of the cosines of multiples
of ϑ, such as
cos2 ϑ = 1
2 cos 2ϑ + 1
2 , cos3 ϑ = 1
4 cos 3ϑ + 3
4 cos ϑ . (1–63)
Therefore, we may also express the Pn (cos ϑ) in this way, obtaining
P2 (cos ϑ) = 34 cos 2ϑ + 14 ,
P4 (cos ϑ) = 35
cos 4ϑ + 5
cos 2ϑ + 9
, (1–64)
64 16 64
63 35 15
P5 (cos ϑ) = 128 cos 5ϑ + 128 cos 3ϑ + 64 cos ϑ ,
··· = ··· .
16 1 Fundamentals of potential theory
P0
1.0
Pn(t)
0.5
P6 P6
0
–1.0 –0.5 0.5 1.0
P4 –0.5 P4
P2
–1.0
t = cos #
1.0
Pn(t) P1
P3 0.5 P5
P7
0
–1.0 –0.5 0.5 1.0
P7 P5 –0.5
P3
P1
–1.0
t = cos #
dm Pn (t)
Pnm (t) = (1 − t2 )m/2 , (1–65)
dtm
which follows from (1–57) and (1–59). Thus, the associated Legendre func-
tions are expressed in terms of the Legendre
√ polynomials of the same degree
n. We give some Pnm , writing t = cos ϑ, 1 − t2 = sin ϑ:
2 cos ϑ − 2 ,
P31 (cos ϑ) = sin ϑ 15 2 3
P11 (cos ϑ) = sin ϑ ,
or associated function):
r
(2n − 2k)!
Pnm (t) = 2−n (1 − t2 )m/2 (−1)k tn−m−2k ,
k! (n − k)! (n − m − 2k)!
k=0
(1–67)
where r is the greatest integer ≤ (n−m)/2; i.e., r is (n−m)/2 or (n−m−1)/2,
whichever is an integer. This formula is convenient for programming.
As this useful formula is seldom found in the literature, we show the
derivation, which is quite straightforward. The necessary information on fac-
torials may be obtained from any collection of mathematical formulas. The
binomial theorem gives
n n
n k n 2n−2k n!
(t − 1) =
2
(−1) t = (−1)k t2n−2k . (1–68)
k k! (n − k)!
k=0 k=0
the quantity n! having been cancelled out. The rth derivative of the power
ts is
dr s s!
r
(t ) = s(s − 1) · · · (s − r + 1) ts−r = ts−r . (1–70)
dt (s − r)!
Setting r = n + m and s = 2n − 2k, we have
Inserting this into the above expression for Pnm (t) and noting that the lowest
possible power of t is either t or t0 = 1, we obtain (1–67).
The surface spherical harmonics are Legendre’s functions multiplied by
cos mλ or sin mλ:
degree 0 P0 (cos ϑ) ;
degree 1 P1 (cos ϑ) ,
P11 (cos ϑ) cos λ , P11 (cos ϑ) sin λ ;
(1–72)
degree 2 P2 (cos ϑ) ,
P21 (cos ϑ) cos λ , P21 (cos ϑ) sin λ ,
P22 (cos ϑ) cos 2λ , P22 (cos ϑ) sin 2λ ;
18 1 Fundamentals of potential theory
and so on.
The geometrical representation of these spherical harmonics is useful.
The harmonics with m = 0 – that is, Legendre’s polynomials – are polyno-
mials of degree n in t, so that they have n zeros. These n zeros are all real
and situated in the interval −1 ≤ t ≤ +1, that is, 0 ≤ ϑ ≤ π (Fig. 1.4).
Therefore, the harmonics with m = 0 change their sign n times in this inter-
val; furthermore, they do not depend on λ. Their geometrical representation
is therefore similar to Fig. 1.5 a. Since they divide the sphere into zones, they
are also called zonal harmonics.
The associated Legendre functions change their sign n − m times in the
interval 0 ≤ ϑ ≤ π. The functions cos mλ and sin mλ have 2m zeros in the
interval 0 ≤ λ < 2π, so that the geometrical representation of the harmonics
for m = 0 is similar to that of Fig. 1.5 b. They divide the sphere into com-
partments in which they are alternately positive and negative, somewhat like
a chess board, and are called tesseral harmonics. “Tessera” means a square
or rectangle, or also a tile. In particular, for n = m, they degenerate into
functions that divide the sphere into positive and negative sectors, in which
case they are called sectorial harmonics, see Fig. 1.5 c.
P6(cos #)
(b) (c)
Fig. 1.5. The kinds of spherical harmonics: (a) zonal, (b) tesseral, (c) sectorial
1.8 Legendre’s functions of the second kind 19
are defined by
n
1 1+t 1
Qn (t) = Pn (t) ln − Pk−1 (t) Pn−k (t) , (1–74)
2 1−t k
k=1
dm Qn (t)
Qnm (t) = (1 − t2 )m/2 . (1–75)
dtm
1 1+t
Q0 (t) = ln = tanh−1 t ,
2 1−t
t 1+t
Q1 (t) = ln − 1 = t tanh−1 t − 1 , (1–76)
2 1−t
3 2 1 1+t 3 3 2 1 3
Q2 (t) = t − ln − t= t − tanh−1 t − t .
4 4 1−t 2 2 2 2
These formulas and Fig. 1.6 show that the functions Qnm are really
quite different from the functions Pnm . From the singularity ±∞ at t = ±1
(i.e., ϑ = 0 or π), we see that it is impossible to substitute Qnm (cos ϑ) for
Pnm (cos ϑ) if ϑ means the polar distance, because harmonic functions must
be regular.
However, we will encounter them in the theory of ellipsoidal harmon-
ics (Sect. 1.16), which is applied to the normal gravity field of the earth
(Sect. 2.7). For this purpose we need Legendre’s functions of the second
20 1 Fundamentals of potential theory
1.0
Q n ( t) Q0
Q2 0.5
0
–1.0 –0.5 0.5 1.0
–0.5
Q4
–1.0
t = cos #
1.0
Q n ( t)
Q3
Q5 0.5
0
–1.0 –0.5 0.5 1.0
–0.5
Q1
–1.0
t = cos #
Fig. 1.6. Legendre’s functions of the second kind: n even (top) and n odd (bottom)
n
1 z+1 1
Qn (z) = Pn (z) ln − Pk−1 (z) Pn−k (z) , (1–77)
2 z−1 k
k=1
1 1+t
ln = tanh−1 t (1–78)
2 1−t
by
1 z+1
ln = coth−1 z . (1–79)
2 z−1
1.9 Expansion theorem and orthogonality relations 21
In particular, we have
1 z+1
Q0 (z) = ln = coth−1 z ,
2 z−1
z z+1
Q1 (z) = ln − 1 = z coth−1 z − 1 , (1–80)
2 z−1
3 2 1 z+1 3 3 2 1 3
Q2 (z) = z − ln − z= z − coth−1 z − z .
4 4 z−1 2 2 2 2
Note that there is no Sn0 , since sin 0λ = 0. In these formulas we have used
the abbreviation 2π π
= (1–85)
λ=0 ϑ=0
σ
for the integral over the unit sphere. The expression
dσ = sin ϑ dϑ dλ (1–86)
denotes the surface element of the unit sphere.
Now we turn to the determination of the coefficients anm and bnm in
(1–81). Multiplying both sides of the equation by a certain Rsr (ϑ, λ) and
integrating over the unit sphere gives
f (ϑ, λ) Rsr (ϑ, λ) dσ = asr [Rsr (ϑ, λ)]2 dσ , (1–87)
σ σ
since in the double integral on the right-hand side all terms except the one
with n = s, m = r will vanish according to the orthogonality relations (1–
83). The integral on the right-hand side has the value given in (1–84), so
that asr is determined. In a similar way we find bsr by multiplying (1–81)
by Ssr (ϑ, λ) and integrating over the unit sphere. The result is
2n + 1
an0 = f (ϑ, λ) Pn (cos ϑ) dσ ;
4π
σ
⎫
2n + 1 (n − m)! ⎪
anm = f (ϑ, λ) Rnm (ϑ, λ) dσ ⎪
⎪
⎪ (1–88)
2π (n + m)! ⎪
⎬
σ
(m = 0) .
2n + 1 (n − m)! ⎪
⎪
bnm = f (ϑ, λ) Snm (ϑ, λ) dσ ⎪
⎪
⎪
2π (n + m)! ⎭
σ
The coefficients anm and bnm can, thus, be determined by integration.
We note that the Laplace spherical harmonics Yn (ϑ, λ) in (1–81) may
also be found directly by the formula
π
2n + 1 2π
Yn (ϑ, λ) = f (ϑ , λ ) Pn (cos ψ) sin ϑ dϑ dλ , (1–89)
4π λ =0 ϑ =0
where ψ is the spherical distance between the points P , represented by ϑ, λ,
and P , represented by ϑ , λ (Fig. 1.7), so that
PN
¸' – ¸
#
#'
P
à P'
⎫
(n − m)! ⎪
⎪
R̄nm (ϑ, λ) = Rnm (ϑ, λ) ⎪
⎪
2(2n + 1) ⎪
⎪ (1–91)
(n + m)! ⎬
(m = 0) .
⎪
⎪
(n − m)! ⎪
⎪
S̄nm (ϑ, λ) = 2(2n + 1) Snm (ϑ, λ) ⎪
⎪
⎭
(n + m)!
The orthogonality relations (1–83) also apply for these fully normalized har-
24 1 Fundamentals of potential theory
This means that the average square of any fully normalized harmonic is
unity, the average being taken over the sphere (the average corresponds to
the integral divided by the area 4π). This formula now applies for any m,
whether it is zero or not.
If we expand an arbitrary function f (ϑ, λ) into a series of fully normalized
harmonics, analogously to (1–81),
∞
n
f (ϑ, λ) = [ānm R̄nm (ϑ, λ) + b̄nm S̄nm (ϑ, λ)] , (1–93)
n=0 m=0
that is, the coefficients are the average products of the function and the
corresponding harmonic R̄nm or S̄nm .
The simplicity of formulas (1–92) and (1–94) constitutes the main ad-
vantage of the fully normalized spherical harmonics and makes them useful
in many respects, even though the functions R̄nm and S̄nm in (1–91) are a
little more complicated than the conventional Rnm and Snm . We have
R̄nm (ϑ, λ) = P̄nm (cos ϑ) cos mλ ,
(1–95)
S̄nm (ϑ, λ) = P̄nm (cos ϑ) sin mλ ,
where
r
√ (2n − 2k)!
P̄n0 (t) = 2n + 1 2−n (−1)k tn−2k (1–96)
k! (n − k)! (n − 2k)!
k=0
for m = 0, and
(n − m)! −n
P̄nm (t) = 2(2n + 1) 2 (1 − t2 )m/2 ·
(n + m)!
(1–97)
r
(2n − 2k)!
(−1)k tn−m−2k
k! (n − k)! (n − m − 2k)!
k=0
1.11 Zonal harmonics and decomposition formula 25
is given by
l2 = r 2 + r 2 − 2r r cos ψ , (1–100)
where ψ is the angle between the radius vectors r and r (Fig. 1.8), so that,
from (1–90),
l
r
P'
r'
Ã
O
1 1 1
= = √ , (1–102)
l
r − 2r r cos ψ + r
2 2 r 1 − 2α u + α2
∞
1
√ = αn Pn (u) = P0 (u)+α P1 (u)+α2 P2 (u)+ · · · . (1–103)
1 − 2α u + α2 n=0
Hence, we obtain
∞
1 r n
= Pn (cos ψ) , (1–104)
l r n+1
n=0
∞ n
(n − m)!
1 Pn (cos ϑ) n
= n+1
r Pn (cos ϑ ) + 2 ·
l r (n + m)!
n=0 m=1
(1–106)
Rnm (ϑ, λ) n Snm (ϑ, λ) n
r Rnm (ϑ , λ ) + r Snm (ϑ , λ ) .
r n+1 r n+1
n
1
Pn (cos ψ) = R̄nm (ϑ, λ)R̄nm (ϑ , λ ) + S̄nm (ϑ, λ)S¯nm (ϑ , λ ) ;
2n + 1
m=0
(1–107)
1.12 Solution of Dirichlet’s problem 27
∞
n
1 1
= ·
l n=0 m=0
2n + 1
(1–108)
R̄nm (ϑ, λ) n S̄nm (ϑ, λ) n
r R̄nm (ϑ , λ ) + r S̄nm (ϑ , λ ) .
r n+1 r n+1
The last formula will be fundamental for the expansion of the earth’s gravi-
tational field in spherical harmonics.
the Yn (ϑ, λ) being determined by (1–89). (This series converges for very
general functions V .) The functions
∞
Vi (r, ϑ, λ) = r n Yn (ϑ, λ) (1–110)
n=0
and
∞
Yn (ϑ, λ)
Ve (r, ϑ, λ) = (1–111)
n=0
r n+1
assume the given values V (1, ϑ, λ) on the surface r = 1. The series (1–109)
converges, and we have for r < 1
r n Yn < Yn (1–112)
Hence, the series (1–110) converges for r ≤ 1, and the series (1–111) con-
verges for r ≥ 1; furthermore, both series have been found to represent
harmonic functions. Therefore, we see that Dirichlet’s problem is solved by
Vi (r, ϑ, λ) for the interior of the sphere r = 1, and by Ve (r, ϑ, λ) for its exte-
rior.
For a sphere of arbitrary radius r = R, the solution is similar. We expand
the given function
∞
V (R, ϑ, λ) = Yn (ϑ, λ) . (1–114)
n=0
The surface spherical harmonics Yn are determined by
π
2n + 1 2π
Yn (ϑ, λ) = V (R, ϑ , λ ) Pn (cos ψ) sin ϑ dϑ dλ . (1–115)
4π
λ =0 ϑ =0
Then the series
∞
r n
Vi (r, ϑ, λ) = Yn (ϑ, λ) (1–116)
R
n=0
solves the first boundary-value problem for the interior, and the series
∞ n+1
R
Ve (r, ϑ, λ) = Yn (ϑ, λ) (1–117)
r
n=0
Poisson’s integral
A more direct solution is obtained as follows. We consider only the exterior
problem, which is of greater interest in geodesy. Substituting Yn (ϑ, λ) from
(1–89) into (1–117), we obtain
Ve (r, ϑ, λ) =
∞ n+1 π
R 2n + 1 2π
V (R, ϑ , λ ) Pn (cos ψ) sin ϑ dϑ dλ .
r 4π
λ =0 ϑ =0
n=0
(1–118)
1.13 Other boundary-value problems 29
The sum in the brackets can be evaluated. We denote the spatial distance
between the points P (r, ϑ, λ) and P (R, ϑ , λ ) by l. Then, using (1–104),
∞ n+1
1 1 1 R
= = Pn (cos ψ) (1–120)
l r 2 + R2 − 2R r cos ψ R
n=0
r
The harmonic function which solves Neumann’s problem for the exterior of
the sphere is then
∞ n+1
R Yn (ϑ, λ)
Ve (r, ϑ, λ) = −R . (1–127)
r n+1
n=0
∞ n+2
∂Ve R
= Yn (ϑ, λ) . (1–128)
∂r n=0
r
Since for the sphere the normal coincides with the radius vector, we have
∂V ∂V
= , (1–129)
∂n r=R ∂r r=R
solves the third boundary-value problem for the exterior of the sphere r = R.
The straightforward verification is analogous to the case of (1–127).
In the determination of the geoidal undulations, the constants h, k have
the values
2
h = − , k = −1 , (1–132)
R
so that
∞ n+1
R Yn (ϑ, λ)
Ve (r, ϑ, λ) = R (1–133)
n=0
r n−1
only to the extensive literature, e.g., the book by Moritz (1995), the inter-
net page www.inas.tugraz.at/forschung/InverseProblems/AngerMoritz.html
or Anger and Moritz (2003).
Forming the radial derivative ∂V /∂r, we note that V (R, ϑ , λ ) does not
depend on r. Thus, we need only to differentiate (r 2 − R2 )/l3 , obtaining
2π π
∂V (r, ϑ, λ) R
= M (r, ψ) V (R, ϑ , λ ) sin ϑ dϑ dλ , (1–135)
∂r 4π λ =0 ϑ =0
where
∂ r 2 − R2 1
M (r, ψ) ≡ = 5 (5R2 r−r 3 −R r 2 cos ψ −3R3 cos ψ) . (1–136)
∂r l3 l
∂V1 R R
= − 2 and V1 (R, ϑ , λ ) = = 1, (1–137)
∂r r R
we obtain
2π π
R R
− 2 = M (r, ψ) sin ϑ dϑ dλ . (1–138)
r 4π λ =0 ϑ =0
Multiplying both sides of this equation by V (r, ϑ, λ) and subtracting it from
(1–135) gives
2π π
∂V R R
+ 2 VP = M (r, ψ) (V − VP ) sin ϑ dϑ dλ , (1–139)
∂r r 4π λ =0 ϑ =0
where
VP = V (r, ϑ, λ) , V = V (R, ϑ , λ ) . (1–140)
In order to find the radial derivative at the surface of the sphere of radius
R, we must set r = R. Then l becomes (Fig. 1.9)
1.14 The radial derivative of a harmonic function 33
P l0 P'
R sin(Ã/2)
R
Ã
ψ
l0 = 2R sin , (1–141)
2
and the function M takes the simple form
1 2R
M (R, ψ) = = . (1–142)
4R2 sin3 ψ2 l03
Differentiation yields
∞
∂V Rn+1
=− (n + 1) n+2 Yn (ϑ, λ) . (1–146)
∂r r
n=0
34 1 Fundamentals of potential theory
const
u=
E
2
# = const
u u2 +
P
u z
# ¯
F1 O E F2 xy - plane
x
z u 2 + E 2 sin # P
¸ ¸ = const
Gr y
ee
nw
i ch
z = u sin β .
Taking u = constant, we find
x2 + y 2 z2
+ = 1, (1–152)
u2 + E 2 u2
which represents an ellipsoid of revolution. For ϑ = constant, we obtain
x2 + y 2 z2
2 − = 1, (1–153)
E 2 sin ϑ E cos2 ϑ
2
The left-hand side depends only on u and ϑ, the right-hand side only on λ.
The two sides cannot be identically equal unless both are equal to the same
constant. Therefore,
h
= −m2 . (1–164)
h
The factor by which h /h is to be multiplied, i.e., the inverse of the main
factor on the left-hand side of (1–163), can be decomposed as follows:
u2 + E 2 cos2 ϑ 1 E2
= − . (1–165)
(u2 + E 2 ) sin2 ϑ sin2 ϑ u2 + E 2
Substituting (1–164) and (1–165) into (1–163) and combining functions of
the same variable, we obtain
1 2 E2 1 m2
[(u + E 2 ) f + 2u f ] + 2 m 2
= − (g + g
cot ϑ) + . (1–166)
f u + E2 g sin2 ϑ
The two sides are functions of different independent variables and must there-
fore be constant. Denoting this constant by n(n + 1), we finally get
E2
(u + E ) f (u) + 2u f (u) − n(n + 1) − 2
2 2 2
m f (u) = 0 ; (1–167)
u + E2
m2
sin ϑ g (ϑ) + cos ϑ g (ϑ) + n(n + 1) sin ϑ − g(ϑ) = 0 ; (1–168)
sin ϑ
h (λ) + m2 h(λ) = 0 . (1–169)
These are the three ordinary differential equations into which the partial
differential equation (1–159) is decomposed by the separation of variables
(1–161).
The second and third equations are the same as in the spherical case,
Eqs. (1–46) and (1–47); the first equation is different. The substitutions
u √
τ =i (where i = −1) and t = cos ϑ (1–170)
E
transform the first and second equations into
¯ ¯ m2
(1 − τ ) f (τ ) − 2τ f (τ ) + n(n + 1) −
2 f¯(τ ) = 0 ,
1 − τ2
(1–171)
m2
(1 − t ) ḡ (t) − 2t ḡ (t) + n(n + 1) −
2 ḡ(t) = 0 ,
1 − t2
where the overbar indicates that the functions f and g are expressed in terms
of the new arguments τ and t. From spherical harmonics we are already
1.16 Ellipsoidal harmonics 39
familiar with the substitution t = cos ϑ and the corresponding equation for
ḡ(t).
Note that f¯(τ ) satisfies formally the same differential equation as ḡ(t),
namely, Legendre’s equation (1–56). As we have seen, this differential equa-
tion has two solutions: Legendre’s function Pnm and Legendre’s function of
the second kind Qnm . For ḡ(t), where t = cos ϑ, the Qnm (t) are ruled out for
obvious reasons, as we have seen in Sect. 1.8. For f¯(τ ), however, both sets of
functions Pnm (τ ) and Qnm (τ ) are possible solutions; they correspond to the
two different solutions f = r n and f = r −(n+1) in the spherical case. Finally,
(1–169) has as before the solutions cos mλ and sin mλ.
We summarize all individual solutions:
u u
f (u) = Pnm i or Qnm i ;
E E
g(ϑ) = P (cos ϑ) ; (1–172)
nm
P
u
b
F1 F2
E
a u2 + E 2
refere
nce ellipsoid
Vi (b, ϑ, λ) = Ve (b, ϑ, λ)
n
∞ (1–176)
= [anm Pnm (cos ϑ) cos mλ + bnm Pnm (cos ϑ) sin mλ] .
n=0 m=0
Formula (1–176) shows that not only functions that are defined on the
surface of a sphere can be expanded into a series of surface spherical har-
monics. Such an expansion is even possible for rather arbitrary functions
defined on a convex surface.
A remark on terminology
The ellipsoidal-harmonic coordinates u, ϑ (or β), λ are the generalization of
spherical coordinates for the sole use of getting closed solutions of Laplace’s
equation, in particular, for the gravity field of the reference ellipsoid in
Sect. 2.7. The brief name “ellipsoidal coordinates” frequently used for u, β, λ
might lead to a confusion with the ellipsoidal coordinates ϕ, λ, h. In the
present book, “ellipsoidal coordinates” will always denote “ellipsoidal ge-
ographic coordinates”, frequently also called “geodetic coordinates”, being
represented by ϕ, λ, h.
2 Gravity field of the earth
2.1 Gravity
The total force acting on a body at rest on the earth’s surface is the resultant
of gravitational force and the centrifugal force of the earth’s rotation and is
called gravity.
Take a rectangular coordinate system whose origin is at the earth’s center
of gravity and whose z-axis coincides with the earth’s mean axis of rotation
(Fig. 2.1). The x- and y-axes are so chosen as to obtain a right-handed
coordinate system; otherwise they are arbitrary. For convenience, we may
assume an x-axis which is associated with the mean Greenwich meridian (it
“points” towards the mean Greenwich meridian). Note that we are assuming
in this book that the earth is a solid body rotating with constant speed
around a fixed axis. This is a rather simplified assumption, see Moritz and
Mueller (1987). The centrifugal force f on a unit mass is given by
f = ω2p , (2–1)
is the distance from the axis of rotation. The vector f of this force has the
z
!
P
p
f
x
y p y
x
f = ω 2 p = [ω 2 x, ω 2 y, 0] . (2–4)
1 2 2
Φ= ω (x + y 2 ) , (2–5)
2
so that
∂Φ ∂Φ ∂Φ
f = grad Φ ≡ , , . (2–6)
∂x ∂y ∂z
If we combine this with Poisson’s equation (1–17) for V , we get the general-
ized Poisson equation for the gravity potential W :
∆W = −4π G + 2ω 2 . (2–9)
with components
∂W x−ξ
gx = = −G dv + ω 2 x ,
∂x l3
v
∂W y−η
gy = = −G dv + ω 2 y , (2–11)
∂y l3
v
∂W z−ζ
gz = = −G dv ,
∂z l3
v
is called the gravity vector; it is the total force (gravitational force plus
centrifugal force) acting on a unit mass. As a vector, it has magnitude and
direction.
The magnitude g is called gravity in the narrower sense. It has the phys-
ical dimension of an acceleration and is measured in gal (1 gal = 1 cm s−2 ),
the unit being named in honor of Galileo Galilei. The numerical value of g
is about 978 gal at the equator, and 983 gal at the poles. In geodesy, another
unit is often convenient – the milligal, abbreviated mgal (1 mgal = 10−3 gal).
In SI units, we have
The direction of the gravity vector is the direction of the plumb line, or the
vertical; its basic significance for geodetic and astronomical measurements
is well known.
In addition to the centrifugal force, another force called the Coriolis force
acts on a moving body. It is proportional to the velocity with respect to the
earth, so that it is zero for bodies resting on the earth. Since in classical
geodesy (i.e., not considering navigation) we usually deal with instruments
at rest relative to the earth, the Coriolis force plays no role here and need
not be considered.
equal within 10−11 , which is a formidable accuracy. He used the same type
of instrument by which experimental physicists have been able to determine
the numerical value of the gravitational constant G only to a poor accuracy
of about 10−4 , as we have seen at the beginning of this book. The coinci-
dence between the inertial and the gravitational mass was far too good to be
a physical accident, but, within classical mechanics, it was an inexplicable
miracle. It was not before 1915 that Einstein made it one of the pillars of
the general theory of relativity!
dW = grad W · dx = g · dx , (2–15)
where
dx = [dx, dy, dz] . (2–16)
If the vector dx is taken along the equipotential surface W = constant, then
the potential remains constant and dW = 0, so that (2–15) becomes
g · dx = 0 . (2–17)
If the scalar product of two vectors is zero, then these vectors are orthogonal
to each other. This equation therefore expresses the well-known fact that the
gravity vector is orthogonal to the equipotential surface passing through the
same point.
The surface of the oceans, after some slight idealization, is part of a
certain level surface. This particular equipotential surface was proposed as
the “mathematical figure of the earth” by C.F. Gauss, the “Prince of Mathe-
maticians”, and was later termed the geoid. This definition has proved highly
suitable, and the geoid is still frequently considered by many to be the fun-
damental surface of physical geodesy. The geoid is thus defined by
W = W0 = constant . (2–18)
2.2 Level surfaces and plumb lines 47
P
H
g
geoid
W = W0
level surface
W = constant
If we look at equation (2–7) for the gravity potential W , we can see that
the equipotential surfaces, expressed by W (x, y, z) = constant, are rather
complicated mathematically. The level surfaces that lie completely outside
the earth are at least analytical surfaces, although they have no simple ana-
lytical expression, because the gravity potential W is analytical outside the
earth. This is not true of level surfaces that are partly or wholly inside the
earth, such as the geoid. They are continuous and “smooth” (i.e., without
edges), but they are no longer analytical surfaces; we will see in the next sec-
tion that the curvature of the interior level surfaces changes discontinuously
with the density.
The lines that intersect all equipotential surfaces orthogonally are not
exactly straight but slightly curved (Fig. 2.2). They are called lines of force,
or plumb lines. The gravity vector at any point is tangent to the plumb line at
that point, hence “direction of the gravity vector”, “vertical”, and “direction
of the plumb line” are synonymous. Sometimes this direction itself is briefly
denoted as “plumb line”.
As the level surfaces are, so to speak, horizontal everywhere, they share
the strong intuitive and physical significance of the horizontal; and they share
the geodetic importance of the plumb line because they are orthogonal to
it. Thus, we understand why so much attention is paid to the equipotential
surfaces.
The height H of a point above sea level (also called the orthometric
height) is measured along the curved plumb line, starting from the geoid
48 2 Gravity field of the earth
(Fig. 2.2). If we take the vector dx along the plumb line, in the direction of
increasing height H, then its length will be
dx = dH (2–19)
and its direction is opposite to the gravity vector g, which points downward,
so that the angle between dx and g is 180◦ . Using the definition of the scalar
product (i.e., for two vectors a and b it is defined as a · b = ab cos ω,
where ω is the angle between the two vectors), we get
g · dx = g dH cos 180◦ = −g dH (2–20)
accordingly, so that Eq. (2–15) becomes
dW = −g dH . (2–21)
This equation relates the height H to the potential W and will be basic for
the theory of height determination (Chap. 4). It shows clearly the insepara-
ble interrelation that characterizes geodesy – the interrelation between the
geometrical concepts (H) and the dynamic concepts (W ).
Another form of Eq. (2–21) is
∂W
g=− . (2–22)
∂H
It shows that gravity is the negative vertical gradient of the potential W , or
the negative vertical component of the gradient vector grad W .
Since geodetic measurements (theodolite measurements, leveling, but
also satellite techniques etc.) are almost exclusively referred to the system
of level surfaces and plumb lines, the geoid plays an essential part. Thus,
we see why the aim of physical geodesy has been formulated as the de-
termination of the level surfaces of the earth’s gravity field. In a still more
abstract but equivalent formulation, we may also say that physical geodesy
aims at the determination of the potential function W (x, y, z). At a first
glance, the reader is probably perplexed about this definition, which is due
to Bruns (1878), but its meaning is easily understood: If the potential W is
given as a function of the coordinates x, y, z, then we know all level surfaces
including the geoid; they are given by the equation
W (x, y, z) = constant. (2–23)
y
P
|| x
y = f (x )
dy d2 y
y = , y = 2 . (2–25)
dx dx
If we use a plane local coordinate system xy in which a parallel to the x-axis
is tangent at the point P under consideration (Fig. 2.3), then this implies
y = 0 and we get simply
1 d2 y
κ= = 2. (2–26)
dx
Level surfaces
Consider now a point P on a level surface S. Take a local coordinate system
xyz with origin at P whose z-axis is vertical, that is, orthogonal to the
surface S (Fig. 2.4). We intersect this level surface
z
plumb
line
y
x
P
level surface
Comparing Fig. 2.4 with Fig. 2.3, we see that z now takes the place of
y. Therefore, instead of (2–26) we have for the curvature of the intersection
of the level surface with the xz-plane:
d2 z
K1 = . (2–29)
dx2
If we differentiate W (x, y, z) = W0 with respect to x, considering that y
is zero and z is a function of x, we get
dz
Wx + Wz = 0,
dx
2 (2–30)
dz dz d2 z
Wxx + 2Wxz + Wzz + Wz 2 = 0 ,
dx dx dx
where the subscripts denote partial differentiation:
∂W ∂2W
Wx = , Wxz = , ... . (2–31)
∂x ∂x ∂z
Since the x-axis is tangent at P , we get dz/dx = 0 at P , so that
d2 z Wxx
=− . (2–32)
dx2 Wz
Since the z-axis is vertical, we have, using (2–22),
∂W ∂W
Wz = = = −g . (2–33)
∂z ∂H
Therefore, Eq. (2–29) becomes
Wxx
K1 = . (2–34)
g
The curvature of the intersection of the level surface with the yz-plane is
found by replacing x with y:
Wyy
K2 = . (2–35)
g
The mean curvature J of a surface at a point P is defined as the arith-
metic mean of the curvatures of the curves in which two mutually perpen-
dicular planes through the surface normal intersect the surface (Fig. 2.5).
Hence, we find
1 Wxx + Wyy
J = − (K1 + K2 ) = − . (2–36)
2 2g
2.3 Curvature of level surfaces and plumb lines 51
surface normal
Here the minus sign is only a convention. This is an expression for the mean
curvature of the level surface.
From the generalized Poisson equation
we find
−2g J + Wzz = −4π G + 2ω 2 . (2–38)
Considering
∂g ∂g
Wz = −g , Wzz = − =− , (2–39)
∂z ∂H
we finally obtain
∂g
= −2g J + 4π G − 2ω 2 . (2–40)
∂H
This important equation, relating the vertical gradient of gravity ∂g/∂H to
the mean curvature of the level surface, is also due to Bruns (1878). It is
another beautiful example of the interrelation between the geometric and
dynamic concepts in geodesy.
Plumb lines
The curvature of the plumb line is needed for the reduction of astronomical
observations to the geoid. A plumb line may be defined as a curve whose
line element vector
dx = [dx, dy, dz] (2–41)
g = [Wx , Wy , Wz ] ; (2–42)
52 2 Gravity field of the earth
that is, dx and g differ only by a proportionality factor. This is best expressed
in the form
dx dy dz
= = . (2–43)
Wx Wy Wz
In the coordinate system of Fig. 2.4, the curvature of the projection of
the plumb line onto the xz-plane is given by
d2 x
κ1 = ; (2–44)
dz 2
this is equation (2–26) applied to the present case. Using (2–43), we have
dx Wx
= . (2–45)
dz Wz
We differentiate with respect to z, considering that y = 0:
d2 x 1 dx dx
= 2 Wz Wxz + Wxx − Wx Wzz + Wzx . (2–46)
dz 2 Wz dz dz
In our particular coordinate system, the gravity vector coincides with the
z-axis, so that its x- and y-components are zero:
Wx = Wy = 0 . (2–47)
where n is the unit vector along the plumb line (its unit tangent vector) and
n1 is the unit vector along the principal normal to the plumb line. This may
be easily verified. Using the local xyz-system, we have
n = [0, 0, 1] ,
(2–54)
n1 = [cos α, sin α, 0] ,
where α is the angle between the principal normal and the x-axis (Fig. 2.6).
The z-component of (2–53) yields Bruns’ equation (2–40), and the horizontal
components yield
∂g ∂g
= g κ cos α , = g κ sin α . (2–55)
∂x ∂y
These are identical to (2–50) and (2–51), since κ1 = κ cos α and κ2 = κ sin α,
as differential geometry shows. Equation (2–53) is called the generalized
Bruns equation.
More about the curvature properties and the “inner geometry” of the
gravitational field will be found in books by, e.g., Hotine (1969: Chaps. 4–
20), Marussi (1985) and Moritz and Hofmann-Wellenhof (1993: Chap. 3).
z
plumb
n line
sin ® y
®
®
1
cos
n1
x
unit sphere
N
vertical
n
P F
G L
F
eridian
Greenwich m
equator
earth
which follow from integrating (2–21). The integral is taken along the plumb
line of point P , starting from the geoid, where H = 0 and W = W0 (see also
Fig. 2.8).
The quantities
Φ, Λ, W or Φ, Λ, H (2–57)
are called natural coordinates. They are the real-earth counterparts of the
ellipsoidal coordinates. They are related in the following way to the geocen-
tric rectangular coordinates x, y, z of Sect. 2.1. The x-axis is associated with
the mean Greenwich meridian; from Fig. 2.7 we read that the unit vector of
the vertical n has the xyz-components
g = [Wx , Wy , Wz ] . (2–59)
earth's surface
P
level surface
W = constant
H
sea level geoid
0 W = W0
so that
g = −g n . (2–61)
This equation, together with (2–58) and (2–59), gives
−Wz
Φ = tan−1 ,
Wx2 + Wy2
Wy (2–63)
Λ = tan−1 ,
Wx
W = W (x, y, z) .
r
l
Ã
O r'
dM
where we now denote the mass element by dM ; the integral is extended over
the entire earth. Into this integral we substitute the expression (1–104):
∞
1 r n
= Pn (cos ψ) , (2–65)
l r n+1
n=0
These formulas are very symmetrical and easy to remember: the coefficient,
multiplied by 2n + 1, of the solid harmonic
R̄nm (ϑ, λ)
(2–73)
r n+1
is the integral of the solid harmonic
the actual evaluation of the integrals requires that the density be expressed
as a function of r , ϑ , λ . Although no such expression is available at present,
this fact does not diminish the theoretical and practical significance of spher-
ical harmonics, since the coefficients Anm , Bnm can be determined from the
boundary values of gravity at the earth’s surface. This is a boundary-value
problem (see Sect. 1.13) and will be elaborated later.
Recalling the relations (1–91) and (1–98) between conventional and fully
normalized spherical harmonics, we can also write equations (2–71) and (2–
72) in terms of conventional harmonics, readily obtaining
n
∞
Rnm (ϑ, λ) Snm (ϑ, λ)
V = Anm + Bnm , (2–76)
r n+1 r n+1
n=0 m=0
where
An0 =G r n Pn (cos ϑ ) dM ;
earth
⎫
(n − m)! ⎪
⎪
Anm = 2 G r n Rnm (ϑ , λ ) dM ⎪
⎪ (2–77)
(n + m)! ⎪
⎬
earth
(m = 0) .
(n − m)! ⎪
⎪
Bnm = 2 G r n Snm (ϑ , λ ) dM ⎪
⎪
(n + m)! ⎪
⎭
earth
Anm = GM an Cnm
(n = 0) . (2–79)
Bnm = GM an Snm
Distinguish the coefficient Snm and the function Snm ! The coefficient Cn0
has formerly been denoted by −Jn . Note that C is related to cosine and S
is related to sine.
60 2 Gravity field of the earth
1
C̄n0 = √ Cn0 ,
2n + 1
⎫
(n + m)! ⎪
⎪
⎪
⎪
C̄nm = Cnm ⎪
⎪ (2–80)
2(2n + 1)(n − m)! ⎬
(m = 0)
⎪
⎪
(n + m)! ⎪
⎪
S̄nm = Snm ⎪
⎪
2(2n + 1)(n − m)! ⎭
r = r0
r0
the same dimensions, then the series for V would indeed still converge at the
surface of the earth. Owing to the mass irregularities, however, the series of
the actual potential V of the earth can be divergent or also convergent at the
surface of the earth. Theoretically, this makes the use of a harmonic expan-
sion of V at the earth’s surface somewhat difficult; practically, it is always
safe to regard it as convergent. For a detailed discussion see Moritz (1980 a:
Sects. 6 and 7) and Sect. 8.6 herein.
It need hardly be pointed out that the spherical-harmonic expansion,
always expressing a harmonic function, can represent only the potential out-
side the attracting masses, never inside.
r 2 S22 = 6r 2 sin2 ϑ sin λ cos λ = 6(r sin ϑ cos λ)(r sin ϑ sin λ) = 6xy . (2–82)
r R10 = z , r S10 = 0 ,
r R11 = x , r S11 = y ,
(2–83)
r 2 R20 = − 12 x2 − 12 y 2 + z 2 , r 2 S20 = 0 ,
r 2 R21 = 3x z , r 2 S21 = 3y z ,
Substituting these functions into the expression (2–77) for the coefficients
Anm and Bnm yields for the zero-degree term
A00 = G dM = GM ; (2–84)
earth
that is, the product of the mass of the earth times the gravitational constant.
For the first-degree coefficients, we get
A10 = G z dM , A11 = G x dM , B11 = G y dM ;
earth earth earth
(2–85)
and for the second-degree coefficients
1
A20 = G (−x2 − y 2 + 2z 2 ) dM ,
2
earth
A21 = G x z dM , B21 = G y z dM , (2–86)
earth earth
1 1
A22 = G (x2 − y 2 ) dM , B22 = G x y dM .
