Tidal Deformation of the Solid Earth
Tidal Deformation of the Solid Earth
Tidal Deformation of the Solid Earth
Transform your PDFs into Flipbooks and boost your revenue!
Leverage SEO-optimized Flipbooks, powerful backlinks, and multimedia content to professionally showcase your products and significantly increase your reach.
F A C U L T Y O F S C I E N C E<br />
U N I V E R S I T Y O F C O P E N H A G E N<br />
Master Thesis in Geophysics, March 16, 2009<br />
<strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
- A Finite Difference Discretization<br />
STINE KILDEGAARD POULSEN<br />
Niels Bohr Institute, University <strong>of</strong> Copenhagen<br />
SUPERVISORS:<br />
Klaus Mosegaard<br />
Carl Christian Tscherning
Abstract<br />
<strong>Tidal</strong> forces are a source <strong>of</strong> noise in many geophysical field observations, but <strong>the</strong>y are<br />
useful in revealing information <strong>of</strong> <strong>the</strong> interior <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. <strong>Tidal</strong> forces arise because<br />
<strong>of</strong> <strong>the</strong> gravitational attraction from external bodies varying over <strong>the</strong> volume <strong>of</strong> a body.<br />
The <strong>Earth</strong>’s response to <strong>the</strong> tide generating potential can be derived analytically by <strong>the</strong><br />
driving force, i.e. <strong>the</strong> knowledge <strong>of</strong> <strong>the</strong> orbital motion <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and <strong>the</strong> Moon, <strong>the</strong><br />
Sun and o<strong>the</strong>r external objects, combined with <strong>the</strong> Love numbers, which are related to<br />
<strong>the</strong> rheological properties <strong>of</strong> <strong>the</strong> <strong>Earth</strong>.<br />
By using a two dimensional steady state version <strong>of</strong> Navier’s equation <strong>of</strong> motion,<br />
<strong>the</strong> tidal deformation <strong>of</strong> <strong>the</strong> <strong>Earth</strong> is solved numerically. The Navier equation <strong>of</strong> motions<br />
is discretized in polar coordinates using a finite difference approximation. The<br />
linear system <strong>of</strong> equations proved to be ill-conditioned, and to subdue <strong>the</strong> obstacle a<br />
Tikhonov regularization is applied.<br />
Assuming <strong>the</strong> Moon being <strong>the</strong> only attracting body, and assuming a stationary<br />
<strong>Earth</strong>-Moon relationship, two <strong>Earth</strong> models are developed. The first model, <strong>the</strong> homogeneous<br />
<strong>Earth</strong> model, is assumed to be a pure Poisson solid, where <strong>the</strong> seismic<br />
velocities and <strong>the</strong> elastic parameters are constant throughout <strong>the</strong> <strong>Earth</strong>. The second<br />
model, <strong>the</strong> layered model, includes four layers representing an inner core, an outer<br />
core, a mantle and a crust. The elastic properties are taken from <strong>the</strong> isotropic Preliminary<br />
Reference <strong>Earth</strong> Model, [Dziewonski and Anderson, 1981], and averaged in each<br />
layer. This gives a realistic <strong>Earth</strong> model with a fluid outer core.<br />
The <strong>Earth</strong> models response to <strong>the</strong> tides are given by <strong>the</strong> radial and tangential displacement<br />
fields. Though <strong>the</strong> compression <strong>of</strong> <strong>the</strong> radial displacement field not fully<br />
fulfill <strong>the</strong> picture <strong>of</strong> <strong>the</strong> inner structures <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, <strong>the</strong> results clearly reproduce <strong>the</strong><br />
physical picture <strong>of</strong> <strong>the</strong> solid <strong>Earth</strong> tides, by means <strong>of</strong> an expansive and compressive<br />
<strong>Earth</strong> for <strong>the</strong> radial field and a tractive displacement for <strong>the</strong> tangential field. The radial<br />
displacement field has a maximum expansion <strong>of</strong> 131 mm and 74 mm for <strong>the</strong> homogeneous<br />
and layered <strong>Earth</strong> models, respectively, and <strong>the</strong> maximum compression are<br />
-5 mm and -26 mm for each model. For <strong>the</strong> tangential displacement field <strong>the</strong> maximum<br />
tractions are found to be 41 mm and 45 mm for <strong>the</strong> homogeneous model and <strong>the</strong><br />
layered <strong>Earth</strong> models, respectively. The layerd <strong>Earth</strong> model is similar to <strong>the</strong> analytical<br />
derived, with a factor 1.4 to 4 too small, wea<strong>the</strong>rs <strong>the</strong> homogeneous <strong>Earth</strong> model shows<br />
larger displacements, but as expected, it deviates more from <strong>the</strong> physical picture <strong>of</strong> <strong>the</strong><br />
analytic solution.
Dansk Resumé<br />
Tidekræfter bliver <strong>of</strong>te regnet som støj i forbindelse med ge<strong>of</strong>ysiske målinger og skal<br />
udtrækkes fra data, men tidekræfterne kan afsløre vigtig information om Jordens indre.<br />
Tidekræfter opstår som den varierende kraft på et legeme forårsaget af den gravitationelle<br />
tiltrækning fra eksterne objekter. Jorden er ikke et stift legeme, men elastisk, og<br />
kan derfor deformeres under påvirkning af eksterne kræfter. Tidejorden kan bestemmes<br />
analytisk ved en kombination af bestemmelse af Jordens banebevægelse i forhold<br />
til Månen, Solen og de andre eksterne objekter, og viden om Jordens elasticitet i form<br />
af enhedsløse Love tal.<br />
I dette studie bliver den to-dimensionalle Naviers ligning i ligevægts tilstand løst<br />
numerisk ved at diskretisere Naviers ligning med endelige differencer i et polært koordinat<br />
system. Lignings systemet viser sig at være dårligt stillet (ill-conditioned) og<br />
derfor anvendes en Tikhonov regulering til at løse problemet.<br />
Det antages at Månen er det eneste tiltrækkende legeme, og at Jord-Måne systemet<br />
er stationært i forhold til hinanden. Der er udviklet to Jordmodeller, en homogen og<br />
en lagdelt model. For den homogene model antages det, at Jorden er et fast Poisson<br />
legeme, og de seismisk hastigheder og elastiske parametre antages at være konstante<br />
gennem hele legemet. Den lagdelte model indeholder fire lag; en indre kerne, en ydre<br />
flydende kerne, en kappe og en skorpe, hvilket gør den mere realistisk. Jordmodellen<br />
bruger data fra den isotrope Preliminary Reference <strong>Earth</strong> Model, [Dziewonski and<br />
Anderson, 1981], hvor parametrene er midlet over hvert lag.<br />
Begge Jordmodellernes reaktion på tidekræfterne er givet i form af det radiale og<br />
tangentielle forskydningsfelt. Resultatet er et klart fysiske billede af tidejorden, hvor<br />
det radiale forskydningsfelt er ekspansiv med maksimum ved Ækvator og kompressiv<br />
med maksimum ved Nordpolen, dog er kompressionen af de indre strukture af Jorden<br />
ikke helt opfyldt. Den horisontale forskydning er rent trækkende med maksimum ved<br />
45 ◦ og nul ved både den polære og den ækvatorielle akse.<br />
Den homogene jordmodel resulterer i et radial forskydnings felt med maksimum<br />
ved Ækvator med en 131 mm ekspansion og maksimum kompression ved Nordpolen<br />
på -5 mm. Det tangentielle forskydnings felt for den homogene jordmodel har et maksimum<br />
træk på 41 mm. Den lagdelte jordmodels radiale forskydnings felt er fundet<br />
til ekspandere 74 mm ved Ækvator og blive presset sammen med en forskydning på<br />
-26 mm ved Nordpolen. Det tangentielle forskydnings felt har et maksimalt træk på<br />
45 mm. Den lagdelte jordmodel svarer overens med det analytisk beregnede resultat<br />
med forskydninger på en faktor 1.4 til 4 for små. Den homogene jordmodel viser højere<br />
forskydninger, men det fysiske billede afvigerer som ventet mere fra det analytiske<br />
resultat.
Preface<br />
This <strong>the</strong>sis in geophysics has been carried out at <strong>the</strong> Niels Bohr Institute, University<br />
<strong>of</strong> Copenhagen and applies to a year’s work equivalent to 60 ECTS points. This work<br />
will finish my master’s degree in geophysics.<br />
The reader <strong>of</strong> this text is expected to have a basic geophysical knowledge with a<br />
background in fundamental physics and ma<strong>the</strong>matics.<br />
Acknowledgments<br />
First <strong>of</strong> all I would like to thank my supervisors Klaus Mosegaard and Carl Christian<br />
Tcsherning for <strong>the</strong>ir guidance through <strong>the</strong> whole process and <strong>the</strong>ir patience with all my<br />
questions. A special thanks goes to Klaus for your optimism in <strong>the</strong> frustrating process<br />
<strong>of</strong> modeling and for all your constructive ideas. Thanks to Amir Khan for providing<br />
me with <strong>the</strong> PREM data. I would also like to thank Sebastian Bjerregaard Simonsen<br />
for pro<strong>of</strong>reading, Thomas R. N. Jansson for pro<strong>of</strong>reading and letting me login to your<br />
home computer, when <strong>the</strong> server at NBI failed. Thanks to <strong>the</strong> <strong>of</strong>fice for <strong>the</strong> good<br />
humor. Last, I would like to thank Andreas Rose for pro<strong>of</strong>reding,your unconditional<br />
support and for being my connection to <strong>the</strong> real world.<br />
v<br />
Stine Kildegaard Poulsen<br />
Copenhagen, March 2009.
Contents<br />
Contents vii<br />
1 Introduction 1<br />
1.1 Spherical Coordinate System . . . . . . . . . . . . . . . . . . . . . . 3<br />
2 The Tides 5<br />
2.1 Fundamentals <strong>of</strong> Gravity Field Theory . . . . . . . . . . . . . . . . . 5<br />
2.1.1 Gravitation . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />
2.1.2 Centrifugal Acceleration . . . . . . . . . . . . . . . . . . . . 7<br />
2.1.3 Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />
2.2 <strong>Tidal</strong> Forcing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />
2.2.1 <strong>Tidal</strong> Acceleration and Potential . . . . . . . . . . . . . . . . 10<br />
2.2.2 Observation <strong>of</strong> Tides . . . . . . . . . . . . . . . . . . . . . . 13<br />
3 Theory <strong>of</strong> Elasticity 15<br />
3.1 Stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />
3.2 <strong>Deformation</strong> and Strain . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />
3.3 Equation <strong>of</strong> Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />
3.3.1 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . 19<br />
3.4 <strong>Tidal</strong> <strong>Deformation</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />
3.4.1 Structural Effects on <strong>Earth</strong> Tides . . . . . . . . . . . . . . . . 22<br />
4 Finite Difference Equations 25<br />
4.1 The Centered Finite Difference Scheme . . . . . . . . . . . . . . . . 25<br />
4.2 Error Estimates and Stability <strong>of</strong> FDE . . . . . . . . . . . . . . . . . . 28<br />
4.3 The Finite Difference Formula . . . . . . . . . . . . . . . . . . . . . 30<br />
4.3.1 Mesh Points . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />
4.4 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />
4.4.1 At <strong>the</strong> Surface . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />
4.4.2 Center <strong>of</strong> <strong>Earth</strong> . . . . . . . . . . . . . . . . . . . . . . . . . 35<br />
4.4.3 The Pole and Equator . . . . . . . . . . . . . . . . . . . . . . 37
viii <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
5 Regularization 39<br />
5.1 Ill-conditioning <strong>of</strong> a System . . . . . . . . . . . . . . . . . . . . . . 39<br />
5.2 Tikhonov Regularization . . . . . . . . . . . . . . . . . . . . . . . . 41<br />
5.2.1 Singular Value Decomposition . . . . . . . . . . . . . . . . . 42<br />
5.2.2 The Good Solution . . . . . . . . . . . . . . . . . . . . . . . 43<br />
6 Model Description and Construction 45<br />
6.1 Model Assumptions and Parameters . . . . . . . . . . . . . . . . . . 45<br />
6.2 The Homogeneous <strong>Earth</strong> Model . . . . . . . . . . . . . . . . . . . . 46<br />
6.3 The Layered <strong>Earth</strong> Model . . . . . . . . . . . . . . . . . . . . . . . . 47<br />
6.3.1 Assumptions and Parameters . . . . . . . . . . . . . . . . . . 47<br />
6.4 Model Development . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />
6.4.1 Grid Points and Meshes . . . . . . . . . . . . . . . . . . . . 48<br />
6.4.2 The Coefficient Matrix . . . . . . . . . . . . . . . . . . . . . 49<br />
7 Model Considerations and Tests 53<br />
7.1 Model Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />
7.2 Analyzing <strong>the</strong> System . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />
7.2.1 Singular Value Spectrum for <strong>the</strong> Homogeneous Model . . . . 54<br />
7.2.2 Weighting <strong>of</strong> <strong>the</strong> Surface Boundary Conditions . . . . . . . . 55<br />
7.2.3 Singular Value Spectrum for <strong>the</strong> Layered Model . . . . . . . 57<br />
7.3 Choosing <strong>the</strong> Regularization Parameters . . . . . . . . . . . . . . . . 58<br />
7.3.1 Tikhonov Filter Factors . . . . . . . . . . . . . . . . . . . . . 60<br />
7.4 Testing <strong>the</strong> System . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />
7.4.1 Test <strong>of</strong> Solving Methods . . . . . . . . . . . . . . . . . . . . 61<br />
7.4.2 The General Solution . . . . . . . . . . . . . . . . . . . . . . 61<br />
7.4.3 Testing <strong>the</strong> Weighting <strong>of</strong> Grid Numbers . . . . . . . . . . . . 62<br />
7.4.4 Testing <strong>the</strong> Boundary Conditions . . . . . . . . . . . . . . . . 64<br />
7.4.5 Testing <strong>the</strong> <strong>Tidal</strong> Acceleration . . . . . . . . . . . . . . . . . 65<br />
8 The <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Earth</strong> 67<br />
8.1 <strong>Deformation</strong> in <strong>the</strong> Homogeneous <strong>Earth</strong> Model . . . . . . . . . . . . 67<br />
8.2 <strong>Deformation</strong> in <strong>the</strong> Layered <strong>Earth</strong> Model . . . . . . . . . . . . . . . . 70<br />
8.3 Analytic Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73<br />
8.3.1 Comparison Between Results . . . . . . . . . . . . . . . . . 73<br />
9 Discussion 75<br />
9.1 Results and Errors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />
9.1.1 The Displacement Fields . . . . . . . . . . . . . . . . . . . . 75<br />
9.1.2 Regularization . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />
9.1.3 The Discretization . . . . . . . . . . . . . . . . . . . . . . . 78<br />
9.1.4 Ill-conditioning . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />
9.1.5 Computer Errors . . . . . . . . . . . . . . . . . . . . . . . . 79<br />
9.2 The Reliability <strong>of</strong> <strong>the</strong> Model . . . . . . . . . . . . . . . . . . . . . . 80<br />
9.2.1 Elasticity Assumptions . . . . . . . . . . . . . . . . . . . . . 80<br />
9.2.2 The Real Tides . . . . . . . . . . . . . . . . . . . . . . . . . 80
CONTENTS ix<br />
9.3 Application <strong>of</strong> Tides . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />
9.4 Model Improvements . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />
10 Conclusion 85<br />
Bibliography 87<br />
Appendices 94<br />
A Hyperbolic Function 94<br />
B Matlab Codes 95<br />
B.1 FDE_premiso.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95<br />
B.2 Matlab Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101<br />
B.2.1 ftides_r.m . . . . . . . . . . . . . . . . . . . . . . . . . . 101<br />
B.2.2 ftides_<strong>the</strong>ta.m . . . . . . . . . . . . . . . . . . . . . . . 101<br />
B.2.3 mean_alpha.m . . . . . . . . . . . . . . . . . . . . . . . . . 101<br />
B.2.4 mean_beta.m . . . . . . . . . . . . . . . . . . . . . . . . . 102<br />
C Averaging <strong>of</strong> PREM Layers 103<br />
D Model Tests 104<br />
D.1 Tikhonov Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104<br />
D.2 Regularization Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
Chapter 1<br />
Introduction<br />
For most people tides are a synonym for <strong>the</strong> daily falling and rising <strong>of</strong> sea level. There<br />
is a general understanding that it is due to <strong>the</strong> Moon’s attraction, but what about <strong>the</strong><br />
sea level rising, at <strong>the</strong> side turning away from <strong>the</strong> Moon and <strong>the</strong> influence <strong>of</strong> <strong>the</strong> Sun<br />
and from o<strong>the</strong>r external objects?<br />
<strong>Tidal</strong> forces are more complex than a one dimensional pull and causes many effects<br />
in our solar system; from changing sea level to geological activity on distant moons,<br />
galactic interactions and even <strong>the</strong> final faith <strong>of</strong> some moons. The gravitational force<br />
that arise by <strong>the</strong> pull from external objects varies from one part <strong>of</strong> <strong>the</strong> effected object<br />
to an o<strong>the</strong>r. These differential pulls produces, what are known as tidal forces.<br />
Whenever we touch a body, internal stresses are set up in <strong>the</strong> material in contact.<br />
Elastic materials can be deformed depending <strong>of</strong> <strong>the</strong> applied force and <strong>the</strong> properties <strong>of</strong><br />
<strong>the</strong> material. In short time scales <strong>the</strong> <strong>Earth</strong> can <strong>the</strong>refore be considered to be elastic<br />
and can be deformed by an external force, <strong>the</strong> tidal force.<br />
The tides have been <strong>of</strong> great discussion throughout history, and we know that it causes<br />
many effects throughout <strong>the</strong> solar system. There has been a lot <strong>of</strong> misconception about<br />