4 2
earth earth
are the products of inertia. They are zero if the coordinate axes coincide with
the principal axes of inertia. If the z-axis is identical with the mean rotational
axis of the earth, which coincides with the axis of maximum inertia, at least
the second and third of these products of inertia must vanish. Hence, A21 and
B21 will be zero, but not so B22 , which is proportional to the first product of
2.6 Harmonics of lower degree 63
inertia; B22 would vanish only if the earth had complete rotational symmetry
or if a principal axis of inertia happened to coincide with the Greenwich
meridian.
The five harmonics A10 R10 , A11 R11 , B11 S11 , A21 R21 , and B21 S21 –
all first-degree harmonics and those of degree 2 and order 1 – which must,
thus, vanish in any spherical-harmonic expansion of the earth’s potential,
are called forbidden or inadmissible harmonics.
Introducing the moments of inertia with respect to the x-, y-, z-axes by
the definitions
A= (y 2 + z 2 ) dM ,
B= (z 2 + x2 ) dM , (2–89)
C= (x2 + y 2 ) dM ,
we finally have
A00 = GM ,
A22 = 14 G (B − A) ,
B22 = 12 G D .
Now let the x- and y-axes actually coincide with the corresponding prin-
cipal axes of inertia of the earth. This is only theoretically possible, since
the principal axes of inertia of the earth are only inaccurately known. Then
B22 = 0; taking into account (2–78) and (2–79), we may write explicitly
GM G 1
V = + 3 C − (A + B)/2 (1 − 3 cos2 ϑ) +
r r 2
(2–92)
3
(B − A) sin ϑ cos 2λ + O(1/r 4 ) .
2
4
64 2 Gravity field of the earth
which is obtained by taking into account the relations (1–26) between rect-
angular and spherical coordinates.
Terms of order higher than 1/r 3 may be neglected for larger distances
(say, for the distance to the moon), so that (2–92) or (2–93), omitting the
higher-order terms 0(1/r 4 ), are sufficient for many astronomical purposes,
cf. Moritz and Mueller (1987). Note that the notation 0(1/r 4 ) means terms
of the order of 1/r 4 . For planetary distances even the first term,
GM
V = , (2–94)
r
is generally sufficient; it represents the potential of a point mass. Thus, for
very large distances, every body acts like a point mass.
Using the form (2–78) of the spherical-harmonic expansion of V , then
the coefficients of lower degree are obtained from (2–79) and (2–91). We find
C − (A + B)/2
C20 = − ,
M a2
C21 = S21 = 0 , (2–95)
B−A
C22 = ,
4M a2
D
S22 = .
2M a2
The first of these formulas shows that the summation in (2–78) actually
begins with n = 2; the others relate the coefficients of second degree to the
mass and the moments and products of inertia of the earth.
U = U (x, y, z) , (2–96)
we see that the level ellipsoid, being a surface U = constant, exactly corre-
sponds to the geoid, which is defined as a surface W = constant.
The basic point here is that by postulating that the given ellipsoid be
an equipotential surface of the normal gravity field, and by prescribing the
total mass M , we completely and uniquely determine the normal potential
U . The detailed density distribution inside the ellipsoid, which produces the
potential U , is quite uninteresting and need not be known at all. In fact,
we do not know of any “reasonable” mass distribution for the level ellipsoid
(Moritz 1990: Chap. 5). Pizzetti (1894) unsuccessfully used a homogeneous
density distribution combined with a surface layer of negative density, which
is quite “unnatural”.
This determination is possible by Dirichlet’s principle (Sect. 1.12): The
gravitational potential outside a surface S is completely determined by know-
ing the geometric shape of S and the value of the potential on S. Originally
it was shown only for the gravitational potential V , but it can be applied to
the gravity potential
U = V + 12 ω 2 (x2 + y 2 ) (2–97)
as well if the angular velocity ω is given. The proof follows that in Sect. 1.12,
with obvious modifications. Hence, the normal potential function U (x, y, z)
is completely determined by
1. the shape of the ellipsoid of revolution, that is, its semiaxes a and b,
2. the total mass M , and
3. the angular velocity ω.
The calculation will now be carried out in detail. The given ellipsoid S0 ,
x2 + y 2 z 2
+ 2 = 1, (2–98)
a2 b
66 2 Gravity field of the earth
U (x, y, z) = U0 . (2–99)
This equation applies for all points of S0 , that is, for all values of β. Since
b2 + E 2 = a2 (2–105)
and
cos2 β = 2
3 1 − P2 (sin β) , (2–106)
we have
∞
An Pn (sin β) + 1
3 ω 2 a2 − 13 ω 2 a2 P2 (sin β) − U0 = 0 (2–107)
n=0
2.7 The gravity field of the level ellipsoid 67
or
A0 + 13 ω 2 a2 − U0 P0 (sin β) + A1 P1 (sin β)
∞
+ A2 − 1
3 ω 2 a2 P2 (sin β) + An Pn (sin β) = 0 . (2–108)
n=3
This equation applies for all values of β only if the coefficient of every
Pn (sin β) is zero. Thus, we get
A0 = U0 − 13 ω 2 a2 , A1 = 0 ,
(2–109)
1
A2 = 3 ω 2 a2 , A3 = A4 = . . . = 0 .
Substituting these relations into (2–100) gives
u u
Q0 i Q2 i
V (u, β) = U0 − 13 ω 2 a2 E + 13 ω 2 a2 E P2 (sin β) . (2–110)
b b
Q0 i Q2 i
E E
This formula is basically the solution of Dirichlet’s problem for the level
ellipsoid, but we can give it more convenient forms. It is a closed formula!
First, we determine the Legendre functions of the second kind, Q0 and
Q2 . As
1 1
coth−1 (i x) = cot−1 x = −i tan−1 , (2–111)
i x
we find by (1–80) with z = i u/E:
u E
Q0 i = −i tan−1 ,
E u
(2–112)
u i u2 −1 E u
Q2 i = 1 + 3 2 tan −3 .
E 2 E u E
By introducing in (2–112) the abbreviations
1 u2 −1 E u
q = 1 + 3 2 tan −3 ,
2 E u E
(2–113)
1 b2 −1 E b
q0 = 1 + 3 2 tan −3
2 E b E
and substituting them in equation (2–110), we obtain
E
tan−1
u + q
V (u, β) = U0 − 13 ω 2 a2 1
3 ω 2 a2 q0 P2 (sin β) . (2–114)
E
tan−1
b
68 2 Gravity field of the earth
x2 + y 2 + z 2 = r 2 = u2 + E 2 cos2 β , (2–116)
Expressing P2 as
P2 (sin β) = 3
2 sin2 β − 1
2 (2–125)
and, finally, adding the centrifugal potential Φ = ω 2 (u2 + E 2 ) cos2 β/2 from
(2–102), the normal gravity potential U results as
GM E q
2
U (u, β) = tan−1 + 12 ω 2 a2 sin β − 13 + 1
2 ω 2 (u2 + E 2 ) cos2 β .
E u q0
(2–126)
The only constants that occur in this formula are a, b, GM , and ω. This is
in complete agreement with Dirichlet’s theorem.
where
u2 + E 2 sin2 β
w= (2–128)
u2 + E 2
has been introduced. Thus, along the coordinate lines we have
γ = grad U (2–130)
∂U 1 ∂U
γu = = ,
∂su w ∂u
∂U 1 ∂U
γβ = = √ , (2–131)
∂sβ w u + E ∂β
2 2
∂U 1 ∂U
γλ = =√ = 0.
∂sλ u + E cos β ∂λ
2 2
70 2 Gravity field of the earth
The component γλ is zero because U does not contain λ. This is also evident
from the rotational symmetry.
Performing the partial differentiations, we find
GM ω 2 a2 E q
1
−w γu = 2 2
+ 2 2 2 sin2 β − 16 − ω 2 u cos2 β ,
u +E u + E q0
(2–132)
ω 2 a2 q
−w γβ = − √ 2 2 2
+ ω u + E sin β cos β ,
u2 + E 2 q0
where we have set
u2 + E 2 dq u2 u E
q = − =3 1+ 2 1− tan−1 − 1. (2–133)
E du E E u
Note that q does not mean dq/du; this notation has been borrowed from Hir-
vonen (1960), where q is the derivative with respect to another independent
variable which we are not using here.
For the level ellipsoid S0 itself, we have u = b and get
γβ,0 = 0 . (2–134)
we obtain
GM
γ= ·
a a2 sin2 β + b2 cos2 β
(2–140)
m e q0 m e q0
· 1+ sin β + 1 − m −
2 2
cos β .
3 q0 6 q0
a γb sin2 β + b γa cos2 β
γ= . (2–144)
a2 sin2 β + b2 cos2 β
We finally introduce the ellipsoidal latitude on the ellipsoid, ϕ, which is the
angle between the normal to the ellipsoid and the equatorial plane (Fig. 2.11).
Using the formula from ellipsoidal geometry,
72 2 Gravity field of the earth
b
¯
' '
a O
N
b
tan β = tan ϕ , (2–145)
a
we obtain
a γa cos2 ϕ + b γb sin2 ϕ
γ= . (2–146)
a2 cos2 ϕ + b2 sin2 ϕ
The computation is left as an exercise for the reader. This rigorous formula
for normal gravity on the ellipsoid is due to Somigliana from 1929.
We close this section with a short remark on the vertical gradient of
gravity at the reference ellipsoid, ∂γ/∂su = ∂γ/∂h. Bruns’ formula (2–40),
applied to the normal gravity field with the corresponding ellipsoidal height
h and with = 0, yields
∂γ
= −2γ J − 2ω 2 . (2–147)
∂h
The mean curvature of the ellipsoid is given by
1 1 1
J= + , (2–148)
2 M N
where M and N are the principal radii of curvature: M is the radius in the
direction of the meridian, and N is the normal radius of curvature, taken
in the direction of the prime vertical. From ellipsoidal geometry, we use the
formulas
c c
M= 2 2 3/2
, N= 2
, (2–149)
(1 + e cos ϕ) (1 + e cos2 ϕ)1/2
2.9 Expansion of the normal potential in spherical harmonics 73
where
a2
c= (2–150)
b
is the radius of curvature at the pole. The normal radius of curvature, N ,
admits a simple geometrical interpretation (Fig. 2.11). It is, therefore, also
known as the “normal terminated by the minor axis” (Bomford 1962: p. 497).
r cos ϑ = u sin β .
The longitude λ is the same in both systems. We easily find from these
equations
u
cot ϑ = √ tan β ,
u + E2
2
(2–153)
2 2 2
r = u + E cos β .
The direct transformation of (2–151) by expressing u and β in terms of
r and ϑ by means of equations (2–153) is extremely laborious. However, the
problem can be solved easily in an indirect way.
We expand tan−1 (E/u) into the well-known power series
E E 1 E 3 1 E 5
tan−1 = − + − ... . (2–154)
u u 3 u 5 u
The substitution of this series into the first equation of formula (2–113), i.e.,
1 u2 −1 E u
q= 1 + 3 2 tan −3 , (2–155)
2 E u E
74 2 Gravity field of the earth
Pn (1) = 1 (2–163)
(see also Fig. 1.4). Comparing the coefficients in both expressions for V , we
2n
find
n GM E 2n m e
A2n = (−1) 1− . (2–164)
2n + 1 2n + 3 3q0
Equations (2–160) and (2–164) give the desired expression for the potential
of the level ellipsoid as a series of spherical harmonics.
The second-degree coefficient A2 is
A2 = G (A − C) . (2–165)
Here we have introduced the first eccentricity e = E/a. For n = 1 this gives
the important formula
C −A
C20 = − (2–171)
M a2
or, equivalently,
C−A
J2 = , (2–172)
M a2
which is in agreement with the respective relation in (2–95) when taking into
account the rotational symmetry causing A = B.
Finally, we note that on eliminating q0 = (1/i) Q2 (i(b/E)) by using
Eq. (2–167), and U0 by using Eq. (2–122), we may write the expansion of V
in ellipsoidal harmonics, Eq. (2–110), in the form
i u
V (u, β) = GM Q0 i
E E
15i
u
+ G C − A − 1
3 M E 2 Q
2 i P2 (sin β) .
2E 3 E
(2–173)
This shows that the coefficients of the ellipsoidal harmonics of degrees zero
and two are functions of the mass and of the difference between the two
principal moments of inertia. The analogy to the corresponding spherical-
harmonic coefficients (2–91) is obvious. This is a closed formula, not a trun-
cated series!
and similar parameters that characterize the deviation from a sphere are
small. Therefore, series expansions in terms of these or similar parameters
will be convenient for numerical calculations.
Linear approximation
In order that the readers may find their way through the subsequent practical
formulas, we first consider an approximation that is linear in the flattening f .
Here we get particularly simple and symmetrical formulas which also exhibit
plainly the structure of the higher-order expansions.
It is well known that the radius vector r of an ellipsoid is approximately
given by
r = a (1 − f sin2 ϕ) . (2–175)
As we will see subsequently, normal gravity may, to the same approximation,
be written
γ = γa (1 + f ∗ sin2 ϕ) . (2–176)
For ϕ = ±90◦ , at the poles, we have r = b and γ = γb . Hence, we may write
b = a (1 − f ) , γb = γa (1 + f ∗ ) , (2–177)
and solving for f and f ∗ , we obtain
a−b
f = ,
a
(2–178)
γb − γa
f∗ = ,
γa
so that f is the flattening defined by (2–174), and f ∗ is an analogous quantity
which may be called gravity flattening.
To the same approximation, (2–143) becomes
f + f∗ = 5
2 m, (2–179)
where
. ω2a centrifugal force at equator
m= = . (2–180)
γa gravity at equator
This is Clairaut’s theorem in its original form. It is one of the most striking
formulas of physical geodesy: the (geometrical) flattening f in (2–178) can
be derived from f ∗ and m, which are purely dynamical quantities obtained
by gravity measurements; that is, the flattening of the earth can be obtained
from gravity measurements.
Clairaut’s formula is only a first approximation and must be improved,
first by the inclusion of higher-order ellipsoidal terms in f , and secondly by
taking into account the deviation of the earth’s gravity field from the normal
gravity field. But the principle remains the same.
78 2 Gravity field of the earth
Second-order expansion
We now expand the closed formulas of the two preceding sections into series
in terms of the second numerical eccentricity e and the flattening f , in
general up to and including e4 or f 2 . Terms of the order of e6 or f 3 and
higher will usually be neglected.
We start from the series
E E 1 E 3 1 E 5 1 E 7
tan−1 = − + − + ··· ,
u u 3 u 5 u 7 u
3 5 7
1 E 2 E 3 E
q=2 − + − ··· , (2–181)
3·5 u 5·7 u 7·9 u
3 5 7
1 E 1 E 1 E
q = 6 − + − ··· .
3·5 u 5·7 u 7·9 u
The first two series have already been used in the preceding section in (2–
154) and (2–156), respectively; the third is obtained by substituting the
tan−1 series into the closed formula (2–133) for q .
On the reference ellipsoid S0 , we have u = b and
E E
= = e , (2–182)
u b
so that
tan−1 e = e − 13 e3 + 1
5 e5 · · · ,
q0 = 2
15 e3 1 − 67 e2 · · · ,
(2–183)
q0 = 25 e2 1 − 37 e2 · · · ,
e q0
= 3 1 + 37 e2 · · · .
q0
a
b= √ = a 1 − 12 e2 + 38 e4 · · · . (2–184)
1+e2
2.10 Series expansions for the normal gravity field 79
ω2a 3
=m+ 2 m2 , (2–188)
γa
which gives the mass in terms of equatorial gravity. Using this equation, we
can express GM in Eq. (2–185) in terms of γa , obtaining
U0 = a γa 1 − 13 e2 + 11 1 4 2 2
6 m+ 5e − 7e m+ 4 m .
11 2 (2–190)
(2–198)
1 2 5
+ − 2 f + 2 f m sin ϕ . 4
so that we have
f2 = −f + 52 m + 12 f 2 − 26
7 fm+ 15
4 m2 ,
(2–200)
f4 = − 12 f2 + 5
2 f m.
By substituting
sin4 ϕ = sin2 ϕ − 1
4 sin2 2ϕ , (2–201)
2.10 Series expansions for the normal gravity field 81
we finally obtain
γ = γa (1 + f ∗ sin2 ϕ − 1
4 f4 sin2 2ϕ) , (2–202)
where
γb − γa
f∗ = = f2 + f4 (2–203)
γa
is the “gravity flattening”.
C −A 1 2 m e
= − . (2–204)
M E2 3 45 q0
Expanding q0 by means of (2–183), we find
C −A 1
2
= 2 13 e2 − 13 m − 27 e2 m . (2–205)
ME e
Substituting this into (2–170) yields
C − A 1 2
−C20 = J2 = =3e − 1
3 m − 13 e4 + 1
21 e2 m
M E2 (2–206)
= f − m−
2
3
1
3
1
3 f2 + 2
21 f m,
−C40 = J4 = − 15 e4 + 2
7 e2 m = − 45 f 2 + 47 f m . (2–207)
The higher C or J, respectively, are already of an order of magnitude that
we have neglected.
1 b
3/2 b
= 2 1 + e2 cos2 ϕ = 2 1 + 32 e2 cos2 ϕ · · · ,
M a a
(2–210)
1 b
1/2 b
= 2 1 + e2 cos2 ϕ = 2 1 + 12 e2 cos2 ϕ · · · ,
N a a
we have
1 1 b
2b
+ = 2 2 + 2e2 cos2 ϕ = 2 (1 + 2f cos2 ϕ) . (2–211)
M N a a
Here we have limited ourselves to terms linear in f , since the elevation h is
already a small quantity. Thus, we find from (2–209) after simple manipula-
tions:
∂γ 2γ
= − (1 + f + m − 2f sin2 ϕ) . (2–212)
∂h a
The second derivative ∂ 2 γ/∂h2 may be taken from the spherical approxima-
tion, obtained by neglecting e2 or f :
Using Eq. (2–198) for γ, we may also write the difference γh − γ in the form
2γa
3γa
γh − γ = − 1 + f + m + − 3f + 52 m sin2 ϕ) h + 2 h2 . (2–216)
a a
The symbol γh denotes the normal gravity for a point at latitude ϕ, situated
at height h above the ellipsoid; γ is the gravity at the ellipsoid itself, for the
same latitude ϕ, as given by (2–202) or equivalent formulas.
Second-order series developments for the inner gravity field are found in
Moritz (1990: Chap. 4); this is the main reason for such a development here,
because today one uses the closed formulas wherever possible.
2.11 Reference ellipsoid – numerical values 83
γ = 978.0490 (1 + 0.005 2884 sin2 ϕ − 0.000 0059 sin2 2ϕ) gal , (2–219)
b = 6 356 912 m ,
E = 522 976 m ,
(2–220)
e2 = 0.006 7682 ,
m = 0.003 4499 .
a = 6 378 245 m ,
(2–223)
f = 1/298.3 .
Contemporary data
After the start of Sputnik, the first artificial satellite, in 1957, the Interna-
tional Astronomical Union, in 1964, adopted a new set of constants, among
them a = 6 378 160 m and f = 1/298.25. The value of a, which is consid-
erably smaller than that for the International Ellipsoid, incorporates astro-
geodetic determinations; the change in the value of J2 , and consequently of
f , is due to the results from artificial satellites.
In 1967, these values were taken by the International Union of Geodesy
and Geophysics (IUGG) as the Geodetic Reference System 1967.
This decision was soon seen to be wrong; especially the value of a was
recognized to be too large: now we believe to be on the order of 6 378 137 m,
the value of the Geodetic Reference System 1980 (GRS 1980) and, based on
it, the World Geodetic System 1984 (WGS 84). More details of these two
systems are given below.
Table 2.1. Note that these parameters, as given in the table, are defined
as exact! Note also that GM , the “geocentric gravitational constant” of the
earth, may also more figuratively be denoted as “product of the (Newtonian)
gravitational constant and the earth’s mass”.
On the basis of these defining parameters and by the computational
formulas given in Moritz (1980 b), the geometrical and physical constants of
Table 2.2 may be derived.
The GRS 1980 is still (2005) valid as the official reference system of the
IUGG and it forms the fundamental basis of the WGS 84.
Physical constants
U0 = 62 636 860.850 m2 s−2 normal potential at the ellipsoid
J4 = −0.000 002 370 912 22 spherical-harmonic coefficient
J6 = 0.000 000 006 083 47 spherical-harmonic coefficient
J8 = −0.000 000 000 014 27 spherical-harmonic coefficient
m = 0.003 449 786 003 08 m = ω 2 a2 b/(GM )
γa = 9.780 326 7715 m s−2 normal gravity at the equator
γb = 9.832 186 3685 m s−2 normal gravity at the pole
On the basis of this information, the former U.S. Defense Mapping Agency
88 2 Gravity field of the earth
(DMA) has proposed to replace the value of G in the WGS 84 by the standard
IERS value and to refine the coordinates of the GPS tracking stations. The
revised WGS 84, valid since January 2, 1994, has been given the designation
WGS 84 (G 730), where the ‘G’ indicates that the respective coordinates
used were obtained through GPS and the following number 730 indicates
the GPS week number when DMA has implemented the refined system.
In 1996, the U.S. National Imagery and Mapping Agency (NIMA) – the
successor of DMA – has implemented a revised version of the frame denoted
as WGS 84 (G 873) and being valid since September 29, 1996. The frame is
realized by monitor stations with refined coordinates. The associated ellip-
soid and its gravity field are now defined by the four parameters a, f, GM, ω,
which are slightly different compared to the respective ITRF values, e.g., the
current WGS 84 (G 873) frame and the ITRF97 show insignificant systematic
differences of less than 2 cm. Hence, they are virtually identical.
Note that the refinements applied to the WGS 84 reference frame have
reduced the uncertainties in the coordinates of the frame, the uncertainty
of the gravitational model, and the uncertainty of the geoid undulations;
however, they have not changed the WGS 84 coordinate system in the sense
of definition !
More general, the relationship between the WGS 84 and the ITRF is
characterized by two statements: (1) WGS 84 and ITRF are consistent; (2)
the differences between WGS 84 and ITRF are in the centimeter range world-
wide (National Imagery and Mapping Agency 2000).
However, if a transformation between reference frames is required, this
is accomplished by a datum transformation (see Sect. 5.7).
Physical constants
U0 = 62 636 851.7146 m2 s−2 normal potential at the ellipsoid
γa = 9.780 325 3359 m s−2 normal gravity at the equator
γb = 9.832 184 9378 m s−2 normal gravity at the pole
γ̄ = 9.797 643 2222 m s−2 mean value of normal gravity
M = 5.973 3328 · 1024 kg mass of the earth (includes atmosphere)
m = 0.003 449 786 506 84 m = ω 2 a2 b/(GM )
Some history (even if only some years old) is important here because
the parameters selected to originally define the WGS 84 reference ellipsoid
were the semimajor axis a, the product of the earth’s mass and the grav-
itational constant GM (also denoted as “geocentric gravitational constant
of the earth”), the normalized second-degree zonal gravitational coefficient
C̄20 , and the earth’s angular velocity ω. Due to significant refinements of
these original defining parameters, the DMA recommended, e.g., a refined
value for the GM parameter.
Anyway, a decision was made to retain the original WGS 84 reference
ellipsoid values for the semimajor axis a = 6 378 137 m and for the flattening
f = 1/298.257 223 563. For this reason, the four defining parameters were
chosen to be a, f, GM, ω.
Readers who like some confusion may continue right here; otherwise skip
this short paragraph. Due to this new choice of the defining parameters,
there are in addition two distinct values for the C̄20 term, one is dynami-
cally derived and the other geometrically by the defining parameters. The
90 2 Gravity field of the earth
n n'
P geoid
W = W0
N gP
Q reference
ellipsoid
°Q U = W0
∆g = gP − γ Q . (2–227)
∆g = gP − γQ ; (2–228)
al
line
orm
||z
ln
mb
n
da
n'
plu
so i
ip
L–¸
ell
ian
L–¸ »
ri d
co
me
s'
´ F
wich
Green
P
'
¸
L
define the direction of the plumb line n or of the gravity vector g, can be
determined by astronomical measurements. The ellipsoidal coordinates (or
geodetic coordinates in the sense of geographical coordinates on the ellip-
soid) given by the direction of the ellipsoidal normal n have been denoted
by ϕ and λ – these coordinates should not be confused with the ellipsoidal-
harmonic coordinates of Sect. 1.15 ! It is evident that this λ is identical with
the geocentric longitude (and also with the ellipsoidal-harmonic longitude).
Thus,
ξ = Φ − ϕ,
(2–230)
η = (Λ − λ) cos ϕ .
It is also possible to compare the vectors g and γ at the same point P . Then
we get the gravity disturbance vector
δg = gP − γ P . (2–231)
2.12 Anomalous gravity field 93
δg = gP − γP . (2–232)
The difference in direction – i.e., the deflection of the vertical – is the same
as before, since the directions of γP and γQ coincide virtually.
The gravity disturbance is conceptually even simpler than the gravity
anomaly, but it has not been that important in terrestrial geodesy. The
significance of the gravity anomaly is that it is given directly: the gravity
g is measured on the geoid (or reduced to it), see Chap. 3, and the normal
gravity γ is computed for the ellipsoid.
Relations
There are several basic mathematical relations between the quantities just
defined. Since
∂U
UP = UQ + N = UQ − γ N , (2–233)
∂n Q
we have
WP = UP + TP = UQ − γ N + TP . (2–234)
Because
WP = UQ = W0 , (2–235)
we find
T =γN (2–236)
(where we have omitted the subscript P on the left-hand side) or
T
N= . (2–237)
γ
This is the famous Bruns formula, which relates the geoidal undulation to
the disturbing potential.
94 2 Gravity field of the earth
g = grad W ,
(2–238)
γ = grad U ,
Then
∂W ∂U . ∂U
g=− , γ=− =− , (2–240)
∂n ∂n ∂n
because the directions of the normals n and n almost coincide. Therefore,
the gravity disturbance is given by
∂W ∂U . ∂W ∂U
δg = gP − γP = − − =− − (2–241)
∂n ∂n ∂n ∂n
or
∂T
δg = − . (2–242)
∂n
Since the elevation h is reckoned along the normal, we may also write
∂T
δg = − . (2–243)
∂h
Comparing (2–242) with (2–239), we see that the gravity disturbance δg, be-
sides being the difference in magnitude of the actual and the normal gravity
vector, is also the normal component of the gravity disturbance vector δg.
We now turn to the gravity anomaly ∆g. Since
∂γ
γP = γQ + N, (2–244)
∂h
we have
∂T ∂γ
− = δg = gP − γP = gP − γQ − N. (2–245)
∂h ∂h
Remembering the definition (2–228) of the gravity anomaly and taking into
account Bruns’ formula (2–237), we find the following equivalent relations:
∂T ∂γ
− = ∆g − N, (2–246)
∂h ∂h
∂T ∂γ
∆g = − + N, (2–247)
∂h ∂h
2.12 Anomalous gravity field 95
∂T 1 ∂γ
∆g = − + T, (2–248)
∂h γ ∂h
∂γ
δg = ∆g − N, (2–249)
∂h
1 ∂γ
δg = ∆g − T, (2–250)
γ ∂h
relating different quantities of the anomalous gravity field.
Another equivalent form is
∂T 1 ∂γ
− T + ∆g = 0 . (2–251)
∂h γ ∂h
This expression has been called the fundamental equation of physical geodesy,
because it relates the measured quantity ∆g to the unknown anomalous
potential T . In future, the relation
∂T
+ δg = 0 (2–252)
∂h
may replace it.
It has the form of a partial differential equation. If ∆g were known
throughout space, then (2–251) could be discussed and solved as a real par-
tial differential equation. However, since ∆g is known only along a surface
(the geoid), the fundamental equation (2–251) can be used only as a bound-
ary condition, which alone is not sufficient for computing T . Therefore, the
name “differential equation of physical geodesy”, which is sometimes used
for (2–251), is rather misleading.
One usually assumes that there are no masses outside the geoid. This is
not really true. But neither do we make observations directly on the geoid;
we make them on the physical surface of the earth. In reducing the measured
gravity to the geoid, the effect of the masses outside the geoid is removed by
computation, so that we can indeed assume that all masses are enclosed by
the geoid (see Chaps. 3 and 8).
In this case, since the density is zero everywhere outside the geoid, the
anomalous potential T is harmonic there and satisfies Laplace’s equation
∂2T ∂2T ∂2T
∆T ≡ + + = 0. (2–253)
∂x2 ∂y 2 ∂z 2
This is a true partial differential equation and suffices, if supplemented by
the boundary condition (2–251), for determining T at every point outside
the geoid. If we write the boundary condition in the form
∂T 1 ∂γ
− + T = ∆g , (2–254)
∂n γ ∂n
96 2 Gravity field of the earth
where ∆g is assumed to be known at every point of the geoid, then we see that
a linear combination of T and ∂T /∂n is given upon that surface. According
to Sect. 1.13, the determination of T is, therefore, a third boundary-value
problem of potential theory. If it is solved for T , then the geoidal height,
which is the most important geometric quantity in physical geodesy, can be
computed by Bruns’ formula (2–237).
Therefore, we may say that the basic problem of physical geodesy, the
determination of the geoid from gravity measurements, is essentially a third
boundary-value problem of potential theory.
we get √
3
R= a2 b . (2–257)
In a similar way, we may define a mean value of gravity, γ0 , as normal gravity
at latitude ϕ = 45◦ (Moritz 1980b: p. 403). Numerical values of about
Since the normal to the sphere is the direction of the radius vector r, we
have to the same approximation
∂ ∂ ∂
= = . (2–260)
∂n ∂h ∂r
In Bruns’ theorem (2–237) we may replace γ by γ0 , and Eqs. (2–246)
through (2–250) and (2–251) become
∂T 2γ0
− = ∆g + N, (2–261)
∂h R
∂T 2γ0
∆g = − − N, (2–262)
∂r R
∂T 2
∆g = − − T, (2–263)
∂r R
2γ0
δg = ∆g + N, (2–264)
R
2
δg = ∆g + T , (2–265)
R
∂T 2
+ T + ∆g = 0 . (2–266)
∂r R
The last equation is the spherical approximation of the fundamental bound-
ary condition.
Remark
The meaning of this spherical approximation should be carefully kept in
mind. It is used only in equations relating the small quantities T, N, ∆g, δg,
etc. The reference surface is never a sphere in any geometrical sense, but
always an ellipsoid. As the flattening f is very small, the ellipsoidal formulas
can be expanded into power series in terms of f , and then all terms containing
f, f 2 , etc., are neglected. In this way one obtains formulas that are rigorously
valid for the sphere, but approximately valid for the actual reference ellipsoid
as well. However, normal gravity γ in the gravity anomaly ∆g = g − γ
must be computed for the ellipsoid to a high degree of accuracy. To speak
of a “reference sphere” in space, in any geometric sense, may be highly
misleading.
∞
T = T (R, ϑ, λ) = Tn (ϑ, λ) . (2–268)
n=0
∞ n+1
∂T 1 R
δg = − = (n + 1) Tn (ϑ, λ) . (2–269)
∂r r r
n=0
∞
∂T 1
δg = − = (n + 1) Tn (ϑ, λ) . (2–270)
∂r R
n=0
∂T 2
∆g = − − T. (2–271)
∂r r
Its exact meaning will be discussed at the end of the following section. The
substitution of (2–269) and (2–267) into this equation yields
∞ n+1
1 R
∆g = (n − 1) Tn (ϑ, λ) . (2–272)
r n=0 r
∞
1
∆g = (n − 1) Tn (ϑ, λ) . (2–273)
R n=0
By omitting the terms of degrees one and zero, we get a new function
2 ∞ n+1
R R R
H = H − H0 − H1 = Hn . (2–277)
r r r
n=2
r l
à R 2d ¾
Ã
d¾ R
unit sphere
r=1
terrestrial sphere
r=R
The reason for this modification of Poisson’s integral is that the formulas
of physical geodesy are simpler if the functions involved do not contain har-
monics of degrees zero and one. It is therefore convenient to split off these
terms. This is done automatically by the modified Poisson integral (2–279).
We now apply these formulas to the gravity anomalies outside the earth.
Equation (2–272) yields at once
∞ n+1
R
r ∆g = (n − 1) Tn (ϑ, λ) . (2–280)
r
n=0
or
R2 r 2 − R2 1 3R
∆gP = − − 2 cos ψ ∆g dσ . (2–282)
4π r l3 r r
σ
This is the formula for the computation of gravity anomalies outside the
earth from surface gravity anomalies, or for the upward continuation of grav-
ity anomalies.
Finally, we discuss the exact meaning of the gravity anomaly δgP outside
the earth. We start with a convenient definition. The level surfaces of the
actual gravity potential, the surfaces
W = constant , (2–283)
2.14 Gravity anomalies outside the earth 101
geopotential
P W = WP = const.
surface
NP
Q spheropotential
U = WP = const.
surface
P0
geoid W = W0
N
Q0 ellipsoid U = W0
are often called geopotential surfaces; the level surfaces of the normal gravity
field, the surfaces
U = constant , (2–284)
are called spheropotential surfaces.
We consider now the point P outside the earth (Fig. 2.15) and denote
the geopotential surface passing through it by
W = WP . (2–285)
U = WP (2–286)
of the same constant WP . The normal plumb line through P intersects this
spheropotential surface at the point Q, which is said to correspond to P .
We see that the level surfaces W = WP and U = WP are related to each
other in exactly the same way as are the geoid W = W0 and the reference
ellipsoid U = W0 . If, therefore, the gravity anomaly is defined by
∆gP = gP − γQ , (2–287)
as in Sect. 2.12, then all derivations and formulas of that section also apply
for the present situation, the geopotential surface W = WP replacing the
geoid W = W0 , and the spheropotential surface U = WP replacing the
ellipsoid U = W0 . This is also the reason why (2–271) applies at P as well
as at the geoid.
Note that P in Sect. 2.12 is a point at the geoid, which is denoted by P0
in Fig. 2.15.
This situation will be taken up again in Chap. 8, in the context of Molo-
densky’s problem.
102 2 Gravity field of the earth
Therefore,
R3 R4
2
lim (r T ) = lim T2 + 2 T3 + · · · = 0, (2–293)
r→∞ r→∞ r r
so that r
r T = r 2 T − lim (r 2 T ) = r 2 T
2
(2–294)
∞ r→∞
and r
r T =−
2
r 2 ∆g(r) dr . (2–295)
∞
2.15 Stokes’ formula 103
2r 2
= − 3l − 3R cos ψ ln(r − R cos ψ + l) + r + 3R cos ψ ln r .
l
(2–298)
The reader is advised to perform this integration, taking into account (2–
275), or at least to check the result by differentiating the right-hand side
with respect to r.
For large values of r, we have
R
l =r 1− cos ψ · · · = r − R cos ψ · · · (2–299)
r
and, hence, we find that as r → ∞, the right-hand side of the above indefinite
integral approaches
5R cos ψ − 3R cos ψ ln 2 . (2–300)
If we subtract this from the indefinite integral, we get the definite integral,
since infinity is its lower limit of integration. Thus,
r
r 3 − R2 r 3R
− +1+ cos ψ dr
l3 r
∞ (2–301)
2r 2 r − R cos ψ + l
= + r − 3l − R cos ψ 5 + 3 ln .
l 2r
Hence, we obtain Pizzetti’s formula
R
T (r, ϑ, λ) = S(r, ψ) ∆g dσ , (2–302)
4π
σ
104 2 Gravity field of the earth
where
2R R R l R2 r − R cos ψ + l
S(r, ψ) = + − 3 2 − 2 cos ψ 5 + 3 ln . (2–303)
l r r r 2r
where
1 ψ ψ ψ
S(ψ) = −6 sin +1−5 cos ψ−3 cos ψ ln sin + sin2 (2–305)
sin(ψ/2) 2 2 2
ψ
r = R and l = 2R sin . (2–306)
2
By Bruns’ theorem, N = T /γ0 , we finally get
R
N= ∆g S(ψ) dσ . (2–307)
4π γ0
σ
This formula was published by G.G. Stokes in 1849; it is, therefore, called
Stokes’ formula, or Stokes’ integral. It is by far the most important formula
of physical geodesy because it performs to determine the geoid from gravity
data. Equation (2–304) is also called Stokes’ formula, and S(ψ) is known as
Stokes’ function.
Using formula (2–302), which was derived by Pizzetti (1911) and later on
by Vening Meinesz (1928), we can compute the anomalous potential T at any
point outside the earth. Dividing T by the normal gravity at the given point
P (Bruns’ theorem), we obtain the separation NP between the geopotential
surface W = WP and the corresponding spheropotential surface U = WP ,
which, outside the earth, takes the place of the geoidal undulation N (see
Fig. 2.15 and the explanations at the end of the preceding section).
We mention again that these formulas are based on a spherical approx-
imation; quantities of the order of 3 · 10−3 N are neglected. This results in
an error of probably less than 1 m in N , which can be neglected for many
practical purposes. Sagrebin (1956), Molodenskii et al. (1962: p. 53), Bjer-
hammar, and Lelgemann have developed higher approximations, which take
into account the flattening f of the reference ellipsoid; see Moritz (1980 a:
Sect. 39).
2.16 Stokes’ integral and Stokes’ function 105
PN
north pole
®
P
Ã
d¾
dÃ
sinÃd®
Hence, we find
2π π
R
N= ∆g(ψ, α) S(ψ) sin ψ dψ dα (2–310)
4π γ0 α=0 ψ=0
The functions S(ψ) and F (ψ) are shown in Fig. 2.17. Alternatively, we may
2.16 Stokes’ integral and Stokes’ function 107
0
30° 60° 90° 120° 150° 180°
F(Ã)
–1
–2
S(Ã)
ϑ = 90◦ − ϕ . (2–315)
Hence, we have
2π π/2
dσ = cos ϕ dϕ dλ , (2–316)
λ=0 ϕ=−π/2
σ
Comparing this with Stokes’ formula (2–304), we find the expression for
Stokes’ function in terms of Legendre polynomials (zonal harmonics):
∞
2n + 1
S(ψ) = Pn (cos ψ) . (2–326)
n=2
n−1
where
∞
T (ϑ, λ) = Tn (ϑ, λ) . (2–329)
n=2
know the exact mass of the earth, how can we make M rigorously equal to
M?
Subsequently, we will see that the first-degree harmonic can always be
assumed to be zero. Under this assumption, we can substitute (2–331) into
(2–328) and express T by the conventional Stokes formula (2–304). Thus we
obtain
G δM R
T = + ∆g S(ψ) dσ . (2–333)
R 4π
σ
This is the generalization of Stokes’ formula for T . It holds for an arbitrary
reference ellipsoid whose center coincides with the center of the earth.
First-degree terms
The coefficients of the first-degree harmonic in the potential W are, according
to (2–85) and (2–87), given by
GM xc , GM yc , GM zc , (2–334)
GM xc , GM yc , GM zc . (2–335)
As xc , yc , zc are very small in any case, these are practically equal to
They are zero, and there is no first-degree harmonic T1 (ϑ, λ) if and only if
the center of the reference ellipsoid coincides with the center of gravity of
the earth. This is usually assumed.
In the general case, we find from the first-degree term of (2–76), on
putting r = R and using the coefficients (2–85) together with (2–87),
GM
T1 (ϑ, λ) = 2 (zc − zc ) P10 (cos ϑ) + (xc − xc ) P11 (cos ϑ) cos λ
R
(2–338)
+ (yc − yc ) P11 (cos ϑ) sin λ .
If the origin of the coordinate system is taken to be the center of the reference
ellipsoid, then xc = yc = zc = 0. With P10 (cos ϑ) = cos ϑ, P11 (cos ϑ) = sin ϑ,
2.17 Generalization to an arbitrary reference ellipsoid 111
and GM/R2 = γ0 we then obtain the following expression for the first-degree
harmonic of T :
N1 (ϑ, λ) = xc · e , (2–343)
are the equations of the geoid and the ellipsoid, where in general the con-
stants W0 and U0 are different. As in Sect. 2.12, we have, using Fig. 2.12,
WP = UQ − γ N + T , but now UQ = U0 = W0 = WP , so that
γ N = T − (W0 − U0 ) . (2–345)
δW = W0 − U0 , (2–346)
T − δW
N= . (2–347)
γ
We also need the extension of Eqs. (2–246) through (2–250). Those for-
mulas which contain N instead of T are easily seen to hold for an arbitrary
reference ellipsoid as well, but the transition from N to T is now effected by
means of (2–347). Hence, Eq. (2–247), i.e.,
∂T ∂γ
∆g = − + N, (2–348)
∂h ∂h
remains unchanged, but (2–248) becomes
∂T 1 ∂γ 1 ∂γ
∆g = − + T− δW . (2–349)
∂h γ ∂h γ ∂h
∂T 1 ∂γ 1 ∂γ
− + T = ∆g + δW . (2–350)
∂h γ ∂h γ ∂h
T − δW
N= (2–351)
γ0
and
∂T 2 2
∆g = − − T + δW (2–352)
∂r R R
and
∂T 2 2
− − T = ∆g − δW . (2–353)
∂r R R
2.17 Generalization to an arbitrary reference ellipsoid 113
where, by (1–89),
1
∆g0 = ∆g dσ . (2–362)
4π
σ
A final note
A direct consequence of Eq. (2–356) is that N0 has an immediate geometrical
meaning: if a is the equatorial radius (semimajor axis) of the given reference
ellipsoid, then
aE = a + N0 (2–366)
is the equatorial radius of an ellipsoid whose normal potential U0 is equal to
the actual potential W0 of the geoid, and which encloses the same mass as
that of the earth, the flattening f remaining the same. The reason is that
for such a new ellipsoid E the new N0 = 0 by (2–356) with δM = 0 and
δW = 0.
A small additive constant N0 is equivalent to a change of scale for a nearly
spherical earth. To see this, imagine a nearly spherical orange. Increasing the
thickness of the peel of an orange everywhere by 1 mm (say) is equivalent
to a similarity transformation (uniform increase of the size) of the orange’s
surface.
So, the usual Stokes formula, without N0 , gives a global geoid that is
determined only up to the scale which implicitly is contained in N0 . It is,
however, geocentric, at least in theory, because it contains no spherical har-
monic of first degree, T1 (ϑ, λ). It would be exactly geocentric if the earth
were covered uniformly by gravity measurements. The scale was formerly de-
termined astrogeodetically, historically by grade measurements dating back
2.18 Gravity disturbances and Koch’s formula 115
A historical remark
This remark is due to Mrs. M. I. Yurkina, Moscow. Mathematically, the
above is the solution of Neumann’s problem (the second boundary-value
problem of potential theory) for the sphere, cf. Sect. 1.13. It is a classical
problem of potential theory, with a history of at least 150 years, similarly
to Stokes’ formula. “Neumann’s problem” is named after the mathematician
Carl Neumann, who edited his father’s (Franz Neumann) lectures from the
1850s (Neumann 1887: see especially p. 275). The external spherical Neu-
mann problem also occurs in Kellogg (1929: p. 247). It is again found in
Hotine (1969: pp. 311, 318).
Their basic significance for modern physical geodesy with a known earth
surface was recognized and elaborated by Koch (1971). So the present inte-
gral formula should perhaps be called F. Neumann – C. Neumann – Kellogg –
Hotine – Koch formula. For brevity, we refer to it as Koch’s formula.
dN = −ε ds (2–372)
or
dN
ε=− ; (2–373)
ds
the minus sign is a convention, its meaning will be explained later.
ds
s ellipsoid
Fig. 2.18. The relation between the geoidal undulation and the
deflection of the vertical
2.19 Deflections of the vertical and formula of Vening Meinesz 117
north pole
¸'– ¸
®
P
Ã
d¾
Fig. 2.19. Relation between geographical and polar coordinates on the sphere
We now introduce the azimuth α, as shown in Fig. 2.16. From the spherical
triangle of Fig. 2.19 we get, using well-known formulas of spherical trigonom-
etry,
sin ψ cos α = cos ϕ sin ϕ − sin ϕ cos ϕ cos(λ − λ) ,
(2–382)
sin ψ sin α = cos ϕ sin(λ − λ) .
Substituting these relations into the preceding equations, we find the simple
expressions
∂ψ ∂ψ
= − cos α , = − cos ϕ sin α , (2–383)
∂ϕ ∂λ
so that
∂S(ψ) dS(ψ) ∂S(ψ) dS(ψ)
=− cos α , =− cos ϕ sin α . (2–384)
∂ϕ dψ ∂λ dψ
These are substituted into (2–379) and the corresponding formula for ∂N/∂λ
and from equations (2–377) we finally obtain
2π π/2
1 dS(ψ)
ξ(ϕ, λ) = ∆g(ϕ , λ ) cos α cos ϕ dϕ dλ ,
4π γ0 λ =0 ϕ =−π/2 dψ
2π π/2
1 dS(ψ)
η(ϕ, λ) = ∆g(ϕ , λ ) sin α cos ϕ dϕ dλ
4π γ0 λ =0 ϕ =−π/2 dψ
(2–385)
or, written in the usual abbreviated form,
1 dS(ψ)
ξ= ∆g cos α dσ ,
4π γ0 dψ
σ
(2–386)
1 dS(ψ)
η= ∆g sin α dσ .
4π γ0 dψ
σ
2.20 The vertical gradient of gravity 119
cos ϕ sin(λ − λ)
tan α = , (2–388)
cos ϕ sin ϕ − sin ϕ cos ϕ cos(λ − λ)
The reader can easily verify that these equations give the deflection compo-
nents ξ and η with the correct sign corresponding to the definition (2–230);
see also Fig. 2.13. This is the reason why we introduced the minus sign in
(2–373).
We note that the formula of Vening Meinesz is valid as it stands for an
arbitrary reference ellipsoid, whereas Stokes’ formula had to be modified by
adding a constant N0 . If we differentiate the modified Stokes formula with
respect to ϕ and λ, to get Vening Meinesz’ formula, then this constant N0
drops out and we get Eqs. (2–386).
For the practical application of Stokes’ and Vening Meinesz’ formulas
and problems, the reader is referred to Sect. 2.21 and to Chap. 3.
∂g
= −2g J − 2ω 2 , (2–390)
∂H
120 2 Gravity field of the earth
∂g ∂γ ∂∆g
= + . (2–391)
∂H ∂H ∂H
The normal gradient ∂γ/∂H is given by (2–147) and (2–148). The anomalous
.
part, ∂∆g/∂H = ∂∆g/∂r, will be considered now.
Expression in terms of ∆g
Equation (2–272) may be written as (note that r ∆g is harmonic and the
factor must be 1 for r = R)
∞ n+2
R
∆g(r, ϑ, λ) = ∆gn (ϑ, λ) . (2–392)
r
n=0
Expression in terms of N
By differentiating equation (2–271),
∂T 2
∆g = − − T, (2–396)
∂r r
with respect to r, we get
∂∆g ∂2T 2 ∂T 2
=− 2 − + 2 T. (2–397)
∂r ∂r r ∂r r
To this formula we add Laplace’s equation ∆T = 0, which in spherical
coordinates has the form
∂2T 2 ∂T tan ϕ ∂T 1 ∂2T 1 ∂2T
+ − + + = 0; (2–398)
∂r 2 r ∂r r 2 ∂ϕ r 2 ∂ϕ2 r 2 cos2 ϕ ∂λ2
where γ0 is a global mean value as usual. This equation expresses the vertical
gradient of the gravity anomaly in terms of the geoidal undulation N and
its first and second horizontal derivatives. It can be evaluated by numerical
differentiation, using a map of the function N . However, it is less suited for
practical application than (2–394) because it requires an extremely accurate
and detailed local geoidal map, which is hardly ever available; inaccuracies
of N are greatly amplified by forming the second derivatives.
∂N ∂N
= −R ξ , = −R η cos ϕ , (2–401)
∂ϕ ∂λ
so that
∂2N ∂ξ ∂2N ∂η
2
= −R , 2
= −R cos ϕ . (2–402)
∂ϕ ∂ϕ ∂λ ∂λ
122 2 Gravity field of the earth
®1
=
®
P q
® = ®2
à = Ã2
à = Ã1
Fig. 2.20. A template
45°30'
q
20'
'= 45°10'
R
N= ∆g S(ψ) dσ (2–407)
4π γ0
σ
with its explicit forms (2–310) for the template method and (2–317) for the
method that uses fixed blocks.
124 2 Gravity field of the earth
The final integral is simply the area Ak of the compartment and we obtain
Ak S(ψk )
ck = . (2–412)
4π γ0 R
The advantage of the template method is its great flexibility. The influ-
ence of the compartments near the computation point P is greater than that
of the distant ones, and the integrand changes faster in the neighborhood
of P . Therefore, a finer subdivision is necessary around P . This can easily
be provided by templates. Yet, the method is completely old-fashioned and
thus obsolete.
The advantage of the fixed system of blocks formed by a grid of ellipsoidal
coordinates lies in the fact that their mean gravity anomalies are needed for
many different purposes. These mean anomalies of standard-sized blocks,
once they have been determined, can be easily stored and processed by a
computer. Also, the same subdivision is used for all computation points,
whereas the compartments defined by a template change when the template
is moved to the next computation point. The flexibility of the method of
standard blocks is limited; however, one may use smaller blocks (5 × 5 , for
example) in the neighborhood of P and larger ones (1◦ × 1◦ , for example)
farther away. With current electronic computation, this method is the only
one used in practice. The theoretical usefulness of polar coordinates will be
shown now.
2.21 Practical evaluation of the integral formulas 125
N = Ni + Ne , (2–417)
where
2π ψ0
R
Ni = ∆g S(ψ) dσ ,
4π γ0 α=0 ψ=0
(2–418)
2π π
R
Ne = ∆g S(ψ) dσ .
4π γ0 α=0 ψ=ψ0
The radius ψ0 of the inner zone corresponds to a linear distance of a few
kilometers. Within this distance, we may treat the sphere as a plane, using
polar coordinates s, α, where
. . . ψ
s = R ψ = R sin ψ = 2R sin , (2–419)
2
so that the element of area becomes
R2 dσ = s ds dα . (2–420)
126 2 Gravity field of the earth
2π s0
1 ∆g
ξi = − cos α s ds dα ,
2π γ0 α=0 s=0 s2
(2–423)
2π s0
1 ∆g
ηi = − sin α s ds dα ,
2π γ0 α=0 s=0 s2
s0
∂∆g 1 2π
∆g − ∆gP
= s ds dα . (2–424)
∂H i 2π α=0 s=0 s3
In order to evaluate these integrals, we expand ∆g into a Taylor series at
the computation point P :
1
2
∆g = ∆gP + x gx + y gy + x gxx + 2x y gxy + y 2 gyy + · · · . (2–425)
2!
Inserting this into the above integrals, we can easily evaluate them. Perform-
ing the integration with respect to α first and noting that
2π
0 dα = 2π ,
2π 2π 2π
0 sin α dα = 0 cos α dα = 0 sin α cos α dα = 0 , (2–429)
2π 2π
0 sin2 α dα = 0 cos2 α dα = π ,
we find
s0
1 s2
Ni = ∆gP + (gxx + gyy ) + · · · ds , (2–430)
γ0 0 4
s0
1
ξi = − (gx + · · ·) ds ,
2γ0 0
s0 (2–431)
1
ηi = − (gy + · · ·) ds ,
2γ0 0
∂∆g 1 s0
= (gxx + gyy + · · ·) ds . (2–432)
∂H i 4 s=0
We now perform the integration over s, retaining only the lowest nonvanish-
ing terms. The result is
s0
Ni = ∆gP , (2–433)
γ0
s0 s0
ξi = − gx , ηi = − gy , (2–434)
2γ0 2γ0
∂∆g s0
= (gxx + gyy ) . (2–435)
∂H i 4
We see that the effect of the innermost circular zone on Stokes’ formula
depends, to a first approximation, on the value of ∆g at P ; the effect on
Vening Meinesz’ formula depends on the first horizontal derivatives of ∆g;
and the effect on the vertical gradient depends on the second horizontal
derivatives.
Note that the contribution of the innermost zone to the total deflection
of the vertical has the same direction as the line of steepest inclination of
the “gravity anomaly surface”, because the plane vector
ϑ = [ξi , ηi ] (2–436)
Dg = 40 mgal
Dg = 30 mgal
Dg = 20 mgal
The direction of grad ∆g defines the line of steepest descent (Fig. 2.22).
The values of gx and gy can be obtained from a gravity map. They are the
inclinations of north-south and east-west profiles through P . Values for gxx
and gyy may be found by fitting a polynomial in x and y of second degree
to the gravity anomaly function in the neighborhood of P .
A remark on accuracy
Deflections of the vertical ξ, η, if combined with astronomical observations
of astronomical latitude Φ and astronomical longitude Λ, furnish positions
on the reference ellipsoid, expressed by ellipsoidal coordinates
ϕ=Φ−ξ,
(2–438)
λ = Λ − η sec ϕ ,
Unfortunately, to get the same precision for horizontal as for vertical po-
sition, is much more difficult, keeping in mind the relation 1 ∼ = 30 m on
the earth’s surface. So to get an accuracy of 1 m, which is not too difficult
with Stokes’ formula, means an accuracy better than 0.03 in both Φ and ξ
(analogously to Λ and η), which is almost impossible to achieve practically.
3 Gravity reduction
3.1 Introduction
Gravity g measured on the physical surface of the earth must be distin-
guished from normal gravity γ referring to the surface of the ellipsoid. To
refer g to sea level, a reduction is necessary. Since there are masses above
sea level, the reduction methods differ depending on the way how to deal
with these topographic masses. Gravity reduction is essentially the same for
gravity anomalies ∆g and gravity disturbances δg.
Gravity reduction serves as a tool for three main purposes:
Only the first two purposes are of a direct geodetic nature. The third is of
interest to theoretical geophysicists and geologists, who study the general
structure of the crust, and to exploration geophysicists.
The use of Stokes’ formula for the determination of the geoid requires
that the gravity anomalies ∆g represent boundary values at the geoid. This
implies two conditions: first, gravity g must refer to the geoid; second, there
must be no masses outside the geoid (Sect. 2.12). Hence, figuratively speak-
ing, gravity reduction consists of the following steps:
H earth's surface
geoid
P0
Fig. 3.1. Gravity reduction
130 3 Gravity reduction
The first step requires knowledge of the density of the topographic masses,
which is somewhat problematic.
By such a reduction procedure certain irregularities in gravity due to
differences in height of the stations are removed so that interpolation and
even extrapolation to unobserved areas become easier (Sect. 9.7).
P outside cylinder
Assume first that P is above the cylinder, c > b. Then the potential is given
by the general formula (1–12),
U =G dv . (3–1)
l
we have
l= s2 + (c − z)2 (3–3)
and
dv = dx dy dz = s ds dα dz . (3–4)
z z
P P
l
a c
s a
b dm z b
xy
so that we have
b
U = 2π G −c+z+ a2 + (c − z)2 dz . (3–7)
0
∂U
A=− . (3–10)
∂c
P on cylinder
In this case we have c = b, and Eqs. (3–9) and (3–11) become
√
b + a2 + b2
U0 = π G −b + b a2 + b2 + a ln
2 2
, (3–12)
a
A0 = 2π G a + b − a2 + b2 . (3–13)
P inside cylinder
We assume that P is now inside the cylinder, c < b. By the plane z = c we
separate the cylinder into two parts, 1 and 2 (Fig. 3.3), and compute U as
the sum of the contributions of these two parts:
Ui = U1 + U2 , (3–14)
where the subscript i denotes that P is now inside the cylinder. The term
U1 is given by (3–12) with b replaced by c, and U2 by the same formula with
b replaced by b − c. Their sum is
√
Ui = π G − c2 − (b − c)2 + c a2 + c2 + (b − c) a2 + (b − c)2
√ (3–15)
c+ a2 + c2 b−c+ a2 + (b − c)2
+ a2 ln + a2 ln .
a a
2 P b -c
b
1 c
Circular disk
Let the thickness b of the cylinder go to zero such that the product
κ = b (3–17)
remains finite. The quantity κ may then be considered as the surface density
with which matter is concentrated on the surface of a circle of radius a. We
need potential and attraction for an exterior point. By setting
κ
= (3–18)
b
2π
α= , (3–20)
n
P ®
a = a1
a = a2
Since Ae and Ai differ only by a constant, this constant drops out in the
second equation of (3–21), and we obtain from (3–11) and (3–16)
2π
∆Ae = ∆Ai = G a22 + (c − b)2 − a21 + (c − b)2
n
(3–22)
− a22 + c2 + a21 + c2 .
g0 = g + F , (3–24)
where
∂g
F =− H (3–25)
∂H
is the free-air reduction to the geoid. Note that the assumption of no masses
above the geoid may be interpreted in the sense that such masses have been
mathematically removed beforehand, so that this reduction is indeed carried
out “in free air”.
For many practical purposes it is sufficient to use instead of ∂g/∂H the
normal gradient of gravity (associated with the ellipsoidal height h) ∂γ/∂h,
obtaining
. ∂γ .
F =− H = +0.3086 H [mgal] (3–26)
∂h
for H in meters.
3.4 Bouguer reduction 135
AB = 2π G H (3–27)
for H in meters.
Removing the plate is equivalent to subtracting its attraction (3–27) from
the observed gravity. This is called incomplete Bouguer reduction. Note that
this is the usual “plane” Bouguer plate; for a truly “spherical” Bouguer plate
we would have 4π instead of 2π (Moritz 1990: p. 235).
To continue and complete our gravity reduction, we must now apply the
free-air reduction F as given in (3–26). This combined process of removing
the topographic masses and applying the free-air reduction is called complete
Bouguer reduction. Its result is Bouguer gravity at the geoid:
gB = g − AB + F . (3–29)
P0
gravity measured at P g
minus Bouguer plate − 0.1119 H
plus free-air reduction + 0.3086 H (3–30)
Since gB now refers to the geoid, we obtain genuine gravity anomalies in the
sense of Sect. 2.12 by subtracting normal gravity γ referred to the ellipsoid:
∆gB = gB − γ . (3–31)
Terrain correction
This simple procedure is refined by taking into account the deviation of the
actual topography from the Bouguer plate of P (Fig. 3.6). This is called
terrain correction or topographic correction. At A the mass surplus ∆m+ ,
which attracts upward, is removed, causing g at P to increase. At B the
mass deficiency ∆m− is made up, causing g at P to increase again. The
terrain correction is always positive.
The practical determination of the terrain correction At is carried out
by means of a template, similar to that shown in Fig. 2.20, using (3–22) and
adding the effects of the individual compartments:
At = ∆A . (3–32)
A earth's surface
Dm +
H - HP
B P
Bouguer plate
Dm –
H
HP
geoid
P0
Again, we can use a template in polar coordinates (Fig. 2.20) for theoretical
considerations or a rectangular grid (Fig. 2.21) for numerical computations.
For a surplus mass ∆m+ , H > Hp , we have
b = H − HP , c = 0; (3–33)
b = c = HP − H . (3–34)
gB = g − AB + At + F (3–35)
Unified procedure
It is also possible to compute the total effect of the topographic masses,
AT = AB − At , (3–36)
in one step by using columns with base at sea level (Fig. 3.7), again sub-
HP
P0
gB = g − AT + F . (3–38)
plumb line
Q
H = HP
earth's surface
z = HQ
geoid
P0 W = W0
provided that the actual gravity gradient ∂g/∂H inside the earth were known.
It can be obtained by Bruns’ formula (2–40),
∂g
= −2g J + 4π G − 2ω 2 , (3–40)
∂H
if the mean curvature J of the geopotential surfaces and the density are
known between P and Q.
The normal free-air gradient is given by (2–147):
∂γ
= −2γ J0 − 2ω 2 , (3–41)
∂h
where J0 is the mean curvature of the spheropotential surfaces. If the ap-
proximation
.
g J = γ J0 (3–42)
is sufficient, then we get from (3–40) and (3–41)
∂g ∂γ
= + 4π G . (3–43)
∂H ∂h
Numerically, neglecting the variation of ∂γ/∂h with latitude, we find for the
density = 2.67 g cm−3 and (truncated) G = 6.67 · 10−11 m3 kg−1 s−2
∂g
= −0.3086 + 0.2238 = −0.0848 gal km−1 , (3–44)
∂H
so that (3–39) becomes
with g in gal and H in km. This simple formula, although being rather crude,
is often applied in practice.
The accurate way to compute gQ would be to use (3–39) and (3–40) with
the actual mean curvature J of the geopotential surfaces, but this would
require knowledge of the detailed shape of these surfaces far beyond what is
attainable today.
Another way of computing gQ , which is more practicable at present, is
the following. It is similar to the usual reduction of gravity to sea level (see
Sect. 3.4) and consists of three steps:
1. Remove all masses above the geopotential surface W = WQ , which
contains Q, and subtract their attraction from g at P .
2. Since the gravity station P is now “in free air”, apply the free-air
reduction, thus moving the gravity station from P to Q.
140 3 Gravity reduction
3. Restore the removed masses to their former position, and add alge-
braically their attraction to g at Q.
gravity measured at P gP
1. remove Bouguer plate − 0.1119 (HP − HQ )
2. free-air reduction from P to Q + 0.3086 (HP − HQ )
3. restore Bouguer plate − 0.1119 (HP − HQ )
together: gravity at Q gQ = gP + 0.0848 (HP − HQ ) .
(3–46)
This is the same as (3–45), which is, thus, confirmed independently. We see
now that the use of (3–43) or (3–45) amounts to replacing the terrain with
a Bouguer plate.
Finally, we note that the reduction of Poincaré and Prey, abbreviated
as Prey reduction, yields the actual gravity which would be measured inside
the earth if this were possible. Its purpose is, thus, completely different from
the purpose of the other gravity reductions which give boundary values at
the geoid.
It cannot be directly used for the determination of the geoid but is needed
to obtain orthometric heights as will be discussed in Sect. 4.3. Actual gravity
g0 at a geoidal point P0 is related to Bouguer gravity gB , Eq. (3–38), by
g0 = gB − AT, P0 . (3–47)
Pratt–Hayford system
This system of compensation was outlined by Pratt and put into a mathe-
matical form by J.F. Hayford, who used it systematically for geodetic pur-
poses.
The principle is illustrated in Fig. 3.9. Underneath the level of compen-
sation there is uniform density. Above, the mass of each column of the same
cross section is equal. Let D be the depth of the level of compensation, reck-
oned from sea level, and let 0 be the density of a column of height D. Then
the density of a column of height D + H (H representing the height of the
topography) satisfies the equation
(D + H) = D 0 , (3–48)
142 3 Gravity reduction
4 km 6 km
2 km 3 km
5 km
H
H'
.
D = 100 km 2.67 2.62 2.57 2.59 2.52 2.67 2.76
level of compensation
are assumed.
For a spheroidal earth, the columns will converge slightly towards its
center, and other refinements may be introduced. We may postulate either
equality of mass or equality of pressure; each postulate leads to somewhat
different spherical refinements. It may be mentioned that for computational
reasons Hayford used still another, slightly different model; for instance, he
reckoned the depth of compensation D from the earth’s surface instead of
from sea level.
Airy–Heiskanen system
Airy proposed this model, and Heiskanen gave it a precise formulation for
geodetic purposes and applied it extensively. Figure 3.10 illustrates the prin-
ciple. The mountains of constant density
The higher they are, the deeper they sink. Thus, root formations exist under
mountains, and “antiroots” under the oceans.
We denote the density difference 1 − 0 by ∆. On the basis of assumed
numerical values, we have
Denoting the height of the topography by H and the thickness of the cor-
responding root by t (Fig. 3.10), then the condition of floating equilibrium
is
t ∆ = H 0 , (3–58)
so that
0
t= H = 4.45 H (3–59)
∆
results. For the oceans, the corresponding condition is
t ∆ = H (0 − w ) , (3–60)
where H and w are defined as above and t is the thickness of the antiroot
(Fig. 3.10), so that we get
0 − w
t = H = 2.73 H (3–61)
1 − 0
144 3 Gravity reduction
2 km
H
H'
T = 30 km density 2.67
t'
8.9 km
density 3.27
T +H +t (3–63)
regional compensation
local compensation
1. removal of topography,
2. removal of compensation,
3. free-air reduction to the geoid.
Steps 1 and 3 are known from Bouguer reduction, so that the techniques of
Sect. 3.4 can be applied to them. Step 2 is new and will be discussed now
for the two main topographic-isostatic systems.
Pratt–Hayford system
The method is the same as for the terrain correction, Sect. 3.4, Eq. (3–32).
The attraction of the (negative) compensation is again computed by
AC = ∆A , (3–65)
H
∆ = 0 , (3–68)
D
which differs slightly from (3–50), in order to restore equality of mass ac-
cording to
(0 − ∆) D + 0 H = 0 D . (3–69)
The first term on the left-hand side represents the mass of the layer between
the level of compensation and sea level; the second term represents the mass
of the topography, now assumed to have a density 0 .
Airy–Heiskanen system
Again we use
AC = ∆A , (3–70)
b = t, c = HP + T + t , (3–71)
P
HP H
P0
T
HP + T + t
T+t
root t
Total reduction
In analogy with (3–38), the topographic-isostatically reduced gravity on the
geoid becomes
gTI = g − AT + AC + F , (3–72)
148 3 Gravity reduction
Oceanic stations
Here the terms AT and F of (3–72) are zero, since the station is situated on
the geoid, but the term AC is more complicated.
In the Pratt–Hayford model, the procedure is as follows. The mass sur-
plus (3–53) of a suboceanic column of height D − H (Fig. 3.9) is removed
and used to fill the corresponding oceanic column of height H to the proper
density 0 . In mathematical terms, this is
AC = −A1 + A2 , (3–73)
where both A1 and A2 are of the form (3–32), ∆A is given by (3–22). For
A1 we have
b = D − H , c = D, (3–74)
b = c = H (3–75)
and density 0 − w .
In the Airy–Heiskanen model, the mass surplus of the antiroot, 1 − 0 ,
is used to fill the oceans to the proper density 0 . The corresponding value
is again given by (3–73), where for A1 we now have
b = t , c=T, (3–76)
b = c = H (3–77)
and density 0 − w .
In both models, Eq. (3–72) reduces for oceanic stations to
Topographic-isostatic anomalies
The topographic-isostatic gravity anomalies are – in analogy to the Bouguer
anomalies – defined by
∆gTI = gTI − γ . (3–79)
If any of the topographic-isostatic systems were rigorously true, then the
topographic-isostatic reduction would fulfil perfectly its goal of complete reg-
ularization of the earth’s crust, which would become level and homogeneous.
Then, with a properly chosen reference model for γ, the topographic-isostatic
gravity anomalies (3–79) would be zero.
The actual topographic-isostatic compensation occurring in nature can-
not completely conform to such abstract models. As a consequence, nonzero
topo-graphic-isostatic gravity anomalies will be left, but they will be small,
smooth, and more or less randomly positive and negative. On account of
this smoothness and independence of elevation, they are better suited for in-
terpolation or extrapolation than any other type of anomalies; see Chap. 9,
particularly Sect. 9.7.
It may be stressed again that for geodetic purposes the topographic-
isostatic model used must be mathematically precise and self-consistent, and
the same model must be used throughout. Refinements include the consid-
eration of irregularities of density of the topographic masses and the consid-
eration of the anomalous gradient of gravity.
N = N c + δN (3–80)
δWB = UT (3–82)
δWTI = UT − UC , (3–83)
where the relevant formulas are the first equation of (3–21), (3–9), (3–12),
and (3–15). The point U refers to is always the point P0 at sea level (Fig. 3.1).
For UT we use U0 , see (3–12), with b = H and density 0 (see Fig. 3.12). For
UC in the continental case, we use Ue , see (3–9), with the following values:
Pratt–Hayford,
H
b = c = d , density 0 ; (3–85)
D
Airy–Heiskanen,
The corresponding considerations for the oceanic case are left as an exercise
for the reader.
The indirect effect with Bouguer anomalies is very large, of the or-
der of ten times the geoidal undulation itself. See the map at the end of
Helmert (1884: Tafel I), where the maximum value is 440 m! The reason is
that the earth is in general topographic-isostatically compensated. There-
fore, the Bouguer anomalies cannot be used for the determination of the
geoid.
With topographic-isostatic gravity anomalies, as might be expected, the
indirect effect is smaller than N , of the order of 10 m. It is necessary, how-
ever, to compute the indirect effect δNI carefully, using exactly the same
topographic-isostatic model as for the gravity reductions.
Furthermore, before applying Stokes’ formula, the topographic-isostatic
gravity anomalies must be reduced from the geoid to the cogeoid. This is
done by a simple free-air reduction, using (3–26), by adding to ∆gI the
correction
δ = +0.3086 δN [mgal] , (3–87)
3.8 The inversion reduction of Rudzki 151
1 ∂ δN
δξ = − ,
R ∂ϕ
(3–88)
1 ∂ δN
δη = − .
R cos ϕ ∂λ
Then
δW = UT − UC = 0 . (3–90)
This procedure was given by M. P. Rudzki in 1905. For the present purpose,
we may consider the geoid to be a sphere of radius R (Fig. 3.13). Let the
mass element dm at Q be replaced by a mass element dm at a certain point
Q inside the geoid situated on the same radius vector. The potential due to
152 3 Gravity reduction
P0
l
R l'
r
à Q
Q' (dm)
r' (dm')
dm = dm . (3–95)
P Q
HP H
P0
Q0
Q0Q = Q0Q' H
Q'
gR = g − AT + AC + F , (3–96)
#
where AC = ∆A with b = H, c = H + HP , the density being equal to
that of topography.
Since the indirect effect is zero, the cogeoid of Rudzki coincides with the
actual geoid, but the gravity field outside the earth is changed, which to-
day is in the center of attention. In addition, the Rudzki reduction does not
correspond to a geophysically meaningful model. Nevertheless, it is impor-
tant conceptually. Regard it an interesting historic curiosity, but never even
consider to use it!
H
%
geoid
· = %H
∆gF = g + F − γ (3–100)
∆gH = gH − γ . (3–101)
1. The term F above has been seen to be part of every gravity reduction
rather than a full-fledged gravity reduction itself.
2. Approximately, free-air anomalies may be identified with Helmert’s
condensation anomalies as we have seen above.
3.9 The condensation reduction of Helmert 155
These are the main methods that have been proposed for the reduction of
gravity. A simple overview is given by Fig. 3.16.
H H H H
T H
D t
δHAB = l1 − l2 . (4–1)
If we measure a circuit, that is, a closed leveling line where we finally return
to the initial point, then the algebraic sum of all measured differences in
height will not in general be rigorously zero, as one would expect, even if we
had been able to observe with perfect precision. This misclosure indicates
that leveling is more complicated than it appears at first sight.
Let us look into the matter more closely. Figure 4.2 shows the relevant
geometrical principles. Let the points A and B be so far apart that the pro-
cedure of Fig. 4.1 must be applied repeatedly. Then the sum of the leveled
height differences between A and B will not be equal to the difference in the
orthometric heights HA and HB . The reason is that the leveling increment
δn, as we henceforth denote it, is different from the corresponding increment
δHB of HB (Fig. 4.2), due to the nonparallelism of the level surfaces. De-
noting the corresponding increment of the potential W by δW , we have by
(2–21)
−δW = g δn = g δHB , (4–2)
_ _
A B
l2
l1
B
±HAB
A
±n ±HB
A HB
HA
geoid
where g is the gravity at the leveling station and g is the gravity on the
plumb line of B at δHB . Hence,
g
δHB = δn = δn . (4–3)
g
There is, thus, no direct geometrical relation between the result of lev-
eling and the orthometric height, since (4–3) expresses a physical relation.
What, then, if not height, is directly obtained by leveling? If gravity g is also
measured, then
δW = −g δn (4–4)
is determined, so that we obtain
B
WB − WA = − g δn . (4–5)
A
Note that this integral is independent of the path of integration; that is,
different leveling lines connecting the points A and B (Fig. 4.3) should give
the same result. This is evident because W is a function of position only;
therefore, to every point there corresponds a unique value W . If the leveling
line returns to A, then the total integral must be zero:
$
g dn = −WA + WA = 0 . (4–7)
4.2 Geopotential numbers and dynamic heights 159
A
Fig. 4.3. Two different leveling lines connecting A and B (taken to-
gether, they form a circuit)
%
The symbol denotes an integral over a circuit.
On the other hand, the measured height difference, that is, the sum of
the leveling increments
B
B
∆nAB = δn = dn , (4–8)
A A
depends on the path of integration and is, thus, not in general zero for a
circuit: $
dn = misclosure = 0 . (4–9)
which is the difference between the potential at the geoid and the potential
at the point A, has been introduced as the geopotential number of A in
Sect. 2.4. It is defined so as to be always positive.
As a potential difference, the geopotential number C is independent of
the particular leveling line used for relating the point to sea level. It is the
same for all points of a level surface; it can, thus, be considered as a natural
measure of height, even if it does not have the dimension of a length.
The geopotential number C is measured in geopotential units (g.p.u.),
where
1 g.p.u. = 1 kgal m = 1000 gal m. (4–11)
.
Using g = 0.98 kgal in (4–10), we get
. .
C = g H = 0.98 H , (4–12)
so that the geopotential numbers in g.p.u. are almost equal to the height
above sea level in meters.
The geopotential numbers were adopted at a meeting of a Subcommission
of the IAG at Florence in 1955. Formerly, the dynamic heights were used,
defined by
C
H dyn = , (4–13)
γ0
where γ0 is normal gravity for an arbitrary standard latitude, usually 45◦ :
γ45◦ = 9.806 199 203 m s−2 = 980.6 199 203 gal (4–14)
for the GRS 1980. Just note and keep in mind that 1 gal = 10−2 m s−2 and,
accordingly, 1 mgal = 10−5 m s−2 .
The dynamic height differs from the geopotential number only in the
scale or the unit: The division by the constant γ0 in (4–13) merely con-
verts a geopotential number into a length. However, the dynamic height has
no geometrical meaning whatsoever, so that the division by an arbitrary γ0
somehow obscures the true physical meaning of a potential difference. Hence,
the geopotential numbers are, for reasons of theory and for practically es-
tablishing a national or continental height system, preferable to the dynamic
heights.
Dynamic correction
It is sometimes convenient to convert the measured height difference ∆nAB
of (4–8) into a difference of dynamic height by adding a small correction.
Using Eqs. (4–13) and (4–10) gives
B
1 1
∆HAB = HB − HA =
dyn dyn dyn
(CB − CA ) = g dn , (4–15)
γ0 γ0 A
4.3 Orthometric heights 161
so that
dyn
∆HAB = ∆nAB + DCAB , (4–17)
where
B
. g − γ0
B
g − γ0
DCAB = dn = δn (4–18)
A γ0 γ0
A
C = W0 − W , (4–20)
and H its orthometric height, that is, the length of the plumb-line segment
between P0 and P . Perform the integration in (4–10) along the plumb line
P0 P . This is permitted because the result is independent of the path. We
then get
H
C= g dH . (4–21)
0
H earth's surface
geoid
P0
or
ḡ = g + 0.0424 H (g in gal, H in km) . (4–31)
The factor 0.0424 refers to the normal density = 2.67 g/cm3 . The corre-
sponding formula for arbitrary constant density is, by (3–43),
1 ∂γ
ḡ = g − + 2π G H . (4–32)
2 ∂h
C
H= (4–33)
g + 0.0424 H
This presupposes that gravity g varies linearly along the plumb line. This
can usually be assumed with sufficient accuracy, even in extreme cases, as
shown by Mader (1954) and by Ledersteger (1955).
Orthometric correction
The orthometric correction is added to the measured height difference, in
order to convert it into a difference in orthometric height.
We let the leveling line connect two points A and B (Fig. 4.5) and apply
a simple trick first:
∆HAB = HB − HA = HB − HA − HB
dyn dyn
+ HA dyn
+ (HB − HA
dyn
)
(4–35)
dyn
= ∆HAB + (HB − HB
dyn
) − (HA − HA
dyn
).
164 4 Heights
earth's
B surface
geoid
A0 B0
Consider now the differences between the orthometric and dynamic heights,
HA − HA dyn
and HB − HB dyn
. Imagine a fictitious leveling line leading from
point A0 at the geoid to the surface point A along the plumb line. Then the
measured height difference would be HA itself: ∆nA0A = HA , so that
dyn
DCA0 A = ∆HA 0A
− ∆nA0 A = HA
dyn
− HA (4–37)
and
HA − HA
dyn
= −DCA0 A ,
(4–38)
HB − HB
dyn
= −DCB0 B .
Substituting (4–36) and (4–38) into (4–35), we finally have
or
∆HAB = ∆nAB + OCAB , (4–40)
where
OCAB = DCAB + DCA0 A − DCB0 B (4–41)
is the orthometric correction. This is a remarkable relation between the or-
thometric and dynamic corrections (Ledersteger 1955). We may write this
where we have reversed the sequence of the subscripts of the last term and,
consequently, the sign. With DCB0 A0 = 0 (why?), we may write
where ḡA and ḡB are the mean values of gravity along the plumb lines of A
and B. Thus, the orthometric correction (4–41) becomes
B
g − γ0 ḡA − γ0 ḡB − γ0
OCAB = δn + HA − HB . (4–46)
γ0 γ0 γ0
A
Here again we need the mean values of gravity along the plumb lines, ḡA and
ḡB ; γ0 is an arbitrary constant for which we always take normal gravity at
45◦ latitude.