tides, and what causes <strong>the</strong>m. In 1912 Alfred Wegener suggested that tidal forces broke<br />
up <strong>the</strong> continents, [Wegener, 1966].<br />
The discovery <strong>of</strong> tides dates back to <strong>the</strong> beginning <strong>of</strong> our calendar, when Pliny <strong>the</strong><br />
Elder in his Historia Naturalis states that near <strong>the</strong> temple <strong>of</strong> Hercules <strong>the</strong> ocean rises<br />
and falls. The ma<strong>the</strong>matical basis for <strong>the</strong> <strong>the</strong>ory <strong>of</strong> tides was founded in 1687 by Isaac<br />
Newton in Philosophiae Naturalis Principia Ma<strong>the</strong>matica, (for a translated version<br />
see [Newton, 1833]), when he developed <strong>the</strong> laws <strong>of</strong> motion and <strong>the</strong> universal law <strong>of</strong><br />
gravitation. This is shortly described in Section 2.1. In 1776, Pierre-Simon Laplace<br />
formulated <strong>the</strong> first <strong>the</strong>oretical description <strong>of</strong> <strong>the</strong> ocean tides, by a system <strong>of</strong> partial<br />
differential equation, known as <strong>the</strong> Laplace tidal equation (for a translated version see<br />
[Laplace, 1832]). It was also this year <strong>the</strong> first experiment with a pendulum, to observe<br />
<strong>the</strong> tides, was carried out. The Laplace tidal equation is still used today, and gives <strong>the</strong><br />
tidal acceleration on <strong>the</strong> surface in parts <strong>of</strong> long-, daily- and half daily periodic variations.<br />
At this point in history, <strong>the</strong> <strong>Earth</strong> was assumed to be rigid, but about 1876,<br />
Lord Kelvin stated that <strong>the</strong> <strong>Earth</strong> is not completely rigid, it can deform for example as
2 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
a result <strong>of</strong> tidal forces. G. H. Darwin, assumed in 1883 that for a perfectly rigid <strong>Earth</strong>,<br />
<strong>the</strong> observed amplitude <strong>of</strong> <strong>the</strong> ocean tide would equal <strong>the</strong> <strong>the</strong>oretical value. If <strong>the</strong> solid<br />
part also was deformed, <strong>the</strong> measured amplitude would equal <strong>the</strong> difference between<br />
oceanic and solid <strong>Earth</strong> tides. He applied it to observations <strong>of</strong> long periodic oceanic<br />
tides, and found that <strong>the</strong> amplitude was only two-third <strong>of</strong> <strong>the</strong> <strong>the</strong>oretical tides, i.e. <strong>the</strong><br />
<strong>Earth</strong> is not completely rigid, but deformable. The <strong>Earth</strong>’s rigidity is <strong>of</strong>ten compared<br />
with steel. Steel is not easy to bend, however under <strong>the</strong> right circumstances, it is possible.<br />
Toge<strong>the</strong>r with seismological data and polar motion, this new data led to <strong>the</strong> first<br />
discoveries <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s elastic properties, [Melchior, 1978]. In 1909, E. H. Love<br />
introduced <strong>the</strong> Love numbers, which are dimensionless constants <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s elastic<br />
parameters, (see Love [1944]). Fur<strong>the</strong>r definition <strong>of</strong> Love numbers will be given in<br />
Section 3.4.<br />
This <strong>the</strong>sis presents two simple models <strong>of</strong> <strong>the</strong> deformation <strong>of</strong> <strong>the</strong> <strong>Earth</strong> due to <strong>the</strong><br />
solid <strong>Earth</strong> tides, based on a finite difference discretization <strong>of</strong> <strong>the</strong> Navier equation <strong>of</strong><br />
motion. The two <strong>Earth</strong> models are a homogeneous <strong>Earth</strong> model and a layered <strong>Earth</strong><br />
model, where <strong>the</strong> later is an expansion <strong>of</strong> <strong>the</strong> first. The <strong>the</strong>sis opens with <strong>the</strong> basic<br />
<strong>the</strong>ory <strong>of</strong> tides in Chapter 2, followed by <strong>the</strong> <strong>the</strong>ory <strong>of</strong> elasticity in Chapter 3. This<br />
chapter also deals with <strong>the</strong> equation <strong>of</strong> motion and <strong>the</strong> tidal deformation <strong>of</strong> <strong>the</strong> <strong>Earth</strong>.<br />
In Chapter 4, <strong>the</strong> finite difference equations are introduced and a discretization <strong>of</strong><br />
<strong>the</strong> equation <strong>of</strong> motion is developed in a polar coordinate system to match <strong>the</strong> tidal<br />
acceleration. Solving <strong>the</strong> linear system was a cumbersome process and a lot <strong>of</strong> <strong>the</strong><br />
work in this master <strong>the</strong>sis has gone into regularizing <strong>the</strong> linear system <strong>of</strong> equations to<br />
get <strong>the</strong> most suitable results. The regularization is described in Chapter 5.<br />
Chapter 6 describes and sets up <strong>the</strong> assumption for <strong>the</strong> two <strong>Earth</strong> models and an<br />
analysis <strong>of</strong> <strong>the</strong> system is performed. Chapter 7 dealing with <strong>the</strong> modeling <strong>of</strong> <strong>the</strong> system.<br />
The construction <strong>of</strong> <strong>the</strong> model is described, <strong>the</strong> regularization parameters are<br />
chosen and <strong>the</strong> system is tested in varies ways. In Chapter 8, <strong>the</strong> results from <strong>the</strong> modeling<br />
are presented and compared to an analytic result. Chapter 9 discuss <strong>the</strong> given<br />
results, where also suggestions for improvements are proposed, and Chapter 10 concludes<br />
this master <strong>the</strong>sis.
Spherical Coordinate System 3<br />
1.1 Spherical Coordinate System<br />
Throughout <strong>the</strong> <strong>the</strong>sis several angles are defined and to keep a track <strong>of</strong> <strong>the</strong>m from <strong>the</strong><br />
beginning, <strong>the</strong>y are depicted in Figure 1.1.1. The figure shows <strong>the</strong> relation between <strong>the</strong><br />
Cartesian coordinates X, Y, Z placed in <strong>the</strong> <strong>Earth</strong>’s center <strong>of</strong> mass, where <strong>the</strong> XY-axes<br />
spans <strong>the</strong> equatorial plane, and <strong>the</strong> spherical coordinates r, θ, λ. θ is <strong>the</strong> colatitude,<br />
λ is <strong>the</strong> geocentric longitude. θ is <strong>the</strong> geocentric latitude, and ψ is simply <strong>the</strong> angle<br />
between two points. The <strong>Tidal</strong> force is calculated using ψ and <strong>the</strong> polar coordinate<br />
system (r, θ) is used in <strong>the</strong> Navier equation <strong>of</strong> motion.<br />
X<br />
0<br />
Z<br />
λ<br />
θ<br />
r<br />
θ<br />
Figure 1.1.1: Relation between coordinates. The Cartesian coordinates X, Y, Z is placed in<br />
<strong>the</strong> <strong>Earth</strong>’s center <strong>of</strong> mass, where <strong>the</strong> XY-axes spans <strong>the</strong> equator plan. θ is <strong>the</strong> colatitude, θ<br />
is <strong>the</strong> geocentric latitude, λ is <strong>the</strong> geocentric longitude and ψ is simply <strong>the</strong> angle between two<br />
points. The <strong>Tidal</strong> force is calculated with ψ and <strong>the</strong> polar coordinate system used in <strong>the</strong> Navier<br />
equation <strong>of</strong> motion, normally uses <strong>the</strong> colatitude θ.<br />
ψ<br />
r ′<br />
P<br />
P ′<br />
Y
Chapter 2<br />
The Tides<br />
This chapter aims to describe <strong>the</strong> basic <strong>the</strong>ory about tides. First, in Section 2.1, <strong>the</strong><br />
classic mechanics <strong>of</strong> gravitation and centrifugal force is shortly described. Secondly<br />
followed a definition <strong>of</strong> tides in Section 2.2, where also <strong>the</strong> tidal acceleration in polar<br />
coordinates is written out. The methods used to measure tides, rounds <strong>of</strong>f this chapter.<br />
2.1 Fundamentals <strong>of</strong> Gravity Field Theory<br />
For a rotating body, <strong>the</strong> force <strong>of</strong> gravity (Section 2.1.3) is <strong>the</strong> combination <strong>of</strong> gravitational<br />
attraction (Section 2.1.1) arising from <strong>the</strong> mass distributed in a planet for<br />
example <strong>the</strong> <strong>Earth</strong> or o<strong>the</strong>r celestial bodies, as well as <strong>the</strong> centrifugal force (Section<br />
2.1.2) arising from <strong>the</strong> body’s rotation about its own axis. The gravity field reveal<br />
information <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s interior, (Section 2.1.1).<br />
Gravity changes with time, appearing at recognizable frequencies, are called tides,<br />
(Section 2.2). Gravity can also appear at time scales ranging from secular to abrupt.<br />
Secular changes includes subduction <strong>of</strong> continental plates, deformation <strong>of</strong> <strong>the</strong> coremantle<br />
boundary, post glacial rebound, melting <strong>of</strong> ice caps and glaciers. Abrupt<br />
changes includes volcanic and earthquake activity and only gives local effects. Gravity<br />
changes arise due to 1.) <strong>the</strong> time dependent gravitational constant (this is only a<br />
<strong>the</strong>oretical consideration in cosmology, ant <strong>the</strong> time dependency is so fare not verified<br />
for <strong>the</strong> Moon), 2.) variation and direction <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s rotation, which changes <strong>the</strong><br />
centrifugal force, 3.) variation <strong>of</strong> terrestrial mass displacements which includes vertical<br />
crustal movements that changes <strong>the</strong> distance to <strong>the</strong> center <strong>of</strong> mass, and it includes<br />
density variations in <strong>the</strong> crust and mantle and 4.) tidal accelerations, [Torge, 2001].<br />
The first three point are out <strong>of</strong> scoop <strong>of</strong> this <strong>the</strong>sis and will not be discussed in fur<strong>the</strong>r<br />
details, however <strong>the</strong> fourth point, <strong>the</strong> tidal acceleration, is <strong>the</strong> emphasis <strong>of</strong> this <strong>the</strong>sis<br />
and will be covered in <strong>the</strong> following, (Section 2.2).
6 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
2.1.1 Gravitation<br />
In an initial reference frame, two point masses m1 and m2, see Figure 2.1.1, with <strong>the</strong><br />
distance l between <strong>the</strong> masses, attract each o<strong>the</strong>r with <strong>the</strong> gravitational force given by<br />
Newton’s law <strong>of</strong> gravitation<br />
−11 m3<br />
F = −G m1m2<br />
l 2<br />
l<br />
l<br />
, (2.1.1)<br />
where G = 6.673 · 10<br />
kgs2 is <strong>the</strong> universal constant <strong>of</strong> gravitation, l is <strong>the</strong> distance<br />
between <strong>the</strong> masses, l = r − r ′ . The vectors F and l points in opposing directions, see<br />
Figure 2.1.1.<br />
Letting <strong>the</strong> mass, m1, at <strong>the</strong> attracting point P, go towards unity, <strong>the</strong> formula becomes<br />
a = −G m2<br />
l 2<br />
The gravitation is now an acceleration, where a initiate at P, directed towards P ′ . The<br />
gravitational field is invariant to rotation and can be represented by <strong>the</strong> gradient <strong>of</strong> <strong>the</strong><br />
potential V<br />
where <strong>the</strong> scalar potential is given by<br />
X<br />
0<br />
Z<br />
l<br />
l<br />
.<br />
a = ∇V , (2.1.2)<br />
r ′<br />
ψ<br />
m2<br />
Figure 2.1.1: Gravitation <strong>of</strong> two point masses m1 and m2 in a Cartesian coordinate system.<br />
The points P and p ′ are separated by a distance l and angle ψ.<br />
P ′<br />
r<br />
l<br />
m1<br />
P<br />
Y
Fundamentals <strong>of</strong> Gravity Field Theory 7<br />
V = GM<br />
l<br />
. (2.1.3)<br />
At <strong>the</strong> surface <strong>of</strong> <strong>the</strong> <strong>Earth</strong> (l = r, where r is <strong>the</strong> radius <strong>of</strong> <strong>the</strong> <strong>Earth</strong>) <strong>the</strong> potential<br />
for a point mass is 6.26 × 107 m2<br />
s2 and <strong>the</strong> gravitation is on an average a = 9.82 m<br />
s2 .<br />
[Torge, 2001].<br />
2.1.2 Centrifugal Acceleration<br />
In an accelerated reference frame fictitious forces arise. The centrifugal force is one <strong>of</strong><br />
<strong>the</strong>m, initiating from <strong>the</strong> rotation <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, about its own axis. The centrifugal force<br />
acts on every element <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, in Figure 2.1.2 <strong>the</strong> centrifugal force is displaced for<br />
an element <strong>of</strong> mass P near <strong>the</strong> <strong>Earth</strong>’s surface.<br />
With a constant angular velocity, ω, <strong>the</strong> centrifugal force is given by<br />
z = m(ω × r) × ω = m ω 2 ρ , (2.1.4)<br />
where ρ = |ρ| = r cos θ, and r is <strong>the</strong> distance to <strong>the</strong> point in question at geocentric<br />
latitude θ calculated positive from <strong>the</strong> Equator towards <strong>the</strong> Pole. See also Figure 1.1.1<br />
in Section 1.1, where <strong>the</strong> different angles are depicted in a spherical coordinate system.<br />
ω<br />
0<br />
Figure 2.1.2: Centrifugal force in a rotating reference frame. ω is <strong>the</strong> rotation vector, ρ is <strong>the</strong><br />
distance to <strong>the</strong> point P from <strong>the</strong> rotating axis. r is <strong>the</strong> distance to <strong>the</strong> point from <strong>the</strong> center <strong>of</strong><br />
mass and z is <strong>the</strong> centrifugal force, θ is <strong>the</strong> latitude.<br />
ρ<br />
θ<br />
r<br />
P<br />
z
8 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
The centrifugal potential written in <strong>the</strong> vector field yields<br />
Z = 1<br />
2 m ω2 ρ 2 . (2.1.5)<br />
5 m2<br />
At <strong>the</strong> Equator <strong>the</strong> centrifugal force has a value <strong>of</strong> 1.1 · 10<br />
s2 and <strong>the</strong> centrifugal<br />
acceleration is 0.03 m<br />
s2 , which is 0.3% <strong>of</strong> <strong>the</strong> gravitation. The potential and <strong>the</strong> acceleration<br />
are both zero at <strong>the</strong> Pole. The <strong>Earth</strong>’s rotation vector ω can change <strong>the</strong> centrifugal<br />
acceleration because <strong>of</strong> its secular, periodic and irregular variations, [Knudsen<br />
and Hjorth, 2000; Torge, 2001].<br />
2.1.3 Gravity<br />
The gravity g is <strong>the</strong> result <strong>of</strong> gravitation aG and centrifugal acceleration az<br />
g = aG + az ,<br />
and <strong>the</strong> same can be written for <strong>the</strong> potentials. For an ellipsoidal model, like <strong>the</strong> <strong>Earth</strong>,<br />
gravitation decreases at equator and increases at <strong>the</strong> poles. This results in a varying<br />
gravity between 9.78 m<br />
s2 at <strong>the</strong> Equator to 9.83 m<br />
s2 at <strong>the</strong> Poles, [Torge, 2001].<br />
2.2 <strong>Tidal</strong> Forcing<br />
As already mentioned <strong>the</strong> gravitational force on a body, arises by an external orbiting<br />
object, working with different force throughout <strong>the</strong> body, see Figure 2.2.1(a). The<br />
variation <strong>of</strong> <strong>the</strong> force over <strong>the</strong> volumes <strong>of</strong> material is defined as <strong>the</strong> differential force,<br />
tide-generating force or simply <strong>the</strong> tidal force. Due to <strong>the</strong> acceleration <strong>of</strong> <strong>the</strong> external<br />
object this is a non-inertial reference frame. In consequence <strong>of</strong> Newton’s 2. law,<br />
fictitious forces have to be included, working against <strong>the</strong> gravitation, this is <strong>the</strong> centrifugal<br />
force. These two forces cancels out in <strong>the</strong> center <strong>of</strong> mass <strong>of</strong> a homogeneous<br />
sphere. The tidal acceleration is obtained by subtracting <strong>the</strong> centrifugal force from<br />
<strong>the</strong> gravitation, see Figure 2.2.1(b). The total tidal force on a sphere is depicted on<br />
Figure 2.2.1(c), and it shows how <strong>the</strong> force varies with power and direction all over<br />
<strong>the</strong> volume. The coordinate system is placed at <strong>the</strong> center <strong>of</strong> mass in <strong>the</strong> body. The<br />
tidal force pulls two bodies, with significant masses, from each o<strong>the</strong>r by <strong>the</strong> simple<br />
approximation<br />
Ft =<br />
≈<br />
G M m<br />
R2 G M m<br />
−<br />
− r2 R2 (2.2.1)<br />
2 G M m r<br />
R3 . (2.2.2)<br />
Where M and m are <strong>the</strong> mass <strong>of</strong> <strong>the</strong> body and <strong>the</strong> mass <strong>of</strong> <strong>the</strong> external object, respectively.<br />
R is <strong>the</strong> distance between <strong>the</strong> two bodies and r is <strong>the</strong> radius <strong>of</strong> <strong>the</strong> body. R ≫ r,<br />
[de Pater and Lissauer, 2001].<br />
From Equation (2.2.1), <strong>the</strong> Moon is affected by 1.82 · 10 18 N from <strong>the</strong> <strong>Earth</strong>, and<br />
<strong>the</strong> <strong>Earth</strong> is affected by 6.69·10 18 N from <strong>the</strong> Moon and 3.02·10 18 N from <strong>the</strong> Sun, i.e.