Accuracy
Let us first evaluate the effect on H of an error in the mean gravity ḡ. From
H = C/ḡ, we obtain by differentiation
C H
δH = − 2
δḡ = − δḡ . (4–47)
ḡ ḡ
Since ḡ is about 1000 gal, we have, neglecting the minus sign, the simple
formula
.
δH[mm] = δḡ[mgal] H[km] , (4–48)
the subscripts denoting the units; δH is the error in H, caused by an error
δḡ in ḡ.
For H = 1 km,
.
δH[mm] = δḡ[mgal] , (4–49)
166 4 Heights
which shows that an error δḡ in the order of 100 mgal falsifies an elevation
of 1000 m by only 10 cm.
Let us now estimate the effect of an error of the density on ḡ. Differ-
entiating (4–32) and omitting the minus sign we find
δḡ = 2π G H d . (4–50)
C
∗ dC
H = , (4–54)
0 γ
C = γ̄ H ∗ , (4–55)
where H∗
1
γ̄ = ∗ γ dH ∗ (4–56)
H 0
is the mean normal gravity along the plumb line.
As the normal potential U is a simple analytic function, these formulas
can be evaluated straightforwards; but since the potential of the earth is
evidently not normal, what does all this mean? Consider a point P on the
physical surface of the earth. It has a certain potential WP and also a certain
normal potential UP , but in general WP = UP . However, there is a certain
point Q on the plumb line of P , such that UQ = WP ; that is, the normal
potential U at Q is equal to the actual potential W at P . The normal height
H ∗ of P is nothing but the ellipsoidal height of Q above the ellipsoid, just
as the orthometric height of P is the height of P above the geoid.
For more details the reader is referred to Sect. 8.3; Fig. 8.2 illustrates
the geometric relations.
We now give some practical formulas for the computation of normal
heights from geopotential numbers. Writing (4–56) in the form
H∗
1
γ̄ = ∗ γ(z) dz (4–57)
H 0
corresponding to (4–28), then we can express γ(z) by (2–215) as
2 3 2
γ(z) = γ 1 − 1 + f + m − 2f sin ϕ z + 2 z ,
2
(4–58)
a a
where γ is the gravity at the ellipsoid, depending on the latitude ϕ but not
on z. Thus, straightforward integration with respect to z yields
1 2 z2 3 z 3 H ∗
γ̄ = ∗ γ z − 1 + f + m − 2f sin ϕ
2
+ 2
H a 2 a 3 0
(4–59)
1 ∗ 1 ∗2 1 ∗3
= ∗γ H − 1 + f + m − 2f sin ϕ H + 2 H
2
H a a
or H ∗2 H ∗2
γ̄ = γ 1 − 1 + f + m − 2f sin ϕ 2
+ 2 . (4–60)
a a
This formula may be used for computing H ∗ by the formula
C
H∗ = . (4–61)
γ̄
168 4 Heights
where γ is normal gravity at the ellipsoid, for the same latitude ϕ. The
accuracy of this formula will be sufficient for almost all practical purposes;
still more accurate expressions are given in Hirvonen (1960).
Corresponding to the dynamic and orthometric corrections, there is a
normal correction NC of the measured height differences. Equation (4–46)
immediately yields, on replacing ḡ by γ̄ and H by H ∗ :
B
g − γ0 γ̄A − γ0 ∗ γ¯B − γ0 ∗
NCAB = δn + HA − HB , (4–64)
γ0 γ0 γ0
A
so that
∗ ∗ ∗
∆HAB = HB − HA = ∆nAB + NCAB . (4–65)
The normal heights were introduced by Molodensky in connection with his
method of determining the physical surface of the earth; see Chap. 8.
we can write the different kinds of height in a common form which is very
instructive:
C
height = , (4–67)
G0
4.5 Comparison of different height systems 169
where the height systems differ according to how the gravity value G0 in the
denominator is chosen. We have:
dynamic height: G0 = γ0 = constant ,
normal height: G0 = γ̄ .
It is seen that one can devise an unlimited number of other height systems
by selecting G0 in a different way.
The geopotential number C is, in a way, the most direct result of leveling
and is of great scientific importance. However, it is not a height in a geomet-
rical or practical sense. While the dynamic height has at least the dimension
of a height, it has no geometrical meaning. One advantage is that points of
the same level surface have the same dynamic height; this corresponds to the
intuitive feeling that if we move horizontally, we remain at the same height.
Note that the orthometric height differs for points of the same level surface
because the level surfaces are not parallel. This gives rise to the well-known
paradoxes of “water flowing uphill”, etc.
The dynamic correction can be very large, because gravity varies from
equator to pole by about 5000 mgal. Take, for instance, a leveling line of
.
1000 m difference of height at the equator, where g = 978.0 gal, computed
.
with γ0 = γ45◦ = 980.6 gal. Then (4–18) gives a dynamic correction of
approximately
978.0 − 980.6
DC = · 1000 m = −2.7 m . (4–69)
980.6
Because of these large corrections, dynamic heights are not suitable as prac-
tical heights, and the geopotential numbers are preferable for scientific pur-
poses.
Orthometric heights are the natural “heights above sea level”, that is,
heights above the geoid. Therefore, they have an unequalled geometrical
and physical significance. Their computation is relatively laborious, unless
Helmert’s simple formula (4–33) is used, which is sufficient in most cases.
The orthometric correction is rather small. In the Alpine leveling line of
Mader (1954), leading from an elevation of 754 m to 2505 m, the orthometric
correction is about 15 cm per 1 km of measured height difference. See also
Sect. 8.15.
The physical and geometrical meaning of the normal heights is less ob-
vious; they depend on the reference ellipsoid used. Although they are basic
in the new theories of physical geodesy, they have a somewhat artificial
character as compared to the orthometric heights. They are, however, easy
170 4 Heights
1. misclosures be eliminated,
2. corrections to the measured heights be as small as possible.
Empirical height systems have been devised to give smaller corrections than
either the orthometric or the normal heights. They have no clear physical
significance, however, and are beyond the scope of this book.
Accuracy
Leveling is one of the most accurate geodetic measurements. A standard
error of ±0.1 mm per km distance is possible; it increases with the square
root of the distance.
If the error of measurement and interpolation, etc., of gravity is negligible,
then the differences in the geopotential number C can be determined with
an accuracy of ±0.1 gal m per km distance; this corresponds to ±0.1 mm
in measured height. Referring to gravity measurements, it is sufficient to
measure at distances of some kilometers.
Dynamic heights and normal heights are clearly as accurate as the geopo-
tential numbers, because normal gravity γ is errorless. Orthometric heights,
however, are also affected by imperfect knowledge of density, etc., but only
slightly; see the end of Sect. 4.3.
Triangulated heights
Historically and for the sake of completeness, the determination of heights
by triangulation, that is, by means of zenith angles, should be mentioned.
The main problem is the atmospheric refraction affecting the zenith an-
gles. Thus, the accuracy of triangulated heights is much less than that of
leveling. Consequently, triangulated heights are not considered any longer
here.
For small distances (e.g., < 1 km), trigonometric height measurements,
referred to the local plumb line, have the character of a leveled height dif-
ference δn This fact may be used (with care!) to fill small gaps in a leveling
network.
4.6 GPS leveling 171
Remark on misclosures
All misclosures in any acceptable system of heights denoted for the moment
by h (not to be confused with ellipsoidal heights) must be zero:
$
dh = 0 (4–70)
_ _
A B
l2
l1 B
earth's surface
HB – HA
A
HB
HA hB
hA geoid
NA NB
ellipsoid
With GPS leveling, δhAB is obtained, so that with a known geoid, i.e., known
δNAB , the orthometric height difference δHAB may be computed accord-
ing to (4–75). This is a tremendous advantage since otherwise the classical
leveling together with gravity measurements is required to determine the
orthometric height difference, see Eqs. (4–40) and (4–46).
Note that only the difference of the geoidal undulations impacts the
result.
5 The geometry of the earth
5.1 Overview
This chapter consists of three parts.
5.2 Introduction
Geodesy, as the theory of size and shape of the earth, is not a purely geo-
metrical science since the earth’s gravity field, a physical entity, is involved
in many geodetic measurements, especially terrestrial ones.
The gravimetric methods are usually considered to constitute physical
geodesy in the narrower sense. The measurements of triangulation, leveling,
and geodetic astronomy, all make essential use of the plumb line, which,
being the direction of the gravity vector, is no less physically defined by
nature than its magnitude, that is, the gravity g. All determinations of the
geoid by various methods and its use as well as the use of deflections of the
vertical belong to physical geodesy, quite as well as the gravimetric methods.
Even in the age of GPS, we have many previous geodetic data which
continue to be useful and have to be understood in order to be optimally
combined with the new satellite data. In precise operations of engineering
geodesy such as tunnel surveying, the plumb line and deflections of the ver-
tical must be taken into account.
For an optimal understanding and use of local (or rather regional) geo-
detic datums, we must know their relation to a global geodetic system as
used in GPS. Therefore, it is appropriate to start with global geometry in a
rather elementary way.
A few introductory ideas may help in comprehending this subject. To
fix the position of a point in space, we need three coordinates. We can use,
and have used, a rectangular Cartesian coordinate system. This is the basic
geometric coordinate system. It may be easily converted computationally to
ellipsoidal coordinates ϕ, λ, h referred to any given reference ellipsoid.
For many special purposes, however, it is preferable to take what we have
called the natural coordinates: Φ (astronomical latitude), Λ (astronomical
longitude), and H (orthometric height), which directly refer to the gravity
field of the earth (Sect. 2.4). The height H may be obtained by geometric
leveling, combined with gravity measurements, and Φ and Λ are determined
by astronomical measurements.
As long as the geoid can be identified with an ellipsoid, the use of these
coordinates for computations is very simple. Since this identification is suf-
ficient only for results of rather low accuracy, the deviations of the geoid
from an ellipsoid must be taken into account. As we have seen, the geoid has
rather disagreeable mathematical properties. It is a complicated surface with
discontinuities of curvature. Thus, it is not suitable as a surface on which to
perform mathematical computations directly, as on the ellipsoid.
176 5 The geometry of the earth
ϕ and λ are the ellipsoidal coordinates on the ellipsoid, sometimes also called
geodetic latitude and geodetic longitude to distinguish them from the astro-
nomical latitude Φ and the astronomical longitude Λ. Astronomical and el-
lipsoidal coordinates differ by the deflection of the vertical (components ξ
and η). The quantity h is the geometric height above the ellipsoid; it differs
from the orthometric height H above the geoid by the geoidal undulation N .
Geodetic measurements (angles, distances) are treated similarly. The
principle of triangulation is well known: historically, distances were obtained
indirectly by measuring the angles in a suitable network of triangles; only
one baseline was necessary in principle to furnish the scale of the network.
Triangulation was indispensable in former times, because angles could be
measured much more easily than long distances.
Nowadays, however, long distances can be measured directly just as eas-
ily as angles by means of electronic instruments, so that triangulation, using
angular measurements, is often replaced or supplemented by trilateration,
using distance measurements. The computation of triangulations and trilat-
erations on the ellipsoid is easy. It is, therefore, convenient to reduce the
measured angles, baselines, and long distances to the ellipsoid, in much the
same way as the astronomical coordinates are treated. Then the ellipsoidal
coordinates ϕ, λ obtained (1) by reducing the astronomical coordinates and
(2) by computing triangulations or trilaterations on the ellipsoid can be
compared; they should be identical for the same point.
Today, of course, GPS is the best method for determining ϕ, λ, and h
directly.
Satellites categories
Essentially, the GPS satellites provide a platform for radio transceivers,
atomic clocks, computers, and various ancillary equipment. The electronic
equipment of each satellite allows the user to measure a pseudorange to the
satellite, and each satellite broadcasts a message which allows the user to
determine the spatial position of the satellite for arbitrary instants. The aux-
iliary equipment of each satellite, among others, consists of solar panels for
power supply and a propulsion system for orbit and stability control.
There are several classes or types of GPS satellites. These are the Block I,
Block II, Block IIA, Block IIR, Block IIR-M, and the future Block IIF and
Block III satellites. An up-to-date description is difficult because new nota-
tions are introduced in a rather arbitrary way; an example is the recently
introduced notation Block IIR-M.
Eleven Block I satellites were launched in the period between 1978 to
1985. Today, none of them is in operation anymore.
The essential difference between Block I and Block II satellites is related
to U.S. national security. Block I satellite signals were fully available to
civilian users. Starting with Block II, satellite signals may be restricted for
civilian use. The Block II satellites are equipped with mutual communication
capability. Some of them carry retroreflectors and can be tracked by laser
ranging.
The Block IIR satellites (“R” denotes replenishment or replacement)
have a design life of 10 years. They are equipped with improved facilities for
communication and intersatellite tracking. Block IIR-M satellites incorpo-
rate two new military signals and a second civil signal. The first Block IIR-M
was launched on September 25, 2005.
Currently (April 2006), the first launch of a Block IIF satellite (“F”
denotes follow on) is scheduled for 2008 (instead of the previously projected
dates mid of 2006 and 2007). These satellites will broadcast a third civil
signal on L5 (see Sect 5.3.5).
Presently, the DOD undertakes studies for the next generation of GPS
satellites, called Block III satellites. Preliminary dates (likely to change) are
2011/12 for first launches and on-orbit tests (Civil GPS Service Interface
5.3 The Global Positioning System 179
Satellite signal
The key to the accuracy of the system is the fact that all signal compo-
nents are precisely controlled by atomic clocks. These highly accurate fre-
quency standards of GPS satellites produce the fundamental frequency of
10.23 MHz. Coherently derived from this frequency are (presently) two sig-
nals in the L-band, the L1 and the L2 carrier waves generated by multiplying
the fundamental frequency by 154 and 120, respectively, yielding
L1 = 1575.42 MHz ,
L2 = 1227.60 MHz .
These dual frequencies are essential for eliminating the major source of error,
i.e., the ionospheric refraction.
The pseudoranges that are derived from measured travel times of the
signal from each satellite to the receiver use two pseudorandom noise (PRN)
codes that are modulated onto the two carriers.
The C/A-code (coarse/acquisition-code) is available for civilian use. Each
C/A-code is a unique sequence of 1023 bits, called chips, which is repeated
each millisecond. The duration of each C/A-code chip is about 1 µs. Equiv-
alently, the chip length – denoted also as wavelength or chip width (Misra
and Enge 2001: Sect. 2.3.1) – is about 300 m. The C/A-code is presently
modulated upon L1 only and is purposely omitted from L2. This omission
allows the JPO to control the information broadcast by the satellite and,
thus, denies full system accuracy to nonmilitary users.
The P-code (precision-code) has been reserved for U.S. military and other
authorized users. This is achieved by using the W-code to encrypt the P-
code to the Y-code (anti-spoofing). The P-code has an effective chip length
of about 30 m. The P-code is modulated on both carriers L1 and L2.
In addition to the PRN codes, a data message is modulated onto the
carriers consisting of status information, satellite clock bias, and satellite
ephemerides. The orbit data are given as Kepler-like elements and are de-
noted as broadcast ephemerides. The full set of elements is given in, e.g.,
Montenbruck and Gill (2001: Sect. A.2.2). It is worth noting that the present
signal structure will be improved in the near future (see Sect. 5.3.5).
Control segment
The operational control system (OCS) consists of a master control station,
monitor stations, and ground control stations. The main tasks of the OCS
180 5 The geometry of the earth
are tracking of the satellites for the orbit and clock determination and predic-
tion, time synchronization of the satellites, and upload of the data message
to the satellites.
Monitor stations
There are five monitor stations located at Hawaii, Colorado Springs, As-
cension Island in the South Atlantic Ocean, Diego Garcia in the Indian
Ocean, and Kwajalein in the North Pacific Ocean. Each of these stations
is equipped with a precise atomic time standard and receivers which con-
tinuously measure pseudoranges to all satellites in view. Pseudoranges are
measured every 1.5 seconds and, using ionospheric and meteorological data,
they are smoothed to produce 15-minute interval data which are transmitted
to the master control station.
User segment
The diversity of the military and civilian users is matched by the type of
receivers available today.
On the basis of the type of observables (i.e., code pseudoranges or phase
pseudoranges) and of the availability of codes (i.e., C/A-code, P-code, or
Y-code), GPS receivers can be classified. For the majority of navigation ap-
plications, C/A-code pseudorange receivers will suffice. With this type of
receiver, only code pseudoranges using the C/A-code on L1 are measured.
Typical devices output the three-dimensional position either in latitude, lon-
gitude, and height or in some map projection systems, e.g., universal trans-
verse Mercator (UTM) coordinates and height.
5.3 The Global Positioning System 181
Observables
In concept, the GPS observables are ranges which are deduced from mea-
sured time or phase differences based on a comparison between received
signals and receiver-generated signals. As mentioned earlier, the ranges are
biased by satellite and receiver clock errors and, consequently, they are de-
noted as pseudoranges. Essentially, pseudoranges differ from distances by an
unknown additive constant.
Apart from the satellite and the receiver clock bias, further error sources
can be classified into three groups, i.e., satellite-related errors (e.g., orbital
errors), signal propagation medium-related errors (e.g., ionospheric and tro-
pospheric refraction), and receiver-related errors (e.g., antenna phase center
variation, multipath), but are omitted in the subsequent simplified models.
Extended models are given in Hofmann-Wellenhof et al. (2003: Sect. 10.2.2).
Code pseudoranges
The measured time difference ∆t is affected by the satellite clock error δS
and the receiver clock error δ. The error δS of the satellite clock can be
modeled by a polynomial with the coefficients being transmitted in the nav-
igation message. Assuming the δS correction is applied, the time interval ∆t
multiplied by the speed of light c yields the code pseudorange R and, hence,
R = c ∆t . (5–2)
Assuming a common time reference for satellite and receiver, e.g., GPS time,
the term ∆t may be decomposed into the run time ∆t(GPS) and the receiver
clock errors δ leading to
R = c ∆t(GPS) + c δ = + c δ , (5–3)
where is the geometric range between the satellite and the receiver. The
receiver module responsible for code pseudorange measurements is denoted
as delay lock loop (DLL). Details on the DLL functionality are given in Misra
and Enge (2001: Sect. 9.5).
Phase pseudoranges
Assuming again that the satellite clock error correction is applied, the phase
pseudorange Φ is modeled by
λΦ = + cδ + λN , (5–4)
where the carrier wavelength λ has been introduced. The range represents
the distance between the satellite at emission epoch t and the receiver at
reception epoch t + ∆t. Phase measurements are ambiguous, since the initial
integer number N of cycles between satellite and receiver is unknown. As
long as the tracking of a satellite is not interrupted, the ambiguity remains
constant within the tracking loop of the receiver. The responsible receiver
hardware is denoted as phase lock loop (PLL). Compared to (5–3), the phase
pseudorange differs from the code pseudorange only by the phase ambiguity
term λ N . Dividing the above equation by λ scales the phase to cycles.
As mentioned previously, the majority of navigation applications does not
need carrier phase measurements. Only for increased accuracy requirements
(e.g., relative positioning; see below), phase measurements become relevant.
Doppler data
Some of the first solution models proposed for GPS were to use the Doppler
observable. Considering Eq. (5–4), the equation for the observed Doppler
5.3 The Global Positioning System 183
D = λ Φ̇ = ˙ + c δ̇ , (5–5)
where the derivatives with respect to time are indicated by a dot. The raw
Doppler shift is less accurate than integrated Doppler.
The Doppler shift is measured in the carrier tracking loop of a GPS re-
ceiver (Misra and Enge 2001: Sect. 9.6). Assuming a known satellite velocity,
the Doppler shift can be used to estimate the velocity of the user.
Point positioning
When using a single receiver, usually point positioning with code pseudo-
ranges is performed. The concept of point positioning is simple (Fig. 5.2).
Without clock errors, trilateration in space (i.e., using three ranges) solves
the task to determine the point coordinates. Using pseudoranges, four ob-
servations are necessary to account for the three coordinate components and
the receiver clock error. For point positioning, GPS provides two levels of
service: the standard positioning service (SPS) with access for civilian users
and the precise positioning service (PPS) with access for authorized users.
SPS performance standards are based on signal-in-space performance.
Contributions of ionosphere, troposphere, receiver, multipath, topography, or
interference are not included. Furthermore, SPS is provided on the L1 signal
only; the L2 signal is not part of the SPS (Department of Defense 2001).
The global average positioning domain accuracy amounts to 13 m horizontal
error (95% probability level) and 22 m vertical error (95% probability level).
184 5 The geometry of the earth
The PPS has access to both codes and provides accuracies down to the
meter level.
Differential GPS
Selective availability (SA), the deliberate degradation of the point position-
ing accuracy by “dithering” (i.e., distorting on purpose) the satellite clock
(called δ-process) and manipulating the ephemerides (called ε-process), has
led to the development of differential GPS (DGPS). Only the basic idea is
explained here.
DGPS is based on the use of two (or more) receivers, where one (station-
ary) reference or base receiver is located at a known point and the position
of the (mostly moving) remote receiver is to be determined. Using code pseu-
doranges, at least four common satellites must be tracked simultaneously at
both sites. The known position of the reference receiver is used to calculate
corrections to the observed pseudoranges. These corrections are then trans-
mitted via telemetry (i.e., controlled radio link) to the roving receiver and
allow the computation of the rover position with far more accuracy than for
the single-point positioning mode.
Using DGPS based on C/A-code pseudoranges, real-time accuracies at
the 1–5 m level can be routinely achieved. Phase-smoothed code ranges yield
the submeter level (Lachapelle et al. 1992). Even higher accuracies can be
reached by the use of carrier phases (precise DGPS). For ranges up to some
20 km, accuracies at the subdecimeter level can be obtained in real time (De-
Loach and Remondi 1991). To achieve this accuracy, the ambiguities must
be resolved “on the fly” and, therefore, (generally) dual-frequency receivers
are required. Furthermore, five satellites per epoch are required.
After the deactivation of SA in May 2000, DGPS must be seen from a
different viewpoint. The increased point positioning accuracy achieved with
a single receiver may suffice for some kinds of applications.
Relative positioning
At present, highest accuracies are achieved in the relative-positioning mode
with observed carrier phases. Relative positioning is associated with base-
lines, i.e., the three-dimensional vector between a known reference station
and the location to be determined. Processing a baseline requires that the
phases are simultaneously observed at both baseline endpoints (Fig. 5.1).
Originally, relative positioning was only possible by postprocessing data.
Today, (near) real-time data transfer over short baselines is routinely possi-
ble, which enables real-time computation of baseline vectors and has led to
the real-time kinematic (RTK) technique.
5.3 The Global Positioning System 185
satellite
vector
baseline
unknown station
reference station
where f0 = 10.23 MHz denotes the basic GPS frequency. The carrier L5,
placed in a protected aeronautical radio navigation service band, was re-
cently allocated by the World Radio Conference organized regularly by the
International Telecommunication Union (Vorhies 2000).
Note that both new civil GPS signals will have two codes. L5 will not
share with military signals and use two equal-length codes in phase quadra-
ture, each clocked at 10.23 MHz. L2 is shared between civil and military
signals. The new L2c signal provides two codes by time multiplexing. The
two codes are of different length (Fontana et al. 2001). The existing military
Y-code will be replaced by new (split) M-codes.
The linear carrier phase combination of L2 with L5 results in a signal
with a wavelength of about 5.9 m. Long wavelengths facilitate ambiguity
resolution. By contrast, the linear combination of L1 with L5 will be used
as ionosphere-free combination because large frequency differences are ad-
vantageous for calculating ionospheric corrections. The common processing
of phase data from all three carriers will be performed in the three-carrier
ambiguity resolution approach (Vollath et al. 1999).
A perspective for the implementation is given in the 2001 Federal Radio-
navigation Plan: IOC (18 satellites in orbit with the new L2c signal and
M-code capability) is planned for 2008 and FOC (24 satellites in orbit) is
planned for 2010. At least one satellite is planned to be operational with the
new L5 capability no later than 2005, with IOC planned for 2012 and FOC
planned for 2014.
where the WGS 84 (World Geodetic System 1984, see Sect. 2.11) coordinates
X j (t), Y j (t), Z j (t) are the components of the geocentric position vector of
the satellite at epoch t, and XA , YA , ZA are the three unknown WGS 84 co-
ordinates of the observing site, which might be denoted (XA , YA , ZA )WGS 84
or, which means the same, (XA , YA , ZA )GPS .
How many unknowns are involved? Note that the satellite coordinates
X j (t), Y j (t), Z j (t) may always be assumed known (more precisely, are cal-
k
j l
%Ak (t) m
Z WGS-84 %Al (t)
j
%A (t) %A
m
(t)
Y WGS-84
ZA
XA
X WGS-84
YA
culable) from the information broadcast by the satellite. Therefore, there re-
main the three unknown station coordinates XA , YA , ZA and the unknown
receiver clock error δA (t). In other terms, at least four satellites are required
to set up four equations of type (5–6). Denoting the satellites by j, k, l, m,
the corresponding system of equations
j
RA (t) = jA (t) + c δA (t) ,
k (t) = k (t) + c δ (t) ,
RA A A
(5–8)
l (t) = l (t) + c δ (t) ,
RA A A
k
l
j
%Ak (t) %Al (t)
j %A
m
(t) m
Z WGS-84 %A (t)
j
%Bk (t) %Bl (t)
%B (t)
%Bm(t)
B
A bAB
ZA ZB
Y WGS-84
YB XB
X WGS-84 XA
YA
the relation
XB = XA + bAB (5–10)
may be formulated, and the components of the baseline vector bAB are
⎡ ⎤ ⎡ ⎤
XB − XA ∆XAB
bAB = ⎣ YB − YA ⎦ = ⎣ ∆YAB ⎦ . (5–11)
ZB − ZA ∆ZAB
The coordinates of the reference point must be given in the WGS 84 and are
usually approximated by a code pseudorange solution. Relative positioning
can be performed with code pseudoranges (cf. Eq. (5–3)) or with phase
pseudoranges (cf. Eq. (5–4)). Subsequently, only phase pseudoranges are
explicitly considered. We repeat (5–4),
λΦ = + cδ + λN , (5–12)
where we have already explained the wavelength λ, the phase Φ, the distance
(which is the same as for the code pseudorange model), the speed of light
c, the receiver clock error δ, and the ambiguity N in Sect. 5.3.3.
Introducing f , the frequency of the corresponding satellite signal, and
taking into account the relation f = c/λ, we may divide (5–12) by λ obtain-
ing
1
Φ= +fδ+N. (5–13)
λ
This may be generalized to
1 j
Φji (t) = (t) + f δi (t) + Nij , (5–14)
λ i
5.4 From GPS to coordinates 191
where Φji (t) is the measured carrier phase expressed in cycles referred to
station i and satellite j at epoch t. The time-independent phase ambiguity
Nij is an integer number and, therefore, often called integer ambiguity or
integer unknown or simply ambiguity.
Relative positioning requires simultaneous observations at both the ref-
erence and the unknown point. This means that the observation time tags
for the two points must be the same. Assuming such observations (5–14) at
the two points A and B to satellite j and another satellite k simultaneously
at epoch t, the following measurement equations may be set up:
1 j
ΦjA (t) = (t) + f δA (t) + NAj ,
λ A
1
ΦkA (t) = kA (t) + f δA (t) + NAk ,
λ (5–15)
1
ΦB (t) = jB (t) + f δB (t) + NBj ,
j
λ
1
ΦkB (t) = kB (t) + f δB (t) + NBk .
λ
Introducing the short-hand notations
Φjk k j k j
AB (t) = ΦB (t) − ΦB (t) − ΦA (t) + ΦA (t) ,
jk k j k j
AB (t) = B (t) − B (t) − A (t) + A (t) ,
(5–16)
jk
NAB = NBk − NBj − NAk + NAj ,
the known station A to achieve the high accuracy. Note that the resulting
coordinates are obtained in the WGS 84.
This concludes the short introduction how the user of GPS gets WGS 84
coordinates, i.e., geocentric rectangular coordinates X, Y, Z or, computed
from them, ellipsoidal coordinates ϕ, λ, h; see Sect. 5.6.1.
" earth's
surface
±
hH
P0 geoid
N
Q Q0 ellipsoid
H
ϕHelmert = ϕPizzetti + ξ,
R
(5–18)
H
λHelmert = λPizzetti + η sec ϕ ,
R
.
which can be read from Fig. 5.4, since QQ0 = H ε; R = 6371 km is the
mean radius of the earth. Even if ε = 1 arc minute and H = 1000 m, the
distance QQ0 is only about 30 cm and the ellipsoidal coordinates differ by
less than 0.01 , which is below the accuracy of astronomical observations.
For most purposes, we may, therefore, neglect the difference between the two
projections.
Pizzetti’s projection is better adapted to the geoid, because there is an
exact correspondence between a geoidal point P0 and an ellipsoidal point
Q0 . Helmert’s projection has overwhelming practical advantages, notably
the straightforward conversion of the ellipsoidal coordinates ϕ, λ, h into rect-
angular coordinates x, y, z; it is also simpler in other respects. The decisive
advantage of Helmert’s projection is its direct relation to GPS. It is, there-
fore, exclusively used now in practice.
x2 + y 2 z 2
+ 2 = 1. (5–19)
a2 b
The representation of this ellipsoid in terms of ellipsoidal coordinates is given
by
x = N cos ϕ cos λ ,
z
P
p
h
nQ
X Q
x
'
x,y plane
N
a2
N= . (5–21)
a2 cos2 ϕ + b2 sin2 ϕ
These equations are known from ellipsoidal geometry; it may also be verified
by direct substitution that a point with xyz-coordinates (5–20) satisfies the
equation of the ellipsoid (5–19) and so lies on the ellipsoid. The components
of the unit normal vector n are
n = cos ϕ cos λ , cos ϕ sin λ , sin ϕ , (5–22)
because ϕ is the angle between the ellipsoidal normal and the xy-plane,
which is the equatorial plane (Fig. 5.5). Now let the coordinates of a point
P outside the ellipsoid form the vector
X = [X , Y , Z ] ; (5–23)
x = [x , y , z ] . (5–24)
that is
X = x + h cos ϕ cos λ ,
Z = z + h sin ϕ .
X = (N + h) cos ϕ cos λ ,
Y = (N + h) cos ϕ sin λ ,
(5–27)
2
b
Z = N + h sin ϕ .
a2
These equations are the basic transformation formulas between the ellip-
soidal coordinates ϕ, λ, h and the rectangular coordinates X, Y, Z of a point
outside the ellipsoid. The origin of the rectangular coordinate system is the
center of the ellipsoid, and the z-axis is its axis of rotation; the x-axis has
the Greenwich longitude 0◦ and the y-axis has the longitude 90◦ east of
Greenwich (i.e., λ = +90◦ ).
A possible source of confusion is that the normal radius of curvature of
the ellipsoid and the geoidal undulation are both denoted by the symbol N ;
in (5–27), N is, of course, the normal radius of curvature. Generally, let the
context decide between quantities of such different magnitude (6000 km and
60 m).
Equations (5–27) permit the computation of rectangular coordinates
X, Y, Z from the ellipsoidal coordinates ϕ, λ, h.
The inverse procedure, the computation of ϕ, λ, h from given X, Y, Z, is
frequently performed iteratively, although a solution in closed form exists.
A possible iterative
√ procedure is as follows.
Denoting X + Y 2 by p, we get from the first two equations of (5–27)
2
so that
p
h= −N. (5–29)
cos ϕ
The third equation of (5–27) may be transformed into
a2 − b2
Z= N− N + h sin ϕ = (N + h − e2 N ) sin ϕ , (5–30)
a2
196 5 The geometry of the earth
where e2 = (a2 − b2 )/a2 . Dividing this equation by the above expression for
p, we find
Z 2 N
= 1−e tan ϕ , (5–31)
p N +h
so that −1
Z N
tan ϕ = 1 − e2 . (5–32)
p N +h
Given X, Y, Z, and hence p, Eqs. (5–29) and (5–32) may be solved iteratively
for h and ϕ. As a first approximation, we set h = 0 in (5–32), obtaining
Z
tan ϕ(1) = (1 − e2 )−1 . (5–33)
p
Z + e2 b sin3 θ
ϕ = arctan ,
p − e2 a cos3 θ
Y (5–36)
λ = arctan ,
X
p
h= −N,
cos ϕ
where
Za
θ = arctan (5–37)
pb
is an auxiliary quantity and
are
√ first and second numerical eccentricity. As introduced in (5–28), p =
X 2 + Y 2 . Actually, there is no reason why these formulas are less popular
than the iterative procedure since there is no significant difference between
the two methods. Computation methods with neither iteration nor approx-
imation are, e.g., given by Sünkel (1977) and Zhu (1993).
• ellipsoidal coordinates: ϕ, λ, h;
• ellipsoidal-harmonic coordinates: β, λ, u,
alternatively: ϑellipsoidal-harmonic , λ, u;
• spherical coordinates: ϕ̄, λ, r, alternatively: ϑspherical , λ, r.
The longitude λ is the same in all triples. The ellipsoidal coordinates latitude
ϕ and longitude λ are sometimes also denoted geodetic latitude and geodetic
z
2 2
radius u + E
ere of
sp h
hP
throug P
oid
l lips
le
a soid
oc ellip h
nf c e #ell-har.
ren Q
co
efe
#sph. r
r
¯
F1 ' ' F2
x,y plane
E = a2 – b 2 u2 + E 2
ϑellipsoidal-harmonic = 90◦ − β ,
(5–39)
ϑspherical = 90◦ − ϕ̄ .
Note, however, that we did not use these indications to distinguish be-
tween the spherical and the ellipsoidal-harmonic ϑ! Thus, the reader is chal-
lenged to attentively distinguish between these quantities. Wherever possi-
ble, we tried to avoid conflicts.
Some examples: we used the spherical coordinates r, ϑ, λ in Sects. 1.4,
1.11, 1.12, 1.14, 2.5, 2.6, 2.13, 2.18, etc. We used the ellipsoidal-harmonic
coordinates u, ϑ, λ in Sects. 1.15, 1.16; we used the ellipsoidal-harmonic co-
ordinates u, β, λ in Sects. 2.7, 2.8, and we used the spherical coordinates
r, ϑ, λ as well as the ellipsoidal-harmonic coordinates u, β, λ in Sect. 2.9.
The following equations express the rectangular coordinates in these
three systems:
√
X = (N + h) cos ϕ cos λ = u2 + E 2 cos β cos λ = r cos ϕ̄ cos λ ,
√
Y = (N + h) cos ϕ sin λ = u2 + E 2 cos β sin λ = r cos ϕ̄ sin λ ,
(5–40)
2
b
Z = N + h sin ϕ = u sin β = r sin ϕ̄ .
a2
These relations, which follow from combining Eqs. (1–26), (1–151), and (5–
27), can be used if we wish to compute u and β from h and ϕ or from r and
ϕ̄, etc.
flattening f ) and (2) its position with respect to the earth or the geoid.
This relative position is most simply defined by the coordinates x0 , y0 , z0
of the center of the reference ellipsoid with respect to the geocenter. Since
the geocenter was not accessible to classical geodetic measurements before
the satellite era, a fundamental or initial point P1 on the earth surface was
chosen, such as Meades Ranch for North America and Potsdam for Central
Europe. It turns out that a convenient but conventional choice of the el-
lipsoidal coordinates ϕ1 , λ1 , h1 of the fundamental point P1 is equivalent to
x0 , y0 , z0 of the geocenter.
Thus, we have 5 defining parameters:
• 2 parameters a (semimajor axis) and f (flattening) as form parameters,
and
• 3 parameters x0 , y0 , z0 or ϕ1 , λ1 , h1 as position parameters.
Later on we shall also admit a scale factor and small rotations around the
three coordinate axes.
A (geodetic) datum transformation defines the relationship between a
global (geocentric) and a local (in general nongeocentric) three-dimensional
Cartesian coordinate system; therefore, a datum transformation transforms
one coordinate system of a certain type to another coordinate system of the
same type. This is one of the primary tasks when combining GPS data with
terrestrial data, i.e., the transformation of geocentric WGS 84 coordinates
to local terrestrial coordinates. The terrestrial system is usually based on a
locally best-fitting ellipsoid, e.g., the Clarke ellipsoid or the GRS-80 ellipsoid
in the U.S. and the Bessel ellipsoid in many parts of Europe. The local
ellipsoid is linked to a nongeocentric Cartesian coordinate system, where the
origin coincides with the center of the ellipsoid.
ZT
Z
"3
XT X Y
"2
x0
YT
"1
X
XT
account for the coordinates of the origin of the X system in the XT system.
Note that a single scale factor is considered. More generally (but with GPS
not necessary), three scale factors, one for each axis, could be used. The
rotation matrix is an orthogonal matrix which is composed of three successive
rotations
R = R 3 {ε3 } R2 {ε2 } R1 {ε1 } . (5–43)
Explicitly,
⎡ ⎤
cos ε2 cos ε3 cos ε1 sin ε3 sin ε1 sin ε3
⎢ + sin ε1 sin ε2 cos ε3 − cos ε1 sin ε2 cos ε3 ⎥
⎢ ⎥
⎢ ⎥
R=⎢
⎢− cos ε2 sin ε3 cos ε1 cos ε3 sin ε1 cos ε3 ⎥
⎥
⎢ − sin ε1 sin ε2 sin ε3 + cos ε1 sin ε2 sin ε3 ⎥
⎣ ⎦
sin ε2 − sin ε1 cos ε2 cos ε1 cos ε2
(5–44)
is obtained.
In the case of known transformation parameters x0 , µ, R, a point from
the X system can be transformed into the XT system by (5–41).
If the transformation parameters are unknown, they can be determined
with the aid of common (identical) points, also denoted as control points.
This means that the coordinates of the same point are given in both systems.
Since each common point (given by XT and X) yields three equations, two
common points and one additional common component (e.g., height) are
sufficient to solve for the seven unknown parameters. In practice, redun-
dant common point information is used and the unknown parameters are
calculated by least-squares adjustment.
Since the parameters are mixed nonlinearly in Eq. (5–41), a linearization
must be performed, where approximate values x 0 approx , µapprox , R approx are
required.
5.7 Geodetic datum transformations 201
µ = µapprox + δµ = 1 + δµ (5–45)
is obtained. Furthermore, the rotation angles εi in (5–44) are small and may
be treated as differential quantities. Introducing these quantities into (5–44),
setting cos εi = 1 and sin εi = εi , and considering only first-order terms gives
⎡ ⎤
1 ε3 −ε2
R = ⎣ −ε3 1 ε1 ⎦ = I + δR , (5–46)
ε2 −ε1 1
x 0 approx = X T − X (5–48)
follows by substituting the approximations for the scale factor and the rota-
tion matrix into Eq. (5–41).