<strong>Tidal</strong> Forcing 9<br />
<strong>the</strong> <strong>Earth</strong> feels <strong>the</strong> tidal force 3.5 times stronger than <strong>the</strong> Moon feels from <strong>the</strong> <strong>Earth</strong>,<br />
which is equivalent to <strong>the</strong> difference <strong>of</strong> <strong>the</strong> radius. This is due to <strong>the</strong> <strong>Earth</strong>’s greater<br />
mass. The mass <strong>of</strong> <strong>the</strong> Sun, is much greater than <strong>the</strong> mass <strong>of</strong> <strong>the</strong> Moon, but it only<br />
affect <strong>the</strong> <strong>Earth</strong> half <strong>of</strong> what <strong>the</strong> Moon does, which is <strong>the</strong> impact <strong>of</strong> <strong>the</strong> distance given<br />
by <strong>the</strong> inverse cube. In fact <strong>the</strong> <strong>Earth</strong> tides are regulated by <strong>the</strong> orbits <strong>of</strong> <strong>the</strong> Moon and<br />
<strong>the</strong> <strong>Earth</strong>-Moon system about <strong>the</strong> Sun and all o<strong>the</strong>r planetary components.<br />
The tidal force affects <strong>the</strong> whole <strong>Earth</strong> system for example <strong>the</strong> atmosphere, <strong>the</strong> oceans<br />
and <strong>the</strong> solid <strong>Earth</strong>. The atmospheric tide is a global periodic oscillation <strong>of</strong> <strong>the</strong> atmosphere<br />
exited by insolation <strong>of</strong> <strong>the</strong> atmosphere, gravity field and non-linear interactions<br />
between tides and planetary waves, [Torge, 2001]. Tides can also influence deep moonquakes,<br />
which are triggered by <strong>the</strong> tides from <strong>the</strong> <strong>Earth</strong> [Lammlein, 1977], earthquakes<br />
(see e.g. [Heaton, 1975]), <strong>the</strong> periodicity <strong>of</strong> volcanic activity (see e.g. [Sottili et al.,<br />
2007]), and geyser activity (see e.g. [Rinehart, 1972]), ice quakes, where <strong>the</strong> seismicity<br />
rate can be compared with <strong>the</strong> tidal forcing <strong>of</strong> an ice shelf, see [Anandakrishnan<br />
and Alley, 1997], glaciers (see e.g. [Kulessa et al., 2003]), Climate, (see e.g. [Munk<br />
and Dzieciuch, 2001]).<br />
It is important to distinguish between ocean tides and body tides. Ocean tides are<br />
variations <strong>of</strong> sea level strongly dependent <strong>of</strong> coast and bottom topography and ocean<br />
currents. Body tides or solid earth tides is <strong>the</strong> deformation <strong>of</strong> ocean and land only<br />
dependent <strong>of</strong> <strong>the</strong> gravitational forces <strong>of</strong> <strong>the</strong> Moon, Sun and o<strong>the</strong>r external objects, and<br />
it works in all parts <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and deforms <strong>the</strong> <strong>Earth</strong> to an ellipsoid aligned with <strong>the</strong><br />
<strong>Earth</strong>-Moon (Sun) axis, [Wang, 1997]. The tidal accelerations are within ±1µm/s 2 <strong>of</strong><br />
<strong>the</strong> <strong>Earth</strong>’s gravity, [Wenzel, 1997]. The ocean tides will not be covered in <strong>the</strong> <strong>the</strong>sis,<br />
and from this point forward, <strong>the</strong> tides will only refer to <strong>the</strong> solid <strong>Earth</strong> tides, if nothing<br />
else is mentioned.<br />
B<br />
(a)<br />
A<br />
B<br />
A<br />
(b) (c)<br />
Figure 2.2.1: (a) The gravitational acceleration due to an external object is depicted for two<br />
points A and B at <strong>the</strong> surface <strong>of</strong> a spherical body and in <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. The external<br />
object is placed to <strong>the</strong> right <strong>of</strong> <strong>the</strong> figure on <strong>the</strong> line from <strong>the</strong> bodies center <strong>of</strong> mass. (b) After<br />
eliminating <strong>the</strong> fictive acceleration in <strong>the</strong> non-inertial reference frame, <strong>the</strong> tidal acceleration<br />
arises in <strong>the</strong> points A and B. The forces try to deform <strong>the</strong> body along <strong>the</strong> body-external-objectline.<br />
(c) <strong>the</strong> same as previous figure, but here <strong>the</strong> tidal acceleration is calculated for all point<br />
in <strong>the</strong> <strong>Earth</strong>. It is seen that <strong>the</strong> tidal acceleration varies in power and direction over <strong>the</strong> sphere.<br />
Figure (c) is adapted from [wikimedia.org, 2006].
10 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
F2<br />
Figure 2.2.2: The total torque <strong>of</strong> a planets north pole is shown in <strong>the</strong> figure. When a planet is<br />
rotating about its own axis, rises an equatorial bulge around <strong>the</strong> Equator. The planet is also<br />
affected by an external object or moon, which is orbiting <strong>the</strong> planet in <strong>the</strong> positive direction.<br />
The planet experience an acceleration from <strong>the</strong> moon which creates an tidal bulge. The bulge<br />
is pressed in front <strong>of</strong> <strong>the</strong> movement and it is lowering <strong>the</strong> rotation speed <strong>of</strong> <strong>the</strong> planet. F1 and<br />
F2 are <strong>the</strong> gravitational force at <strong>the</strong> given points.<br />
Deformable bodies rotating about <strong>the</strong>ir own axis, will experience a distortion <strong>of</strong> <strong>the</strong><br />
body. Looking at a spherical solid body entirely covered with water, rotating about its<br />
own axis, it will experience an equatorial bulge <strong>of</strong> both solid and water. This is not<br />
to be mistaken by <strong>the</strong> tidal bulge. The equatorial bulge is not created by <strong>the</strong> gravitational<br />
force and has no periodic variation, but is only a consequence <strong>of</strong> rotation. The<br />
equatorial bulge is <strong>the</strong> baseline, for which real tidal effects are measured. The tidal<br />
bulge is on <strong>the</strong> o<strong>the</strong>r hand a result from gravitation alone. There arises two bulges due<br />
to <strong>the</strong> conservation <strong>of</strong> angular momentum. The total torque <strong>of</strong> <strong>the</strong> sphere with a moon<br />
moving prograde about <strong>the</strong> sphere is depicted in Figure 2.2.2. The sphere is shown<br />
from <strong>the</strong> North Pole. The satellite’s orbiting period is longer than <strong>the</strong> sphere’s rotations<br />
period, and due to <strong>the</strong> non perfect elasticity, <strong>the</strong> bulge would be pushed in front<br />
<strong>of</strong> <strong>the</strong> movement. The gravitational force <strong>of</strong> <strong>the</strong> moon is greater on <strong>the</strong> near-side bulge<br />
than <strong>the</strong> far-side bulge. This will lower <strong>the</strong> rotation speed <strong>of</strong> <strong>the</strong> sphere, [de Pater and<br />
Lissauer, 2001; Tsantes, 1972].<br />
The <strong>Earth</strong> moves prograde, as seen above <strong>the</strong> North Pole, and a little faster than <strong>the</strong><br />
Moon revolves around <strong>the</strong> <strong>Earth</strong> (also prograde). Therefore <strong>the</strong> <strong>Earth</strong> drags <strong>the</strong> high<br />
tide bulge 3 ◦ ahead <strong>of</strong> its movement, [Tsantes, 1972].<br />
2.2.1 <strong>Tidal</strong> Acceleration and Potential<br />
Considering a revolving geocentric coordinate system without rotation, all points in<br />
<strong>the</strong> same distance from <strong>the</strong> <strong>Earth</strong>’s center, experiences an equal orbital acceleration.<br />
The gravitation <strong>of</strong> <strong>the</strong> celestial bodies have to cancel out in <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>.<br />
The tidal acceleration is <strong>the</strong> difference between <strong>the</strong> gravitation a, depending on <strong>the</strong><br />
position <strong>of</strong> <strong>the</strong> point P, and <strong>the</strong> constant acceleration a0 at <strong>the</strong> <strong>Earth</strong>’s center<br />
at = a − a0 = GMi<br />
|r − ri| 2<br />
|r − ri| GMi<br />
−<br />
|r − ri| r2 ri<br />
ri<br />
i<br />
F1<br />
, (2.2.3)<br />
where Mi is <strong>the</strong> mass <strong>of</strong> <strong>the</strong> celestial body attracting <strong>the</strong> <strong>Earth</strong>. The calculations are<br />
carried out for each two-body system with <strong>the</strong> <strong>Earth</strong> being one part and <strong>the</strong> Moon, <strong>the</strong>
<strong>Tidal</strong> Forcing 11<br />
Sun or a planet are <strong>the</strong> second part. |r − ri| and ri are <strong>the</strong> distances to <strong>the</strong> celestial body<br />
in question from <strong>the</strong> point P and <strong>the</strong> center <strong>of</strong> gravity, respectively. The total tidal<br />
acceleration is <strong>the</strong> result <strong>of</strong> adding all <strong>the</strong> tidal accelerations toge<strong>the</strong>r.<br />
Using spherical harmonics <strong>the</strong> tidal potential is given by<br />
Φn = GMi<br />
∞ n r<br />
Pn(cos ψ) , (2.2.4)<br />
ri ri<br />
n=2<br />
where Pn(cos ψ) are <strong>the</strong> Legendre polynomials, n = 0 and n = 1 terms cancels out in<br />
<strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. ψ is defined as <strong>the</strong> angle between <strong>the</strong> object and <strong>the</strong> point in<br />
question. Looking only at <strong>the</strong> n = 2 term, which is <strong>the</strong> main contribution (≈ 98%), and<br />
applying P2(cos ψ) = 3<br />
2 cos2 ψ − 1<br />
2 . The main tidal potential can be calculated from<br />
Φ2 = GMi<br />
2<br />
r 2<br />
r 3 i<br />
3 cos 2 ψ − 1 , (2.2.5)<br />
Equation (2.1.2) states that <strong>the</strong> tidal acceleration is <strong>the</strong> gradient <strong>of</strong> <strong>the</strong> tidal potential.<br />
Assuming rotation symmetry and choosing <strong>the</strong> coordinate system with origin at <strong>the</strong><br />
center <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and <strong>the</strong> radial axis pointing towards <strong>the</strong> Moon. The components <strong>of</strong><br />
<strong>the</strong> tidal acceleration, in vector coordinates atides = (ar(r, ψ, λ), aψ, (r, ψ, λ), aλ(r, ψ, λ)),<br />
can <strong>the</strong>n be written as<br />
a tides<br />
r (r, ψ, λ) = ∂Φ2<br />
∂r<br />
a tides<br />
ψ (r, ψ, λ) = − ∂Φ2<br />
r∂ψ<br />
= GMi<br />
r<br />
r 3 i<br />
= 3 GMi<br />
2<br />
3 cos 2 ψ − 1 <br />
r<br />
r 3 i<br />
(2.2.6)<br />
sin 2ψ (2.2.7)<br />
a tides<br />
λ (r, ψ, λ) = − ∂Φ2<br />
= 0 . (2.2.8)<br />
r cos ψ∂λ<br />
is <strong>the</strong> radial component and is both compressive and expansive in <strong>the</strong> <strong>Earth</strong>-Moon<br />
is <strong>the</strong> latitudinal component parallel to<br />
<strong>the</strong> tangent <strong>of</strong> <strong>the</strong> latitudinal direction towards south and atides λ is <strong>the</strong> longitudinal component<br />
and is zero because <strong>of</strong> assumption <strong>of</strong> rotation symmetry. The two tangential<br />
accelerations are tractive and dominates <strong>the</strong> large oceans, [Torge, 2001].<br />
atides r<br />
line and dominates <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong> tides, atides ψ<br />
The Periodic Tides<br />
The position <strong>of</strong> <strong>the</strong> attacking body is dependent <strong>of</strong> time. The tides are periodic; diurnal<br />
and semi-diurnal (12 hours) depending on <strong>the</strong> rotation <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, fortnightly (14<br />
days) and monthly due to <strong>the</strong> motion <strong>of</strong> <strong>the</strong> Moon, annual and semi-annual (6 months)<br />
due to <strong>the</strong> orbit <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. There is about a nineteen-yearly period called <strong>the</strong> Metonic<br />
cycle, which is <strong>the</strong> time for recurrence <strong>of</strong> <strong>the</strong> full moon at <strong>the</strong> same date. Depending<br />
on changes <strong>of</strong> <strong>the</strong> Moon, [Love, 1944].
12 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
Following Torge [2001], <strong>the</strong> tidal potential in a geocentric system, also called<br />
Laplace’s tidal equation, can be written as<br />
Vt =<br />
3<br />
4 GMi<br />
r 2<br />
r 3 i<br />
+ sin 2¯θ sin 2δi cos hi<br />
+ cos 2 θ cos 2 δi cos 2hi<br />
<br />
1<br />
3 − sin2 <br />
¯θ (1 − 3 sin 2 δi)<br />
<br />
, (2.2.9)<br />
where θ is <strong>the</strong> geocentric latitude θ = 90 ◦ − θ, index i stands for <strong>the</strong> celestial body<br />
generating <strong>the</strong> gravitational acceleration, ri is <strong>the</strong> distance from <strong>the</strong> attracting body to<br />
<strong>the</strong> <strong>Earth</strong>’s center <strong>of</strong> gravity, δi is <strong>the</strong> declination and hi is <strong>the</strong> hour angle defined as.<br />
hi = LAST − αi,<br />
where LAST is <strong>the</strong> Local Apparent Sidereal Time referring to <strong>the</strong> observer’s local<br />
meridian, and equals <strong>the</strong> hour angle <strong>of</strong> <strong>the</strong> true vernal equinox and αi is <strong>the</strong> right<br />
ascension. The quantities ri, δi and hi in Equation (2.2.9) are all time dependent and<br />
with different periods.<br />
Laplace’s tidal equation is simply included because it explains <strong>the</strong> tidal periods.<br />
The first term <strong>of</strong> Equation (2.2.9) is independent <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s rotation and show 14<br />
days long-periodic variation for <strong>the</strong> moon and half a year for <strong>the</strong> sun. There is also<br />
a non-periodic part<br />
GMir2 <br />
, which only depends on latitude and which result in a<br />
4r 3 i<br />
permanent deformation <strong>of</strong> <strong>the</strong> level surface including <strong>the</strong> geoid. The second term is<br />
due to <strong>the</strong> daily rotation <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and oscillates with diurnal periods. The third term<br />
represents semidiurnal periods where δ and α is controlling <strong>the</strong> long-periodic behaviors,<br />
[Torge, 2001].<br />
It is possible to derive <strong>the</strong> tidal deformation as spherical harmonics <strong>of</strong> second order,<br />
this is especially fruitful when considering <strong>the</strong> disturbance <strong>of</strong> <strong>the</strong> two major external<br />
bodies <strong>the</strong> Moon and <strong>the</strong> Sun. Each term in <strong>the</strong> tide-generating potential corresponds<br />
to a deformation on a spherical harmonic function <strong>of</strong> second order and <strong>the</strong>y can be<br />
divided into three types, which is displayed in Figure 2.2.3. The three spheres are divided<br />
into sectors, where <strong>the</strong> dark colour indicates a rising surface and white indicates<br />
a lowering. The first sphere represents <strong>the</strong> zonal function and it is only dependent <strong>of</strong><br />
<strong>the</strong> latitude. It has a fortnightly period for <strong>the</strong> Moon and semi-annual for <strong>the</strong> Sun.<br />
This function has a maximum at equator and gives a slight flattening <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and<br />
fluctuations in <strong>the</strong> rotation <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. The second sphere in <strong>the</strong> figure represents <strong>the</strong><br />
tesseral function and <strong>the</strong> sectors are changing signs with <strong>the</strong> declination <strong>of</strong> <strong>the</strong> perturbing<br />
body. The corresponding tides are diurnal and <strong>the</strong> amplitude reach maximum at<br />
45 ◦ N and 45 ◦ S and is zero at <strong>the</strong> poles. This function also influence <strong>the</strong> inner core’s<br />
rotation relative to <strong>the</strong> mantle. The last sphere in Figure 2.2.3 is <strong>the</strong> sectorial function.<br />
The tides represented by this function is semi-diurnal and has a maximum at <strong>the</strong><br />
Equator when <strong>the</strong> declination <strong>of</strong> <strong>the</strong> perturbing body is zero, Melchior [1978].
<strong>Tidal</strong> Forcing 13<br />
Figure 2.2.3: The tides on <strong>the</strong> <strong>Earth</strong>’s surface can be represented by three types <strong>of</strong> second<br />
order spherical harmonics. The function are from left to right: zonal, tesseral and sectorial.<br />
The spherical surfaces are divided in sectors, where white represents rising <strong>of</strong> <strong>the</strong> surface and<br />
black lowering. The figure is adapted from Melchior [1978].<br />
2.2.2 Observation <strong>of</strong> Tides<br />
In many geophysical field observations <strong>the</strong> <strong>Earth</strong> tide is a source <strong>of</strong> noise, but it can also<br />
be useful in revealing informations <strong>of</strong> <strong>the</strong> interior <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. The advantage working<br />
with <strong>Earth</strong> tide data is <strong>the</strong> exact knowledge <strong>of</strong> <strong>the</strong> source function, i.e. <strong>the</strong> gravitation,<br />
<strong>the</strong>reby it is possible to detect small <strong>Earth</strong> tide signals by spectrum analysis, [Wang,<br />
1997]. Studying <strong>the</strong> gravity field <strong>of</strong> <strong>the</strong> <strong>Earth</strong> is a widely used area to reveal <strong>the</strong> tidal<br />
fields [Xu et al., 2004]. <strong>Earth</strong> tide observations compared with <strong>the</strong>oretical models may<br />
give important constraints to <strong>the</strong> understanding <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s internal structures. The<br />
<strong>the</strong>oretical tidal parameters (i.e. <strong>the</strong> dimensionless Love numbers which will be explained<br />
in Section 3.4), differ only with a factor <strong>of</strong> ±10 −2 from observations. The tides<br />
also effect <strong>the</strong> rotation and ellipticity <strong>of</strong> <strong>the</strong> <strong>Earth</strong> along with anelasticity and lateral<br />
heterogeneities, [Wang, 1997]. The structural effects <strong>of</strong> tides will be elaborated in fur<strong>the</strong>r<br />
details in Section 3.4.1.<br />
The solid earth tides are investigated by <strong>the</strong>oretical and numerical studies. These studies<br />
have been confirmed by observable measurements which can be both ground-based<br />
and space-based;<br />
Ground-based The ground-based measurements includes gravimeters, tiltmeters and<br />
strainmeters, and <strong>the</strong>y can only observe small-scale terrestrial measurements.<br />
Gravimeters are mainly elastic string gravimeters or superconducting gravimeters.<br />
The common feature is to measure <strong>the</strong> gravity and its variation in space and<br />
time. This is done by balancing <strong>the</strong> mean value <strong>of</strong> <strong>the</strong> gravity with a constant<br />
force equal to <strong>the</strong> mean value, and <strong>the</strong>n observe <strong>the</strong> small deviations. The elastic<br />
gravimeters consist <strong>of</strong> a tensional string which stores <strong>the</strong> energy. A definite<br />
force or torque is applied to support moving or vibrating masses or to indicate<br />
and control a load or torque. Superconducting gravimeters have improved <strong>the</strong><br />
precision <strong>of</strong> tidal measurements, and is today <strong>the</strong> most used instrument, [Latycheva<br />
et al., 2009]. The instrument consist <strong>of</strong> an aluminium sphere coated with<br />
lead. The sphere is placed in a magnetic field <strong>of</strong> two coils in liquid helium temperature,<br />
where <strong>the</strong> coils are carrying a current, and <strong>the</strong> force to keep <strong>the</strong> sphere<br />
in place is measured, [Melchior, 1978]. Obviously <strong>the</strong> tiltmeter and strainmeter
14 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
measures <strong>the</strong> tidal tilt and strain, which both are measurements <strong>of</strong> deformation,<br />
that allow <strong>the</strong> flexure <strong>of</strong> <strong>the</strong> crust to be observed, [Agnew, 1986]. For more<br />
information about modern instruments, see Boudin et al. [2008].<br />
Space-based The space-based observations includes satellites and VLBI 1 . It is maybe<br />
conceiving to call VLBI for space-based, because <strong>the</strong> telescope is based on <strong>the</strong><br />
ground, but it is anyways acting on space-based data. There exist three tidal effects<br />
which have to be considered in precise satellite tracking. They are 1.) direct<br />
perturbation <strong>of</strong> <strong>the</strong> satellite orbit, 2.) tidal deformation <strong>of</strong> oceans and solid <strong>Earth</strong><br />
and last 3.) gravitational effects <strong>of</strong> <strong>the</strong>se deformations on <strong>the</strong> satellite orbit. 1.)<br />
is <strong>the</strong> largest effect and can be model very accurate. The tidal deformation 2.) is<br />
difficult due to <strong>the</strong> insufficient knowledge <strong>of</strong> <strong>the</strong> inner structures <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. In<br />
satellite geodesy is is a pure geometric problem, measured with GPS 2 satellites,<br />
because <strong>of</strong> <strong>the</strong>ir precise positioning. They are very good at determine local displacements<br />
<strong>of</strong> both <strong>the</strong> solid <strong>Earth</strong> tides and <strong>the</strong> ocean tides. The gravitational<br />
effect 3.) can be calculated for near-<strong>Earth</strong> satellites. The degree <strong>of</strong> perturbation<br />
by <strong>the</strong> solid <strong>Earth</strong> tides is much stronger for low orbiting satellites. The satellite<br />
STARLETTE 3 is very suitable for studying <strong>the</strong> solid <strong>Earth</strong> tides, because <strong>of</strong> its<br />
low orbit and small size compared to its mass, which makes it more sensitive<br />
to gravitational attraction. GPS can also be used for this purpose, though because<br />
<strong>of</strong> a larger orbit <strong>the</strong>y are not perturbed as much as STARLETTE. In VLBI<br />
data <strong>the</strong>re are observed diurnal and semidiurnal variations due to <strong>the</strong> solid <strong>Earth</strong><br />
tides, [Seeber, 2003].<br />
Analysis <strong>of</strong> <strong>the</strong> <strong>Earth</strong> tides are normally carried out to determine <strong>the</strong> tidal parameters<br />
i.e. <strong>the</strong> frequency transfer function, see Section 3.4.1, between <strong>the</strong> <strong>Earth</strong>, station and<br />
sensor. The frequency transfer function is usually found by a least squares adjustment<br />
technique, which will not be discussed fur<strong>the</strong>r in this <strong>the</strong>sis. Observations <strong>of</strong>ten<br />
contain a drift, which is a long periodic signal caused by <strong>the</strong> instrument or by meteorological<br />
effects, [Wenzel, 1997].<br />
As it was elucidated in this chapter, <strong>the</strong> celestial bodies attractions or gravitation <strong>of</strong><br />
<strong>the</strong> <strong>Earth</strong> mainly from <strong>the</strong> Moon and <strong>the</strong> Sun changes with <strong>the</strong> inverse square <strong>of</strong> <strong>the</strong><br />
distance to <strong>the</strong> Moon or Sun. The lunar-solar attraction is responsible for <strong>the</strong> orbital<br />
motion <strong>of</strong> <strong>the</strong> <strong>Earth</strong> as well as <strong>the</strong> almost periodic tidal motion. The tidal acceleration,<br />
Equations (2.2.6) and (2.2.7), has been derived, and <strong>the</strong> assumption <strong>of</strong> rotation<br />
symmetry was defined. The next chapter deals with <strong>the</strong> <strong>the</strong>ory <strong>of</strong> elasticity and <strong>the</strong><br />
structural effects in <strong>the</strong> <strong>Earth</strong> due to <strong>the</strong> tidal force.<br />
1 Very Long Baseline Interferometry<br />
2 Global Positioning System<br />
3 French passive satellite
Chapter 3<br />
Theory <strong>of</strong> Elasticity<br />
Elasticity <strong>the</strong>ory is a ma<strong>the</strong>matical framework to understand how materials deform<br />
when <strong>the</strong>y are squeezed, stretched or sheared by an external force. Materials can behave<br />
elastic, plastic or viscose. Elastic materials deform when a force is applied and<br />
return to original shape when <strong>the</strong> force is removed. A fluid’s resistance to deformation<br />
by shear stress or normal stresses is called viscosity, and describes <strong>the</strong> fluid’s resistance<br />
to flow. The study <strong>of</strong> flow and deformation is known as rheology. Viscoelasticity is<br />
defined as <strong>the</strong> capacity <strong>of</strong> materials that both have viscous and elastic characteristics<br />
when <strong>the</strong>y deform. Viscoelastic materials behave as elastic solids on short times scale<br />
and as viscous fluids on long time scales. Viscous materials resist shear and strain,<br />
while elastic materials are affected by strain immediately when <strong>the</strong>y are stretched, but<br />
return to original state when <strong>the</strong> stress is removed. Viscoelastic materials have both<br />
properties. The deformation <strong>of</strong> materials with plastic behavior have a permanent and<br />
nonrecoverable deformation. Most materials are both elastic, plastic and viscose depending<br />
on <strong>the</strong> time scales. All solids flow to a small extent by response <strong>of</strong> shear stress,<br />
this suggest that solids are high viscose liquids, but even for small stresses solids are<br />
elastic. On <strong>the</strong> o<strong>the</strong>r hand fluids are not elastic and <strong>the</strong>y will always be characterized<br />
as having low-stress behavior. In this work all materials in question is assumed to be<br />
purely elastic, [Fowler, 1990].<br />
There are two types <strong>of</strong> forces that applies to a solid body; body forces and surface<br />
forces. The magnitude <strong>of</strong> <strong>the</strong> body force on an element is directly proportional to its<br />
volume (for example <strong>the</strong> tidal force). In <strong>Earth</strong>, rocks become more dense with depth,<br />
as <strong>the</strong> pressure increases, due to a pressure dependency. Surface forces acts across a<br />
surface element. The magnitude is directly proportional to <strong>the</strong> area <strong>of</strong> <strong>the</strong> surface in<br />
action. Surface forces are pressure (normal component) or stress forces (tangentially<br />
component), [Turcotte and Schubert, 1982].<br />
A body can undergo two types <strong>of</strong> motion. The first type is translation and rotation,<br />
which changes <strong>the</strong> whole body. The second form is straining and internal deformation,<br />
[Lay and Wallace, 1995]. It is only <strong>the</strong> second form that is <strong>of</strong> interest here, rotation<br />
symmetry is assumed.
16 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
3.1 Stress<br />
The distribution <strong>of</strong> stress depends on <strong>the</strong> external force applied to <strong>the</strong> body as well as<br />
<strong>the</strong> type <strong>of</strong> <strong>the</strong> material in <strong>the</strong> body. The internal elastic stresses on a solid body acts<br />
both as a pressure-like force along <strong>the</strong> normal to a contact surface and as shear stress<br />
tangential to <strong>the</strong> surface <strong>the</strong>y act on. Toge<strong>the</strong>r <strong>the</strong> forces keeps <strong>the</strong> material from being<br />
pulled apart, [Lautrup, 2005].<br />
Stress is defined as force per unit area. An arbitrary point in a material in space is<br />
attracted by three planes making a total <strong>of</strong> nine component stress tensor. In matrix<br />
notation<br />
dFi =<br />
3<br />
j<br />
τi jds j ,<br />
where dF/ds is <strong>the</strong> force per unit area normal to <strong>the</strong> surface s, <strong>the</strong> first index denotes<br />
<strong>the</strong> surface in which <strong>the</strong> stresses act, <strong>the</strong> second index indicates <strong>the</strong> positive direction in<br />
which <strong>the</strong> stresses act. τi j is <strong>the</strong> stress field also called <strong>the</strong> stress tensor, it is symmetric<br />
and defined as<br />
⎛<br />
⎞<br />
τi j =<br />
⎜⎝<br />
τ11 τ12 τ13<br />
τ21 τ22 τ23<br />
τ31 τ32 τ33<br />
⎟⎠<br />
. (3.1.1)<br />
The diagonal terms τ11, τ22 and τ33 are <strong>the</strong> principal or normal stresses. They are directed<br />
outward, which mean that positive values are tensional and negative values are<br />
τθr<br />
τrλ<br />
τθθ<br />
τθλ<br />
Figure 3.1.1: Geometry <strong>of</strong> <strong>the</strong> stress tensor components <strong>of</strong> a spherical element. τrr, τθθ, τλλ are<br />
<strong>the</strong> normal stresses and are in <strong>the</strong> given figure compressive. τrθ = τθr, τrλ = τλr and τθλ = τθλ<br />
are <strong>the</strong> shear stresses.<br />
τrr<br />
τrθ<br />
τλr<br />
τλθ<br />
θ<br />
τλλ<br />
r<br />
λ
<strong>Deformation</strong> and Strain 17<br />
compressional. The <strong>of</strong>f-diagonal terms are <strong>the</strong> shear stresses. The hydrostatic pressure<br />
is given if <strong>the</strong> normal stresses equal each o<strong>the</strong>r, and <strong>the</strong>n no shear stress will exist. In<br />
spherical coordinates index 1 is <strong>the</strong> radius r, index 2 is <strong>the</strong> colatitude θ and index 3<br />
is <strong>the</strong> longitude λ. In Figure 3.1.1 <strong>the</strong> geometry <strong>of</strong> <strong>the</strong> stress tensor components are<br />
depicted for a spherical square element. The stresses are in this case compressional<br />
and τrθ = τθr, τrλ = τλr and τθλ = τθλ, [Lay and Wallace, 1995; Lautrup, 2005].<br />
3.2 <strong>Deformation</strong> and Strain<br />
<strong>Deformation</strong> within a solid body involves variation <strong>of</strong> <strong>the</strong> displacement field u(x, t) =<br />
ˆx ′ − ˆx, where x and t are <strong>the</strong> actual position and time, i.e. it is described with a<br />
Lagrangian representation. <strong>Deformation</strong> is changes in both length and angular distortions,<br />
this is expressed by spatial gradients <strong>of</strong> <strong>the</strong> displacement field, also called<br />
strains.<br />
Strains are as <strong>the</strong> stresses divided in normal strains and shear strains. Normal<br />
strains are fractional changes in distance and measures how elongated <strong>the</strong> body has<br />
become. Shear strains are rotations with respect to <strong>the</strong> surrounding material. The total<br />
strain <strong>of</strong> a body must be described in three dimensions which is represented with a nine<br />
component tensor. Just like <strong>the</strong> stress tensor. Three normal stresses in <strong>the</strong> diagonal and<br />
six shear stresses in <strong>of</strong>f-diagonal. In matrix notation, <strong>the</strong> strain tensor is given by<br />
ei j = 1<br />
2 (ui, j + u j,i) , (3.2.1)<br />
where i and j goes from 1-3, ui, j = ∂ui<br />
∂x j , and xi represent <strong>the</strong> coordinates, [Lay and<br />
Wallace, 1995].<br />
In spherical coordinates, see Figure 3.2.1, a deformation from point P to Q, can be<br />
given as,<br />
P(r, θ, λ) = Q(r + δr, θ + δθ, λ + δλ) .<br />
The displacement tensor can be represented by three vectors u(r, θ, λ) = (u, v, w),<br />
where u, v and w are <strong>the</strong> deformation in <strong>the</strong> r, θ, λ directions respectively. θ is <strong>the</strong> colatitude<br />
calculated from <strong>the</strong> pole towards equator, and λ <strong>the</strong> geocentric longitude. Following<br />
Hughes and Gaylord [1964] <strong>the</strong> spherical strain tensor components are given
18 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
by<br />
err = ∂u<br />
∂r<br />
(3.2.2)<br />
eθθ = 1 ∂v u<br />
+<br />
r ∂θ r<br />
(3.2.3)<br />
eλλ<br />
erλ<br />
=<br />
=<br />
1 ∂w u cot θ<br />
+ + v<br />
r sin θ ∂λ r r<br />
(3.2.4)<br />
1<br />
erθ =<br />
<br />
<br />
1 ∂u w ∂v<br />
− +<br />
2 r sin θ ∂λ r ∂r<br />
(3.2.5)<br />
1<br />
eλθ =<br />
<br />
1 ∂u v ∂v<br />
− +<br />
2 r ∂θ r ∂r<br />
(3.2.6)<br />
1<br />
<br />
1 ∂w w cot θ<br />
− +<br />
2 r ∂θ r<br />
1<br />
<br />
∂v<br />
r sin θ ∂λ<br />
.<br />
(3.2.7)<br />
Here err, eθθ, eλλ are <strong>the</strong> normal strains and adding <strong>the</strong>m gives <strong>the</strong> dilatation, and<br />
erλ, erθ, eλθ are <strong>the</strong> shear strains. The strain tensor in spherical coordinates can <strong>the</strong>n<br />
be derived to<br />
⎡<br />
e =<br />
⎢⎣<br />
∂u<br />
∂r <br />
1 ∂v/r 1 ∂u<br />
2 r ∂r +<br />
<br />
r ∂θ<br />
1 1 ∂u ∂(w/r<br />
2 r sin θ ∂λ + r ∂r<br />
X<br />
<br />
<br />
1 sin θ<br />
2 r<br />
<br />
1 ∂(v/r)<br />
2 r ∂r<br />
Z<br />
+ 1<br />
r<br />
<br />
∂u<br />
∂θ<br />
1 ∂v u<br />
r ∂θ + r<br />
∂(w/ sin θ)<br />
∂θ + 1<br />
<br />
∂v<br />
r sin θ ∂λ<br />
λ<br />
r sin θ<br />
θ δθ<br />
<br />
<br />
∂(w/r)<br />
+ r<br />
1 1 ∂u<br />
2 r sin θ ∂λ ∂r<br />
1 sin θ ∂(w/ sin θ)<br />
2 r ∂θ + 1<br />
<br />
∂v<br />
r sin θ ∂λ<br />
1 ∂w u cot θ<br />
r sin θ ∂λ + r + r v<br />
Figure 3.2.1: <strong>Deformation</strong> in spherical coordinates. The figure is adapted from Surfaces and<br />
Contact Mechanics [2005], but slightly modified.<br />
r<br />
δλ<br />
P<br />
rδθ<br />
Q<br />
δr<br />
Y<br />
⎤<br />
⎥⎦
Equation <strong>of</strong> Motion 19<br />
3.3 Equation <strong>of</strong> Motion<br />
The displacement for a continuous, homogeneous, isotropic, infinite, elastic medium<br />
is given by <strong>the</strong> Navier equation <strong>of</strong> motion<br />
(λ + µ)∇(∇ · u) + µ∇ 2 u + F = ρ ∂2 u<br />
∂t 2<br />
, (3.3.1)<br />
where u is <strong>the</strong> displacement, λ and µ are <strong>the</strong> Lamé coefficients, F represent <strong>the</strong> forces<br />
acting on a volume element also called <strong>the</strong> body forces, and ρ <strong>the</strong> density <strong>of</strong> <strong>the</strong> media<br />
in question, [Lautrup, 2005]. For an isotropic media, a media with no directional<br />
variation, <strong>the</strong> Lamé parameters are <strong>the</strong> only two independent elastic moduli, [Lay and<br />
Wallace, 1995]. In terms <strong>of</strong> <strong>the</strong> P-wave velocity α2 λ + 2µ<br />
= ρ which is a plane wave<br />
<strong>of</strong> compression and shear and <strong>the</strong> S-wave velocity β2 = µ ρ which is pure shear with no<br />
volume change, S-waves can not propagate in a fluid material. Equation (3.3.1) can be<br />
written as<br />
α 2 ∇(∇ · u) − β 2 ∇ × (∇ × u) + F<br />
ρ = ∂2 u<br />
∂t 2<br />
. (3.3.2)<br />
As described in Section 2, <strong>the</strong> attracting forces from <strong>the</strong> Moon and <strong>the</strong> Sun causes<br />
<strong>the</strong> orbiting <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and <strong>the</strong> tidal force as described in Section 2. The tidal time<br />
variation is a very slow process compared to <strong>the</strong> deformation <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and <strong>the</strong>refore<br />
<strong>the</strong> time dependent term in Equation (3.3.2) disappears. The wave equation becomes<br />
α 2 ∇(∇ · u) − β 2 ∇ × (∇ × u) = − F<br />
ρ<br />
. (3.3.3)<br />
It is important to notice, that <strong>the</strong> right-hand side is a volume force over <strong>the</strong> density<br />
which in fact is an acceleration, given in <strong>the</strong> units <strong>of</strong> m<br />
s2 .<br />
3.3.1 Boundary Conditions<br />
The boundary conditions are describing <strong>the</strong> external influence on <strong>the</strong> <strong>Earth</strong>’s surface<br />
and any boundaries inside <strong>the</strong> <strong>Earth</strong>. The boundary conditions are found by applying<br />
continuity <strong>of</strong> stress and displacement on <strong>the</strong> internal boundaries and <strong>the</strong> free surface.<br />
In this study, internal boundaries are disregarded. At <strong>the</strong> free surface (r = a), only <strong>the</strong><br />
tractions (τrr, τrθ, τrλ) are constrained, not <strong>the</strong> o<strong>the</strong>r components <strong>of</strong> stress, [Hurford<br />
and Greenberg, 2002]. The boundary conditions are<br />
τrr =<br />
τrθ =<br />
τrλ =<br />
E<br />
(1 + σ)(1 − 2 σ) [(1 − σ)err + σ(eθθ + eλλ)] = 0 (3.3.4)<br />
E<br />
1 + σ erθ = 2µerθ = 0 (3.3.5)<br />
E<br />
1 + σ erλ = 2µerλ = 0 , (3.3.6)<br />
where e is <strong>the</strong> strain defined in Section 3.2, σ = λ<br />
2(λ + µ)<br />
is called <strong>the</strong> Poisson ratio<br />
and is <strong>the</strong> ratio <strong>of</strong> <strong>the</strong> lateral contraction strain to <strong>the</strong> longitudinal extension strain.