Introducing Eqs. (5–45), (5–46), (5–47) into (5–41) and skipping de-
tails which can be found, for example, in Hofmann-Wellenhof et al. (1994:
Sect. 3.3) gives the linearized model for a single point i. This model can be
written in the form
X Ti − X i − x 0 approx = A i δp , (5–49)
where the left side of the equation is known and may formally be considered
as an observation. The design matrix A i and the vector δp, containing the
unknown parameters, are given by
⎡ ⎤
1 0 0 Xi 0 −Zi Yi
Ai = ⎣ 0 1 0 Yi Zi 0 −Xi ⎦ ,
0 0 1 Zi −Yi Xi 0 (5–50)
Recall that Eq. (5–49) is now a system of linear equations for point i. For n
common points, the design matrix A is
⎡ ⎤
A1
⎢ A2 ⎥
⎢ ⎥
A=⎢ . ⎥. (5–51)
⎣ .. ⎦
An
1. Transform the plane coordinates (y, x)LS of the common points into
the ellipsoidal surface coordinates (ϕ, λ)LS by using the appropriate
mapping formulas.
4. For network points other than the common points, transform the co-
ordinates (X, Y, Z)GPS into (X, Y, Z)LS via Eq. (5–41) using the trans-
formation parameters determined in the previous step.
6. Map the ellipsoidal surface coordinates (ϕ, λ)LS computed in the pre-
vious step into plane coordinates (y, x)LS by the appropriate mapping
formulas.
Y = y0 + (N + h) cos ϕ sin λ ,
(5–53)
2
b
Z = z0 + N + h sin ϕ .
a2
These equations form the starting point for various important differential
formulas of coordinate transformation.
204 5 The geometry of the earth
b
geocenter
Y
x0 x0
z0 a
y0 ellipsoid center
X
∂X ∂X ∂X ∂X ∂X
δX = δx0 + δa + δf + δϕ + δλ + δh ,
∂a ∂f ∂ϕ ∂λ ∂h
∂Y ∂Y ∂Y ∂Y ∂Y
δY = δy0 + δa + δf + δϕ + δλ + δh , (5–54)
∂a ∂f ∂ϕ ∂λ ∂h
∂Z ∂Z ∂Z ∂Z ∂Z
δZ = δz0 + δa + δf + δϕ + δλ + δh ,
∂a ∂f ∂ϕ ∂λ ∂h
a δϕ = sin ϕ cos λ δx0 + sin ϕ sin λ δy0 − cos ϕ δz0 + 2a sin ϕ cos ϕ δf ,
P
E2
Z
h + ±h
E1 h
' + ±'
±x0 a + ±a
'
a X
Y
Fig. 5.9. A small change of the reference ellipsoid together with a small
parallel shift
5.8 The three-dimensional geodesy of Bruns and Hotine 207
ui
ni
. ei
Pi
Xi
Fi Y
Li
ui (up, zenith)
Xij Pj
x ij = DTi X ij . (5–67)
210 5 The geometry of the earth
eij
tan Aij = ,
nij (5–69)
= uij
cos zij .
n2ij + e2ij + u2ij
Substituting (5–65) for nij , eij and uij , the measurement quantities may be
expressed by the components of the vector X ij in the global system.
is obtained, where (5–64) has also been taken into account, namely, the
fact that n i , e i , u i are unit vectors. Obviously, the second expression arises
immediately from the Pythagorean theorem. Differentiation of (5–70) yields
(5–71)
212 5 The geometry of the earth
where
Xij = Xj − Xi ,
Yij = Yj − Yi , (5–72)
Zij = Zj − Zi
have been introduced accordingly. The relation (5–71) may also be expressed
as
Xij Yij Zij
δsij = (δXj − δXi ) + (δYj − δYi ) + (δZj − δZi )
sij sij sij
(5–73)
if the differentials are replaced by differences.
Azimuths
Again the same principle applies: the measured azimuth Aij as a function of
the local level coordinates is given in (5–69). If nij , eij , uij , the components
of x ij , are substituted by (5–65), the relation
Directions
Measured directions Rij are related to azimuths Aij by the orientation un-
known oi . The relation reads
Zenith angles
The zenith angle zij as function of the local level coordinates is given in
(5–69). If nij , eij , uij , the components of x ij , are substituted by (5–65), the
relation
= u /s
cos zij ij ij
is obtained, where (5–70) and (5–72) have been used. After a lengthy deriva-
tion, the relation
is obtained.
It is presupposed that the zenith angles are reduced to the chord of the
light path. This reduction may be modeled by
sij
zij = zij meas
+ k, (5–80)
2R
where zijmeas is the measured zenith angle, R is the mean radius of the
earth, and k is the coefficient of refraction. For k either a standard value
may be substituted or the coefficient of refraction is estimated as additional
unknown. In the case of estimation, there are several choices, e.g., one value
for k for all zenith angles or one value for a group of zenith angles or one
value per day. (It is known that measured zenith angles are “weaker” than
other observations, which can be taken into account by giving them lower
weights.)
214 5 The geometry of the earth
Baselines
From relative GPS measurements, baselines X ij(GPS) = X j(GPS) − X i(GPS) in
the WGS 84 are obtained. The position vectors X i(GPS) and X j(GPS) may be
transformed by a three-dimensional (7-parameter) similarity transformation
to a local system indicated by LS. According to Eq. (5–41), the transforma-
tion formula reads
X LS = x0 + µ R X GPS , (5–85)
where the meaning of the individual quantities is the following:
X LS ... position vector in the local system ,
X GPS ... position vector in the WGS 84 ,
x0 ... shift vector ,
R ... rotation matrix ,
µ ... scale factor .
Forming the difference of two position vectors, i.e., the baseline X ij , the
shift vector x0 is eliminated. Using (5–85), there results
X ij(LS) = µ R X ij(GPS) (5–86)
5.10 Combining terrestrial data and GPS 215
where now the vector δp and the design matrix A ij are given by
δp = [δµ ε1 ε2 ε3 ]T ,
⎡ ⎤
Xij 0 −Zij Yij
(5–88)
⎢ ⎥
A ij = ⎣ Yij Zij 0 −Xij ⎦ .
Zij −Yij Xij 0 (GPS)
Note that the rotations εi refer to the axes of the system used in GPS. If
they should refer to the local system, then the signs of the rotations must be
changed, i.e., the signs of the elements of the last three columns of matrix
A ij must be reversed.
The vector X ij(LS) on the left side of (5–87) contains the points X i(LS)
and X j(LS) in the local system. If these points are unknown, then they are
replaced by known approximate values and unknown increments
where the coefficients of these unknown increments (+1 or −1) together with
matrix A ij form the design matrix.
The vector X ij(GPS) in (5–87) is regarded as measurement quantity. Thus,
finally,
h=H +N (5–91)
2. the gravimetric method, using for this purpose gravity anomalies ∆g.
The theories of both methods were known as early as 1850, but what was
lacking were data, especially gravimetric ones. Serious practical applications
started not much before 1950, a hundred years later, just before the advent
of satellites. This will be discussed in detail later in this book.
A reasonable measuring accuracy was achievable, but another difficulty
appeared. Both methods require the evaluation of integrals of the data (ξ and
η, or ∆g) as continuous functions. The data, however, are always measured
at discrete points only. Interpolation is necessary and introduces additional
errors. If the data are distributed uniformly and densely, resulting errors
may be kept small. The fundamental problem exists, however.
5.12 Reduction of the astronomical measurements to the ellipsoid 217
ξ = Φ − ϕ = ∆ϕ ,
(5–93)
η = (Λ − λ) cos ϕ = ∆λ cos ϕ ,
where we have substituted the respective auxiliary quantities. Thus, the con-
version formulas from natural coordinates Φ, Λ, H to ellipsoidal coordinates
ϕ, λ, h are
ϕ=Φ−ξ,
λ = Λ − η/ cos ϕ , (5–94)
h =H +N.
Now we turn to the reduction of the azimuth. Thus, the question is which
∆α arises from ∆ϕ and ∆λ. The answer is found in Eq. (5–75), where we
218 5 The geometry of the earth
only consider the last two terms on the right-hand side (i.e., we do not take
into account changes of the point coordinates). Omitting all subscripts and
introducing the auxiliary quantities of (5–92), we immediately get
∆α = cot z sin α ∆ϕ + (sin ϕ − cos α cos ϕ cot z) ∆λ (5–95)
or, using ∆ϕ = ξ and ∆λ cos ϕ = η, yields
∆α = ξ sin α cot z + sin ϕ ∆λ − η cos α cot z . (5–96)
This equation may be rearranged to
∆α = sin ϕ ∆λ + (ξ sin α − η cos α) cot z . (5–97)
Alternatively, by using ∆λ = η/ cos ϕ, we get
∆α = η tan ϕ + (ξ sin α − η cos α) cot z . (5–98)
In first-order triangulation, the lines of sight are usually almost horizontal
. .
so that z = 90◦ , cot z = 0. Therefore, the corresponding term can in general
be neglected and we get
∆α = η tan ϕ = ∆λ sin ϕ . (5–99)
This is Laplace’s equation in its usual simplified form. It is remarkable that
the differences ∆α = A − α and ∆λ = Λ − λ should be related in such a
simple way. Laplace’s equation is fundamental for the classical astrogeodetic
computation of triangulations (Sect. 5.14).
For later reference we note that the total deflection of the vertical – that
is, the angle ϑ between the actual plumb line and the ellipsoidal normal – is
given by
ϑ = ξ 2 + η2 (5–100)
and that the deflection component ε in the direction of the azimuth α is
ε = ξ cos α + η sin α . (5–101)
It is clear that ϑ in (5–100) has nothing to do with the two different ϑ
used for spherical and ellipsoidal-harmonic coordinates (polar distances).
Returning to the reduction of astronomical to the corresponding ellip-
soidal quantities, we have (5–94) for the reduction of Φ, Λ, H to ϕ, λ, h and,
finally, the formula
α = A − η tan ϕ (5–102)
reduces the astronomical azimuth A to the ellipsoidal azimuth α.
For the application of these formulas, we need the geoidal undulation N
and the deflection components ξ and η with respect to the reference ellipsoid
used. Two points should be noted:
5.12 Reduction of the astronomical measurements to the ellipsoid 219
1996 –0.2
1995.0 1900.0
1998.5
+0.6 +0.4 +0.2 1997
y ['']
1997.0
1998.0
+0.2
x ['']
Fig. 5.13. Polar motion: mean pole displacement 1900–1997 (solid line),
detailed polar motion 1995–1998 (dotted line)
Now Φ, Λ, A are referred to the mean pole; these values are used in geodesy
because they do not vary with time. Longitude, throughout this book, is
reckoned positive to the east, as is usual in geodesy; it should be mentioned
that in the past literature these formulas are often written for west longi-
tude, according to the former practice of astronomers. Since the correction
terms containing x and y are extremely small (of the order of 0.1 ), we may
use either the ellipsoidal values ϕ and λ or the astronomical values Φ and Λ
in these terms. The term containing ϕGr (the latitude of Greenwich) in the
formula for Λ is usually omitted, so that the mean meridian of Greenwich re-
mains fixed as the conventional zero meridian, rather than the astronomical
5.13 Reduction of horizontal and vertical angles and of distances 221
Vertical angles
The relation between the measured zenith angle z and the corresponding
ellipsoidal zenith angle z may be given as
z = z + ε = z + ξ cos α + η sin α , (5–105)
where α is the azimuth of the target.
Spatial distances
Electronic measurement of distance yields straight spatial distances l be-
tween two points A and B (Fig. 5.14). These distances may either be used
directly for computations in the ellipsoidal coordinate system ϕ, λ, h, as in
“three-dimensional geodesy” (see Sect. 5.9), or they may be reduced to the
surface of the ellipsoid to obtain chord distances l0 or geodesic distances s0 .
We again approximate the ellipsoidal arc A0 B0 by a circular arc of radius
R that is the mean ellipsoidal radius of curvature along A0 B0 . By applying
the law of cosines to the triangle OAB, we find
l2 = (R + h1 )2 + (R + h2 )2 − 2(R + h1 )(R + h2 ) cos ψ . (5–106)
222 5 The geometry of the earth
B
A l
h1 s0 h2
A0 l0 B0
R R
Ã
0
With
ψ
cos ψ = 1 − 2 sin2 , (5–107)
2
this is transformed into
h1 h2 ψ
l2 = (h2 − h1 )2 + 4R2 1 + 1+ sin2 ; (5–108)
R R 2
and with
ψ
l0 = 2R sin (5–109)
2
and the abbreviation ∆h = h2 − h1 , we obtain
h1 h2 2
l2 = ∆h2 + 1 + 1+ l . (5–110)
R R 0
l0
s0 = R ψ = 2R sin−1 . (5–112)
2R
Ellipsoidal refinements of these formulas may be found in Rinner (1956).
As a matter of fact, spatial distances are independent of the vertical.
Therefore, the reduction formula (5–111) does not contain the deflection of
the vertical ε.
dN = −ε ds (5–113)
as given in (2–372) is the basic equation (Fig. 5.15). Integrating this relation,
we get B
NB = NA − ε ds , (5–114)
A
where
ε = ξ cos α + η sin α (5–115)
is the component of the deflection of the vertical along the profile AB, whose
azimuth is α (see Eq. (5–101)).
Formula (5–114) expresses the geoidal undulation as an integral of the
vertical deflections along a profile. Since N is a function of position, this
integral is independent of the form of the line that connects the points A and
B. This line need not necessarily be a geodesic on the ellipsoid, and α may in
the general case be variable. In practice, north-south profiles (ε = ξ) or east-
west profiles (ε = η) are often used. The integral (5–114) is to be evaluated
ds
s ellipsoid
that is, by comparing the astronomical and ellipsoidal (or geodetic) coordi-
nates of the same point, then this method is called the astrogeodetic deter-
mination of the geoid.
The astronomical coordinates are directly observed; the ellipsoidal coor-
dinates are obtained in the following way.
P1
rotation axis
ellipsoid
geoid
N1 at the initial point P1 (in our case, N1 = 0), we can finally compute the
geoidal heights N of any point of the triangulation net by repeated applica-
tion of (5–114). These geoidal heights refer to the ellipsoid that was fixed by
prescribing ξ1 , η1 , N1 , and, of course, its semimajor axis a and its flattening
f . To employ a frequently used term, they refer to the given astrogeodetic
datum (a, f ; ξ1 , η1 , N1 ).
By means of N and the orthometric height H, the height h above the
ellipsoid is obtained via h = H + N , so that the rectangular spatial coordi-
nates X, Y, Z can be computed by (5–27). But unless ξ and η are absolute
(geocentric) deflections, the origin of the coordinate system will not be at
the center of the earth (see Sect. 5.7).
A flaw in the procedure described above apparently is that N, ξ, η are
already needed for the reduction of the measured angles and distances to
the ellipsoid. However, for this purpose approximate values of N, ξ, η are
sufficient. These are obtained by performing the process just explained with
unreduced angles and distances. We can also get suitable values for N, ξ, η
in other ways, for instance, by Stokes’ formula.
net. Longitude and azimuth are often measured at the same point. Then
Laplace’s condition
∆α = ∆λ sin ϕ (5–118)
furnishes an important check on the correct orientation of the net and forces
the axis of the ellipsoid to be parallel with the earth’s axis of rotation. Thus
it may be used for adjustment purposes. Astronomical stations with longi-
tude and azimuth observations are, therefore, called Laplace stations. For
these purposes, the measuring accuracy of astronomical field observations is
sufficient, in contrast to the use for directly determining horizontal positions
by ϕ = Φ − ξ, etc. in Sect. 2.21.
The astrogeodetic determination of the geoid, also called astronomical
leveling, was known to Helmert (1880) and even before.
for the gravimetric method. Both methods use the gravity vector g. It is
compared with a normal gravity vector γ. The components ξ = ∆ϕ and
η = ∆λ cos ϕ of the deflection of the vertical represent the differences in
direction, and the gravity anomaly ∆g represents the difference in magnitude
of the two vectors. Helmert’s formula determines the geoidal undulation N
from ξ and η, that is, by means of the direction of g, and Stokes’ formula
determines N from ∆g, that is, by means of the magnitude of g. Both
formulas are somewhat similar: they are integrals which contain ε, or ξ and
η, and ∆g in linear form.
Otherwise, the two formulas show marked differences which are charac-
teristic for the respective method. In Helmert’s formula, the integration is
extended over part of a profile; thus, it is sufficient to know the deflection
of the vertical in a limited area. The position of the reference ellipsoid with
respect to the earth’s center of gravity is unknown, however, and can be
determined only by means of the gravimetric method or, more practically,
the analysis of satellite orbits (Sect. 7.2). Furthermore, the astrogeodetic
method can be used only on land, because the necessary measurements are
impossible at sea.
5.14 The astrogeodetic determination of the geoid 227
or
εA + εB
∆NAB = − sAB . (5–122)
2
Thus, the undulation difference can be computed for the line AB, and simi-
larly for other lines BC and CA in the triangle ABC (Fig. 5.17). The closure
condition
∆NAB + ∆NBC + ∆NCA = 0 (5–123)
must be satisfied and imposed as a condition in the least-squares adjustment
A sAB
B
of the net. Accordingly, the other triangles can be computed as in any other
height network (e.g., leveling net).
It is curious that it may be shown that such closures are mathematically
equivalent to the well-known relation
∂2N ∂2N
= . (5–124)
∂x ∂y ∂y ∂x
Important remark
The principle of reduction of the plumb line is of fundamental theoretical
importance for understanding the geometry of the earth’s gravity field. In
practice, it is usually disregarded if the topography is sufficiently flat, or
replaced by more sophisticated methods in mountainous areas, as we shall
see later (Sects. 8.12 and 8.13). The present section may be skimmed at first
reading, except for the normal curvature of the plumb line at its very end.
Principles
Consider the projection of the plumb line onto the meridian plane. According
to the well-known definition of the curvature of a plane curve, the angle
5.15 Reduction for the curvature of the plumb line 229
dϕ = −κ1 dh , (5–125)
where the minus sign is conventional and the curvature κ1 is given by (2–50):
1 ∂g
κ1 = . (5–126)
g ∂x
The x-axis is horizontal and points northward. Hence, the total change of
latitude along the plumb line between a point on the ground, P , and its
projection onto the geoid, P0 , is given by
P P
δϕ = dϕ = − κ1 dh (5–127)
P0 P0
or P
1 ∂g
δϕ = − dh . (5–128)
P0 g ∂x
Using κ2 of (2–51), we similarly find for the change of longitude
P
1 ∂g
δλ cos ϕ = − dh , (5–129)
P0 g ∂y
Alternative formulas
There is a close relationship between the curvature reduction of astronomical
coordinates and the orthometric reduction of leveling, considered in Sect. 4.3.
The orthometric correction d(OC) has been defined as the quantity that
must be added to the leveling increment dn in order to convert it into the
orthometric height difference dH:
d(OC) = dH − dn . (5–130)
From Fig. 5.18, we see that, for a north-south profile, the curvature reduction
and the orthometric correction are related by the simple formula
∂(OC)
δϕ = . (5–131)
∂x
Similarly, we find
∂(OC)
δλ cos ϕ = . (5–132)
∂y
230 5 The geometry of the earth
earth's surface
dn
dH local horizon
P ±' d (OC)
parallel to geoid
plumb line
±'
Fgeoid Fground
x (north)
P0 geoid
parallel to
equatorial plane
These relations may be used to find computational formulas for the cur-
vature reductions δϕ and δλ. We have
dC C dC
d(OC) = dH − =d −
g ḡ g
(5–136)
dC C dC C g − ḡ dC
= − 2 dḡ − = − 2 dḡ +
ḡ ḡ g ḡ ḡ g
or
H g − ḡ
d(OC) = − dḡ + dn . (5–137)
ḡ ḡ
By substituting this into (5–131) and (5–132), we obtain
H ∂ḡ g − ḡ
δϕ = − + tan β1 ,
ḡ ∂x ḡ
(5–138)
H ∂ḡ g − ḡ
δλ cos ϕ = − + tan β2 ,
ḡ ∂y ḡ
Integrated form
In formula (5–114), the deflection components ξ and η refer to the geoid.
This means that the astronomical observations of Φ and Λ must be reduced
to the geoid.
It is also possible and often more convenient to apply this correction
for plumb-line curvature not to the astronomical coordinates Φ and Λ but
to the geoidal height differences computed from the unreduced deflection
components.
These N values, denoted by N , are obtained by using in (5–116) the
directly observed Φ and Λ, which define the direction of the plumb line at
the station P (Fig. 5.19). The notation N will be reserved for the correct
geoidal heights. Then we read from Fig. 5.19:
dh = dN + dH = dN + dn , (5–141)
where h is the geometric height above the ellipsoid. Thus, we see that the
difference between the unreduced and the correct element of geoidal height,
dN − dN = dH − dn = d(OC) , (5–142)
earth's surface
dn
plumb ellipsoidal dH
line normal local horizon
W = WP
P dN dN '
h
geoid
normal
geoid W = W0
dN
P0
ds ellipsoid
the leveling increment dn, which is the orthometric reduction d(OC). Thus,
The curvature of the normal plumb line in the east-west direction is zero,
owing to the rotational symmetry of the ellipsoid of revolution. The normal
reduction (5–147) is very simple and practically important, see especially
Sect. 8.13.
In this way, we can get only the best-fitting ellipsoid for the region consid-
ered, rather than a general earth ellipsoid. As Fig. 5.20 indicates, a locally
best-fitting ellipsoid may be quite different from the mean earth ellipsoid,
which can be considered a best-fitting ellipsoid for the whole earth.
If a reasonably good approximation of the earth ellipsoid by a local best-
fitting ellipsoid is desired, it is advisable to subtract the effect of the topog-
raphy and of its isostatic compensation from the astrogeodetic deflections of
the vertical before the minimum condition (5–151) is applied. The purpose
of this procedure is to smooth the irregularities of the geoid. In this way,
5.16 Best-fitting ellipsoids and the mean earth ellipsoid 235
best-fitting ellipsoid
for region AB A
geoid
mean earth ellipsoid
Fig. 5.20. A locally best-fitting ellipsoid and the mean earth ellipsoid
Hayford computed the international ellipsoid as ellipsoid that best fits the
isostatically reduced vertical deflections in the United States. Rapp (1963)
made an interesting recomputation.
Please note: Don’t use formula (5–151) in spite of its historical impor-
tance: the determination of local best-fitting ellipsoids is hopelessly obsolete
now!
The previously described method is impaired by unknown density anoma-
lies and by the lack of complete isostatic compensation. Therefore, it is better
to go still one step further and subtract the gravimetrically computed values
ξ g , η g from the astrogeodetic deflections ξ a , η a . Then the minimum condition
(ξ a − ξ g )2 + (η a − η g )2 = minimum (5–152)
results. Thus, we may say that Hayford’s method is equivalent to the use of
(5–152), the gravimetric values ξ g , η g being approximated by deflections that
represent the effect of topography and of its isostatic compensation only. If
the isostatic compensation were complete, and if we had perfect knowledge
of the density above the geoid, both methods would give exactly the same
result if applied properly.
yield results which, to the usual spherical approximation, are identical with
each other and with the physical definition in terms of M , W0 , C − Ā, and
236 5 The geometry of the earth
Varying the size and shape of the reference ellipsoid and its position with
respect to the earth changes only the coefficients δW , a10 , a11 , b11 , and a20 ,
leaving the other coefficients practically invariant. Thus, the minimum of any
of the integrals (5–157), (5–158), (5–159) is obtained if all these coefficients
are equal to zero. Now, δW = 0 means equal potential U0 = W0 ; a10 = a11 =
b11 = 0 means absolute position (coincident centers of gravity); and a20 = 0
means equality of J2 or of C − (A + B)/2.
Therefore, the equivalence of the physical definition by means of M, W0 ,
C − Ā, ω and of the condition of closest approximation in any of the forms
(5–148), (5–149), or (5–153) has been established. (It may be noted that
(5–158) contains no first-degree term, because of the factor (n − 1)2 , and
that (5–159) contains no term of degree zero, so that these equations do not
determine the missing terms.)
6.1 Introduction
The gravity field outside the earth is particularly important at satellite al-
titude; this will be treated mainly in Chap. 7. The considerations of the
present chapter are applicable to gravitational forces also at satellites (see
Sect. 7.2), but their main practical purpose is to compute test values for the
gravity vector, gravity disturbances, and gravity anomalies at flight eleva-
tions for comparison with airborne gravimetry for reference and calibration
purposes. Airborne gravimetry is much faster than both terrestrial and ship-
borne gravimetry, so it is of interest also for geophysical prospecting.
For computational reasons, it is again convenient to split the gravity
potential W and the gravity vector
g = grad W (6–1)
into a normal potential U and a normal gravity vector
γ = grad U , (6–2)
and the disturbing potential T = W − U and the gravity disturbance vector
δg = grad T = g − γ . (6–3)
The normal gravity field is usually taken to be the gravity field of a suit-
able equipotential ellipsoid. This permits closed formulas and offers other
advantages of mathematical simplicity (see Sect. 2.12).
Thus, U and γ are computed first, and W and g are then obtained by
W =U +T ,
(6–4)
g = γ + δg .
For some purposes, we need the vector of gravitation, grad V (pure at-
traction without centrifugal force), rather than the vector of gravity. The
gravitational vector is computed from the gravity vector by subtracting the
vector of centrifugal force:
⎡ 2 ⎤
ω x
grad V = g − grad Φ = g − ⎣ω 2 y ⎦ , (6–5)
0
240 6 Gravity field outside the earth
where the notations of Sect. 2.1 are used. The rectangular coordinate system
x, y, z will be applied in this chapter in the usual sense: it is geocentric, the
x- and y-axes lying in the equatorial plane with Greenwich longitudes 0◦ and
90◦ East, respectively, and the z-axis being the rotation axis of the earth.
The sign of the components of g, γ, δg, etc., will always be chosen so
that they are positive in the direction of increasing coordinates.
z = u sin β .
Then β is given by
√
u2 + E 2
z
tan β = , (6–9)
u x2 + y 2
and for λ we simply have
y
tan λ = . (6–10)
x
With known ellipsoidal-harmonic coordinates, the normal potential U is
given by (2–126):
GM E q
2
U (u, β) = tan−1 + 1
2 ω 2 a2 sin β − 13 + 12 ω 2 (u2 + E 2 ) cos2 β .
E u q0
(6–11)
6.2 Normal gravity vector 241
The components of γ along the coordinate lines are, by (2–131) and (2–132),
1 ∂U 1 GM ω 2 a2 E q
1
γu = =− + sin β − 6 − ω u cos β ,
2 1 2 2
w ∂u w u2 + E 2 u2 + E 2 q0 2
1 ∂U 1 ω 2 a2 q
γβ = √ =− −√ 2 2 2
+ ω u + E sin β cos β ,
w u2 + E 2 ∂β w u2 + E 2 q0
1 ∂U
γλ = √ = 0.
u2 + E 2 cos β ∂λ
(6–12)
To get the components of γ in the xyz-system, we compute
∂U ∂U ∂x ∂U ∂y ∂U ∂z
= + + , etc. (6–13)
∂u ∂x ∂u ∂y ∂u ∂z ∂u
∂U u ∂U u ∂U ∂U
=√ cos β cos λ +√ cos β sin λ + sin β ,
∂u u2 + E 2 ∂x u2 + E 2 ∂y ∂z
∂U √ ∂U 2 ∂U ∂U
= − u2 + E 2 sin β cos λ − u + E 2 sin β sin λ + u cos β ,
∂β ∂x ∂y ∂z
∂U √ ∂U 2 ∂U
= − u2 + E 2 cos β sin λ + u + E 2 cos β cos λ .
∂λ ∂x ∂y
(6–14)
Introducing the components
∂U 1 ∂U
γx = , ··· ; γu = , ··· , (6–15)
∂x w ∂u
we obtain
u u 1
γu = √ cos β cos λ γx + √ cos β sin λ γy + sin β γz ,
2
w u +E 2 2
w u +E 2 w
1 1 u
γβ = − sin β cos λ γx − sin β sin λ γy + √ cos β γz ,
w w w u + E2
2
γλ = − sin λ γx + cos λ γy .
(6–16)
These are the formulas of an orthogonal rectangular coordinate transforma-
tion. The inverse transformation is obtained by interchanging the rows and
242 6 Gravity field outside the earth
u 1
γx = √ cos β cos λ γu − sin β cos λ γβ − sin λ γλ ,
w u2 + E2 w
u 1
γy = √ cos β sin λ γu − sin β sin λ γβ + cos λ γλ , (6–17)
w u2 + E 2 w
1 u
γz = sin β γu + √ cos β γβ .
w w u2 + E 2
r
z
# ' y
¸ r co
x s'
x = r cos ϕ̄ cos λ ,
y = r cos ϕ̄ sin λ , (6–18)
z = r sin ϕ̄
or inversely by
r = x2 + y 2 + z 2 ,
z
ϕ̄ = tan−1 , (6–19)
x2 + y2
y
λ = tan−1 .
x
Now it is convenient to start with the components δgr , δgϕ̄ , δgλ of the
gravity disturbance vector δg, Eq. (6–3), in the spherical coordinates r, ϕ̄, λ.
In analogy to (2–377), we have
∂T 1 ∂T 1 ∂T
δgr = , δgϕ̄ = , δgλ = . (6–20)
∂r r ∂ ϕ̄ r cos ϕ̄ ∂λ
Since we are dealing with the relatively small quantities of the disturbing
field, a spherical approximation may be sufficient (Sect. 2.13), as it was in
the case of Stokes’ formula.
The disturbing potential T may be expressed in terms of the free-air
anomalies at the earth’s surface by the formula of Pizzetti, Eqs. (2–302) and
(2–303),
R
TP = T (r, ϕ̄, λ) = ∆g S(r, ψ) dσ , (6–21)
4π
σ
and
l= r 2 + R2 − 2R r cos ψ . (6–23)
Now we form the derivatives of the extended Stokes function (6–22) with
respect to r and ψ. By differentiating (6–23), we get
∂l r − R cos ψ ∂l Rr
= , = sin ψ . (6–31)
∂r l ∂ψ l
By means of these auxiliary relations, we find
∂S R (r 2 − R2 ) 4R R 6R l
=− 3
− − 2+ 3
∂r rl rl r r
R2 r − R cos ψ + l
+ 3 cos ψ 13 + 6 ln ,
r 2r
(6–32)
∂S 2R2 r 6R2 8R2
= sin ψ − 3 − + 2
∂ψ l rl r
3R2 r − R cos ψ − l r − R cos ψ + l
+ 2 + ln .
r l sin2 ψ 2r
Somewhat more convenient expressions are obtained by substituting
R
t= , (6–33)
r
l
D= = 1 − 2t cos ψ + t2 . (6–34)
r
Then the extended Stokes function (6–22) and its derivatives (6–32) become
2 1 − t cos ψ + D
S(r, ψ) = t + 1 − 3D − t cos ψ 5 + 3 ln , (6–35)
D 2
∂S(r, ψ) t2 1 − t2 4
=− + + 1 − 6D
∂r R D3 D
1 − t cos ψ + D
− t cos ψ 13 + 6 ln ,
2
(6–36)
∂S(r, ψ) 2 6
= −t sin ψ
2 + −8
∂ψ D3 D
1 − t cos ψ − D 1 − t cos ψ + D
−3 − 3 ln .
D sin2 ψ 2
These expressions are used in (6–21) and (6–30) to compute T and δg.
The separation NP of the geopotential surface through P, W = WP , and
the corresponding spheropotential surface U = WP is according to Bruns’
theorem given by
TP
NP = ; (6–37)
γQ
246 6 Gravity field outside the earth
1 ∂NP 1 ∂NP
ξP = − , ηP = − ; (6–38)
r ∂ ϕ̄ r cos ϕ̄ ∂λ
these equations correspond to (2–377). Since γ varies very little with latitude
and is independent of longitude, we have
∂NP ∂ TP 1 ∂TP TP ∂γQ . 1 ∂TP
= = − 2 = (6–39)
∂ ϕ̄ ∂ ϕ̄ γQ γQ ∂ ϕ̄ γQ ∂ ϕ̄ γQ ∂ ϕ̄
and
∂NP 1 ∂TP
= . (6–40)
∂λ γQ ∂λ
Substituting the results of (6–39) and (6–40) into (6–38) and comparing then
with (6–20) shows that
1 1
ξP = − δgϕ̄ , ηP = − δgλ . (6–41)
γQ γQ
We see that NP , ξP , ηP are given by Eqs. (6–21) and (6–30), apart from
the factor ±1/γQ . Hence, these equations are the extensions of Stokes’ and
Vening Meinesz’ formulas for points outside the earth and reduce to these
formulas for r = R, t = 1.
Writing Eqs. (6–41) in the form
we see that the horizontal components of δg are directly related to the de-
flection of the vertical, which is the difference in direction of the vectors
g and γ. The radial component δgr , however, represents the difference in
magnitude of these vectors, since as a spherical approximation
−δgr = δg = gP − γP , (6–43)
H l
s dx dy
xy
F R2d¾
R R
∞
H δgϕ̄
δgϕ̄ = dx dy ,
2π l3
−∞
(6–54)
∞
H δgλ
δgλ = dx dy .
2π l3
−∞
δgϕ̄ = −γ0 ξ ,
(6–56)
δgλ = −γ0 η ,
which follow from (2–264) together with (6–42) and (6–43) applied to sea
level. The symbols R and γ0 denote, as usual, a mean earth radius and a
mean value of gravity on the earth’s surface.
Hence, we may compute T and δg by means of the upward continuation
integral if the geoidal undulations N and the deflection components ξ and η
at the earth’s surface are given.
The plane approximation is sufficient except for very high altitudes (e.g.,
> 250 km). Otherwise, we must use the spherical formula (6–44) for T . For
the radial component δgr , formula (6–44) may also be applied with T re-
placed by r δg, since r δg and r ∆g are harmonic as we know from Sect. 2.14.
The corresponding spherical formulas for the upward continuation of the hor-
izontal components δgϕ̄ and δgλ are not known. The reason why the same
formula, the upward continuation integral, applies for T and the components
of δg in the planar case only is that the derivatives of T are harmonic only
when referred to a Cartesian coordinate system.
Data
For all computations dealing with the external gravity field of the earth,
free-air gravity anomalies must be used for ∆g, since all other types of grav-
ity anomalies correspond to some removal or transport of masses whereby
the external field is changed. If, in addition to ∆g, deflections of the vertical
ξ, η (in the upward continuation) are used, then these quantities should be
computed from free-air anomalies. If, as usually done, the normal free-air
.
gradient ∂y/∂h = 0.3086 mgal/m is used for the free-air reduction, then the
free-air anomalies refer, strictly speaking, to the earth’s physical surface (to
ground level) rather than to the geoid (to sea level). The N values com-
puted from them by Stokes’ formula are height anomalies ζ, referring to the
ground, rather than heights of the actual geoid. However, this distinction is
insignificant and can be ignored in most cases, so that we may consider ∆g
as sea-level anomalies (see Sect. 8.6).
If we cannot neglect this distinction, aiming at highest accuracy in high
and steep mountains for low altitudes H, then we may proceed as follows. We
reduce the free-air anomaly ∆g from the ground point A to the corresponding
point A0 at sea level (Fig. 6.3):
∂∆g
∆gharmonic = ∆g − h, (6–58)
∂h
and use the sea level anomaly ∆gharmonic so obtained. The vertical gradient
∂∆g/∂h may be computed by applying formula (2–394) using the ground-
6.5 Additional considerations 251
H1
A
H
h1 level of F
F h W = W1
sea level
P0 A0 W = W0
δgλ being rigorously the same in both systems. Thus, δgr , δgϕ̄ , δgλ may also
be considered as the components of δg in ellipsoidal-harmonic coordinates.
Then we have
can only be used for δg so that γu and γβ must be computed by the rigorous
formulas (6–12).
The gravity potential W may be computed by the first equation of (6–
4); the gravitational potential V is obtained by subtracting the centrifugal
potential ω 2 (x2 + y 2 )/2; and the vector of gravitation is given by (6–5).
P W = WP
NP
U = WP
Q
H1
F earth's
surface
above ground is measured. This case seems rather to belong to the past. If
the case should arise, gravity anomalies ∆g can be upward continued just as
δg as described in Sect. 6.4.
Again, free-air anomalies referred to ground level or, more accurately,
to some level surface, are to be used. If the ground is elevated above sea
level but reasonably flat, it is somewhat better to regard H as elevation
above ground rather than above sea level, because the ground may then be
considered locally part of a level surface.
The inverse problem, the downward continuation of gravity anomalies or
rather gravity disturbances, occurs in the reduction of gravity measured on
board an aircraft. There is, of course, a relation to harmonic downward con-
tinuation in the solution of Molodensky’s problem as described in Sect. 8.6.
Upward and downward continuation are also tools of geophysical explo-
ration, but here the objective is quite different. Several methods have been
developed in this connection, some of which are also applicable for geodetic
purposes; see, e.g., Dobrin and Savit (1988) or Telfort et al. (1990).
Upward and downward continuation are related as direct and inverse
problems in the theory of inverse problems, see Anger et al. (1993) and also
www.inas.tugraz.at under forschung/InverseProblems/AngerMoritz.html,
where additional references can be found.
7 Space methods
7.1 Introduction
The subject of this chapter is the use of satellite observations for determining
features of the gravity field and the figure of the earth. Only the barest
essentials can be presented within the scope of a chapter. The reader will
find more information in special textbooks such as Hofmann-Wellenhof et
al. (2001), Montenbruck and Gill (2001), and Seeber (2003).
Historical remarks
Immediately after the first launch of artifical satellites (Sputnik 1957, Ex-
plorer 1958), their use for geodetic purposes was initiated, and by now the
Global Positioning System (GPS) has become the most important method
for a fast and precise determination of geodetic positions (see Sect. 5.3).
Historically, the first observational methods were intended to determine the
spatial direction and the distance to the satellite. Most of these methods are
now obsolete, but some principles may be still useful.
Directions
They may be measured by photographing the satellite against the back-
ground of stars, or by means of radio waves transmitted from the satellite,
using the principle of interference. Photography can only achieve an accuracy
of about 0.2 arc seconds and is not used any more in its original sense. The
principle of the photographic method was as follows. On the photographic
plate, the image of the satellite is surrounded by images of stars. The direc-
tions to the surrounding stars are defined by their right ascensions α and
declinations δ, which are known from astronomy. Therefore, by interpolation
we find the right ascension and declination of the satellite representing the
desired direction. This technique is now obsolete.