20 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
µ(3λ + 2µ)<br />
E =<br />
λ + µ<br />
is Young’s modulus or <strong>the</strong> elastic modulus and is <strong>the</strong> stiffness <strong>of</strong><br />
an isotropic, elastic material and is defined as <strong>the</strong> ratio <strong>of</strong> extensional stress over <strong>the</strong><br />
extensional strain <strong>of</strong> for example a cylinder being pulled on both ends, [Hughes and<br />
Gaylord, 1964; Shearer, 1999].<br />
Because <strong>of</strong> <strong>the</strong> constrain <strong>of</strong> rotation symmetry, <strong>the</strong> displacement is confined to <strong>the</strong><br />
rθ-plane, i.e. w = 0 and <strong>the</strong>refore getting ∂i/∂λ = 0. τrr is <strong>the</strong> plane tangent to <strong>the</strong><br />
radius <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and τrθ is <strong>the</strong> plane tangent to <strong>the</strong> latitude <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. This can<br />
also be expressed by <strong>the</strong> displacement coordinates and after some reduction, we get<br />
τrr =<br />
<br />
E<br />
(1 − σ)<br />
(1 + σ)(1 − 2 σ)<br />
∂u<br />
<br />
+ σ 2<br />
∂r u θ<br />
+ vcot<br />
r r<br />
+ 1<br />
r<br />
<br />
∂v<br />
= 0<br />
∂θ<br />
(3.3.7)<br />
<br />
1 ∂u v ∂v<br />
τrθ = µ − + = 0 , (3.3.8)<br />
r ∂θ r ∂r<br />
where u = (u, v) is <strong>the</strong> displacement described in Section 3.2. u is <strong>the</strong> radial displacement<br />
component and v is <strong>the</strong> tangential component <strong>of</strong> <strong>the</strong> displacement. Both<br />
are working in <strong>the</strong> domain r = a, where a is <strong>the</strong> radius <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and 0 ≤ θ ≤ 2π<br />
is <strong>the</strong> colatitude. Notice that <strong>the</strong> angle, θ in Equations (3.3.7) and (3.3.8) is not <strong>the</strong><br />
same as <strong>the</strong> angle ψ in Equation (2.2.4), given in Section 2.2.1. θ is calculated positive<br />
from <strong>the</strong> Pole axis towards <strong>the</strong> Equator, which is opposite <strong>the</strong> angle, ψ. ψ is <strong>the</strong> angle<br />
between <strong>the</strong> tidal force’s attacking point and <strong>the</strong> line <strong>of</strong> <strong>the</strong> attracting body line. In <strong>the</strong><br />
<strong>Earth</strong>-Moon system, where <strong>the</strong> Moon is located in <strong>the</strong> equator line ψ = θ = 90 ◦ − θ.<br />
All angles are depicted in Figure 1.1.1 in Section 1. There can be no normal stresses at<br />
<strong>the</strong> surface, hence τrr = 0, and τrθ = 0 because no shear stress can exist along <strong>the</strong> free<br />
surface <strong>of</strong> a distorted sphere, [Lay and Wallace, 1995]. Fur<strong>the</strong>r note that <strong>the</strong> derivative<br />
with respect to λ has vanished.<br />
These equations can also be explained physically by looking at a simple example<br />
where a traction is applied perpendicular on both ends <strong>of</strong> a rod. The rod will be<br />
stretched out in <strong>the</strong> length and thinned in <strong>the</strong> width. The same is valid for a volume<br />
element in <strong>the</strong> <strong>Earth</strong>, i.e. when <strong>the</strong> <strong>Earth</strong> is affected by a radial traction <strong>of</strong> <strong>the</strong> Moon, it<br />
not only gives a radial displacement on <strong>the</strong> attracted volume element but also a lateral<br />
displacement.
<strong>Tidal</strong> <strong>Deformation</strong> 21<br />
3.4 <strong>Tidal</strong> <strong>Deformation</strong><br />
The solid <strong>Earth</strong>’s response to rotation and ellipticity was first described by Love [1911].<br />
It was a simple model where <strong>the</strong> <strong>Earth</strong> was assume to be incompressible and homogeneous.<br />
The tidal deformation in a homogeneous spherical symmetric <strong>Earth</strong>, can be<br />
written as components <strong>of</strong> <strong>the</strong> displacement vector u = (u, v, w)<br />
u(r, ψ, λ) =<br />
v(r, ψ, λ) =<br />
w(r, ψ, λ) =<br />
1<br />
g(r)<br />
1<br />
g(r)<br />
1<br />
g(r)<br />
∞<br />
Hn(r) Φn<br />
n=2<br />
∞<br />
Ln(r)<br />
n=2<br />
∂Φn<br />
∂ψ<br />
∞ ∂Φn<br />
Ln(r)<br />
sin ψ ∂λ<br />
n=2<br />
(3.4.1)<br />
(3.4.2)<br />
. (3.4.3)<br />
Φn is <strong>the</strong> tidal potential given in Equation (2.2.4), g(r) is <strong>the</strong> acceleration <strong>of</strong> gravity,<br />
described in Section 2.1.3, and Hn and Ln are <strong>the</strong> amplitudes connecting <strong>the</strong> Love numbers<br />
hn, kn, ln at <strong>the</strong> surface <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, [Melchior, 1978]. The Love numbers h, k, l are<br />
dimensionless numbers used as a representation <strong>of</strong> <strong>the</strong> deformation phenomena which<br />
is produced by a potential perturbed by spherical harmonics. The Love numbers are<br />
a measure <strong>of</strong> how solid <strong>the</strong> <strong>Earth</strong> is, and <strong>the</strong>y change as <strong>the</strong> order <strong>of</strong> <strong>the</strong> potential<br />
changes. h describes <strong>the</strong> elastic radial displacements <strong>of</strong> <strong>the</strong> crustal surface relative to<br />
<strong>the</strong> ellipsoid, k describes <strong>the</strong> modification <strong>of</strong> <strong>the</strong> potential caused by <strong>the</strong> deformation<br />
and l describes <strong>the</strong> elastic tangential displacement. Because <strong>of</strong> rotation and ellipticity<br />
it is difficult to verify <strong>the</strong> latitudinal dependencies <strong>of</strong> <strong>the</strong> Love numbers. For a rigid<br />
body h = k = l = 0, but <strong>the</strong> <strong>Earth</strong> is elastic and <strong>the</strong> Love numbers are in <strong>the</strong> range<br />
0.304 ≤ k ≤ 0.312<br />
0.616 ≤ h ≤ 0.624<br />
0.088 ≤ l ≤ 0.084 .<br />
This scale is <strong>of</strong> course depending on <strong>the</strong> <strong>Earth</strong> model. Normally <strong>the</strong> largest tidal contribution<br />
n = 2 is on <strong>the</strong> surface averaged to h = 0.6, k = 0.6 and l = 0.08, [Varga, 1974].<br />
Inserting <strong>the</strong> tidal potential and calculating <strong>the</strong> derivatives, <strong>the</strong> radial and tangential<br />
displacement in polar coordinates, become<br />
u(r, ψ) = h<br />
g<br />
GMr 2<br />
2r 3 i<br />
(3 cos 2 ψ − 1) (3.4.4)<br />
v(r, ψ) = l 3GMr<br />
g<br />
2<br />
2r3 sin 2ψ . (3.4.5)<br />
i<br />
The <strong>Earth</strong>’s radial displacement field is compressive and expansive. In a <strong>Earth</strong>-Moon<br />
system, <strong>the</strong> maximum expansion happens at <strong>the</strong> <strong>Earth</strong>-Moon line, and for <strong>the</strong> Moon<br />
aligned with <strong>the</strong> equatorial axis, <strong>the</strong> maximum expansion is at <strong>the</strong> Equator and maximum<br />
compression at <strong>the</strong> Pole. The horizontal or tangential displacement field is tractive,<br />
with maximum value at mid-latitude. By expanding <strong>the</strong> Legendre polynomials
22 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
in <strong>the</strong> tidal potential, Equation (2.2.4), into spherical harmonics, <strong>the</strong> tides can be expressed<br />
with <strong>the</strong> Love numbers. This is by far <strong>the</strong> most used form, when studying <strong>the</strong><br />
deformation <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. The spherical harmonics can be computed to high degree.<br />
This is among o<strong>the</strong>rs described by Love [1944]; Takeuchi and Saito [1962]; Varga and<br />
Grafarend [1996]; Wenzel [1997].<br />
3.4.1 Structural Effects on <strong>Earth</strong> Tides<br />
Longman [1962] and Farrell [1972] were some <strong>of</strong> <strong>the</strong> first to compute realistic <strong>Earth</strong><br />
models for <strong>the</strong> tidal deformation. The models applied to a non-rotating, spherically<br />
layered, elastic and self-gravitating <strong>Earth</strong>. Self-gravitating states, that when a deformation<br />
in <strong>the</strong> interior occurs, <strong>the</strong> mass is redistributed, and this causes a change in <strong>the</strong><br />
gravity field and influence <strong>the</strong> deformation again, [Wang, 1997]. Longman [1962] used<br />
a Greens function, which is a method to solve inhomogeneous differential equations<br />
with boundary problems, to model <strong>the</strong> surface deformation on <strong>Earth</strong> under a surface<br />
mass load and <strong>the</strong> perturbations in <strong>the</strong> gravity field. The methods were made realistic<br />
by applying <strong>the</strong> Love numbers. Farrell [1972] studied <strong>the</strong> elastic deformation in a<br />
elastic stratified half-space.<br />
The geophysical properties <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s interior are deduced mainly by seismology<br />
and low frequency normal modes. Rheological properties are found by glacial<br />
rebound. Rheology combine elements <strong>of</strong> elasticity, viscosity and plasticity, and it relate<br />
<strong>the</strong> flow and deformation behavior <strong>of</strong> material and its internal structure. <strong>Earth</strong><br />
tides is <strong>the</strong> only tool to explore <strong>the</strong> frequencies in between.<br />
In a model from <strong>the</strong> fifties it was shown by Jeffreys and Vicente [1957], that <strong>the</strong><br />
<strong>Earth</strong>’s fluid core responded with a resonance at diurnal frequencies. This is caused by<br />
a normal mode <strong>of</strong> <strong>the</strong> rotating <strong>Earth</strong>, and can not be recovered from seismological interpretation.<br />
This resonant behavior is called The nearly- diurnal free wobble-resonance,<br />
[Zürn, 1997], because it is in frequency with <strong>the</strong> diurnal tides. The resonance arises<br />
when <strong>the</strong> axis <strong>of</strong> <strong>the</strong> mantle and outer core is slightly separated. The reacting <strong>of</strong> <strong>the</strong><br />
realignment causes <strong>the</strong> wobbling effect. There do also exist an o<strong>the</strong>r wobbling effect<br />
called <strong>the</strong> Chandler wobble named after <strong>the</strong> discoverer S. C Chandler (1891). It arises<br />
because <strong>the</strong> rotational axes deviates by a small angle from <strong>the</strong> symmetry axis, <strong>the</strong> axis<br />
<strong>of</strong> maximum moment <strong>of</strong> inertia. A prograde precession exert as <strong>the</strong> torque acting on<br />
<strong>the</strong> angular momentum and causes <strong>the</strong> pole tides, <strong>the</strong> only tidal effect not arisen by<br />
external bodies. The deformation arising from this effect, adjusts <strong>the</strong> equatorial bulge<br />
towards symmetry, [Stacey and Davis, 2008].<br />
The mantle is also influenced by <strong>the</strong> solid <strong>Earth</strong> tides, by impacting <strong>the</strong> heterogeneity<br />
<strong>of</strong> <strong>the</strong> mantle’s convective flow. The viscoelasticity <strong>of</strong> <strong>the</strong> mantle causes a<br />
frequency dependence <strong>of</strong> <strong>the</strong> tidal parameters. Lateral heterogeneities in <strong>the</strong> mantle<br />
can be estimated by Love-Li or Molodensky-Wang’s perturbations, but that is beyond<br />
<strong>the</strong> scoop <strong>of</strong> this <strong>the</strong>sis. A reliable mantle model by Wang [1997] illustrates that <strong>the</strong><br />
large scale mantle heterogeneity only gives small anomalies in <strong>the</strong> <strong>Earth</strong> tides, and<br />
amplitude increase and phase delay <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s response functions.<br />
Local influences <strong>of</strong> <strong>Earth</strong> tides are not fully understood. Strain and tilt measurements<br />
are vigorously dominated by fault zones and cavities near <strong>the</strong> instruments. To<br />
deduce <strong>the</strong> local inhomogeneities is a hard task because <strong>of</strong> <strong>the</strong> geology and compli-
<strong>Tidal</strong> <strong>Deformation</strong> 23<br />
cated deformation.<br />
This chapter rounds <strong>of</strong>f <strong>the</strong> fundamental <strong>the</strong>ory behind <strong>the</strong> <strong>the</strong>sis. The steady state<br />
version <strong>of</strong> <strong>the</strong> Navier equation <strong>of</strong> motion, Equation (3.3.3) was introduced. Assumptions<br />
about <strong>the</strong> elasticity <strong>of</strong> <strong>the</strong> <strong>Earth</strong> has been defined, and <strong>the</strong> analytic solution <strong>of</strong><br />
<strong>the</strong> tidal displacement at <strong>the</strong> surface layer, <strong>the</strong> Equations (3.4.4) and (3.4.5), has been<br />
introduced.
Chapter 4<br />
Finite Difference Equations<br />
This Section describes <strong>the</strong> numerical methods <strong>of</strong> Finite Difference Equations (FDE)<br />
which is used for solving differential equations. The basic <strong>the</strong>ory is adapted from<br />
Kreyszig [1999], but expanded from Cartesian to polar coordinates. The aim <strong>of</strong> this<br />
section is to perturb <strong>the</strong> two-dimensional Navier equation <strong>of</strong> motion, Equation (3.3.3)<br />
described in Section 3.3, into a FDE formulation.<br />
4.1 The Centered Finite Difference Scheme<br />
The Navier equation <strong>of</strong> motion is a hyperbolic partial differential equation, verified in<br />
Appendix A, and applied on <strong>the</strong> spherical <strong>Earth</strong>, it leads to a boundary value problem.<br />
With <strong>the</strong> vector identity ∇ × (∇ × u) = ∇(∇ · u) − ∇ 2 u, <strong>the</strong> Equation (3.3.3) can be<br />
written as<br />
− F<br />
ρ = (α2 − β 2 )∇(∇·u) + β 2 ∇ 2 u . (4.1.1)<br />
In this case <strong>the</strong> left-hand side equals <strong>the</strong> tidal acceleration. The two dimensional tidal<br />
acceleration components are given by <strong>the</strong> Equations (2.2.6) and (2.2.7). When writing<br />
out <strong>the</strong> coordinates (r, θ, λ), for <strong>the</strong> radial (u) and tangential (v) system, where λ = 0<br />
and ∂λu = 0 (because <strong>of</strong> <strong>the</strong> rotation symmetry).
26 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
− Fu<br />
ρ<br />
− Fv<br />
ρ<br />
= (α 2 − β 2 <br />
)<br />
+β 2<br />
− 1 1 ∂u<br />
u +<br />
r2 r ∂r + ∂2u 1<br />
+<br />
∂r2 r2 ∂u<br />
∂θ<br />
<br />
1 ∂u<br />
r ∂r + ∂2u 1<br />
+<br />
∂r2 r2 ∂2u ∂θ2 <br />
= − α2 − β2 r2 u + α2<br />
r ur + α 2 urr<br />
+ α2 − β2 r2 uθ + β2<br />
r2 uθθ + α2 − β2 uθr<br />
r<br />
= (α 2 − β 2 <br />
) − 1<br />
r2 ∂v 1<br />
+<br />
∂θ r2 ∂2v 1<br />
+<br />
∂θ2 r<br />
+β 2<br />
<br />
1 ∂v<br />
r ∂r + ∂2v 1<br />
+<br />
∂r2 r2 ∂2v ∂θ2 <br />
=<br />
β 2<br />
r vr + β 2 vrr − α2 − β 2<br />
r 2<br />
vθ + α2<br />
vθθ<br />
r2 ∂2v <br />
∂r∂θ<br />
+ 1<br />
r<br />
∂2u <br />
∂θ∂r<br />
(4.1.2)<br />
(4.1.3)<br />
(4.1.4)<br />
+ α2 − β2 vrθ . (4.1.5)<br />
r<br />
The second equality sign is written with index notation, where subscripts r, rr denotes<br />
<strong>the</strong> first and second derivatives <strong>of</strong> u and v with respect to r, subscripts θ, θθ denotes<br />
<strong>the</strong> first and second derivatives <strong>of</strong> u and v with respect to θ and θr, rθ denotes <strong>the</strong> second<br />
order mixed partial derivative <strong>of</strong> u and v with respect to r and θ. Fu<br />
ρ = atides r and<br />
Fv<br />
ρ = atides θ .<br />
The following will only be carried out for <strong>the</strong> radial system, u, because <strong>the</strong> deduction<br />
<strong>of</strong> v would be trivial. Taylor expansion <strong>of</strong> u(r, θ) to <strong>the</strong> second order, gives for <strong>the</strong><br />
radial direction<br />
u(r + hr, θ) = u(r, θ) + hr ur(r, θ) + 1<br />
2 h2 r urr(r, θ) (4.1.6)<br />
u(r − hr, θ) = u(r, θ) − hr ur(r, θ) + 1<br />
2 h2 r urr(r, θ), (4.1.7)<br />
where hr is an infinitesimal change in <strong>the</strong> radial direction. Subtracting Equation (4.1.6)<br />
from Equation (4.1.7), and solving for <strong>the</strong> first derivative <strong>of</strong> r, ur gives<br />
ur(r, θ) = u(r + hr, θ) − u(r − hr, θ)<br />
2hr<br />
. (4.1.8)<br />
This is called a centered finite difference formula for <strong>the</strong> first order derivative <strong>of</strong> r. By<br />
addition <strong>of</strong> Equation (4.1.6) and (4.1.7), <strong>the</strong> second order centered finite difference <strong>of</strong>
The Centered Finite Difference Scheme 27<br />
r, resolves to<br />
urr(r, θ) = u(r + hr, θ) + u(r − hr, θ) − 2 u(r, θ)<br />
h 2 r<br />
. (4.1.9)<br />
The same procedure is carried out for <strong>the</strong> colatitudinal direction. This gives, <strong>the</strong> centered<br />
finite differences for <strong>the</strong> first and second order derivatives <strong>of</strong> θ<br />
uθ(r, θ) = u(r, θ + hθ) − u(r, θ − hθ)<br />
2 hθ<br />
uθθ(r, θ) = u(r, θ + hθ) + u(r, θ − hθ) − 2 u(r, θ)<br />
h2 θ<br />
(4.1.10)<br />
. (4.1.11)<br />
The mixed derivative urθ and uθr equals each o<strong>the</strong>r, [Spiegel and Liu, 1999], and this<br />
continuity was assumed in Section 3.3. The mixed terms are derived by <strong>the</strong> Taylor<br />
expansion at (r, θ) in <strong>the</strong> radial and <strong>the</strong> latitudinal direction. The second order Taylor<br />
expansion for functions <strong>of</strong> two variables gives in total four equations<br />
u(r + hr, θ + hθ) = u(r, θ) + hr ur(r, θ) + hθ uθ(r, θ)<br />
+ 1<br />
2 [h2 r urr(r, θ) + 2hrhθ urθ(r, θ)<br />
+ h 2 θ uθθ(r, θ)] (4.1.12)<br />
u(r + hr, θ − hθ) = u(r, θ) + hr ur(r, θ) − hθ uθ(r, θ)<br />
+ 1<br />
2 [h2 r urr(r, θ) + 2hrhθ urθ(r, θ)<br />
+ h 2 θ uθθ(r, θ)] (4.1.13)<br />
u(r − hr, θ + hθ) = u(r, θ) − hr ur(r, θ) + hθ uθ(r, θ)<br />
+ 1<br />
2 [h2 r urr(r, θ) + 2hrhθ urθ(r, θ)<br />
+ h 2 θ uθθ(r, θ)] (4.1.14)<br />
u(r − hr, θ − hθ) = u(r, θ) − hr ur(r, θ) − hθ uθ(r, θ)<br />
+ 1<br />
2 [h2 r urr(r, θ) + 2hrhθ urθ(r, θ)<br />
+ h 2 θ uθθ(r, θ)] . (4.1.15)<br />
Solving for <strong>the</strong> mixed derivative by adding Equations (4.1.12) and (4.1.15) and subtracting<br />
Equations (4.1.13) and (4.1.14), gives<br />
urθ = u(r + hr, θ + hθ) − u(r + hr, θ − hθ) − u(r − hr, θ + hθ) + u(r − hr, θ − hθ)<br />
4 hr hθ<br />
.<br />
(4.1.16)
28 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
This last centered finite difference scheme closes Section 4.1. Before Section 4.3,<br />
where <strong>the</strong> finite difference equation formulas are discretized, Section 4.2 elucidate <strong>the</strong><br />
errors arising by <strong>the</strong> Taylor expansion.<br />
4.2 Error Estimates and Stability <strong>of</strong> FDE<br />
When writing up <strong>the</strong> central differences, an approximation was made to <strong>the</strong> actual<br />
solution.<br />
Truncation Error is defined as <strong>the</strong> difference between <strong>the</strong> governing differential<br />
equation and <strong>the</strong> approximation to <strong>the</strong> governing equation. This is in contrast to <strong>the</strong><br />
solution error, which is <strong>the</strong> difference between <strong>the</strong> exact solution and <strong>the</strong> approximate<br />
solution.<br />
Following Holmes [2007], <strong>the</strong> local truncation error for <strong>the</strong> central finite difference<br />
schemes for ur and uθ are<br />
τ i r = − h2 r<br />
6 urrr(ηi, θ j) (4.2.1)<br />
τ j<br />
θ = −h2 θ<br />
6 uθθθ(ri, η j) . (4.2.2)<br />
For urr and uθθ <strong>the</strong> truncation error are<br />
τ i rr = − h2 r<br />
12 urrrr( ¯ηi, θ j) (4.2.3)<br />
τ j<br />
θθ = − h2 θ<br />
12 uθθθθ(ri, ¯η j) . (4.2.4)<br />
For urθ and uθr, <strong>the</strong> truncation error is not known. It will in <strong>the</strong> following be called<br />
i j<br />
τ<br />
rθ and τ<br />
i j<br />
θr<br />
. (4.2.5)<br />
The η’s are between ri−1 and ri+1 or between θ j−1 and θ j+1. The total truncation error<br />
τu i <strong>of</strong> <strong>the</strong> inner system <strong>of</strong> u, is given by<br />
τ u i<br />
= −α2<br />
r<br />
− α2 − β 2<br />
r 2<br />
− α2 − β 2<br />
h2 r<br />
6 urrr(ηi, θ j) − α 2 h2r 12 urrrr( ¯ηi, θ j)<br />
h 2 θ<br />
6 uθθθ(ri, η j) − β2<br />
r 2<br />
i j<br />
τrθ h 2 θ<br />
12 uθθθθ(ri, ¯η j)<br />
r<br />
. (4.2.6)<br />
The total truncation error τv i <strong>of</strong> <strong>the</strong> inner system <strong>of</strong> v, is given by<br />
τ v i<br />
= −β2<br />
r<br />
− α2 − β 2<br />
r 2<br />
− α2 − β 2<br />
r<br />
h2 r<br />
6 vrrr(ηi, θ j) − β 2 h2r 12 vrrrr( ¯ηi, θ j)<br />
h 2 θ<br />
6 vθθθ(ri, η j) − α2<br />
r 2<br />
i j<br />
τθr h 2 θ<br />
12 uθθθθ(ri, ¯η j)<br />
. (4.2.7)
Error Estimates and Stability <strong>of</strong> FDE 29<br />
The truncation reduces when hr and hθ reduces. If <strong>the</strong> partial difference equation is<br />
approximated consistently, it is not preposterous to think that <strong>the</strong> solution is stable,<br />
and <strong>the</strong>refore converge to <strong>the</strong> exact solution in a certain time step. The finite difference<br />
scheme is consistent if<br />
lim τ(r, θ) = 0 , where l = max{hr, hθ} .<br />
l→0<br />
Therefore <strong>the</strong> discretization is consistent and second order accurate, but consistency is<br />
not enough, <strong>the</strong> system also needs to be stable to get a reliable solution. The Equations<br />
(4.3.5)-(4.3.6), can be written as <strong>the</strong> linear system<br />
Ax = F.<br />
A is <strong>the</strong> coefficient matrix, x is representing <strong>the</strong> displacements and <strong>the</strong> vector F contains<br />
<strong>the</strong> body forces and <strong>the</strong> boundary constrains with <strong>the</strong> traction free surface and<br />
<strong>the</strong> symmetry at <strong>the</strong> polar and equatorial axes. This equation will be discussed more<br />
throughly in next chapter. If <strong>the</strong> solution is consistent and stable, <strong>the</strong> finite difference<br />
equations is convergent, [Holmes, 2007].<br />
It was easy to see that <strong>the</strong> solution was consistent, but <strong>the</strong> coefficient matrix A has<br />
to be studied more thoroughly to determine <strong>the</strong> stability <strong>of</strong> <strong>the</strong> system, this is carried<br />
out in Chapter 7. This does not describe <strong>the</strong> difference between <strong>the</strong> numerical and <strong>the</strong><br />
exact solution, which <strong>of</strong> course is <strong>the</strong> most direct measure, <strong>of</strong> how good <strong>the</strong> numerical<br />
solution is. A complete numerical solution <strong>of</strong> this system is not derived, but <strong>the</strong><br />
computed results are compared with <strong>the</strong> surface displacements given by <strong>the</strong> Equations<br />
(3.4.4) and (3.4.5) in Section 3.4.
30 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
4.3 The Finite Difference Formula<br />
In this section <strong>the</strong> newly achieved truncation error is dropped, and <strong>the</strong> system is<br />
changing from being exact to being an approximation. Substituting Equations (4.1.8)-<br />
(4.1.11) and (4.1.16) into <strong>the</strong> radial Equation (4.1.3), gives after some calculations<br />
− Fu(r, θ)<br />
ρ<br />
=<br />
<br />
1 α<br />
r<br />
2 (2r + hr)<br />
2 h2 <br />
u(r + hr, θ)<br />
r<br />
<br />
1 α<br />
+<br />
r<br />
2 (2r − hr)<br />
2 h2 <br />
u(r − hr, θ)<br />
r<br />
⎡<br />
+ ⎢⎣ 1<br />
⎤<br />
⎥⎦ u(r, θ + hθ)<br />
+<br />
r 2<br />
⎡<br />
⎢⎣ 1<br />
2β 2 + (α 2 − β 2 ) hθ<br />
2h 2 θ<br />
r2 2β2 − (α2 − β2 ) hθ<br />
2h2 θ<br />
⎤<br />
⎥⎦ u(r, θ − hθ)<br />
+<br />
<br />
1 α<br />
r<br />
2 − β2 <br />
[u(r + hr, θ + hθ) − v(r + hr, θ − hθ)<br />
4 hr hθ<br />
−<br />
−<br />
u(r − hr, θ + hθ) + u(r − hr, θ − hθ)]<br />
⎡<br />
⎢⎣ α2 − β2 r2 + 2 α2<br />
⎤<br />
⎥⎦ u(r, θ) . (4.3.1)<br />
h 2 r<br />
+ 2 β2 )<br />
r 2 h 2 θ<br />
The same is carried out for <strong>the</strong> tangential system, Equation (4.1.5)<br />
− Fv(r, θ)<br />
ρ<br />
=<br />
+<br />
+<br />
+<br />
<br />
1 β<br />
r<br />
2 (2r + hr)<br />
2 h2 <br />
v(r + hr, θ)<br />
r<br />
<br />
1 β<br />
r<br />
2 (2r − hr)<br />
2 h2 <br />
v(r − hr, θ)<br />
r<br />
⎡<br />
⎢⎣ 1<br />
r2 2 α2 − (α2 − β2 )hθ<br />
2h2 ⎤<br />
⎥⎦ v(r, θ + hθ)<br />
θ<br />
⎡<br />
⎢⎣ 1 2 α2 + (α2 − β2 ⎤<br />
)hθ<br />
⎥⎦ v(r, θ − hθ)<br />
r 2<br />
2h 2 θ<br />
+<br />
<br />
1 α<br />
r<br />
2 − β2 <br />
[v(r + hr, θ + hθ) − v(r + hr, θ − hθ)<br />
4 hr hθ<br />
−<br />
−<br />
v(r − hr, θ + hθ) + v(r − hr, θ − hθ)]<br />
⎡<br />
2 β2<br />
⎢⎣ + 2 α2<br />
⎤<br />
⎥⎦ v(r, θ) . (4.3.2)<br />
h 2 r<br />
r 2 h 2 θ<br />
These equations are differential equations corresponding to <strong>the</strong> Navier equation <strong>of</strong><br />
motion, Equation (4.1.1), where hr, and hθ are <strong>the</strong> mesh sizes in <strong>the</strong> radial and <strong>the</strong><br />
latitudinal planes, respectively. At <strong>the</strong> point (r, θ), u equals <strong>the</strong> mean values <strong>of</strong><br />
⎡<br />
⎢⎣ α2 − β2 r2 + 2 α2<br />
h2 +<br />
r<br />
2β2<br />
⎤<br />
⎥⎦<br />
r 2 h 2 θ
The Finite Difference Formula 31<br />
✧ ✧✧✧✧✧✧✧✧✧✧✧✧✧<br />
✡ ✡<br />
✡<br />
✡<br />
✡<br />
✡<br />
✡<br />
✡<br />
✡ ✡<br />
(i + 1, j + 1)<br />
✈<br />
✡<br />
✈(i<br />
+ 1, j)<br />
(i, j + 1) ✈<br />
(i j) ❢<br />
(i − 1, j + 1) ✈<br />
❨<br />
✈ hθ<br />
(i − 1, j)<br />
✈ ❄ ✈<br />
✈<br />
(i − 1, j − 1)<br />
✛<br />
hr<br />
(i, j − 1)<br />
✲<br />
(i + 1, j − 1)<br />
Figure 4.3.1: The figure is showing a nine point stencil <strong>of</strong> <strong>the</strong> discretized Navier equation <strong>of</strong><br />
motion, Equation (4.1.1). The grid points are showed with <strong>the</strong> corresponding index notation.<br />
The hollow dot indicates in which point <strong>the</strong> solution is calculated. This figure shows a central<br />
finite difference scheme, where hr and hθ are <strong>the</strong> meshes for <strong>the</strong> radial and colatitudinal<br />
direction.<br />
and v equals<br />
⎡<br />
2 β2<br />
⎢⎣<br />
h2 +<br />
r<br />
2 α2<br />
, at each <strong>of</strong> <strong>the</strong> eight neighboring points, <strong>the</strong> grid is pictured in Figure 4.3.1. This is<br />
called a nine point stencil.<br />
4.3.1 Mesh Points<br />
In <strong>the</strong> last section <strong>the</strong> mesh points were introduced, now <strong>the</strong>y are used to discrete<br />
<strong>the</strong> equations. The mesh points and <strong>the</strong> corresponding values for <strong>the</strong> solution can be<br />
written with <strong>the</strong> convenient notation<br />
r 2 h 2 θ<br />
Pi j = (ihr, jhθ), ui j = u(ihr, jhθ) .<br />
By using a uniform grid and letting <strong>the</strong> radius cover <strong>the</strong> domain R0 ≤ r ≤ a, where r0<br />
is <strong>the</strong> center <strong>of</strong> <strong>the</strong> sphere, a is at <strong>the</strong> surface <strong>of</strong> <strong>the</strong> sphere, and <strong>the</strong> colatitude is given<br />
by θ0 ≤ θ ≤ 2π. With this new notation, <strong>the</strong> grid and meshes are<br />
ri = R0 + i hr, hr =<br />
a − R0<br />
n<br />
θ j = θ0 + jhθ, hθ = θn − θ0<br />
m<br />
⎤<br />
⎥⎦<br />
for i = 1, . . . , n<br />
for j = 1, . . . , m .