Ranges
They are measured by radar or by laser. Radar is used for measuring ranges
to space probes orbiting in the solar system, which is important to space
sciences rather than to geodesy. Lunar Laser Ranging (LLR) and Satellite
Laser Ranging (SLR) are useful for determining the earth rotation param-
eters because of their high (subcentimeter) accuracy; however, their use is
restricted to a limited number of fundamental stations.
256 7 Space methods
Range rates
Satellite altimetry
Here a short-wave electronic ray is sent, from a satellite flying over the
oceans, vertically down to the ocean surface, reflected there and received by
the satellite again. The measured travel time immediately gives the height
H of the satellite above the ocean surface. Knowing the orbital position of
the satellite with respect to the global reference system, we can compute the
satellite height h above the ellipsoid. Then the difference h− H is the geoidal
height N . This is the case if the sea surface is assumed to coincide with the
geoid. In reality, because of ocean currents, etc., both surfaces are separated
by the “sea surface topography”, which may reach the order of 1 m and is
interesting to oceanography. It can be determined if an accurate ocean geoid
is known from the gravitational field.
The principles of these methods are illustrated in Fig. 7.1, where e indi-
cates the direction observation, s between tracking station and satellite refers
to the range measurement, and, accordingly, ds/dt corresponds to Doppler
observation, whereas ds/dt between the two satellites is obtained by SST;
finally, H is measured by satellite altimetry.
satellite
satellite
s
H
tracking
station e
geoid
land W = W0
ocean
Z0
K'
line of nodes S
À
P
A ! t
bi
or i or
W equa
t
vernal equinox
K
X0
a semimajor axis,
e eccentricity,
i inclination,
(7–3)
Ω right ascension of the node,
ω argument of perigee,
T time of perigee passage .
where r is the distance of the satellite from the earth’s center of mass and
b2
p= = a (1 − e2 ) (7–6)
a
is the length of the radius vector r for v = 90◦ . The radius vector r and
the true anomaly v form a pair of polar coordinates in the orbital plane,
and (7–5) is the well-known polar equation of an ellipse. See Fig. 7.3 for an
illustration of these quantities, where F , the focal point, is the earth’s center
of mass.
According to Kepler’s second law, the area of the elliptical sector swept by
the radius vector r between any two positions of the satellite is proportional
to the time it takes the satellite to pass from one position to the other. In
other words, the time rate of change of the area swept by the radius vector
is constant. Since the element of area of a sector in polar coordinates r and
e2
S
b p
r
A ae À P
a F r0 e1
G M ae n+1
∞
R=− Jn Pn (cos ϑ) (7–12)
ae n=2 r
is a function of r and ϑ only. Note that the main difference between the
perturbing potential R of celestial mechanics and the disturbing potential
T of physical geodesy is that R, but not T , also incorporates the effect
of the flattening through J2 . There are also other perturbing forces acting
on a satellite, such as the resistance of the atmosphere (atmospheric drag),
radiation pressure exerted by the sunlight, etc. These nongravitational per-
turbances must be taken into account separately and will not be considered
here.
Note that the equatorial radius of the earth (the semimajor axis of the
terrestrial ellipsoid) has been denoted by ae , in order to distinguish it from
a, which now denotes the semimajor axis of the orbital ellipse. This notation
will be used in what follows.
Since S is the component of the perturbing force along the radius vector,
we have
∂R
S= . (7–13)
∂r
The components of the perturbing force along the meridian and the prime
vertical are
1 ∂R 1 ∂R
− and . (7–14)
r ∂ϑ r sin ϑ ∂λ
262 7 Space methods
pole
–1 @R
r @#
W
® T
satellite 1 @R
r sin # @¸
! +À ®
90° – #
i
node
t0 +P t0 +P t0 +P
∆a = ȧ dt , ∆e = ė dt , ∆i = ı̇ dt , etc. (7–18)
t0 t0 t0
which follows from the rectangular spherical triangle in Fig. 7.4. The radius
vector r is also a function of v according to (7–5). Finally, Kepler’s second
law (7–7) furnishes the relation between v and the time t:
dt r2
= . (7–20)
dv G M a (1 − e2 )
where
da da dt r2
= = ȧ . (7–22)
dv dt dv G M a (1 − e2 )
difficult, we find
∆a = 0 ,
1 − e2
∆e = − tan i ∆i ,
e
3
ae 5
∆i = 3π e 1 − sin2 i cos i cos ω J3
p 4
4
45 ae 7
+ πe 1 − sin i sin 2i sin 2ω e J4 · · · ,
2
16 p 6
2
ae
∆Ω = −3π cos i J2
p
3
ae 15
+ 3π 1− 2
sin i cot i sin ω e J3
p 4 (7–23)
4
15 ae 7
+ π 1 − sin2 i cos i J4 · · · ,
2 p 4
2
ae 5
∆ω = 6π 1 − sin2 i J2
p 4
3
ae 5
+ 3π 1 − sin i sin i sin ω e J3
2
p 4
4
ae 31 49
− 15π 1− 2
sin i + 4
sin i
p 8 19
3 7
+ − sin2 i sin2 i cos 2ω J4 · · · .
8 16
Terms of the order of e2 J3 and e2 J4 , which are very small, have been ne-
glected in these equations. The proportionality of ∆e and ∆i is more or less
accidental: it applies only with respect to long-periodic disturbances; ė and
di/dt themselves are not proportional. The quantity p is defined by (7–6);
it is hardly necessary to repeat that a, p, e, etc., refer to the orbital ellipse
and not to the terrestrial ellipsoid, of which ae is the equatorial radius.
By integrating over one revolution, we have removed the short-periodic
terms of periods P, 2P, 3P, . . . , such as cos v, cos 2v, etc. What remains are
secular terms, which are constant for one revolution and increase steadily
with the number of revolutions, and the long-periodic terms, which change
very slowly with time in a periodic manner. The argument of perigee ω
increases slowly but steadily, so that the perigee of a satellite orbit also
7.3 Determination of zonal harmonics 265
rotates around the earth, but much slower than the satellite itself; a typical
period of ω is two months. Therefore, terms containing cos ω, sin ω, or sin 2ω
are called long-periodic.
The first equation of (7–23) shows that the semimajor axis of the orbit
does not change secularly or long-periodically. The eccentricity and the in-
clination undergo long-period, but not secular, variations, whereas Ω and ω
change both secularly and long-periodically.
Equations (7–23) are linear in J2 , J3 , J4 , . . . . For practical applications,
nonlinear terms containing J22 , J2 J3 , J2 J4 , etc., must also be taken into ac-
count, since J22 is of the order of J4 . The derivation of these nonlinear terms
is much more difficult, and their expressions are different in the various or-
bital theories that have been proposed. For these reasons, such expressions
will not be given here.
Equations (7–23), supplemented by certain nonlinear terms, can be used
to determine coefficients J2 , J3 , J4 , etc. Since the secular or long-periodic
variations ∆Ω, ∆ω, ∆e, ∆i are known from observation for a sufficient num-
ber of satellites, we obtain equations of the form
Numerical values
Helmert (1884: p. 472) used the regression of the node of the moon’s orbit
to determine J2 , which is the only coefficient to have an appreciable effect
. .
on it. Note that for e = 0 and p = a
ae , the equation for ∆Ω in (7–23)
becomes a 2
e
∆Ω = −3π J2 cos i . (7–27)
a
Helmert found
J2 = 1086.5 · 10−6 (7–28)
by averaging two widely different values. This corresponds to a flattening of
This value is quite close to the recent results but has a much larger uncer-
tainty.
Reliable values by this method can only be obtained from close artificial
satellites. Currently accepted values are, for example,
J2 = 1082.6359 · 10−6 ,
J3 = −2.5324 · 10−6 , (7–30)
J4 = −1.6198 · 10−6 ,
whose standard errors are assumed to be better than ±0.01 · 10−6 . The value
for J2 has been taken from the report of the IAG by Groten (2004), accessible
from www.gfy.ku.dk/∼iag/HB2004/part5/51-groten.pdf. J3 and J4 are from
the recent mission GRACE (see Sect. 7.5).
7.4 Rectangular coordinates of the satellite and perturbations 267
is the representation of the satellite in this system. This result may be trans-
formed into the equatorial system X 0 Y 0 Z 0 by a rotation matrix R and re-
sults in a vector denoted as X0 = [X 0 , Y 0 , Z 0 ]. The transformation is
268 7 Space methods
obtained by ⎡ ⎤ ⎡ ⎤
X0 cos v
⎢ 0 ⎥ ⎢ ⎥
⎢ Y ⎥ = R r ⎢ sin v ⎥ , (7–32)
⎣ ⎦ ⎣ ⎦
Z 0 0
where the matrix R is composed of three successive rotation matrices (see
Figs. 7.2 and 7.3) and is given by
see Hofmann-Wellenhof et al. (2001: p. 43). The column vectors of the or-
thonormal matrix R are the axes of the orbital coordinate system represented
in the equatorial system X0i .
Substituting (7–33) into (7–32) and carrying out the multiplication (Mon-
tenbruck and Gill 2001: Eq. (2.51)) yields
Z 0 = r sin(ω + v) sin i ,
a (1 − e2 )
r= . (7–35)
1 + e cos v
This expresses the rectangular coordinates of the satellite in terms of the
elements of its osculating orbit, the true anomaly v fixing its position as a
function of time.
Since the osculating ellipse does not remain constant, it is convenient
to use a fixed reference orbit – for instance, the osculating ellipse E0 at
a certain instant t0 , having the elements a0 , e0 , i0 , Ω0 , ω0 , T0 . At a later
instant t, the orbital elements will have changed to a0 + ∆t a, e0 + ∆t e, i0 +
∆t i, Ω0 + ∆t Ω, ω0 + ∆t ω, T0 + ∆t T , which corresponds to an osculating
ellipse Et .
7.4 Rectangular coordinates of the satellite and perturbations 269
where again Lnm , L̄nm , Mnm , etc., are functions of the time t.
These perturbations are added to the coordinates computed from (7–34)
using the orbital elements of the reference ellipse E0 . In this way, we obtain
the rectangular coordinates of the satellite in the form
X 0 = X 0 (t; a0 , e0 , i0 , Ω0 , ω0 , T0 ; Cnm , Snm ) ,
Y 0 = Y 0 (t; a0 , e0 , i0 , Ω0 , ω0 , T0 ; Cnm , Snm ) , (7–41)
Z 0 = Z 0 (t; a0 , e0 , i0 , Ω0 , ω0 , T0 ; Cnm , Snm )
as explicit functions of the time t, containing as constant parameters the
orbital element of the reference ellipse E0 and the gravitational coefficients
Cnm and Snm . This is the advantage of (7–41) over the system (7–34), which
formally is much simpler but depends on the variable orbital parameters of
the osculating ellipse.
The actual expressions for (7–41) are very complicated. Therefore, we
have been satisfied with outlining the procedure, referring the reader for
details to the pioneering book by Kaula (1966 a) and to his papers given
there.
7.5 Determination of tesseral harmonics and station positions 271
S S
Z0
P ± ZS0 –ZP0
P ± ±
® ||Y 0 Y0
0 XS0 –XP0 ®
||X 0
Y –Y
S P
0
X0 ®
0
Y
X0
earth
ZP0 = ZP .
The angle θ0 is called Greenwich sidereal time; its value is
θ0 = ω t , (7–45)
where ω is the angular velocity of the earth’s rotation. It is proportional
to the time t and, in appropriate units, measures it. Thus, absolute Green-
wich time is needed to convert the terrestrial coordinates XP , YP , ZP to the
celestial coordinates XP0 , YP0 , ZP0 that are required in (7–42) and (7–43).
As a final step, we substitute the station coordinates, as given by (7–44),
and the satellite coordinates, as symbolized by (7–41), into (7–43), obtaining
expressions of the form
α = α(XP , YP , ZP ; t; a0 , e0 , i0 , Ω0 , ω0 , T0 ; Cnm , Snm ) ,
δ = δ(XP , YP , ZP ; t; a0 , e0 , i0 , Ω0 , ω0 , T0 ; Cnm , Snm ) , (7–46)
s = s(XP , YP , ZP ; t; a0 , e0 , i0 , Ω0 , ω0 , T0 ; Cnm , Snm ) .
7.5 Determination of tesseral harmonics and station positions 273
Z 0= Z
Y
Z 0= Z X
Y0
Y 0 X0
P
Y q0 Y
q0
Y0 X
X0 X0
X
Besides depending on the station coordinates and the time, they also contain
the orbital and gravitational parameters.
Every observation furnishes an equation of type (7–46). Provided we have
a sufficient number of such observation equations, we can solve them for the
station coordinates XP , YP , ZP , for the orbital parameters a0 , e0 , etc., of the
reference ellipse, and for a certain number of gravitational parameters Cnm
and Snm . This is the principle of the orbital method. In practice, differential
formulas will be applied to determine corrections to assume approximate val-
ues by means of a least-squares adjustment. Therefore, the actual analytical
developments are from the outset directed toward obtaining differential for-
mulas corresponding to (7–46). The substitutions indicated above are, thus,
consistently performed in terms of the corresponding differential expressions.
In this way we are able to operate with linear equations and to employ that
efficient tool of linear analysis, matrix calculus. Simple though the princi-
ple of this procedure is, the details when written out are nevertheless so
complicated that the reader must again be referred to the literature, e.g.,
Kaula (1966 a), Montenbruck and Gill (2001). Computer formula manipula-
tion is also used.
Besides these analytical problems, which have been satisfactorily solved,
the geodetic application of (7–46) raises difficulties similar in principle to
those involved in the determination of zonal harmonics by means of (7–24),
but even more serious in practice. Strictly speaking, an infinite number of
unknowns, Cnm , Snm , etc., are to be determined from a finite number of
observations. In order to get a definite solution, it must be assumed that the
effect of higher-degree terms is negligibly small. But even then there are very
many unknowns: coordinates of the observing stations, parameters of the
274 7 Space methods
Present results
At present (2005), several determinations of tesseral harmonics up to the
degree 360 are available from a combination of satellite and terrestrial data.
Soon the degree 1800 will be achieved. These coefficients represent the large-
scale features of the disturbing potential T and, hence, of the geoid, since
the geoidal height is given by N = T /γ. There is a general agreement be-
tween the essential aspects of these determinations as expressed in geoidal
maps, although the details of these maps, and even more so the individual
coefficients, are rather different.
As an example we take the first nonzonal coefficients, C22 and S22 , which,
according to Sect. 2.6, Eq. (2–95), express the inequality of the earth’s prin-
cipal equatorial moments of inertia or, somewhat loosely speaking, its triax-
iality. According to Groten (2004), we have C22 = (1574.5 ± 0.7) · 10−9 and
S22 = (−903.9 ± 0.7) · 10−9 .
Concerning the order of magnitude, J2 is on the order of magnitude of
10 , where all the other coefficients are of order 10−6 . This is why the earth
−3
Geodesy
As mentioned in Sect. 5.3, GPS has revolutionized geodesy in many respects.
Despite the tremendous importance of GPS, in Sect. 5.4 it was shown that
the user of GPS gets only geometric quantities: WGS 84 coordinates, i.e.,
geocentric rectangular coordinates X, Y, Z or, computed from them, ellip-
soidal coordinates ϕ, λ, h (see Sect. 5.6.1). Therefore, the height obtained by
GPS, i.e., the ellipsoidal height h, is purely geometric. To transform these
heights into orthometric heights H by H = h − N , the geoidal undulation N
is required. Using satellites to determine the earth’s gravity field, a globally
uniform height system will result.
Additionally, an accurate knowledge of the earth’s gravity field improves
the orbit determination of satellites.
Oceanography
The sea surface topography (SST), i.e., the difference between the geoid and
the mean sea surface, can be determined when combining satellite altimetry
data and the earth’s gravity field data. From Fig. 7.7 we obtain the relation
h = N + SST + ∆H + a , (7–47)
276 7 Space methods
satellite orbit
a
h
DH
mean sea surface
SST instantaneous
sea surface
geoid
N
ellipsoid
Geophysics
As mentioned earlier, the earth’s gravity field reflects the mass inhomo-
geneities in the interior of the earth. Knowing gravity values on the earth’s
surface and, in addition, complementary data (e.g., magnetic and seismic
data), improved models for the structure and processes in the earth’s inte-
rior may be obtained. These processes may cause the movement of tectonic
plates which are responsible for earthquakes. Thus, we see that the gravity
field is the fundamental link in a chain of interactive processes. Using more
descriptive terms, an improved knowledge on the gravity field may yield
more accurate methods to predict earthquakes. This justifies any effort on
the determination of the earth’s gravity field.
7.6 New satellite gravity missions 277
Before giving some details on the objectives and payloads of the missions,
the different concepts are briefly described.
GPS satellites
LEO satellite
3D accelerometer
earth's surface
The gravity field is derived by inverting (in the sense of inverse problems,
cf. the remark on inverse problems at the end of Sect. 1.13) the information
obtained from the satellite orbit
GPS satellites
LEO satellites
3D accelerometers
earth's surface
GPS satellites
LEO satellite
gradiometer
earth's surface
where the gravitational signal arises from the attracting masses of the earth.
Thus, the measured signal corresponds to the gradients of the component
of the gravity acceleration, i.e., the second derivatives of the gravitational
potential. For instance, in obvious notation we read from Fig. 7.11
Vx2 − Vx1 ∆Vx . ∂Vx
= = = Vxz . (7–48)
∆z ∆z ∂z
Vx1
Dz
x
Vx2
The CHAMP mission was launched on July 15, 2000 from the Russian Ple-
setsk cosmodrome. The main mission parameters of the respective satellite
are the following:
of the CHAMP mission, but refer the reader to the previously mentioned
homepage.
As explained before, the measuring principle for CHAMP is satellite-to-
satellite tracking in high-low mode. The gravity field of the earth perturbes
the CHAMP satellite orbit. These perturbing accelerations correspond to
first derivatives of the gravitational potential V . This implies that the gravity
field of the earth may be derived from observed gravitational satellite orbit
perturbations applying numerical orbit integration (Montenbruck and Gill
2001) or using the energy balance principle (Ilk 1999, Jekeli 1999, Sneeuw
et al. 2002).
For further reading see Reigber et al. (2003), Seeber (2003: Sect. 10.2.2).
As with CHAMP, also the altitude of the GRACE satellites will decrease
in the course of their lifetime primarily because of atmospheric drag. The
7.6 New satellite gravity missions 283
amount of this decrease depends on the solar activity cycle and may accumu-
late in the mission lifetime to some 50 km on low activity, and up to 200 km
on high activity, see http://op.gfz-potsdam.de/grace.
The range between the two satellites must be determined extremely ac-
curately. Its range rate must be known to better than 1 µm s−1 , which is
achieved by intersatellite microwave measurements. The basic idea is that
variations in the gravity field cause variations in the range between the
two satellites; areas of stronger gravity will affect the lead satellite first and,
therefore, accelerate it away from the following satellite (Seeber 2003: p. 479).
GRACE will not only provide a static global gravity field but also its
temporal variations.
To achieve the mission goals, the following payload is on board of the
two satellites:
• The K-band ranging system is the key instrument of GRACE to mea-
sure the range changes between both satellites using dual-band mi-
crowave signals (i.e., two one-way ranges) with a precision of about
1 µm s−1 . The ranges are obtained at a sampling rate of 10 Hz.
• The GPS receiver serves for the precise orbit determination of the
GRACE spacecraft and provides data for atmospheric and ionospheric
profiling. To achieve this, satellite-to-satellite tracking between the
GRACE satellites and the GPS satellites is realized. A navigation solu-
tion comprising position, velocity, and a time mark is derived on board.
The navigation solution is required for the attitude control system. The
precise orbit based on code and carrier pseudoranges is determined on
ground.
• The attitude and orbit control system comprises a cold gas propulsion
system, three magnetic torque rods, star trackers, a three-axis inertial
reference unit to measure angular rates, and a three-axis magnetome-
ter.
• The accelerometer measures all nongravitational accelerations on the
GRACE spacecraft, e.g., due to air drag or solar radiation pressure.
• The laser retroreflector is a passive payload instrument used to reflect
short laser pulses transmitted by ground stations. The distance be-
tween a ground station and a GRACE satellite can be measured with
an accuracy of 1–2 cm. The laser retroreflector data are primarily used
together with the GPS receiver data for the precise orbit determina-
tion.
In 2004, the GRACE science team released to the public a first version
of a new earth gravity field model complete to degree and order 150. The
resulting improved geoid together with satellite altimetry will advance the
284 7 Space methods
Referring to the results, the main output of this mission will be the following:
• spherical-harmonic coefficients for the gravitational potential, see, e.g.,
(2–80),
• corresponding variance-covariance matrix.
Derived products from this main output are geoidal heights, gravity anoma-
lies, and also oceanographic data.
It is important to mention that the GPS orbit analysis of GOCE will
rather yield long-wavelength information of the gravity field, while the satel-
lite gravity gradiometry will yield the short-wavelength information.
GOCE is the first “drag-free” mission, which implies that the satellite
moves in free fall around the earth. Therefore, a drag compensation and
attitude control system is required to compensate for drag forces and torques.
This and more information may be found in Rebhahn et al. (2000),
Drinkwater et al. (2003), Pail (2003), www.esa.int/livingplanet/goce.
Measurements
The basic principle of gradiometry in GOCE is the measurement of accel-
eration differences for a very short baseline. Considering two accelerometers
separated by 50 cm on one axis, Müller (2001) and Pail (2003) write the two
observation equations as
a1 = M + Ω̇ + ΩΩ ∆x + f ng ,
(7–49)
a2 = − M + Ω̇ + ΩΩ ∆x + f ng ,
comprises the components of the angular velocity and is used to describe the
orientation of the gradiometer. Since Ω is skewsymmetric, the tensor ΩΩ
is symmetric. Finally, ∆x in (7–49) is the vector from the intersection of
the three coordinate axes to the respective accelerometer (where the same
length is assumed), and f ng comprises all nongravitational effects (air drag,
solar radiation pressure, etc.).
Now we once add (“common mode”) and once subtract (“differential
mode”) the two accelerations in (7–49) and obtain
(a1 + a2 )/2 = f ng ,
(7–52)
(a1 − a2 )/2 = M + Ω̇ + ΩΩ ∆x ,
Γ = M + Ω̇ + ΩΩ (7–53)
(Γ − ΓT )/2 = Ω̇ ,
(7–54)
(Γ + ΓT )/2 = M + ΩΩ ,
ΓT = M − Ω̇ + ΩΩ . (7–55)
Γ − ΓT = 2Ω̇ (7–56)
and, finally,
(Γ − ΓT )/2 = Ω̇ , (7–57)
which completes our proof for the first relation of (7–54). To prove the second
relation of (7–54), we add Eqs. (7–53) and (7–55):
Γ + ΓT = 2M + 2ΩΩ (7–58)
288 7 Space methods
or
(Γ + ΓT )/2 = M + ΩΩ , (7–59)
which concludes our proof.
Since we have determined Ω̇ in (7–57), we can get Ω by integration:
t
Ω(t) = Ω(t0 ) + Ω̇ dt , (7–60)
t0
where the initial orientation Ω(t0 ) is obtained from the star trackers. Squar-
ing the result for Ω(t) yields ΩΩ, which is needed in (7–59) so that we
find
M = (Γ + ΓT )/2 − ΩΩ (7–61)
as final result for the desired Marussi tensor M. Many more details may be
found in Rummel (1986).
8 Modern views on the
determination of the figure of
the earth
8.1 Introduction
In the preceding chapters we have usually followed what might be called the
conservative approach to the problems of physical geodesy using classical
observations. The geodetic measurements – astronomical coordinates and
azimuths, horizontal angles, gravity observations, etc. – are reduced to the
geoid, and the “geodetic boundary-value problem” is solved for the geoid by
means of Stokes’ integral and similar formulas. The geoid then serves as a
basis for establishing the position of points of the earth’s surface.
The advantage of this approach is that the geoid is a level surface, capable
of a simple definition in terms of the physically meaningful and geodetically
important potential W . The geoid represents the most obvious mathematical
formulation of a horizontal surface at sea level. This is why the use of the
geoid simplifies geodetic problems and makes them accessible to geometrical
intuition.
The disadvantage is that the potential W inside the earth, and hence
the geoid W = constant, depends on the density because of Poisson’s
Eq. (2–9),
∆W = −4π G + 2ω 2 . (8–1)
Therefore, in order to determine or to use the geoid, the density of the
masses at every point between the geoid and the ground must be known, at
least theoretically. This is clearly impossible, and therefore some assumptions
concerning the density must be made, which is unsatisfactory theoretically,
even though the practical influence of these assumptions is usually rather
small.
For this reason it is of basic importance that M.S. Molodensky in 1945 was
able to show that the physical surface of the earth can be determined from
geodetic measurements alone, without using the density of the earth’s crust.
This requires that the concept of the geoid be abandoned. The mathematical
formulation becomes more abstract and more difficult. Both the gravimetric
method and the astrogeodetic method can be modified for this purpose. The
gravity anomalies and the deflections of the vertical now refer to the ground,
290 8 Determination of the figure of the earth
and no longer to sea level; the “height anomalies” at ground level take the
place of the geoidal undulations.
These developments have considerably broadened our insight into the
principles of physical geodesy and have also introduced powerful new meth-
ods for tackling classical problems. Hence their basic theoretical significance
is by no means lessened by the fact that many scientists prefer to retain the
geoid because of its conceptual and practical advantages.
In this chapter, we first give a concise survey of the conventional determi-
nation of the geoid by means of gravity reductions, in order to understand
better the modern ideas. After an exposition of Molodensky’s theory, we
show how the new methods may be applied to classical problems such as
gravity reduction or the determination of the geoid by gravimetric and as-
trogeodetic methods. It should be mentioned that the terms “modern” and
“conventional” merely serve as convenient labels; they do not imply any con-
notation of value or preferability.
earth's
surface
P0 geoid
±N P c
cogeoid
N c N
ellipsoid
Q0
formulas. This is the purpose of the various gravity reductions. They were
considered extensively in Chap. 3; we therefore can limit ourselves to pointing
out those theoretical features that are relevant to our present problem.
If the external masses, the masses outside the geoid, are removed or
moved inside the geoid, then gravity changes. Furthermore, gravity is ob-
served at ground level but is needed at sea level. Thus, the reduction of
gravity involves the consideration of these two effects, in order to obtain
boundary values on the geoid.
This regularization of the geoid by removing the external masses unfortu-
nately also changes the level surfaces and hence, in general, the geoid. This is
the indirect effect; the changed geoid is called the cogeoid or the regularized
geoid.
The principle of this method may be described as follows (Jung 1956:
p. 578); see Fig. 8.1.
2. The gravity station is moved from P down to the geoid, to the point
P0 . Again, the corresponding effect on the gravity is considered.
4. The gravity station is now moved from the geoidal point P0 to the
292 8 Determination of the figure of the earth
cogeoid, to the point P c (hence the notation with upper index c). This
gives the boundary value of gravity at the cogeoid, gc .
5. The shape of the cogeoid is computed from the reduced gravity anoma-
lies
∆gc = gc − γ (8–4)
by Stokes’ formula, which gives N c = QP c .
N = N c + δN . (8–5)
Remark. At first sight it may seem that the masses between the geoid
and the cogeoid should be removed if the cogeoid happens to be below the
geoid, because Stokes’ formula is applied to the cogeoid. However, this is
not necessary, and therefore we need not be concerned with a “secondary
indirect effect”. The argument is a little too technical to be presented here;
see Moritz (1965: p. 26).
In principle, every gravity reduction that gives boundary values at the
geoid is equally suited for the determination of the geoid, provided the in-
direct effect is properly taken into account. Thus, the selection of a good
reduction method should be made from other points of view, such as the
geophysical meaning of the reduced gravity anomalies, the simplicity of com-
putation, the feasibility of interpolation between the gravity stations, the
smallness or even absence of the indirect effect, etc. (see Sect. 3.7).
The Bouguer reduction corresponds to a complete removal of the ex-
ternal masses. In the isostatic reduction, these masses are shifted vertically
downward according to some theory of isostasy. In Helmert’s condensation
reduction, the external masses are compressed to form a surface layer on
the geoid. The Bouguer reduction and especially the isostatic reduction (in
modern terminology topographic-isostatic reduction) are used as auxiliary
quantities for computational purposes, especially to facilitate interpolation.
The free-air anomaly is nowadays used in three senses:
1. at ground level (on the physical surface of the earth) it is simply the
gravity anomaly in the sense of Molodensky (Sect. 8.4);
2. at sea level it may be identified with the analytical continuation of
the Molodensky anomaly from ground down to sea level. This will
be considered in detail in Sect. 8.6. A final review will be found in
Sect. 8.15.
8.2 Gravity reductions and the geoid 293
∂N ∂N c ∂(δN )
ε=− =− − . (8–8)
∂s ∂s ∂s
This means that we must add to the immediate result of Vening Meinesz’
formula, −∂N c /∂s, a term representing the horizontal derivative of δN (see
also Sect. 3.7).
To repeat, the main purpose is to obtain a simple boundary surface.
The geoid approximated by an ellipsoid or even a sphere is a much easier
boundary surface than the physical surface of the earth, to which we turn
now.
which shows that the force g is the gradient vector of the potential.
Let S be the earth’s topographic surface and let W and g be the geopo-
tential and the gravity vector on this surface. Then there exists a relation
g = f (S, W ) , (8–10)
V = W − Φ. (8–11)
W = F2 (S, g) , (8–13)
that is, we get potential from gravity. As we shall see, this is far from
being trivial: we have now a method to replace leveling, a tedious
and time-consuming old-fashioned method, by GPS leveling, a fast and
modern technique (Sect. 4.6).
taken into account according to Sect. 8.2. To apply the above formulas, either
Φ and Λ must be reduced down to the geoid or ξ and η must be reduced up
to the ground. In both cases this involves the reduction for the curvature of
the plumb line (Sect. 5.15), which also depends on the mean value ḡ through
its horizontal derivatives. Hence Prey’s reduction enters here too.
Thus we see that in the conventional approach to the problems of physical
geodesy we must know the density of the outer masses or make assumptions
concerning it. To avoid this, Molodensky proposed a different approach in
1945.
Figure 8.2 shows the geometrical principles of this method, which is es-
sentially a linearization of Eq. (8–10). The ground point P (i.e., point on the
earth’s surface S) is again projected onto the ellipsoid according to Helmert.
However, the ellipsoidal height h is now determined by
h = H∗ + ζ , (8–18)
the normal height H ∗ replacing the orthometric height H, and the height
anomaly ζ replacing the geoidal undulation N .
This will be clear if one considers the surface whose normal potential U at
every point Q is equal to the actual potential W at the corresponding point
P , so that UQ = WP , corresponding points P and Q being situated on the
same ellipsoidal normal. This surface is called the telluroid (Hirvonen 1960,
1961). The vertical distance from the ellipsoid to the telluroid is the normal
height H ∗ (Sect. 4.4), whereas the ellipsoidal height h is the vertical distance
from the ellipsoid to the earth’s surface. Thus, the difference between these
two heights is the height anomaly
ζ = h − H∗ , (8–19)
P
³ earth's
surface S
Q
telluroid S
H* h
ellipsoid E
Q0
where γQ0 is the normal gravity at the ellipsoidal point Q0 . Note that H ∗ is
independent of the density.
The normal height H ∗ of a ground point P is identical with the ellipsoidal
height h, the height above the ellipsoid, of the corresponding telluroid point
Q. If the geopotential function W were equal to the normal potential function
U at every point, then Q would coincide with P , the telluroid would coincide
with the physical surface of the earth, and the normal height of every point
would be equal to its ellipsoidal height. Actually, however, WP = UP ; hence
the difference
ζP = hP − HP∗ = hP − hQ (8–22)
is not zero. This explains the term “height anomaly” for ζ.
The gravity anomaly is now defined as
∆g = gP − γQ ; (8–23)
it is the difference between the actual gravity as measured on the ground
and the normal gravity on the telluroid. The normal gravity on the telluroid,
which we shall briefly denote by γ, is computed from the normal gravity at
the ellipsoid, γQ0 , by the normal free-air reduction, but now applied upward:
∂γ ∗ 1 ∂ 2 γ ∗2
γ ≡ γQ = γQ0 + H + H + ··· . (8–24)
∂h 2! ∂h2
For this reason, the new gravity anomalies (8–23) are called free-air anoma-
lies. They are referred to ground level, whereas the conventional gravity
anomalies have been referred to sea level. Therefore, the new free-air anoma-
lies have nothing in common with a free-air reduction of actual gravity to
sea level, except the name. This distinction should be carefully kept in mind.
A direct formula for computing γ at Q is (2–215),
∗ 2
H ∗ H
γ = γQ0 1 − 2(1 + f + m − 2f sin2 ϕ) +3 , (8–25)
a a
8.4 Molodensky’s approach and linearization 299
T
ζ= , (8–26)
γ
Gravity disturbance
As usual, the gravity disturbance is defined by
δg = gP − γP . (8–27)
Linearization
The linearization applies equally well for the Molodensky problem and the
GPS problem. The geometry is familiar (Fig. 8.2).
We recall the surface Σ, the telluroid, which is defined by the condition
U (Q) = W (P ) . (8–28)
We note that (8–28) is the surface equivalent to the classical relation for sea
level (Fig. 8.3)
U (Q0 ) = W (P0 ) . (8–29)
Equation (8–28) would apply with
if S were an equipotential surface, the geoid, which is the case only over the
oceans with the usual simplifying assumption that the surface of the ocean
is an equipotential surface not changing with time (Fig. 8.3).
Molodensky’s theory does not use the geoid directly but the physical
earth’s surface. We repeat once more that this is Molodensky’s epochal idea
which radically changed the course of physical geodesy since 1945.
We shall, however, use the fictitious case of S being an equipotential sur-
face, but only as a first (or zero-order) assumption in a perturbation approach
for the real earth’s surface (Molodensky series). This first approximation is
the spherical case to be considered in the next section.
Now we consider the linearization in more detail. The ellipsoidal height
h is directly determined by GPS. It may be decomposed into
h = H∗ + ζ . (8–31)
Here, H ∗ is the normal height and ζ is the height anomaly, whose definitions
are seen from Fig. 8.2. In the GPS case we do know the earth’s surface S
directly, but the telluroid Σ and the height anomalies ζ are still required for
formulating the boundary condition, just as the knowledge of the geoid does
not make superfluous the reference ellipsoid.
P0 geoid
N
ellipsoid
Q0
∞
∆g(ϑ, λ) = ∆gn (ϑ, λ) (8–41)
2
∞
∆gn
T =R . (8–42)
n=2
n−1
where
∞
2n + 1
S(ψ) = Pn (cos ψ) , (8–44)
n−1
n=2
where P (cos ψ) are Legendre polynomials. Here ψ denotes the spherical dis-
tance from the point at which T is to be computed.
In exactly the same way, we obtain for the gravity disturbance with the
boundary condition (8–39), summarizing the derivation in Sect. 2.18,
∞
δg(ϑ, λ) = δgn (ϑ, λ) , (8–45)
0
∞
δgn
T (ϑ, λ) = R , (8–46)
n+1
n=0
8.6 Solution by analytical continuation 303
where
∞
2n + 1
K(ψ) = Pn (cos ψ) (8–48)
n+1
n=0
Dg earth's surface S
z
P Dg' h
point level U = UP
hP
ellipsoid U = U0
Fig. 8.4. Analytical continuation from the earth’s surface to point level
surface U = UP , which for our purpose is the same). In the spherical ap-
proximation, both surfaces U = UP and U = U0 are concentric spheres, but
only in the precise sense of the spherical approximation as explained above.
We also use the term “harmonic continuation” because the analytically
continued function satisfies Laplace’s equation. This will be explained in
detail later.
An expansion into a Taylor series gives immediately
∂∆g∗ 1 ∂ 2∆g∗ 1 3 ∂ 3∆g∗
∆g = ∆g∗ + z + z2 + z + ···
∂z 2! ∂z 2 3! ∂z 3
∞ (8–51)
1 n ∂ n∆g∗
= ∆g∗ + z ,
n! ∂z n
n=1
where
z = h − hP (8–52)
is the elevation difference with respect to the computation point P . For
the present, we assume the series (8–51) to be convergent. Here ∆g∗ is the
gravity anomaly at point level (Fig. 8.4). The use of a Taylor series is typical
for analytical continuation. For instance, Taylor series are a standard tool
for analytical continuation of functions of a complex variable.
(a)
Dg
P h
hP
Dg harmonic sea level
(b)
Dg
P h – h0
hP – h0 reference level
h0
(c)
Dg h – hP
P point level Dg *
hP
Dg harmonic
In particular we may take as reference level the level of the point P itself,
so that
h0 = hP , (8–56)
where P is the point at which the height anomaly ζ is computed. If this
choice is made, the last term in the above expression will be zero, because
outside the integral h always means hP , so that h − h0 = hP − hP = 0. Thus
we have
R ∂∆g
ζ= ∆g − (h − hP ) S(ψ) dσ . (8–57)
4π γ0 ∂h
σ
This formula is particularly simple; geometrically it means that the free-
air anomalies are “reduced” (in the sense of “analytically or harmonically
continued”) from the ground to the level of the computation point P (see
Fig. 8.5 b). Thus, the reference level is different for different computation
points.
As we have already indicated at the beginning of Sect. 8.6.1, Fig. 8.5 c
shows that harmonic continuation by Eq. (8–57) is upward for surface points
below the level of P and downward for surface points above the level of P .
Important remark
Equation (8–57) is really a genuine spherical Stokes formula applied to a
“reference sphere”, namely, to the spherical “point level”! An immediate
consequence: this formula can be simply differentiated horizontally to give a
genuine Vening Meinesz formula in the sense of Sect. 2.19 for the deflections
of the vertical. This remark is relevant for Sect. 8.7.
Vertical derivative
The vertical derivative ∂/∂r can be expressed in terms of surface values by
the well-known spherical formula (Sect. 1.14)
∂f 1 R2 f − fQ
=− f+ dσ . (8–58)
∂r R 2π l03
σ
Q is the surface point where ∂f /∂r is computed and to which f in the first
term on the right-hand side refers, σ denotes the unit sphere, and
ψ
l0 = 2R sin . (8–59)
2
This gives ∂∆g/∂r if we put f = ∆g in (8–58). We may also introduce the
linear gradient operator L by
R2 f − fQ
L(f ) = dσ . (8–60)
2π l03
σ
8.6 Solution by analytical continuation 307
(The first term on the right-hand side of (8–58) is much smaller and can be
omitted.)