32 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
n and m are <strong>the</strong> grid points in <strong>the</strong> radial and <strong>the</strong> latitudinal direction, respectively. The<br />
pole singularity (R0 = 0) is discussed later in Section 4.4.1. For any mesh point Pi j,<br />
Equation (4.3.1) and Equation (4.3.2), can be discretized using <strong>the</strong> conservative form,<br />
<strong>the</strong> so called central difference discretization, which can be visualized in Figure 4.3.1<br />
again. The hollow dot indicates in which grid point, <strong>the</strong> solution is calculated. Note<br />
that <strong>the</strong>re are as many solutions as equations. With this notation, <strong>the</strong> Equations (4.3.1)-<br />
(4.3.2), can be written for any mesh point Pi j in <strong>the</strong> form<br />
and<br />
− Fu(ri, θ j)<br />
ρ<br />
− Fv(ri, θ j)<br />
ρ<br />
=<br />
=<br />
1 α2 (2ri + hr)<br />
ri<br />
2 h 2 r<br />
+ 1 α2 (2ri − hr)<br />
ri<br />
+ 1<br />
r2 i<br />
+ 1<br />
r2 i<br />
2 h 2 r<br />
ui+1, j<br />
ui−1, j<br />
2β 2 + (α 2 − β 2 )hθ<br />
2h 2 θ<br />
2β 2 − (α 2 − β 2 )hθ<br />
2h 2 θ<br />
+ 1 α2 − β2 −<br />
ri<br />
4 hr hθ<br />
<br />
α2 − β2 + 2 α2<br />
r 2 i<br />
ui, j+1<br />
ui, j−1<br />
(uui+1, j+1 − ui+1, j−1 − ui−1, j+1 + ui+1, j+1),<br />
h 2 r<br />
1 β2 (2ri + hr)<br />
ri<br />
2 h 2 r<br />
+ 1 β2 (2ri − hr)<br />
ri<br />
+ 1<br />
r2 i<br />
+ 1<br />
r2 i<br />
2 h 2 r<br />
+ 2β2<br />
r 2 i h2 θ<br />
vi+1, j<br />
vi−1, j<br />
2 α 2 − (α 2 − β 2 ) hθ<br />
2h 2 θ<br />
2 α 2 + (α 2 − β 2 ) hθ<br />
2h 2 θ<br />
+ 1 α<br />
ri<br />
2 − β2 ⎡<br />
2 β2<br />
− ⎢⎣ + 2 α2<br />
<br />
ui, j , (4.3.3)<br />
vi, j+1<br />
vi, j−1<br />
(vi+1, j+1 − vi+1, j−1 − vi−1, j+1 + vi−1, j−1)<br />
4 hr hθ<br />
⎤<br />
h 2 r<br />
r 2 i h2 θ<br />
⎥⎦ vi, j , (4.3.4)<br />
where i = 2, . . . , n − 1<br />
j = 2, . . . , m − 1 .<br />
The finite difference scheme at a grid point (ri, θ j) involves nine grid points north,
The Finite Difference Formula 33<br />
northwest, west, southwest, south, sou<strong>the</strong>ast, east, nor<strong>the</strong>ast and <strong>the</strong> center, where<br />
north and south are in <strong>the</strong> radial direction and east and <strong>the</strong> west are in <strong>the</strong> latitudinal<br />
direction. The center (ui j and vi j) are called <strong>the</strong> master grid points and are <strong>the</strong> approximation<br />
to <strong>the</strong> solution (u(ri, θ j) and v(ri, θ j)), respectively. The coordinates can been<br />
seen in Figure 4.3.1, where north is <strong>the</strong> point in <strong>the</strong> direction <strong>of</strong> radius.<br />
By defining<br />
r i− 1 2<br />
r i+ 1 2<br />
= ri − hr<br />
2<br />
= ri + hr<br />
2 ,<br />
and separating <strong>the</strong> equations in radial, latitudinal and mixed terms, this more convenient<br />
notation becomes<br />
− Fu(ri, θ j)<br />
ρ<br />
− Fv(ri, θ j)<br />
ρ<br />
=<br />
=<br />
1<br />
ri<br />
α 2 (r i+ 1 2<br />
− (α2 − β 2 ) ui j<br />
r 2 i<br />
+ 1<br />
r 2 i<br />
+ 1<br />
ri<br />
and<br />
ui+1, j − (r 1 i+ 2 , j + ri− 1 ) ui j + r 1<br />
2 i−<br />
ui−1, j)<br />
2<br />
+ 1<br />
r 2 i<br />
(α 2 − β 2 ) (ui, j−1 − ui, j+1)<br />
2 hθ<br />
h2 r<br />
β2 (ui, j+1 − ui j + ui, j−1)<br />
(α 2 − β 2 )(ui+1, j+1 − ui+1, j−1 − ui−1, j+1 + ui−1, j−1)<br />
4 hr hθ<br />
1 β<br />
ri<br />
2 (r 1 i+<br />
vi+1, j − (r 1<br />
2<br />
i+ 2<br />
+ r 1 i− ) vi j + r 1<br />
2 i−<br />
vi−1, j)<br />
2<br />
h2 r<br />
+ 1<br />
r2 α<br />
i<br />
2 (vi, j+1 − vi j + vi, j−1)<br />
h2 θ<br />
+ 1<br />
r 2 i<br />
+ 1<br />
ri<br />
(α 2 − β 2 ) (vi, j−1 − vi, j+1)<br />
2 hθ<br />
(α 2 − β 2 ) (vi+1, j+1 − vi+1, j−1 − vi−1, j+1 + vi−1, j−1)<br />
h 2 θ<br />
4 hr hθ<br />
,<br />
.<br />
(4.3.5)<br />
(4.3.6)<br />
These last two equations are <strong>the</strong> final FDE’s corresponding to <strong>the</strong> steady state version<br />
<strong>of</strong> <strong>the</strong> Navier equation <strong>of</strong> motion, Equation (4.1.1), where ui j and vi j are <strong>the</strong> approximated<br />
to <strong>the</strong> solution u(ri, θ j) and v(ri, θ j). In principal <strong>the</strong> constants should also have<br />
been discretized, but is here neglected.
34 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
4.4 Boundary Conditions<br />
To apply <strong>the</strong> FDE equations on areal <strong>Earth</strong>, <strong>the</strong>re have to be applied seven boundary<br />
conditions two at <strong>the</strong> free surface, depending <strong>of</strong> <strong>the</strong> surface tractions, one at <strong>the</strong> center<br />
<strong>of</strong> <strong>Earth</strong>, two at <strong>the</strong> polar axis and last two at <strong>the</strong> equatorial axis.<br />
4.4.1 At <strong>the</strong> Surface<br />
To apply <strong>the</strong> traction free conditions at <strong>the</strong> free surface, i = a, where a is <strong>the</strong> surface<br />
<strong>of</strong> <strong>Earth</strong>, a first order backward difference scheme is used for for <strong>the</strong> r-derivative, to<br />
avoid dealing with fictitious points. A centered second order scheme is used for <strong>the</strong><br />
θ-derivatives. This is an implicit method. See <strong>the</strong> stencil for this problem in Figure<br />
4.4.1, <strong>the</strong> figure is true for both both displacements u and v. The boundary conditions<br />
at <strong>the</strong> free surface was described in Section 3.3.1 with <strong>the</strong> Equations (3.3.7)-(3.3.8).<br />
After discretization, <strong>the</strong> equations becomes<br />
0 = τrr<br />
E<br />
<br />
=<br />
(1 − σ)<br />
(1 + σ)(1 − 2 σ)<br />
ua j − ua−1, j<br />
hr<br />
+ σ<br />
<br />
2 ui j +<br />
a<br />
va,<br />
<br />
j+1 −<br />
<br />
va, j−1<br />
+ cot θ j va j<br />
2 hθ<br />
(4.4.1)<br />
0 = τrθ<br />
<br />
1 ua, j+1 − ua, j−1<br />
= µ<br />
+<br />
a 2 hθ<br />
va j − va−1, j<br />
−<br />
hr<br />
1<br />
a va<br />
<br />
j . (4.4.2)<br />
The constants are not canceled out because <strong>of</strong> <strong>the</strong> computer precision. When values are<br />
close to zero <strong>the</strong> constants have an important contribution to <strong>the</strong> solution. The equations<br />
(4.4.1) and (4.4.2) are both dependent <strong>of</strong> <strong>the</strong> radial and tangential displacement,<br />
<strong>the</strong>refore <strong>the</strong> two FDEs, Equations (4.3.5)-(4.3.6), have to be combined and solved as<br />
one system.<br />
The local truncation errors for <strong>the</strong> backward scheme for ur and uθ are<br />
τ i r = − hr<br />
2 urr( ˜ηi, θ j) (4.4.3)<br />
τ j<br />
θ<br />
= −hθ<br />
2 uθθ(ri, ˜η j) . (4.4.4)<br />
These truncation errors add to <strong>the</strong> total truncation error at <strong>the</strong> boundaries.
Boundary Conditions 35<br />
i<br />
i − 1<br />
i − 2<br />
j − 1<br />
❣<br />
✲ ❣ ✛<br />
✇ ✇<br />
<br />
✒<br />
<br />
<br />
✻<br />
j j + 1<br />
❣<br />
✇<br />
✇ ✇ ✇<br />
✛ i = a, Free surface<br />
Figure 4.4.1: The diagram is showing <strong>the</strong> free surface at i = a. This is an implicit method. In<br />
<strong>the</strong> radial direction a backward finite difference scheme is used, and in <strong>the</strong> latitudinal direction<br />
a central finite difference scheme is used.<br />
4.4.2 Center <strong>of</strong> <strong>Earth</strong><br />
The case <strong>of</strong> <strong>the</strong> boundary condition R0 = 0 needs special attention due to <strong>the</strong> pole singularity.<br />
By using a staggered grid <strong>the</strong> coefficient matrix gets a satisfactory structure.<br />
ri =<br />
<br />
i − 1<br />
<br />
hr, where hr =<br />
2<br />
a<br />
n − 1<br />
2<br />
for i = 1, 2, . . . , n<br />
[Li, 2004]. Using <strong>the</strong> non-conservative form <strong>of</strong> <strong>the</strong> discretization at i = 1<br />
− Fu(r1, θ j)<br />
ρ<br />
=<br />
1 α2 (u0, j − 2 u1, j + u2, j)<br />
r1<br />
h 2 r<br />
+ 1 α<br />
r1<br />
2 (u2, j − u0, j)<br />
2hr<br />
+ (α2 − β2 ) u1, j<br />
r2 i<br />
+ 1<br />
r2 β<br />
1<br />
2 (u1, j+1 − 2 u1 j + u1, j−1)<br />
h2 θ<br />
+ 1<br />
r2 (α<br />
1<br />
2 − β2 ) (u1, j−1 − u1, j+1)<br />
2 hθ<br />
(α2 − β2 ) (u2, j+1 − u2, j−1 − u0, j+1 + u0, j+1)<br />
+ 1<br />
r1<br />
4 hr hθ<br />
(4.4.5)<br />
, (4.4.6)
36 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
and<br />
− Fv(r1, θ j)<br />
ρ<br />
=<br />
1 β2 (v0, j − 2v1, j + v2, j)<br />
r1<br />
h 2 r<br />
+ 1 β<br />
r1<br />
2 (v2, j − v0, j)<br />
2hr<br />
+ 1<br />
r2 1 h2 θ<br />
α 2 (v1, j+1 − 2 v1 j + v1, j−1)<br />
+ 1<br />
r2 (α<br />
1<br />
2 − β2 ) (v1, j−1 − v1, j+1)<br />
2 hθ<br />
(α2 − β2 ) (v2, j+1 − v2, j−1 − v0, j+1 + v0, j+1)<br />
+ 1<br />
r1<br />
4 hr hθ<br />
(4.4.7)<br />
By definition <strong>of</strong> i it is determined that U0 j = 0. The finite difference equation for i = 1<br />
is simplified to<br />
The radial field<br />
− Fu(r1, θ j)<br />
ρ<br />
and for <strong>the</strong> tangential field<br />
− Fv(r1, θ j)<br />
ρ<br />
=<br />
=<br />
1 α2 (u2, j − 2 u1, j)<br />
r1<br />
h 2 r<br />
+ 1 α<br />
r1<br />
2 u2, j<br />
2hr<br />
+ (α2 − β2 ) u1, j<br />
r2 i<br />
+ 1<br />
r2 β<br />
1<br />
2 (u1, j+1 − 2 u1 j + u1, j−1)<br />
h2 θ<br />
+ 1<br />
r2 (α<br />
1<br />
2 − β2 ) (u1, j−1 − u1, j+1)<br />
2 hθ<br />
(α2 − β2 ) (u2, j+1 − v2, j−1)<br />
+ 1<br />
r1<br />
4 hr hθ<br />
1 β<br />
r1<br />
2 (v2, j − 2 v1, j)<br />
h2 +<br />
r<br />
1 β<br />
r1<br />
2 v2, j<br />
2hr<br />
+ 1<br />
r2 α<br />
1<br />
2 (v1, j+1 − 2 v1 j + v1, j−1)<br />
h2 θ<br />
+ 1<br />
r2 (α<br />
1<br />
2 − β2 ) (v1, j−1 − v1, j+1)<br />
2 hθ<br />
(α2 − β2 ) (v2, j+1 − v2, j−1)<br />
+ 1<br />
r1<br />
4 hr hθ<br />
(4.4.8)<br />
, (4.4.9)<br />
(4.4.10)
Boundary Conditions 37<br />
4.4.3 The Pole and Equator<br />
The displacement governed from <strong>the</strong> tidal acceleration applied on a sphere is equally<br />
sized in <strong>the</strong> four quadrants, i.e. <strong>the</strong> boundary problem in <strong>the</strong> latitudinal direction is<br />
periodic in <strong>the</strong> domain [0, 2π], and because <strong>of</strong> axes symmetry, we are only looking at<br />
one quadrant, <strong>the</strong> interval [0, π<br />
2 ] i.e. from <strong>the</strong> polar axis to <strong>the</strong> equatorial axis.<br />
For <strong>the</strong> radial displacement <strong>the</strong> most simple choice, to apply boundary conditions<br />
on <strong>the</strong> free Pole and Equator axes, is to exploit <strong>the</strong> symmetry <strong>of</strong> <strong>the</strong> displacements on<br />
both sides <strong>of</strong> <strong>the</strong> axes, <strong>the</strong> displacement vector is horizontal at <strong>the</strong> Pole and vertical at<br />
<strong>the</strong> Equator. In practice this is carried by smoothing <strong>the</strong> partial derivatives<br />
ui,2 − ui,1 = 0 (4.4.11)<br />
ui,m−1 − ui,m = 0 , (4.4.12)<br />
where i = 1, 2, . . . , n. Equation (4.4.11) is <strong>the</strong> boundary condition for <strong>the</strong> polar axis and<br />
Equation (4.4.12) is for <strong>the</strong> equatorial axis. The latitudinal displacement is symmetric<br />
about <strong>the</strong> center <strong>of</strong> mass and <strong>the</strong>refore zero at <strong>the</strong> polar and equatorial axes. This yields<br />
vi,1 = 0 (4.4.13)<br />
vi,m = 0 . (4.4.14)
Chapter 5<br />
Regularization<br />
This chapter deals with <strong>the</strong> problem <strong>of</strong> solving <strong>the</strong> two dimensional Navier equation<br />
<strong>of</strong> motion and how to treat an ill-conditioned system. The chapter is not in chronological<br />
order with <strong>the</strong> rest <strong>of</strong> <strong>the</strong> <strong>the</strong>sis, but <strong>the</strong> <strong>the</strong>ory <strong>of</strong> regularization was important to<br />
outline, as <strong>the</strong> linear system <strong>of</strong> equations proved to be ill-conditioned.<br />
Let A ∈ R n×m be a rectangular matrix, F a vector with n data points and x a vector<br />
with m model parameters. We are aiming to solve <strong>the</strong> linear system<br />
Ax = F. (5.0.1)<br />
This equation corresponds to <strong>the</strong> FDE, Equations (4.3.5)-(4.3.6), derived in Section<br />
4.3. A is <strong>the</strong> coefficient matrix with <strong>the</strong> coefficient <strong>of</strong> <strong>the</strong> differentials, F is <strong>the</strong> lefthand<br />
side <strong>of</strong> <strong>the</strong> Equations (4.3.5) and (4.3.6). This is <strong>the</strong> tidal acceleration given in <strong>the</strong><br />
Equations (2.2.6)-(2.2.7) in Section 2.2.1 toge<strong>the</strong>r with <strong>the</strong> boundary constrains given<br />
in Section 4.4. x corresponds to <strong>the</strong> displacements u and v from Chapter 4, which are<br />
going to be modeled.<br />
If A is nonsingular, Equation (5.0.1) has a unique solution for every F. On <strong>the</strong> o<strong>the</strong>r<br />
hand if A is singular <strong>the</strong> equation has zero or infinite solutions, [Aster et al., 2005].<br />
This is not a regular inverse problem, because <strong>the</strong> data vector F is exactly known from<br />
<strong>the</strong>ory. The connection and stability between data and model spaces and <strong>the</strong> solution<br />
uniqueness and ability to fit data is examined in <strong>the</strong> next section.<br />
5.1 Ill-conditioning <strong>of</strong> a System<br />
An ill-conditioned system is one without a solution, i.e. a system that does not have<br />
a unique solution or where <strong>the</strong> solution is discontinuously dependent <strong>of</strong> <strong>the</strong> data. A<br />
common characteristic <strong>of</strong> <strong>the</strong> problem is a large condition number. A condition number<br />
states a measure <strong>of</strong> instability <strong>of</strong> <strong>the</strong> solution, [Aster et al., 2005]. When <strong>the</strong>re is no<br />
errors in <strong>the</strong> data vector, <strong>the</strong> condition number can more accurately be defined as <strong>the</strong><br />
maximum ratio <strong>of</strong> <strong>the</strong> relative error in <strong>the</strong> true solution x, that is x − ˆx/x, divided<br />
by <strong>the</strong> relative error in A, and that is A − Â/A. ˆx is <strong>the</strong> actual and  is <strong>the</strong> actual
40 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
data. The relative error bound <strong>of</strong> <strong>the</strong> model can be computed to<br />
x − ˆx<br />
x ≤ AA−1 A − Â<br />
<br />
A<br />
The constant is <strong>the</strong> condition number <strong>of</strong> A<br />
.<br />
cond(A) = AA −1 . (5.1.1)<br />
[Parker, 1994]. A condition number <strong>of</strong> A near 1 indicates a well-conditioned matrix,<br />
and a matrix with high condition number is said to be ill-conditioned. The rank <strong>of</strong> A<br />