The term ∂ζ/∂r no longer occurs in (8–57) as it did in (8–53) and (8–55),
and will not be needed.
∆g∗ = ∆g + g1 , (8–61)
Important remark
Please note carefully that we are using “linear”, or “first-order”, in two very
different senses:
∆g∗ = ∆g + g1 + g2 + g3 + · · · (8–67)
and
ζ = ζ0 + ζ1 + ζ2 + ζ3 + · · · . (8–68)
Generalizing (8–66), we have
R
ζi = gi S(ψ) dσ , (8–69)
4π γ0
σ
starting from
g0 = ∆g . (8–71)
Here the operator Ln is also defined recursively:
starting with
L1 = L (8–73)
with the gradient operator L defined above, (8–60), and z given by (8–52).
you are not satisfied with 1 mgal, prescribe 10−3 mgal or 10−1000 mgal!
Bjerhammar has pointed out that the assumption of a complete contin-
uous gravity coverage at every point of the earth’s surface, from which the
above negative answer follows, is unrealistic because we can measure gravity
only at discrete points. If the purpose of physical geodesy is understood as
the determination of a gravity field that is compatible with the given discrete
observations, then it is always possible to find a potential that can be an-
alytically continued down to the ellipsoid. This is the theoretical basis for
least-squares collocation.
Here we need only one result: Do not worry about analytical contin-
uation! It is always possible with an arbitrarily small error being
not equal to 0 (though not for one being 0).
So, in the same year 1969, Marych and Moritz independently found an el-
ementary solution by analytical continuation in the form of an infinite series
denoted as “Molodensky series”. Details can be found in Moritz (1980 a):
The original form of Molodensky’s series obtained by solving an integral
equation is found in Sect. 45. Pellinen’s equivalence proof that the simple
“analytical continuation solution” and Molodensky’s integral equation solu-
tion are equivalent (that means, the series are termwise equal!) is found in
Sect. 46.
We remark that analytical continuation is a purely mathematical concept
independent of the density of the topographic masses. Thus, it is not an
“introduction of gravity reduction through the backdoor”, which would be
contrary to the spirit of Molodensky’s theory.
P
Dg
hP ground
h
sea level
Dg harmonic
Fig. 8.6. Free-air anomalies at ground level, ∆g, and at sea level, ∆g harmonic
312 8 Determination of the figure of the earth
with the actual gravity anomalies outside the earth, since the function r ∆g
is harmonic according to Sect. 2.14.
(Remark: we are consistently using the notation ∆g for ground level,
∆gharmonic for sea level, and ∆g∗ for point level; see Fig. 8.5.)
It follows that the harmonic function T that is produced by ∆gharmonic
according to Pizzetti’s generalization (2–302) of Stokes’ formula
R
T (r, ϑ, λ) = ∆gharmonic S(r, ψ) dσ (8–74)
4π
σ
is identical with the actual disturbing potential of the earth outside and on
its surface.
Applications
Assume that we got in some way (e.g., by the Taylor series mentioned above
or by collocation to be treated in Chap. 10 or by a high-resolution gravita-
tional field from satellite observations) the downward continuation ∆gharmonic
to sea level. Then we can compute the external gravity field, its spherical
harmonics, etc., rigorously by means of the conventional formulas of Chaps. 2
and 6, provided we use ∆gharmonic rather than ∆g in the relevant formulas.
For instance, the coefficients of the spherical harmonics of the gravitational
potential may be obtained by expanding the function ∆gharmonic according to
Sect. 1.9 together with Sect. 1.6. If we wish to compute the height anomaly
ζ at a point P at ground level, we must remember that P lies above the el-
lipsoid, so that the formulas for the external gravity field are to be applied.
By Bruns’ formula ζ = T /γ0 (8–50), we get
R
ζ= ∆gharmonic S(r, ψ) dσ , (8–75)
4π γ0
σ
C'
B
A
∂∆g
∆gharmonic = ∆g − (h − hP ) . (8–78)
∂h
8.7 Deflections of the vertical 315
Since these anomalies refer to a level surface, Vening Meinesz’ formula can
now be applied directly and gives (8–77).
ξ = Φ − ϕ∗ ,
(8–79)
η = (Λ − λ∗ ) cos ϕ .
ellipsoidal normal
normal
plumb line P
earth's
surface
parallel to
'* equator
' ''
Q0 Q00 ellipsoid
λ = λ , (8–81)
316 8 Determination of the figure of the earth
since Q0 and Q00 lie on the same ellipsoidal meridian. Furthermore, even
in extreme cases the distance between Q0 and Q00 can never exceed a few
centimeters. For this reason, we may also set
ϕ = ϕ (8–82)
Therefore, ξ and ξHelmert differ by the normal reduction for the curvature of
the plumb line,
h
−δϕnormal = f ∗ sin 2ϕ . (8–86)
R
The deflection components ξHelmert and ηHelmert are used in astrogeodetic com-
putations; ξ and η are those obtained gravimetrically from formulas such as
(8–77) and (8–88) below.
These relations are mathematically quite analogous to the corresponding
equations (5–138) for the conventional method using the geoid, but now,
with the use of the normal curvature, the once formidable obstacle of the
correction for plumb-line curvature practically belongs to the past.
Remark on accuracy
With Molodensky’s theory, the accuracy problem mentioned at the end of
Sect. 2.21 even aggravates, because in a mountainous terrain it is almost
impossible to compute the Molodensky corrections with an accuracy of 0.03
(say), so that these observations cannot be directly used for precise horizontal
positions.
8.8 Gravity disturbances: the GPS case 317
1∞
1 dK dK
ξ= δg cos α dσ + gn cos α dσ ,
4π γ0 dψ 4π γ0 dψ
σ n=1 σ
∞ (8–88)
1 dK 1 dK
η= δg sin α dσ + gn sin α dσ .
4π γ0 dψ 4π γ0 dψ
σ n=1 σ
T =γζ. (8–92)
Then
W =U +T (8–93)
C = W0 − W (8–94)
is the geopotential number, the physical measure of height above sea level,
conventionally obtained by the cumbersome method of leveling, but now
computed in a direct way from gravity data. This is the physical, more
general, equivalent of the geometric determination of the normal height by
H ∗ = h − ζ, according to Eq. (8–31).
It can be shown that, in the linear approximation, the Molodensky cor-
rection for the gravity disturbance has the same form as for the gravity
anomaly and can for each quantity be computed using either ∆g or δg.
The formulas for the Molodensky corrections and their numerical values
are the same to the linear approximation.
All this shows the power of Molodensky’s approach even in problems he
never treated himself.
Let the masses outside the geoid be removed or moved inside the geoid, as
described in Sect. 8.2, and consider the effect of this procedure on quantities
referred to the ground.
We denote the changes in potential and in gravity by δW and δg; then
the new values at ground will be
W c = W − δW ,
(8–95)
gc = g − δg .
(It is clear that δg here is not the gravity disturbance!) The disturbing
potential T = W − U becomes
T c = T − δW . (8–96)
The physical surface S as such will remain unchanged, but the telluroid Σ will
change, because its points Q are defined by UQ = WP , and the potential W at
any surface point P will be affected by the mass displacements according to
(8–95). The distance Q Qc between the original telluroid Σ and the changed
telluroid Σc (Fig. 8.9) is given by
δW
Q Qc = (8–97)
γ
according to Bruns’ theorem. This is identical with the variation of the height
anomaly ζ, so that
δW
δζ = ζ − ζ c = . (8–98)
γ
Normal gravity γ on the telluroid Σ becomes on the changed telluroid Σc
∂γ 1 ∂γ
γc = γ + δζ = γ + δW , (8–99)
∂h γ ∂h
³c
³ earth's surface S
Qc
±³
Q changed telluroid Sc
telluroid S
ellipsoid E
1. the direct effect, −δg, of the shift of the outer masses on g; and
Let us repeat once more that all these anomalies ∆gc refer to the physical
surface of the earth, to “ground level”!
If the masses outside the geoid are completely removed, then ∆gc is
a Bouguer anomaly; if the outer masses are shifted vertically downward
according to some isostatic hypothesis, then ∆gc is an isostatic anomaly, etc.
In this way we may get a “ground equivalent” for each conventional gravity
reduction. The two are always related by analytical continuation. See below
for the isostatic anomalies; for analytical continuation see Sect. 8.6.
Now we may describe the determination of the height anomalies ζ in a
way that is similar to the corresponding procedure for the geoidal undula-
tions N of Sect. 8.2:
ζ = ζ c + δζ . (8–103)
according to (8–75) and (8–103), and on the other hand the geoidal undula-
tions by
R c∗ δW
N= ∆g S(ψ) dσ + (8–105)
4π γ0 γ geoid
σ
according to the ordinary Stokes formula applied to ∆gc∗ and (8–5). Since
the height anomalies refer to the elevation h, the function S(R+h, ψ) replaces
in (8–104) the original function of Stokes S(ψ) ≡ S(R, ψ), which occurs in
(8–105) because the geoidal undulation refers to zero elevation. We could
use γ0 in (8–104) as well. Summarizing, we have the following steps:
1. Computation of the free-air anomaly at ground level, ∆g, according to
(8–23).
2. Computation of the isostatic anomaly at ground level, ∆gc , according
to (8–101).
3. Downward continuation of ∆gc by (8–54), where ∆g and ∆gharmonic are
replaced by ∆gc and ∆gc∗ . The resulting isostatic anomalies at sea
level, ∆gc∗ , may now be used for two purposes: either for
4a. the determination of the physical surface of the earth according to
(8–104), or for
4b. the determination of the geoid according to (8–105).
An error in the assumed density of the masses below the earth’s surface
affects the geoidal undulations as determined from (8–105) but does not
influence the height anomalies resulting from (8–104). This is clear because
a wrong guess of the density means only that the masses above sea level are
not completely removed, which is no worse than not removing them at all
when using free-air anomalies.
This method is of particular interest for practical computations, as we
will see later. It has become popular by the name “remove-restore method”,
invented by K. Colic and others, see Sect. 11.1.
the interior of the earth in such a way that the exterior potential remains
unchanged. This is not unlike the Rudzki reduction, where the geoid remains
unchanged. Whereas the Rudzki reduction is, however, “constructive” in the
sense that a way of performing it can be described, our present interpreta-
tion of free-air reduction as harmonic continuation is nonconstructive, it is
an “improperly posed” inverse problem; cf. Anger and Moritz (2003) and
www.inas.tugraz.at/forschung/InverseProblems/AngerMoritz.html, as well
as Fig. 8.10.
S topographic masses
(a)
geoid
W W
(DW = 0) (DW = –4¼G%)
W =W0
S S
(b) cogeoid (c) harmonic
geoid
W harmonic
Wc (DW harmonic = 0)
Wc (DW = 0)
c W
(DW = 0)
(DW c = 0)
W c =W0 W harmonic =W0
Fig. 8.10. (a) Geoid and topographic masses, (b) mass displacement in gravity
reduction, (c) “ill-defined” mass displacement in free-air reduction as harmonic
continuation
Important remark
The isostatic gravity anomalies and the topographic-isostatically reduced
deflections of the vertical (Sect. 8.14) are fundamental for least-squares col-
location in mountain areas (Sect. 11.2). Thus, the spatial approach due
to Molodensky is basic even for least-squares collocation!
324 8 Determination of the figure of the earth
Exercise
Collecting all these remarks into a separately readable paper on the various
aspects of free-air reduction would be a nice task for a seminar work. The
present authors offer a prize of Euro 500, the “Molodensky Prize”, to the
first excellent review paper on this topic.
h = H∗ + ζ . (8–107)
N − ζ = H∗ − H . (8–108)
P
³
Q earth's
surface
H
telluroid
H*
P0 geoid
N
ellipsoid
Q0
This means that the difference between the geoidal undulation N and the
height anomaly ζ is equal to the difference between the normal height H ∗ and
the orthometric height H. Since ζ is also the undulation of the quasigeoid,
this difference is also the distance between geoid and quasigeoid.
According to Sect. 4.5, the two heights are defined by
C C
H= , H∗ = , (8–109)
ḡ γ̄
where C is the geopotential number, ḡ is the mean gravity along the plumb
line between geoid and ground, and γ̄ is the mean normal gravity along the
normal plumb line between ellipsoid and telluroid. By eliminating C between
these two equations, we readily find
ḡ − γ̄
H∗ − H = H, (8–110)
γ̄
which is also the distance between the geoid and the quasigeoid, see (8–108);
hence
ḡ − γ̄
N =ζ+ H. (8–111)
γ̄
The height anomaly ζ may be expressed, for instance, by Molodensky’s
formula (8–57). Then we obtain
R R ḡ − γ̄
N= ∆g S(ψ) dσ + g1 S(ψ) dσ + H , (8–112)
4π γ0 4π γ0 γ̄
σ σ
. 1 ∂γ
γ̄ = γ − H. (8–114)
2 ∂h
The quantity γ̄ in the denominator can be replaced by our usual constant
γ0 . Since the Bouguer anomaly is rather insensitive to local topographic
irregularities, the coefficient is locally constant so that there is approximately
a linear relation between ζ and the local irregularities of the height H. In
other words, the quasigeoid mirrors the topography (Fig. 8.12).
8.11 A first balance 327
earth's
surface
quasigeoid
³
ellipsoid
ζ − N = 0.1 m . (8–117)
geoid. The avoidance of such assumptions has been the guiding idea of Molo-
densky’s research. However, orthometric heights are but little affected by er-
rors in density. The error in H due to the imperfect knowledge of the density
hardly ever exceeds 1–2 decimeters even in extreme cases (Sect. 4.3). It is
presumably smaller than the inaccuracy of the corresponding ζ even with
very good gravity coverage, because of inevitable errors of interpolation, etc.
If, therefore, the method of Sect. 8.10 is used, the geoid can be determined
with virtually the same accuracy as the quasigeoid. Note that it is theoret-
ically even possible to eliminate completely the errors arising from the use
of the geoid (Moritz 1962, 1964). Thus, we may well retain the geoid with
its physical significance and its other advantages.
How much do Molodensky’s formulas differ from the corresponding equa-
tions of Stokes and Vening Meinesz? The deviation of ζ from the result of
the original Stokes formula is given by the equivalent expressions
R R ∂∆g
ζ1 = g1 S(ψ) dσ or ζ1 = − (h − hP ) S(ψ) dσ
4π γ0 4π γ0 ∂h
σ σ
(8–120)
according to Eqs. (8–62) and (8–66). This correction may even be smaller
than the difference ζ − N (see Sect. 11.3).
It is appropriate again to point out that the deflection of the vertical
is relatively more affected by the Molodensky correction than is the height
anomaly. In extreme cases, this correction may attain values of a few seconds,
as studies of models by Molodensky (Molodenski et al. 1962: pp. 217–225)
indicate. This is considerable, since 1 in the deflection corresponds to 30 m
in position. Numerical estimates will be found in Chap. 11.
We may summarize the result of applying Stokes’ and Vening Meinesz’
formulas to free-air anomalies directly, without any corrections. Stokes’ for-
mula yields height anomalies ζ with high accuracy; for many practical pur-
poses, we may, in addition, identify these height anomalies with the corre-
sponding geoidal undulations N . Vening Meinesz’ formula gives deflections
of the vertical at ground level that are relatively less accurate but often
acceptable.
An advantage of the modern theory is its direct relation to the external
gravity field of the earth, which is particularly important nowadays for the
computation of the effect of gravitational disturbances on spacecraft trajec-
tories and satellite orbits. It is immediately clear that ground-level quan-
tities, such as free-air gravity anomalies, are better suited for this purpose
than the corresponding quantities referred to the geoid, which is separated
from the external field by the outer masses. For the computation of the ex-
ternal field and of spherical harmonics, the method described in Sect. 8.6.5
8.12 Some background 329
does not work over land areas. The classical method for a detailed geoid
determination on the continents has been the gravimetric method, in spite
of the fact that it is severely handicapped by lack of an adequate gravity
coverage (or lack of information on such a coverage). Thus, we have the
paradoxical situation that on the oceans, long a stepchild of geodesy, the
geoid is now in general known much better than on the continents.
Still, the gravimetric method has continued to fascinate theoreticians
because it gives rise to very interesting and deep mathematical problems,
related to the geodetic boundary-value problem discussed above in this chap-
ter.
These enormous practical and theoretical developments concerning global
satellite and gravimetric gravity field determination have somewhat over-
shadowed the determination of detailed geoids in smaller areas, in partic-
ular, astrogeodetic geoids. Especially in mountainous regions, local geoid
determinations are difficult. The gravimetric method does not work very
well in high mountains. The astrogeodetic method, using astronomical ob-
servations of latitude and longitude, does work well there but is considered
time-consuming and somewhat old-fashioned, perhaps also because work-
ing during the night is not very popular nowadays. An appropriate use
of gravity and astrogeodetic data in high mountains must involve some
topographic-isostatic reduction. Furthermore, the theory behind the astro-
geodetic method is not nearly as attractively difficult as the theory of Molo-
densky’s problem. Last but not least, high-mountain areas are exceptional
and, apart from such countries as Switzerland and Austria, are frequently
regions of little economic interest. For these and similar reasons, the main-
stream of geodetic practice and theory has flown with grand indifference
around high mountains, ignoring such trivial obstacles.
Still, a country such as Switzerland has made a virtue out of necessity
and has traditionally been very active in local astrogeodetic geoid determi-
nation (Elmiger 1969, Gurtner 1978, Gurtner and Elmiger 1983). Austria
has followed up (Österreichische Kommission für die Internationale Erdmes-
sung 1983). It has been found that, even besides the problem of getting the
required observations, the underlying theory is not so trivial as one might
think and shows quite interesting features.
Concerning measurements, astronomical observations have again proved
very feasible in mountains; see the articles by Erker, Bretterbauer and Gerst-
bach, Lichtenegger and Chesi in Chap. 2 of Österreichische Kommission für
die Internationale Erdmessung (1983), followed by Sünkel et al. (1987). The
main advantages of astrogeodetic versus gravimetric data for local geoid
determination in mountainous regions may be summarized as follows:
1. It is sufficient to have astrogeodetic deflections of the vertical in the
8.12 Some background 331
earth's surface
telluroid
P
actual level surface W = WP
³
Q normal level surface U = WP
"
–± "
normal plumb line
(curved)
actual plumb line
ellipsoidal normal
(straight)
P0 P0' P0'' geoid W = W0
N "0
ellipsoid U = W0
Q0 Q0' Q0''
U (Q) = W (P ) . (8–121)
That is, Q is defined such that its normal potential U equals the actual
potential W of P .
This corresponds to the classical relation
ζ = QP , (8–123)
8.12 Some background 333
N = Q0 P0 . (8–124)
T =W −U, (8–125)
where g denotes gravity and γ normal gravity. So far, g(P0 ) denotes the ac-
tual gravity on the geoid; we are not yet here considering mass-transporting
gravity reductions.
Analogously we have according to Molodensky:
Generally we will, as far as feasible, use the subscript “0” to designate quan-
tities referred to sea level, to distinguish them from quantities referred to
the earth’s surface, which do not carry such a subscript. For instance, ∆g0
334 8 Determination of the figure of the earth
refers to sea level and ∆g to the earth’s surface. With GPS we have gravity
disturbances
∆g = g(P ) − γ(P ) . (8–129)
Regarding plumb line definition, we must distinguish three lines (Fig. 8.13):
1. the straight ellipsoidal normal Q0 P ,
2. the actual plumb line P0 P ,
3. the normal plumb line P0 P .
Geometrically, the ellipsoidal normal is defined as the straight line through
P perpendicular to the ellipsoid. The (actual) plumb line is defined by the
condition that, at each point of the line, the tangent coincides with the
gravity vector g at that point; the plumb line is very slightly curved, but its
curvature is irregular, being determined by the irregularities of topographic
masses. The normal plumb line, at each of its points, is tangent to the normal
gravity vector γ; it possesses a curvature that is even smaller and completely
regular.
The points P0 , P0 , and P0 coincide within a few decimeters, and we will
not distinguish them in what follows. The reason is that the distance, in arc
seconds, between P0 and P0 is much smaller than the effect of plumb line
curvature (Sect. 5.15). The same applies for Q0 , Q0 , and Q0 .
The direction of the gravity vector g is the direction of (the tangent to)
the plumb line. It is determined by two angles, the astronomical latitude Φ
and the astronomical longitude Λ. Let Φ, Λ be referred to the earth’s surface
(to point P ) and Φ0 , Λ0 to the geoid (strictly speaking, to point P0 ). The
differences
δϕ = Φ0 − Φ , δλ = Λ0 − Λ (8–130)
express the effect of plumb line curvature (Fig. 8.14). You may also wish to
refer back to Fig. 5.18. Hence, we have
Φ0 = Φ + δϕ , Λ0 = Λ + δλ . (8–131)
Knowing the plumb line curvature δΦ, δΛ, we could use these simple formulas
to compute the sea-level values Φ0 , Λ0 from the observed surface values Φ, Λ.
In the same way as Φ, Λ are related to the actual plumb line, the ellip-
soidal latitude ϕ and the ellipsoidal longitude λ refer to the straight ellip-
soidal normal. The quantities
ξ = Φ − ϕ, η = (Λ − λ) cos ϕ (8–132)
are the components of the deflection of the vertical in a north-south and an
east-west direction. For an arbitrary azimuth α, the vertical deflection ε is
given by
ε = ξ cos α + η sin α . (8–133)
8.12 Some background 335
±'
parallel to F0
equator F
P0''
where δϕ, δλ express the normal plumb line curvature. These equations are
the “normal equivalent” to (8–131): the “normal surface values” ϕ̄, λ̄ cor-
respond to the “actual surface values” Φ, Λ and the ellipsoidal values ϕ, λ
correspond to the geoidal values Φ0 , Λ0 . To make the analogy complete, we
should replace ϕ = ϕ(P0 ) by ϕ(P0 ), but we have consistently neglected such
differences.
In contrast to the actual plumb line curvature, it is very easy to compute
the normal curvature of the plumb line: from (5–147) we have
earth’s surface. On the other hand, the normal plumb line is physically (or
dynamically) defined by means of the external gravity field of an equipo-
tential ellipsoid. Hence also ϕ̄, λ̄ are dynamically defined. The quantities
obtained by replacing ϕ, λ by ϕ̄, λ̄ so that
ξ̄ = ξ + δϕnormal , η̄ = η , (8–139)
Compare ε and ε̄ in Fig. 8.13 and note that in this figure δ denotes the
curvature of the normal plumb line for the azimuth α given by the analogous
formula
dN = −ε0 ds , (8–142)
where ε0 denotes the deflection of the vertical at the geoid. Integration be-
tween two points A and B yields the difference between their geoidal heights:
B
NB − NA = − ε0 ds , (8–143)
A
N
ds ellipsoid
or, approximately,
ε0A + ε0B
NB − NA = − sAB , (8–144)
2
where sAB denotes the horizontal distance between A and B. The minus
sign is conventional. Cf. Sect. 5.14.
A corresponding relation to height anomalies according to Molodensky
is found as follows (Molodensky et al. 1962: p. 125):
∂ζ ∂ζ
dζ = ds + dh , (8–145)
∂s ∂h
notations following Fig. 8.16. Since the earth’s surface is not a level surface,
we also have a vertical part (∂ζ/∂h) h in addition to the usual horizontal
part (∂ζ/∂s) ds. The vertical part arises from change in height and is usually
smaller than the horizontal part.
In analogy to (8–142), the horizontal part is given by
∂ζ
= −ε̄ , (8–146)
∂s
where ε̄ denotes the dynamical deflection of the vertical at the earth’s surface;
cf. (8–140) and Fig. 8.13. For the vertical part we have from (8–126):
∂ζ ∂ T 1 ∂T 1 ∂γ
= = − T (8–147)
∂h ∂h γ γ ∂h γ ∂h
or
∂ζ ∆g g−γ
=− =− (8–148)
∂h γ γ
according to the fundamental equation of physical geodesy (8–36).
Hence (8–145) becomes
g−γ
dζ = −ε̄ ds − dh . (8–149)
γ
earth's surface
dh
W = WP
P @³
" @s
ds
ds U = WP
gc = g − δg . (8–151)
upward from Q0 to Q. Whereas for the first process the use of the normal
gradient ∂γ/∂h is problematic, it is fully justified for the second process.
In a similar way, we might interpret δϕnormal as a reduction of ϕ for
normal curvature of the plumb line upwards, say, from P0 to P . This is
possible because in (8–136) ϕ could be said to refer to P0 (because P0 and
P0 practically coincide), and because ϕ̄ denotes the latitude of the tangent
to the normal plumb line at P . This interpretation is instructive because of
the analogy with gravity reduction, though regarding ϕ and ϕ̄ as ellipsoidal
and dynamic latitude of the same point P appears more natural. Refer again
to our key figure (Fig. 8.13).
As pointed out above, the present interpretation of ξ c and η c as isostati-
cally reduced deflections of the vertical at the earth’s surface is conceptually
rigorous and therefore also practically more accurate, but this decisive ad-
vantage implies a computational drawback if integration along a profile is
used: Since this integration must now be performed along the earth’s surface
and not along a level surface such as the geoid, computation will be more
complicated. Instead of the simple Helmert formula (8–143), we now must
use the Molodensky formula (8–150):
B B
gc − γ
c
ζB − c
ζA =− ε ds −
c
dh (8–158)
A A γ
with
εc = ξ c cos α + η c sin α , (8–159)
and ∆gc = gc −γ, where gc is the isostatically reduced surface value of gravity
(measured value g minus attraction of the topographic-isostatic masses).
From the isostatic height anomalies ξ c obtained in this way, we then get
the actual height anomalies ζ by applying the indirect effect:
ζ = ζ c + δζ (8–160)
with
TTI
δζ = . (8–161)
γ
This is completely analogous to (8–5) and (8–3), but now TTI is the potential
of the topographic-isostatic masses at the surface point P . As a matter of
fact, normal gravity in (8–3) refers to the ellipsoid, and in (8–161) to the
telluroid, but the difference is generally small.
For higher mountains, the isostatic reduction procedure described in the
present section is preferable in practice to a direct application of Moloden-
sky’s formula (8–150) because the isostatically reduced vertical deflections
8.15 The meaning of the geoid 341
are much smoother and easier to interpolate. It is, however, extremely la-
borious from a computational point of view since the integration must be
performed along the earth’s surface (or, what is practically the same, along
the telluroid).
We remark that the computational drawback of the present method,
the Molodensky integration along the earth’s surface, can be completely
avoided if we perform our computations in space: instead of integrating
along a surface, we perform collocation in space. This modern procedure,
to be described in the next chapter, permits a simple and computationally
convenient use of surface deflections and also their combination with gravi-
metric and other data. Still, the present developments are necessary for a
full understanding of the collocation approach.
Final remarks
In these last sections we tried to apply the same principle for topographic-
isostatic reduction (the “remove-restore method”) at point level to all terres-
trial data related to the gravity vector: gravity anomalies and disturbances
(Sect. 8.9) and deflections of the vertical (Sect. 8.14). This unified view of
isostatically reduced data thus makes them directly suitable for combined
solutions by least-squares collocation to be treated in Chap. 10.
9.1 Introduction
Some of the most important problems of gravimetric geodesy are formulated
and solved in terms of integrals extended over the whole earth. An example
is Stokes’ formula. Thus, in principle, we need the gravity g at every point
of the earth’s surface. As a matter of fact, even in the densest gravity net
we measure g only at relatively few points so that we must estimate g at
other points by interpolation. In large parts of the oceans we have made no
observations at all; these gaps must be filled by some kind of extrapolation.
Mathematically, there is no difference between interpolation and extrap-
olation; therefore they are denoted by the same term, prediction.
Prediction (i.e., interpolation or extrapolation) cannot give exact values;
hence, the problem is to estimate the errors that are to be expected in the
gravity g or in the gravity anomaly ∆g. As usual, gravity disturbances δg
are appropriately comprised whenever we speak of gravity anomalies.
Since ∆g is further used to compute other quantities, such as the geoidal
undulation N or the deflection components ξ and η, we must also investigate
the influence of the prediction errors of ∆g on N, ξ, η, etc. This is called
error propagation, which will play a basic role.
It is also important to know which prediction method gives highest ac-
curacy, either in ∆g or in derived quantities N, ξ, η, etc. To be able to find
these “best” prediction methods, it is necessary to have solved the previous
problem, to know the prediction error of ∆g and its influence on the derived
quantities.
Summarizing, we have the following problems:
Since we are interested in the average rather than the individual errors,
we are led to a statistical treatment. This will be the topic of the present
chapter.
346 9 Statistical methods in physical geodesy
The symbol M stands for the average over the whole earth (over the unit
sphere); this average is equal to the integral over the unit sphere divided
by its area 4π. The integral is zero if there is no term of degree zero in the
expansion of the gravity anomalies ∆g into spherical harmonics, that is, if a
reference ellipsoid of the same mass as the earth and of the same potential
as the geoid is used. This will be assumed throughout this chapter.
Note that if this is not the case, that is, if M {∆g} = m = 0, then we
may form new gravity anomalies ∆g∗ = ∆g − m by subtracting the average
value m. Then M {∆g∗ } = 0 and all the following developments apply to the
“centered” anomalies ∆g∗ .
Clearly, the quantity M {∆g}, which is zero, cannot be used to charac-
terize the average size of the gravity anomalies. Consider then the average
square of ∆g,
1
var{∆g} ≡ M {∆g2 } = ∆g2 dσ . (9–2)
4π
σ
It is called the variance of the gravity anomalies. Its square root is the root
mean square (rms) anomaly:
rms{∆g} ≡ var{∆g} = M {∆g2 } . (9–3)
The rms anomaly is a very useful measure of the average size of the gravity
anomalies; it is usually given in the form
the sign ± expresses the ambiguity of the sign of the square root and sym-
bolizes that ∆g may be either positive or negative. The rms anomaly is very
intuitive; but the variance of ∆g is more convenient to handle mathemati-
cally and admits an important generalization.
9.2 The covariance function 347
The average is to be extended over all pairs of points P and P for which
P P = s = constant.
The covariances characterize the statistical correlation of the gravity
anomalies ∆g and ∆g , which is their tendency to have about the same
size and sign. If the covariance is zero, then the anomalies ∆g and ∆g are
uncorrelated or independent of one another (note that in the precise lan-
guage of mathematical statistics, zero correlation and independence are not
quite the same, but we may neglect the difference here!); in other words,
the size or sign of ∆g has no influence on the size or sign of ∆g . Gravity
anomalies at points that are far apart may be considered uncorrelated or
independent because the local disturbances that cause ∆g have almost no
influence on ∆g and vice versa.
If we consider the covariance as a function of distance s = P P , then we
get the covariance function C(s) mentioned at the beginning:
For s = 0, we have
C(0) = M {∆g2 } = var{∆g} (9–7)
according to (9–2). The covariance for s = 0 is the variance.
A typical form of the function C(s) is shown in Fig. 9.1. For small dis-
tances s (1 km, say), ∆g is almost equal to ∆g, so that the covariance is
almost equal to the variance; in other words, there is a very strong corre-
lation. The covariance C(s) decreases with increasing s because then the
C (s)
anomalies ∆g and ∆g become more and more independent. For very large
distances, the covariance will be very small but not in general exactly zero
because the gravity anomalies are affected not only by local mass distur-
bances but also by regional factors. Therefore, we may expect an oscillation
of the covariance between small positive and negative values.
Note that positive covariances mean that ∆g and ∆g tend to have the
same size and the same sign; negative covariances mean that ∆g and ∆g
tend to have the same size and opposite sign. The stronger this tendency,
the larger is C(s); the absolute value of C(s) can, however, never exceed the
variance C(0).
The practical determination of the covariance function C(s) is somewhat
problematic. If we were to determine it exactly, we should have to know grav-
ity at every point of the earth’s surface. This we obviously do not know; and
if we knew it, then the covariance function would have lost most of its signif-
icance because then we could solve our problems rigorously without needing
statistics. As a matter of fact, we can only estimate the covariance function
from samples distributed over the whole earth. But even this is not quite
possible at present because of the imperfect or completely missing gravity
data over the oceans. For a discussion of sampling and related problems see
Kaula (1963, 1966 b).
The first comprehensive estimate of the covariance function was made by
Kaula (1959). Some of his values are given in Table 9.1 for historical interest.
They refer to free-air anomalies. The argument is the spherical distance
s
ψ= (9–8)
R
corresponding to a linear distance s measured on the earth’s surface; R is a
mean radius of the earth. The rms free-air anomaly is
√
rms{∆g} = 1201 = ± 35 mgal . (9–9)
We see that C(s) decreases with increasing s and that, for s/R > 30◦ , very
small values oscillate between plus and minus.
For some purposes we need a local covariance function rather than a
global one; then the average M is extended over a limited area only, instead
of over the whole earth as above. Such a local covariance function is useful
for more detailed studies in a limited area – for instance, for interpolation
problems. As an example we mention that Hirvonen (1962), investigating the
local covariance function of the free-air anomalies in Ohio, found numerical
values that are well represented by an analytical expression of the form
C0
C(s) = , (9–10)
1 + (s/d)2
9.2 The covariance function 349
where
C0 = 337 mgal2 , d = 40 km . (9–11)
∞
n−1
Cg (P, Q) = A sn+2 Pn (cos ψ) , (9–12)
(n − 2)(n + B)
n=3
where ∆gn (ϑ, λ) is the Laplace surface harmonic of degree n; or, more ex-
plicitly,
∞
n
∆g(ϑ, λ) = anm Rnm (ϑ, λ) + bnm Snm (ϑ, λ) , (9–15)
n=2 m=0
where
Rnm (ϑ, λ) = Pnm (cos ϑ) cos mλ ,
(9–16)
Snm (ϑ, λ) = Pnm (cos ϑ) sin mλ
are the conventional spherical harmonics; or in terms of fully normalized
harmonics (see Sect. 1.10):
∞
n
∆g(ϑ, λ) = ānm R̄nm (ϑ, λ) + b̄nm S̄nm (ϑ, λ) . (9–17)
n=2 m=0
Substituting (9–18) and taking into account the orthogonality relations (1–
83) and the normalization (1–91), we easily find
n
M {∆gn2 } = (ā2nm + b̄2nm ) . (9–21)
m=0
Consider now the average product (9–19) of two Laplace harmonics of differ-
ent degree, n = n. Owing to the orthogonality of the spherical harmonics,
the integral in (9–19) is zero:
In statistical terms this means that two Laplace harmonics of different de-
grees are uncorrelated or, broadly speaking, statistically independent.
In a way similar to that used for the gravity anomalies, we may also
expand the covariance function C(s) into a series of spherical harmonics. Let
us take an arbitrary, but fixed, point P as the pole of this expansion. Thus
spherical polar coordinates ψ (angular distance from P ) and α (azimuth)
are introduced (Fig. 9.2). The angular distance ψ corresponds to the linear
distance s according to (9–8). If we expand the covariance function, with
argument ψ, into a series of spherical harmonics with respect to the pole P
and coordinates ψ and α, we have
∞
n
C(ψ) = cnm Rnm (ψ, α) + dnm Snm (ψ, α) , (9–23)
n=2 m=0
north pole
à = const. ®
P Ã
P'
equ
ator
which is of the same type as (9–15). But since C depends only on the distance
ψ and not on the azimuth α, the spherical harmonics cannot contain any
terms that explicitly depend on α. The only harmonics independent of α are
the zonal functions
Rn0 (ψ, α) ≡ Pn (cos ψ) , (9–24)
so that we are left with
∞
C(ψ) = cn Pn (cos ψ) . (9–25)
n=2
The cn ≡ cn0 are the only coefficients that are not equal to zero. We also
use the equivalent expression in terms of fully normalized harmonics:
∞
C(ψ) = c̄n P̄n (cos ψ) . (9–26)
n=2
The coefficients in these series, according to Sects. 1.9 and 1.10, are given
by
2n + 1 2π π
cn = C(ψ) Pn (cos ψ) sin ψ dψ dα
4π α=0 ψ=0
(9–27)
2n + 1 π
= C(ψ) Pn (cos ψ) sin ψ dψ
2 ψ=0
and
cn
c̄n = √ . (9–28)
2n + 1
We now determine the relation between the coefficients cn of C(ψ) in
(9–25) and the coefficients ānm and b̄nm of ∆g in (9–18). For this purpose
we need an expression for C(ψ) in terms of ∆g, which is easily obtained by
writing (9–27) more explicitly. Take the two points P (ϑ, λ) and P (ϑ , λ ) of
Fig. 9.2. Their spherical distance ψ is given by
Here ψ and the azimuth α are the polar coordinates of P (ϑ , λ ) with respect
to the pole P (ϑ, λ).
The symbol M in (9–6) denotes the average over the unit sphere. Two
steps are required to find it. First, we average over the spherical circle of
radius ψ (denoted in Fig. 9.2 by a broken line), keeping the pole P fixed and
letting P move along the circle so that the distance P P remains constant.
This gives 2π
∗ 1
C = ∆g(ϑ, λ) ∆g(ϑ , λ ) dα , (9–30)
2π α=0
354 9 Statistical methods in physical geodesy
According to (9–22), only the term with n = n is different from zero so that
from (9–21) we finally obtain
n
cn = M {∆gn2 } = (ā2nm + b̄2nm ) . (9–39)
m=0
Hence, cn is the average square of the Laplace harmonic ∆gn (ϑ, λ) of degree
n, or its variance. For these reasons the cn are also called degree variances.
The “degree covariances” are zero because of (9–22).
Equation (9–39) relates the coefficients ānm and b̄nm of ∆g and cn of C(s)
in the simplest possible way. Note that ānm and b̄nm are coefficients of fully
normalized harmonics, whereas cn are coefficients of conventional harmonics.
As a matter of fact, we may also use the anm and bnm (conventional) or
the c̄n (fully normalized); but then (9–39) will obviously become slightly
more complicated. It should be mentioned that the mathematics behind the
statistical description of the gravity anomalies is the theory of stochastic
processes. The gravity anomaly field is treated as a stationary stochastic
process on a sphere; the spherical-harmonic expansions of this section are
nothing but the spectral analysis of that process. A comprehensive treatment
of this topic is found in Moritz (1980 a).
distinction between these two kinds of prediction and the mathematical for-
mulation is the same in both cases.
In order to predict a gravity anomaly at P , we must have information
about the gravity anomaly function. The values observed at certain points
are the most important information. In addition, we need some information
on the form of the anomaly function. If the gravity measurements are very
dense, then the continuity or “smoothness” of the function is sufficient – for
instance, for linear interpolation. Otherwise we may try to use statistical
information on the general structure of the gravity anomalies. Here we must
consider two kinds of statistical correlation: the autocorrelation – the corre-
lation between each other – of gravity anomalies and the correlation of the
gravity anomalies with height.
Correlation with height will for the moment be disregarded; Sect. 9.7
will be devoted to this topic. The autocorrelation is characterized by the
covariance function considered in Sect. 9.2.
Mathematically, the purpose of prediction is to find a function of the
observed gravity anomalies ∆g1 , ∆g2 , . . . , ∆gn in such a way that the un-
known anomaly ∆gP at P is approximated by the function
.
∆gP = F (∆g1 , ∆g2 , . . . , ∆gn ) . (9–40)
The coefficients αP i depend only on the relative position of P and the grav-
ity stations 1, 2, . . . , n; they are independent of the ∆gi . Depending on the
way we choose these coefficients, we obtain different interpolation or extrap-
olation methods. Here are some examples.
Geometric interpolation
The “gravity anomaly surface”, as represented by a gravity anomaly map,
may be approximated by a polyhedron by dividing the area into triangles
whose corners are formed by the gravity stations and passing a plane through
the three corners of each triangle (Fig. 9.3). This is approximately what is
done in constructing the contour lines of a gravity anomaly map by means
of graphical interpolation.
Analytically, this interpolation may be formulated as follows. Let point
P be situated inside a triangle with corners 1, 2, 3 (Fig. 9.3). To each point
9.4 Interpolation and extrapolation of gravity anomalies 357
2
P
1
3
Representation
Often the measured anomaly of a gravity station 1 is made to represent the
whole neighborhood so that
- P ≡ ∆g1
∆g (9–44)
αP 1 = 1 , αP 2 = αP 3 = . . . = αP n = 0 . (9–45)
This method is rather crude but simple and accurate enough for many pur-
poses.
358 9 Statistical methods in physical geodesy
Zero anomaly
If there are no gravity measurements in a large area – for instance, on the
oceans –, then the estimate
-P ≡ 0
∆g (9–46)
is used in this area. In this trivial case all αP i are zero.
If all known gravity stations are far away, and if we know of nothing
better, then this primitive extrapolation method is applied, although the
accuracy is poor. At best, this method may work with isostatic anomalies.
None of these three methods gives optimum accuracy. In the next section we
investigate the accuracy of the general prediction formula (9–41) and find
those coefficients αP i that yield the most accurate results.
By squaring we find
ε2P = ∆gP − αP i ∆gi ∆gP − αP k ∆gk
i k
(9–49)
= ∆gP2 − 2 αP i ∆gP ∆gi + αP i αP k ∆gi ∆gk .
i i k
Let us now form the average M of this formula over the area considered
(either a limited region or the whole earth). Then we have from (9–6),
M {∆gi ∆gk } = C(i k) ≡ Cik ,
M {∆gP2 } = C(0) ≡ C0 .
9.5 Accuracy of prediction methods 359
σP Q = M {εP εQ } . (9–55)
360 9 Statistical methods in physical geodesy
Thus we recognize the basic role of the covariance function in accuracy stud-
ies. The error function, on the other hand, is fundamental for problems of
error propagation.
∂m2P
≡ −2CP i + 2αP k Cik = 0 (i = 1, 2, . . . , n) (9–63)
∂αP i
or
Cik αP k = CP i . (9–64)
This is a system of n linear equations in the n unknowns αP k ; the solution
is
(−1)
αP k = Cik CP i , (9–65)
(−1)
where Cik denote the elements of the inverse of the symmetric matrix
[Cik ].
Substituting (9–65) into (9–41) gives
We see that for optimal prediction we must know the statistical behavior of
the gravity anomalies through the covariance function C(s).
There is a close connection between this optimal prediction method
and the method of least-squares adjustment. Although they refer to some-
what different problems, both are designed to give most accurate results.
The linear equations (9–64) correspond to the “normal equations” of ad-
justment computations. Prediction by means of formula (9–67) is therefore
called “least-squares prediction”. A generalization to heterogeneous data is
“least-squares collocation” to be treated in Chap. 10. In its most general
362 9 Statistical methods in physical geodesy
(−1)
= C0 − Cik CP i CP k .
Thus, the standard error of least-squares prediction is given by
(−1)
m2P = C0 − Cik CP i CP k
⎡ ⎤−1 ⎡ ⎤
C11 C12 . . . C1n CP 1
⎢
⎢ C21 C22 . . . C2n ⎥ ⎢ ⎥
⎥ ⎢ CP 2 ⎥
= C0 − CP 1 , CP 2 , . . . , CP n ⎢ . .. .. ⎥ ⎢ .. ⎥ .
⎣ .. . . ⎦ ⎣ . ⎦
Cn1 Cn2 . . . Cnn CP n
(9–73)
9.7 Correlation with height 363
In the same way we find the error covariance in the points P and Q:
(−1)
σP Q = CP Q − Cik CP i CQk
⎡ ⎤−1 ⎡ ⎤
C11 C12 . . . C1n CQ1
⎢ C
⎢ 21 C22 . . . C2n ⎥ ⎢
⎥ ⎢ CQ2 ⎥
⎥
= CP Q − CP 1 , CP 2 , . . . , CP n ⎢ . .. .. ⎥ ⎢ .. ⎥ .
⎣ .. . . ⎦ ⎣ . ⎦
Cn1 Cn2 . . . Cnn CQn
(9–74)
These two formulas give the error covariance function for least-squares pre-
diction. Both formulas have a form similar to that of (9–67) and are equally
well suited for computations so that ∆g - and its accuracy can be calculated
at the same time.
It is clear that, after appropriate slight changes, this theory applies au-
tomatically to gravity disturbances δg.
Practical considerations
Geometric interpolation (Sect. 9.4) is suited for the interpolation of point
anomalies in a dense gravity net, with station distances of 10 km or less. If
mean anomalies for blocks of 5 × 5 or larger are needed rather than point
anomalies, then some kind of representation, such as that considered in the
previous section, may be simpler and hardly less accurate.
Least-squares prediction is, by its very definition, more accurate than
either geometric interpolation or representation, but the improvement in ac-
curacy is not striking. The main advantage of least-squares prediction is
that it permits a systematic, purely numerical, digital processing of gravity
data; gravity anomalies are stored in data bases, and gravity anomaly maps,
if necessary, are generated automatically. The same formula applies to both
interpolation and extrapolation so that gaps in the gravity data make no dif-
ference in the method of computation, which becomes completely schematic
(Moritz 1963). For practical and computational details see Rapp (1964) and
many other papers published since.
For larger station distances, of 50 km or more, prediction of individual
point values becomes meaningless. In this case we must work with mean
anomalies of, say, 1◦ × 1◦ blocks.
h [m] h [m]
1500 1500
1000 1000
500 500
®
Dg [mgal] Dg [mgal]
0 0
–150 –100 –50 0 +50 –50 0 +50 +100 +150
which is important in many cases. Therefore our formulas were valid only
for gravity anomalies uncorrelated with height, such as isostatic or, to a
certain extent, Bouguer anomalies; or for free-air anomalies in moderately
flat areas. Free-air anomalies in mountains must be treated differently.
Figure 9.4 due to U.A. Uotila shows the correlation of free-air anomalies
with height. The gravity anomalies ∆g are plotted against the height h. If
there were an exact functional dependence between ∆g and h, then all points
would lie on a straight line (or, more generally, on a curve). In reality, there
is only an approximate functional relation, a general trend or tendency of
the free-air anomalies to increase linearly with height; exceptions, even large
ones, are possible. This shows very well the meaning of correlation.
We have characterized the mutual correlation of the gravity anomalies
by the “autocovariance function” (9–6),
∆h = h − M {h} , (9–78)
9.7 Correlation with height 365
where the symbol M {h} denotes the mean height of the whole area consid-
ered.
If ∆g and ∆h are not correlated, then the function B(s) is identically
zero. If this is not the case, then we should also take the height into account
in our interpolation.
It is easy to extend the prediction formula (9–41) for this purpose, but
this has turned out to be of little practical importance.
z = ∆g − b ∆h , (9–79)
b = 2π G . (9–80)
Let us form the covariance function Z(s) between the “Bouguer anomaly” z
of (9–79) and height difference ∆h
elevation ®
Bouguer anomaly ®
% = 2.2 g/cm3
% = 2.4 g/cm3
% = 2.6 g/cm3
It may be shown that this is equivalent to the condition that the points of
Fig. 9.4 lie approximately on a straight line. The coefficient b is then given
by
b = tan α (9–85)
as the inclination of the line towards the h-axis.
In practice these conditions are very often fulfilled to a good approx-
imation. Furthermore, by computing b from Eq. (9–84) or determining it
graphically by means of (9–85), we often get a value that is close to the
normal Bouguer gradient (9–81).
If we assume that b depends only on the rock density , then we obtain a
means for determining the average density, which is often difficult to measure
directly. This is the “Nettleton method”, used in geophysical prospecting: the
coefficient b is found statistically by means of Eqs. (9–84) or (9–85), and
the rock density is then computed from (9–80). Figure 9.5 illustrates the
principle of this method; see also Jung (1956: p. 600).
If the condition (9–83) is fulfilled, then we may consider the “Bouguer
anomaly” z as a gravity anomaly that is completely uncorrelated with height;
we can directly apply to it the whole theory of the preceding sections. But
even when this condition is not quite satisfied, Bouguer anomalies will in
general be far less correlated with height than free-air anomalies. The fact
that in (9–79) gravity is reduced to a mean height and not to sea level,
is quite irrelevant in this connection because this is only a question of an
9.7 Correlation with height 367
Interpolation
Let errorless values of T be given at q spatial points P1 , P2 , . . . , Pq ; these
points may lie on the earth’s surface or in space above the earth’s surface.
We put
T (Pi ) = fi , i = 1, 2, . . . , q (10–4)
370 10 Least-squares collocation
or in matrix notation
Ab = f . (10–7)
If the square matrix A is regular, then the coefficients bk are uniquely de-
termined by
b = A−1 f . (10–8)
This model is suitable, for instance, for a determination of the geoid by
satellite altimetry, since this method, rather directly, yields geoidal heights
Ni and hence, by Bruns’ theorem (2–236), T (Pi ) = γi Ni . For the astro-
geodetic geoid determination, we must generalize this model, which leads us
to collocation.
Collocation
Here we wish to reproduce, by means of the approximation (10–2), q mea-
sured values which again are assumed to be errorless (this assumption is
not essential and will be dropped later). These measured values are assumed
to be linear functionals L1 T, L2 T, . . . , Lq T of the anomalous potential T .
“Linear functional” means nothing else than a quantity LT that depends
linearly on T but need not be an ordinary function but may, say, also con-
tain a differentiation or an integral; essentially, it is the same as a “linear
operator”.
In fact, deflections of the vertical,
1 ∂T 1 ∂T
ξ=− , η=− , (10–9)
γ ∂x γ ∂y
but also gravity anomalies,
∂T 2
∆g = − − T, (10–10)
∂z R
and gravity disturbances
∂T
δg = − (10–11)
∂z
10.1 Principles of least-squares collocation 371
1 ∂
Li = (10–12)
γ ∂x
q
Bik bk = i with Bik = Li ϕk , (10–14)
i=1
Least-squares interpolation
Let us consider a function
K = K(P, Q) , (10–16)
in which two points P and Q are the independent variables. Let this function
K be
• symmetric with respect to P and Q,
• harmonic with respect to both points, everywhere outside a certain
sphere, and
• positive-definite (the positive definitiveness of a function is defined
similarly as in the case of a matrix).
Then the function K(P, Q) is called a (harmonic) kernel function (Moritz
1980 a: p. 205). A kernel function K(P, Q) may serve as “building material”
from which we can construct base functions. Taking for the base functions
the form
ϕk (P ) = K(P, Pk ) , (10–17)
where P denotes the variable point and Pk is a fixed point in space, we obtain
least-squares interpolation already treated by a quite different approach in
Chap. 9.
This name originates from the statistical interpretation of the kernel
function as a covariance function (Sect. 9.2); then least-squares interpola-
tion has some minimum properties (least-error variance, similarly as in least-
squares adjustment). This interpretation is not essential, however; one may
also work with arbitrary analytical kernel functions, considering the proce-
dure as a purely analytical mathematical approximation technique. Normally
one tries to combine both aspects in a reasonable way.
Substituting (10–17) into (10–6), we get
Aik = K(Pi , Pk ) = Cik ; (10–18)
this square matrix now is symmetric (in the general case, Aik is not sym-
metric!) and positive definite because of the corresponding properties of the
function K(P, Q). Then the coefficients bk follow from (10–8) and may be
substituted into (10–2). With the notation
ϕk (P ) = K(P, Pk ) = CP k , (10–19)
the result may be written in the form
⎡ ⎤−1 ⎡ ⎤
C11 C12 . . . C1q f1
⎢ C
⎢ 21 C 22 . . . C ⎥
2q ⎥
⎢ f
⎢ 2⎥
⎥
f (P ) = CP 1 CP 2 . . . CP q ⎢ .
.. ⎦ ⎣ .. ⎥
⎥ ⎢
. . . , (10–20)
⎣ .. .. ⎦
Cq1 Cq2 . . . Cqq fq
10.1 Principles of least-squares collocation 373
Least-squares collocation
Here we again derive the base functions from a kernel function K(P, Q), but
in a way slightly different from (10–17): we put
ϕk (P ) = LQ
k K(P, Q) , (10–21)
Bik = LPi LQ
k K(P, Q) = Cik , (10–22)
which gives a matrix which again is symmetric. Solving (10–14) for bk and
substituting into (10–2) gives with
ϕk (P ) = LQ
k K(P, Q) = CP k (10–23)
the formula
⎡ ⎤−1 ⎡ ⎤
C11 C12 . . . C1q 1
⎢
⎢ C21 C 22 . . . C 2q
⎥ ⎢ 2 ⎥
⎥ ⎢ ⎥
f (P ) = CP 1 CP 2 . . . CP q ⎢ . .. .. ⎥ ⎢ .. ⎥ . (10–24)
⎣ .. . . ⎦ ⎣ . ⎦
Cq1 Cq2 . . . Cqq q
cov(NP , NQ ) = LPi LQ
j cov(∆gP , ∆gQ ) , (10–27)
where Ni denotes the geoidal height at point i and ∆g is the gravity anomaly
at point P , and L denotes the Stokes formula. Explicit expressions are found
in Moritz (1980 a: Sect. 15).
In this statistical interpretation, we take the kernel function K(P, Q)
as the covariance function C(P, Q). Then f (P ) is an optimal estimate (in
the sense of least variance) for the anomalous potential T and hence for
the height anomaly ζ = T /γ, on the basis of arbitrary measurement data.
For geoid determination in mountainous areas, relevant terrestrial measure-
ment data primarily are ξ, η, and ∆g. The covariances Cik and CP i are
given by known analytical expressions, see Tscherning and Rapp (1974) or
Moritz (1980 a: Sect. 15). A general computer program for collocation is
described in Sünkel (1980).
Least-squares collocation may easily be generalized to observational data
affected by random errors; systematic effects may also be taken into consid-
eration. In addition to the estimated quantities (f in our present case) we
may also compute their standard error by a formula similar to (10–24). A
comprehensive presentation of a least-squares collocation may be found in
Moritz (1980 a). You cannot learn collocation from this slight chapter only!
(Moritz 1980 a: Sect. 23, Eq. (32-1)). The point P (r, θ, λ) is the computation
point, and Q(r , θ , λ ) is a current data point; ψ is the spherical distance
between (θ, λ) and (θ , λ ), and R is the mean radius of the earth. The de-
pendence on r is given by the factor
r −(n+2) (10–30)
10.2 Application of collocation to geoid determination 375
not only the harmonic character of the anomalous potential T outside the
earth’s surface is preserved, but in addition, the computational removal of
the topographic masses above sea level makes the function T harmonic down
to sea level. Hence, the collocation formula (10–24) can be applied also at
sea level, giving cogeoid heights N c . By applying the inverse reduction (the
indirect effect) to the computed height anomalies ζ c and cogeoid heights N c ,
we get actual ζ and N . It can be expected that errors in the isostatic model
used (e.g., an Airy–Heiskanen model) will largely cancel in this combined
procedure of reduction and “anti-reduction” (remove-restore technique; see
Sect. 11.1).
The procedure is theoretically optimal and practically well suited for
computer use. The integrability conditions, which in Helmert integration
are represented by the closures of the individual triangles (see Sect. 5.14),
are automatically taken into account. The fact that the deflections of the
vertical are given only in a certain region has the effect that the geoid can
only be computed in that region. Since, even by collocation, differences in
geoidal heights between two neighboring stations A and B depend essentially
only on the deflections in those two stations, the lack of data outside the
region under consideration will hardly cause a noticeable distortion. Note,
however, that the addition of a constant to all geoidal heights N will not
affect the deflections of the vertical; hence, astrogeodetic data determine
the geoidal heights only up to an additive constant. This constant may be
chosen such that the average value of the computed N is zero, and the result
of collocation comes near to this case.
To get immediately almost geocentric geoidal heights, it is appropriate to
take into consideration a global trend which mainly affects ζ and N itself, by
subtracting the effect of a suitable global gravity field, e.g., the gravity earth
model given as a spherical-harmonic expansion up to degree 180◦ × 180◦ of
Rapp (1981), say, following Sünkel (1983). This will be described in the next
section; in the present section we limit ourselves to the isostatic reduction.
Computational procedure
The computational procedure consists of the following steps:
1. Transformation of the astrogeodetic surface deflections ξ, η from the
local datum used for the geocentric Geodetic Reference System 1980
by the well-known differential formulas of Vening Meinesz (see Heiska-
nen and Moritz 1967: Eq. (5-59)). This is necessary since collocation
requires a reference system which is as realistic as possible.
ξ̄, η̄ by (8–136).
6. By applying the indirect effect (10–2) and (8–153), we get actual height
anomalies ζ and geoidal heights N .
11 Computational methods
T = L() (11–1)
or
output = L(input) , (11–2)
where L denotes the linear operation of least-squares collocation (not to be
confused with a linear functional L as used, e.g., in Eq. (10–13)).
In Sect. 8.9 we have introduced gravity reduction from the point of view
of the modern theory. To repeat, immediately specializing to topographic-
isostatic reduction, we have
and
δT = L(). (11–7)
We have only slightly generalized from ∆g to .
Now we proceed an important step further. The remove-restore principle
has only two requirements:
Thus, we use simultaneously the earth model EM for the longer wave-
lengths and the topographic-isostatic geological model TI for the shorter
wavelengths. Since the spherical-harmonic expansions are generalizations,
for the sphere, of Fourier series for the circle, we can speak of wavelengths.
Denoting the maximum degree of the spherical-harmonic expansion with N ,
this can be associated with a shortest resolvable wavelength λ according to
2π 360◦
λ= = . (11–8)
N N
For an expansion to degree N = 180 (say), we have λ = 360◦ /180 = 2◦ ,
which roughly corresponds to 200 km on a meridian or on the equator. In
many cases, the half wavelength λ/2 is considered (see Seeber 2003: p. 469).
Since EM (approximately) takes care of the long waves up to a certain
maximum degree N , it is resonable to represent the remaining short waves
from N to infinity. This sequence N + 1, N + 2, . . . , ∞ will be denoted by
CN , where CN is the abbreviation of the “complement” of the sequence
from 2 to N .
Thus, we may write for the residuals
N − T CN ,
δT = T − TEM TI
(11–9)
δ = − N CN
EM − TI .
Remark
As we have noted at the beginning of Sect. 10.2, the remove-restore process
aims at removing all known major trends:
• the local topography produces Bouguer anomalies,
• the regional features (i.e., their isostatic compensation), in addition to
the Bouguer effect, produce topographic-isostatic anomalies,
• the global irregularities are expressed by an earth model and lead to
what is modestly called the “residual anomalies”.
It is clear that what is “removed” before the computation, must be fully
“restored” after the computation.
The pioneering work has been done by Sünkel (1983). Later work, espe-
cially by Sünkel et al. (1987), Kühtreiber (1998, 2002 a, 2002 b), and Erker
et al. (2003) has refined, extended and perfected the gravity field in Austria,
but the 1983 work is good for an introduction.
Sünkel (1983) used least-squares collocation to calculate the geoid for the
main part of Austria from a very good material of deflections of the vertical.
Gravity anomalies of a comparable quality were not yet available in 1983. In
addition to an isostatic reduction (Sect. 8.14) according to Airy–Heiskanen
(T = 30 km), he also removed a global trend by means of an earth gravity
model, represented by a spherical-harmonic expansion up to a certain degree
N . In particular, he used the model of Rapp (1981) with N = 180.
After removing the topographic-isostatic trend TTI and this global trend
N (remember, EM denotes earth model), there remains a residual anoma-
TEM
lous potential δT , given by
N N
δT = T − TTI − TEM + TTI . (11–10)
Data
The topography in Austria is rather varied, with elevations up to 3800 m.
The density of astrogeodetic stations was 10 to 20 km; the total number of
deflections data used was 521. No gravity anomalies were used in this first
computation.
The topographic-isostatic reduction of the deflections of the vertical was
made using a rather crude digital terrain model consisting of mean elevations
for 20 × 20 rectangles. It has been obtained by digitizing a map 1 : 500 000.
11.2 Geoid in Austria by collocation 383
The standard error of this model is on the order of 100 m. Investigations have
shown that, in spite of its poor accuracy, the model is reasonably adequate for
reduction of deflections of the vertical; it is, however, totally inadequate for
gravity! In fact, the reduction error for ξ, η is approximately proportional
to the terrain inclination; it is thus very small if the deflection station is
situated in an area of inclination zero. This is the case not only if the station
lies in a horizontal plane but also if it lies on the top of a mountain, as most
deflection stations do.
Results
It turned out that almost all of the signal (T, N, ζ) comes from the topo-
graphic-isostatic model and the N = 180 gravity model used. This part,
TI + EM, lies between 41.5 m and 49.5 m. The contribution of collocation
(γ −1 T ) lies between −0.5 m and 1.5 m, after removal of a pronounced trend
on the order of 3 m.
The efficiency of topographic-isostatic reduction can also be seen from
the fact that it has reduced the variance of the deflections of the vertical in
384 11 Computational methods
Austria (the square of the average size of ξ and η) from 30 (arc second)2 to
5 (arc second)2 .
So we may say that we can determine the Austrian geoid to 1–2 m without
measurements (deflections of the vertical) and without collocation, knowing
only a topographic map! This is even more surprising since Austria is not
particularly well isostatically compensated.
Of considerable interest is the effect of analytical continuation on the
isostatically (plus earth model) reduced anomalous potential TTI . It is ex-
pressed by the difference γ −1 T at the earth’s surface minus γ −1 T at sea
level. This difference reaches a maximum of 13 cm in the Central Alps and is
otherwise positive and negative. In the terminology of the present book, this
is the separation between the real geoid and the harmonic geoid (Sect. 8.15).
Of the same interest is the difference between the height anomalies ζ
(= γ −1 T at the earth’s surface) and the geoidal heights N (= γ −1 T at sea
level). The maximum of 35 cm for ζ − N is reached at the Grossglockner
mountain (the highest peak in Austria, H = 3797 m). The results are in
excellent agreement with the approximate formula
where ∆gB is the Bouguer anomaly in gal and H is the elevation in the
same units as ζ and N . The agreement may easily be verified, since the
Bouguer anomalies in the investigated area range from 10 mgal to −170 mgal,
corresponding to topographic heights from 200 m to 3000 m (Sünkel 1983:
p. 140). In Sünkel et al. (1987: p. 69), the differences ζ − N for the whole of
Austria range between −2 cm and +56 cm.
All this has been computed only from the measured deflections of the
vertical. Gravity observations have been included by Kühtreiber (2002 a,
2002 b) and Erker et al. (2003), leading to what might be a “few-centimeter
geoid”.
Important: the astrogeodetic geoid and the gravimetric geoid are com-
pared and finally combined after systematic trends have been eliminated by
Kühtreiber (2002 b) and Erker et al. (2003).
ζ = ζ0 + ζ1 + ζ2 + ζ3 + · · · , (11–14)
11.3 Molodensky corrections 385
R
ζi = gi S(ψ) dσ , (11–15)
4π γ
σ
∗
∆g = ∆g + g1 + g2 + g3 + · · · . (11–16)
The correction terms gn are evaluated recursively by
n
gn = − z r Lr (gn−r ) , (11–17)
r=1
starting from
g0 = ∆g . (11–18)
Here the operator Ln is also defined recursively:
starting with
L1 = L (11–20)
with the gradient operator L defined by the integral (8–60), that is,
R2 f − fQ
L(f ) = dσ . (11–21)
2π l03
σ
conclusions:
1. The method of Molodensky corrections depends very much on the de-
tails of numerical integration (data density, smoothing, etc.).
4. At the end of Sects. 2.21 and 8.8, we have remarked a curious phe-
nomenon. Using the same data, gravimetric methods seem to furnish
the vertical position (expressed by ζ or N ) roughly by one order of
magnitude better than the horizontal position (as expressed by ξ, η).
If we take the old astronomer’s rule that 1 ∼
= 30 m in position, then
1 m corresponds to 0.03 . Assume that we get 1 m in vertical position
and wish to get the same accuracy for horizontal position. This would
mean that we have to get the astronomical measurements Φ, Λ and
the deflections of the vertical ξ, η with better than 0.03 . This also
seems to apply with the order of magnitude of the Molodensky correc-
tions, where a Molodensky correction ζ1 = 0.41 m comes along with a
ξ1 = η1 = 2 , which corresponds to 60 m.
In this sense, gravimetry is weaker by one order of magnitude in deter-
mining the horizontal than the vertical position. This is an admittedly
388 11 Computational methods
Final remark
The computation of Molodensky reductions is heavy work. So in mountain-
ous areas, least-squares collocation is definitely preferable to integration,
except for certain test computations (Sideris 1987, 1990).
Collocation also permits comparison and combination of astrogeodetic
and gravimetric data; a key paper is Kühtreiber (2002 b).
All this, however, builds on the fundamental ideas of M.S. Molodensky.
In his landmark publication (Krarup 1969) one clearly sees the transition
from Molodensky’s problem to least-squares collocation.
Fontana RD, Cheung W, Stansell T (2001): The modernized L2 civil signal. GPS
World, 12(9): 28–34.
Forsberg R, Tscherning CC (1981): The use of height data in gravity field approxi-
mation by collocation. Journal of Geophysical Research, 86 (B9): 7843–7854.
Forsberg R, Tscherning CC (1997): Topographic effects in gravity field modelling
for BVP. Available at www.gfy.ku.dk/∼cct/comored.htm.
Frank P, Mises R von (eds) (1930): Die Differential- und Integralgleichungen der
Mechanik und Physik, 2nd edition, part 1: Mathematischer Teil. Vieweg,
Braunschweig (reprint 1961 by Dover, New York and Vieweg, Braunschweig).
Galle A (1914): Das Geoid im Harz. Veröffentlichung des Geodätischen Instituts
Potsdam, vol 61.
Grafarend E, Offermanns G (1975): Eine Lotabweichungskarte Westdeutschlands
nach einem geodätisch konsistenten Kolmogorov-Wiener-Modell. Deutsche
Geodätische Kommission bei der Bayerischen Akademie der Wissenschaften,
Reihe A: Theoretische Geodäsie, vol 82.
Groten E (2004): Fundamental parameters and current (2004) best estimates of the
parameters of common relevance to astronomy, geodesy, and geodynamics.
Available at www.gfy.ku.dk/∼iag/HB2004/part5/51-groten.pdf.
Gurtner W (1978): Das Geoid in der Schweiz. Institut für Geodäsie und Photogram-
metrie der ETH Zürich, vol 20.
Gurtner W, Elmiger H (1983): Computation of geoidal heights and vertical deflec-
tions in Switzerland. Presented at the XVIII General Assembly of the IUGG
at Hamburg, August 15–27.
Heiskanen W (1924): Untersuchungen über Schwerkraft und Isostasie. Finnish Geo-
detic Institute, Helsinki, vol 4.
Heiskanen W (1928): Ist die Erde ein dreiachsiges Ellipsoid? Gerlands Beiträge zur
Geophysik, 19: 356–377.
Heiskanen WA (1957): The Columbus geoid. EOS, Transactions, American Geo-
physical Union, 38: 841–848.
Heiskanen WA, Moritz H (1967): Physical geodesy. Freeman, San Francisco London.
Heiskanen WA, Vening Meinesz FA (1958): The earth and its gravity field. McGraw-
Hill, New York.
Helmert FR (1880): Die mathematischen und physikalischen Theorien der höheren
Geodäsie, part 1. Teubner, Leipzig (reprint 1962).
Helmert FR (1884): Die mathematischen und physikalischen Theorien der Höheren
Geodäsie, part 2. Teubner, Leipzig (reprint 1962).
Hirvonen RA (1960): New theory of the gravimetric geodesy. Publications of the
Isostatic Institute of the International Association of Geodesy, Helsinki, vol
32.
Hirvonen RA (1961): The reformation of geodesy. Journal of Geophysical Research,
66: 1471–1478.
Hirvonen RA (1962): On the statistical analysis of gravity anomalies. Publications
of the Isostatic Institute of the International Association of Geodesy, Helsinki,
vol 37.
Hofmann-Wellenhof B, Kienast G, Lichtenegger H (1994): GPS in der Praxis. Sprin-
ger, Wien New York.
References 391
Molodenskii MS, Eremeev VF, Yurkina MI (1962): Methods for study of the external
gravity field and figure of the earth. Israel Program of Scientific Translations,
Jerusalem (Russian original 1960).
Montenbruck O, Gill E (2001): Satellite orbits – models, methods, and applications,
corrected 2nd printing. Springer, Berlin.
Moritz H (1962): Studies on the accuracy of the computation of gravity in high ele-
vations. Publications of the Isostatic Institute of the International Association
of Geodesy, Helsinki, vol 38.
Moritz H (1963): Interpolation and prediction of point gravity anomalies. Publica-
tions of the Isostatic Institute of the International Association of Geodesy,
Helsinki, vol 40.
Moritz H (1964): Zur Bestimmung des Geoides und seiner Verwendung als Reduk-
tionsfläche. Zeitschrift für Vermessungswesen, 89: 200–202.
Moritz H (1965): Schwerevorhersage und Ausgleichungsrechnung. Zeitschrift für
Vermessungswesen, 90: 181–184.
Moritz H (1980 a): Advanced physical geodesy. Wichmann, Karlsruhe (reprint 2001
by Civil and Environmental Engineering and Geodetic Science, Ohio State
University, Columbus, Ohio).
Moritz H (1980 b): Geodetic Reference System 1980. Bulletin Géodésique, 54: 395–
405.
Moritz H (1990): The figure of the earth – theoretical geodesy and the earth’s
interior. Wichmann, Karlsruhe.
Moritz H (1995): Science, mind and the universe – an introduction to natural phi-
losophy. Wichmann, Heidlberg.
Moritz H, Hofmann-Wellenhof B (1993): Geometry, relativity, geodesy. Wichmann,
Karlsruhe.
Moritz H, Mueller II (1987): Earth rotation – theory and observation. Ungar, New
York.
Moritz H, Yurkina MI (eds) (2000): M.S. Molodensky – in memoriam. Mitteilungen
der geodätischen Institute der Technischen Universität Graz, vol 88.
Mueller II (1985): Reference coordinate systems and frames: concepts and realiza-
tion. Bulletin Géodésique, 59: 181–188.
Müller J (2001): Die Satellitengradiometriemission GOCE – Theorie, technische Re-
alisierung und wissenschaftliche Nutzung. Deutsche Geodätische Kommission
bei der Bayerischen Akademie der Wissenschaften, Reihe C, vol 541.
National Imagery and Mapping Agency (2000): Department of Defense World Geo-
detic System 1984 – its definition and relationships with local geodetic sys-
tems, 3rd edition, amendment 1. NIMA Technical Report TR 8350.2, Bethes-
da, Maryland. Available as PDF file at www.nima.mil.
Neumann F (1887): Vorlesungen über die Theorie des Potentials und der Kugel-
funktionen (edited by C. Neumann). Teubner, Leipzig.
Österreichische Kommission für die Internationale Erdmessung (ed) (1983): Das
Geoid in Österreich. Geodätische Arbeiten Österreichs für die Internationale
Erdmessung. Neue Folge, vol III.
Pail R (2003): Satellitengeodäsie. Lecture manuscript available at the Institute of
Navigation and Satellite Geodesy of the Graz University of Technology.
394 References
correlation 346, 347, 350, 356, 359, ellipsoid, see best-fitting –, gravity field
363–366, 375 of –, mean earth –
correlation length 350 ellipsoidal coordinates 1, 37, 41, 55, 92,
correlation with elevation 346, 375 107, 119, 122–124, 128, 175, 176, 192,
correlation with height 363 193, 195, 197, 199, 202–206, 208,
covariance 342, 343, 346–350, 352, 354, 214, 217, 224, 275, 296, 299, 315, 316
356, 359–361, 363, 365, 372–374, 380 ellipsoidal harmonic coordinates 10,
covariance function 342, 346–350, 352, 34, 197
354, 356, 359–361, 363, 365, 372, 374 ellipsoidal harmonics 3, 19, 34, 37, 73,
covariance propagation 350, 373, 380 76
curvature of level surfaces 48, 53 ellipsoidal height 72, 81, 134, 167, 171,
curvature of normal plumb line 233 192, 193, 202, 203, 214, 217, 275, 276,
curvature of plumb line 48, 228 284, 296–298, 300, 301
curvature parameter 350 ellipsoidal normal 91, 92, 192, 194, 218,
297, 332–335
Eötvös unit 350
Datum, see geodetic –, local –, local
equipotential ellipsoid 84, 239, 240, 336
astrogeodetic –, three-dimensional
equipotential surface 1, 2, 46, 47, 65,
transformation of –
66, 70, 284, 300, 341, 343
datum transformations 203
error covariance function 360, 363
decomposition formula 22, 25, 26, 58 error propagation 345, 373
deflection of the vertical 91–93, 116, Euler angles 174
117, 122, 127, 141, 151, 176, 218, 221, expansion theorem 21
223, 224, 226, 234, 246, 293, 308, extrapolation of gravity anomalies 355
314, 328, 334, 336–338, 339
density 6, 7, 31
First boundary-value problem 27, 28,
DGPS, see differential GPS
31, 294
differential formulas for datum fixed boundary-value problem 295
transformations 203 flattening 1, 77, 78, 80, 86, 88, 89, 96,
differential GPS 184 97, 104, 114, 199, 205, 225, 249–251,
Dirichlet’s problem 27–29, 67 257, 261, 266, 267, 301
discontinuity 7 forbidden spherical harmonics 63
disturbing potential 91, 93, 96, 215, free-air anomaly 154, 155, 250, 292,
236, 239, 243, 248, 249, 261, 274, 290, 293, 307, 320, 322, 326, 348, 385
299, 310, 312, 319 free-air reduction 134–136, 139, 140,
Doppler 87, 182, 183, 256, 271, 331 146, 148, 150, 151, 154, 250, 293, 298,
downward continuation 253, 311–313, 322–324, 338, 344
321, 322, 342, 343, 375 free boundary-value problem 295
dynamic correction 161, 164, 165, 169 fully normalized spherical harmonics
dynamic height 159, 160, 169, 170 23, 24, 59
function, see analytic –, harmonic –
Earth’s crust 2, 129, 144, 146, 149, 289 functional 373
earth’s rotation 1, 2, 43, 272 fundamental equation of physical
eccentricity 34, 40, 66, 71, 74, 76, 78, geodesy 95, 337, 371
80, 86, 89, 197, 258, 265, 266, 280,
282 Gal 45, 84, 139, 160, 163, 169, 227,
Einstein summation convention 362 327, 384
Subject index 399
GRS 1980, see Geodetic Reference isostatic reduction 141, 154, 292, 293,
System 1980 318, 321, 329, 340, 375, 376, 382, 386
ITRF, see International Terrestrial
Harmonic continuation 305, 306, 309, Reference Frame
322, 323, 344
harmonic function 7, 8, 12, 28, 30, 32, Kepler’s laws 258
56, 61, 97, 99, 100, 247, 248, 290, kernel function 372–374
309, 310, 312, 369 Koch’s formula 93, 115, 116
harmonics 61, see ellipsoidal –, Krarup–Runge theorem 310
spherical –, tesseral –, zonal – Krassowsky’s ellipsoid 84
height 157, see comparison of different
– systems, dynamic –, ellipsoidal –, Laplace’s equation 7–9, 10–12, 14, 31,
Helmert –, normal –, orthometric – 37, 39, 41, 95, 102, 121, 218, 224,
height above sea level 160, 318, 321, 290, 294, 304, 309, 342, 375
343 Laplace’s equation in
Helmert height 163, 166 ellipsoidal-harmonic coordinates 34,
Helmert reduction 153 37
Helmert’s formula 226 Laplace’s equation in spherical
Helmert’s projection 192 coordinates 9
Laplace spherical harmonics 22, 302
IERS 85–88, 219 latitude, see astronomical –, geocentric
IERS reference pole 86, 87 –, reduced –
least-squares collocation 265, 274, 309,
indirect effect 149–151, 153, 154,
311, 323, 329, 331, 341, 362, 369,
291–294, 296, 320, 321, 329, 340, 376,
374, 375, 379, 382, 388
377
least-squares interpolation 372, 375
inertial mass 45
least-squares prediction 361
initial point 157, 199, 224, 225
Legendre, see associated – functions
inner zone effects 125
Legendre’s differential equation 13, 14,
International Association of Geodesy 19
83, 388 Legendre’s functions 13, 14
International Ellipsoid 83, 84, 235 Legendre’s functions of the second kind
International Geoid Service 388 19
International Gravity Bureau 388 Legendre’s polynomials 15, 16, 20, 26
International Terrestrial Reference level ellipsoid 65, 67, 70, 75
Frame 87, 88 leveling 48, 157–160, 163, 164, 169–172,
International Union of Geodesy and 175, 216, 226, 228, 229, 232, 233, 284,
Geophysics 84, 310 295, 296, 298, 318, 331, 336, 337
interpolation of gravity anomalies 355 leveling, see astronomical –, GPS –
inverse problem 31, 253, 323, 342, 343 level surface 46–48, 53, 54, 100, 101,
inversion reduction of Rudzki, see 120, 157, 169, 216, 228, 291, 333, 341
Rudzki reduction linear approximation 308, 309, 313,
isostasy 141, see Airy–Heiskanen –, 318, 329
Vening Meinesz regional system, linear eccentricity 34, 66, 86, 89
Pratt–Hayford model linearization 189, 191, 200, 296, 297,
isostatic anomalies 321 300, 307
isostatic compensation 234, 235, 381 local astrogeodetic datum 224
Subject index 401