is <strong>the</strong> dimension <strong>of</strong> <strong>the</strong> column space or range <strong>of</strong> A. The column space is <strong>the</strong> set <strong>of</strong> all<br />
vectors F that can be written as a linear combination <strong>of</strong> <strong>the</strong> columns in A. A has full<br />
rank if k = rank(A) = min(m, n). If <strong>the</strong> rank is smaller than min(m, n) <strong>the</strong> matrix is<br />
rank-deficient, [Aster et al., 2005]. When measurement errors or approximation errors<br />
are present or discretization is applied on <strong>the</strong> system, <strong>the</strong>re do not, in a ma<strong>the</strong>matical<br />
frame work, exist a rank <strong>of</strong> matrix A. If <strong>the</strong>re exist some columns <strong>of</strong> A (with some<br />
error level), which are practically linearly independent, <strong>the</strong>re exist a pseudo rank. This<br />
very loose formulation is <strong>the</strong> definition <strong>of</strong> a numerical rank, [Hansen, 1996].<br />
According to Hansen [1996], ill-conditioned systems can be divided into two classes<br />
<strong>of</strong> problems based on <strong>the</strong> properties <strong>of</strong> <strong>the</strong> coefficient matrix.<br />
Rank deficient problems The coefficient matrix has <strong>the</strong> problems<br />
• Cluster <strong>of</strong> small singular values.<br />
• Large well-defined gap between large and small singular values.<br />
This implies that <strong>the</strong>re exist one or more linear combinations in <strong>the</strong> coefficient<br />
matrix. There usually exist a reformulation that eliminate <strong>the</strong> ill-conditioning by<br />
extracting <strong>the</strong> linearly independent information. There exist a numerical rank,<br />
and <strong>the</strong> problem involves matrices which are exactly or almost rank deficient.<br />
This definition do not contemplate how little a small singular value is.<br />
Discrete ill-posed problems Come from <strong>the</strong> discretization <strong>of</strong> continuous ill-posed<br />
problems, and <strong>the</strong> properties <strong>of</strong> <strong>the</strong>ir coefficient matrices are:<br />
• Large cluster <strong>of</strong> small singular values which increases with <strong>the</strong> dimension<br />
<strong>of</strong> <strong>the</strong> problem.<br />
• The singular values decay sequentially to zero with no gap, and <strong>the</strong> Picard<br />
condition is satisfied, see Section 5.2.2.<br />
Usually <strong>the</strong>re is no reformulation <strong>of</strong> <strong>the</strong> problem, which can eliminate <strong>the</strong> illconditioning.<br />
Because <strong>of</strong> <strong>the</strong> lack <strong>of</strong> gap in <strong>the</strong> singular values, <strong>the</strong>re do not<br />
exist a notion for <strong>the</strong> numerical rank. The goal <strong>of</strong> <strong>the</strong> solution is to find a good<br />
balance <strong>of</strong> <strong>the</strong> residual norm and <strong>the</strong> solution norm.
Tikhonov Regularization 41<br />
If an ill-conditioned matrix does not apply to ei<strong>the</strong>r <strong>of</strong> <strong>the</strong> classes, <strong>the</strong>n <strong>the</strong> regularization<br />
is not possible and <strong>the</strong> system has to be solved as good as possible without regularization<br />
e.g. by iterative clarification. A well-conditioned problem is one where <strong>the</strong><br />
components <strong>of</strong> <strong>the</strong> solution’s SVD, increases on an average and <strong>the</strong> condition number<br />
can be a measure <strong>of</strong> <strong>the</strong> solution’s sensitivity towards <strong>the</strong> data space, [Hansen, 1996].<br />
Rank deficient problems arises in applications where noise or o<strong>the</strong>r unwanted disturbance<br />
are suppressed for example in signal processing. Discrete ill-posed problems<br />
normally arises doing <strong>the</strong> numerical treatment <strong>of</strong> inverse problems. Oil exploration is a<br />
typical field, where ill-posed problem arises when trying to determine <strong>the</strong> composition<br />
<strong>of</strong> <strong>the</strong> subsurface with seismic inversion techniques, o<strong>the</strong>r fields are image restoration<br />
and tomography.<br />
5.2 Tikhonov Regularization<br />
In order to solve and stabilize <strong>the</strong> ill-conditioned problem, a Tikhonov regularization<br />
program is used. The Matlab code is to be downloaded from <strong>the</strong> Matlab Central at<br />
Hansen [2008a], and <strong>the</strong> code is a part <strong>of</strong> a larger regularization package, regutools,<br />
documented in Hansen [2008b].<br />
Tikhonov regularization is one <strong>of</strong> <strong>the</strong> oldest and probably most used regularization<br />
methods for solving inverse problems. The idea is to find <strong>the</strong> best regularized solution<br />
xε that minimizes <strong>the</strong> damped least square problem<br />
xε = min{Ax − F 2 2 + ε2Lx 2 2 } , (5.2.1)<br />
where ε is <strong>the</strong> damping or regularization parameter, which controls <strong>the</strong> regularization,<br />
a large ε yields a large damping on <strong>the</strong> system. See more about choosing <strong>the</strong> good<br />
solution in Section 5.2.2. L is in <strong>the</strong> standard form <strong>the</strong> identity matrix or in <strong>the</strong> generalized<br />
form L Im where L is a p × m matrix with full rank. · 2 denotes <strong>the</strong> second<br />
norm, l2. Tikhonov regularization is a direct method, because <strong>the</strong> solution xε given by<br />
Equation (5.2.1) is <strong>the</strong> solution to<br />
<br />
<br />
<br />
<br />
min A<br />
<br />
ε I<br />
<br />
F<br />
x −<br />
0<br />
2<br />
, (5.2.2)<br />
where L = I. When using Equation (5.2.1) it is expected that <strong>the</strong> errors in <strong>the</strong> righthand<br />
side is unregularized and that <strong>the</strong>ir covariance matrix is identical to <strong>the</strong> identity<br />
matrix, [Hansen, 2008b].
42 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
5.2.1 Singular Value Decomposition<br />
In Hansen [2008b]’s Tikhonov regularization a singular value decomposition (SVD)<br />
is used to analysis <strong>the</strong> ill-posed problem.<br />
Letting A ∈ R n×m be a rectangular matrix with n ≥ m. The singular value decomposition<br />
<strong>of</strong> A can be written as<br />
A = UΣV T =<br />
k<br />
i=1<br />
ui σi v T . (5.2.3)<br />
U = (u1, . . . , un) and V = (v1, . . . , vm) are orthonormal matrices i.e. <strong>the</strong> transposed<br />
matrix multiplied with <strong>the</strong> non-transposed matrix gives <strong>the</strong> identity matrix. p is <strong>the</strong><br />
number <strong>of</strong> nonzero singular values, <strong>the</strong> rank <strong>of</strong> A. U spans <strong>the</strong> data space R n and V<br />
spans <strong>the</strong> data space R m . The vectors ui and vi are called left and right singular values<br />
<strong>of</strong> A, respectively. Σ = diag(σ1, . . . , σk) is a diagonal matrix with non-negative<br />
entries. σ1 ≥ σ2 ≥ . . . ≥ σk ≥ 0 and <strong>the</strong> numbers σi is called <strong>the</strong> singular value <strong>of</strong> A.<br />
k = min(m, n). If p = k, A has full rank.<br />
In l2 <strong>the</strong> condition number <strong>of</strong> A is given by <strong>the</strong> ratio<br />
cond(A) = σ1<br />
σk<br />
. (5.2.4)<br />
This is exactly true for A, if it has full rank and all <strong>the</strong> singular values are used in <strong>the</strong><br />
pseudoinverse solution (p = k). The SVD is unique for a given matrix A, aside from<br />
singular vectors connected to multiple singular values, [Aster et al., 2005; Hansen,<br />
2008b]. This can be stated because SVD is coupled to <strong>the</strong> eigenvalue decomposition<br />
by <strong>the</strong> relation<br />
and<br />
A T A = VΣ T U T UΣV T = VΣ 2 V T<br />
(5.2.5)<br />
AA T = UΣV T VΣ T U T = UΣ 2 U T . (5.2.6)<br />
The right-hand side describe <strong>the</strong> eigenvalue decomposition <strong>of</strong> <strong>the</strong> left-hand side, [Trefe<strong>the</strong>n<br />
and Bau, 1997].<br />
SVD is effective in <strong>the</strong> Tikhonov regularization because it smoo<strong>the</strong>ns <strong>the</strong> solution by<br />
less weighting <strong>of</strong> small singular values, as mentioned small singular values implies an<br />
almost rank deficient matrix, [Hansen, 2008b]. With SVD it can be shown<br />
p uT i F<br />
vi , (5.2.7)<br />
xLSQ =<br />
i=1<br />
where p is <strong>the</strong> rank <strong>of</strong> <strong>the</strong> matrix. From this equation it is very clear how very small<br />
singular values in <strong>the</strong> dominator can give large values in <strong>the</strong> model space, which will<br />
dominate <strong>the</strong> solution and it can become very unstable. By dropping <strong>the</strong> small singular<br />
values <strong>the</strong> system is regularized and <strong>the</strong> solution is less sensitive to noise from <strong>the</strong> data<br />
space. The cost <strong>of</strong> <strong>the</strong> regularization, is <strong>the</strong> fact that <strong>the</strong> solution is no longer unbiased<br />
and <strong>the</strong> resolution is reduced, [Aster et al., 2005] .<br />
σi
Tikhonov Regularization 43<br />
5.2.2 The Good Solution<br />
The essence about choosing <strong>the</strong> good solution is a question <strong>of</strong> balance. Because <strong>the</strong><br />
system is ill-conditioned, it has to be smoo<strong>the</strong>d to give a suitable solution, but it is not<br />
wanted that to smoo<strong>the</strong>n to solution too much that you lose important information.<br />
Filter Factor and Discrete Picard Condition<br />
As mentioned, regularization is a processes <strong>of</strong> damping or filtering out small singular<br />
values. The Tikhonov regularized solution in terms <strong>of</strong> SVD, Equation (5.2.7) gets <strong>the</strong><br />
form<br />
xreg =<br />
k<br />
i=1<br />
u<br />
fi<br />
T i F<br />
σi<br />
vi , (5.2.8)<br />
where fi is <strong>the</strong> filter factors. In <strong>the</strong> used Tikhonov regularization <strong>the</strong> filter factor for<br />
(L = I) is<br />
fi =<br />
σ 2 i<br />
σ 2 i<br />
+ ε2 . (5.2.9)<br />
The filter factor is especially effective when σi > ε. This also shows if ε > σ1, <strong>the</strong>n<br />
<strong>the</strong> problem is unregularized. The filter factor is usually close to one for large singular<br />
values. For values in between, <strong>the</strong> filter factors decreases monotonic when <strong>the</strong> singular<br />
values decreases.<br />
Equation (5.2.8) shows that <strong>the</strong> decay <strong>of</strong> |uT i F| compared to <strong>the</strong> decay <strong>of</strong> σi plays<br />
an important role in regularization. In real world <strong>the</strong> right-hand side F is always con-<br />
taminated with an error. The discrete Picard condition states that if |u T i<br />
F| in a discrete<br />
ill-posed problem, on an average, decay faster to zero than <strong>the</strong> singular values, <strong>the</strong>n<br />
<strong>the</strong> unperturbed right-hand side satisfy <strong>the</strong> Picard condition. If this is not true, <strong>the</strong><br />
Tikhonov regularization cannot give a useful solution, [Hansen, 1996].<br />
Regularization Parameter<br />
ε controls <strong>the</strong> sensitivity <strong>of</strong> <strong>the</strong> regularized solution to perturbations in A and F. The<br />
threshold <strong>of</strong> <strong>the</strong> perturbation is proportional to ε −1 . There exist several methods to<br />
estimate <strong>the</strong> best regularization parameter. There are <strong>the</strong> L-curve criterion, generalized<br />
cross-validation (GCV), discrepancy principle or <strong>the</strong> Quasi-optimality criterion. In<br />
most cases, <strong>the</strong> L-curve gives <strong>the</strong> best estimate <strong>of</strong> <strong>the</strong> regularization parameter, [Rojas,<br />
1996; Aster et al., 2005; Hansen, 2008b].<br />
The L-curve criterion plots <strong>the</strong> solution norm ||Lxreg|| versus <strong>the</strong> misfit norm ||Axreg−<br />
F|| on a double logarithmic scale. It is called <strong>the</strong> L-curve because it almost always has<br />
a characteristic L shaped curve, with a distinct corner, see Figure 5.2.1. The L-curve<br />
is piecewise linear with break points at<br />
(xi, yi) = (log ||Axi − F||, log ||Lxi||),<br />
where i = 1, . . . , p. The optimal value ɛ, <strong>the</strong> regularization parameter, is considered to<br />
be at <strong>the</strong> corner. The corner <strong>of</strong> <strong>the</strong> L-curve marks this transition, since it represents a
44 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
compromise between <strong>the</strong> minimization <strong>of</strong> <strong>the</strong> norm <strong>of</strong> <strong>the</strong> residual and <strong>the</strong> semi-norm<br />
<strong>of</strong> <strong>the</strong> solution. The horizontal branch <strong>of</strong> <strong>the</strong> L-curve is dominated by <strong>the</strong> regularization<br />
error and <strong>the</strong> vertical branch shows <strong>the</strong> sharp increase <strong>of</strong> <strong>the</strong> semi-norm caused by<br />
propagation errors. The ill-conditioning <strong>of</strong> A causes a strong growth in <strong>the</strong> weighted<br />
semi-norm when ɛ exceeds a certain threshold, which is in fact <strong>the</strong> numerical rank <strong>of</strong><br />
A for a well-chosen ɛ.<br />
In <strong>the</strong> Regularization package, [Hansen, 2008a], <strong>the</strong> Matlab codes lcurve.m and<br />
lcorner.m plots respectively <strong>the</strong> L-curve and calculates <strong>the</strong> L-corner in relation to<br />
<strong>the</strong> Tikhonov regularization where <strong>the</strong> filter factor is incorporated. For Tikhonov regularization<br />
<strong>the</strong> L-curve is a smooth function, and this method selects <strong>the</strong> value, ɛ which<br />
maximizes <strong>the</strong> curvature <strong>of</strong> <strong>the</strong> L-curve, [Hansen, 2008b]. If <strong>the</strong> discrete Picard condition<br />
do not hold, <strong>the</strong> L-curve method may fail to give a good estimate <strong>of</strong> <strong>the</strong> regularization<br />
parameter, [Vogel, 1996].<br />
Figure 5.2.1: Example <strong>of</strong> <strong>the</strong> L-curve. Adapted from Hansen [2008b]
Chapter 6<br />
Model Description and Construction<br />
The following two chapters are concerning <strong>the</strong> development <strong>of</strong> two <strong>Earth</strong> models. The<br />
two models in question are a completely homogeneous <strong>Earth</strong> and a layered <strong>Earth</strong>. They<br />
are designed to determine <strong>the</strong> displacement field <strong>of</strong> <strong>the</strong> <strong>Earth</strong> due to <strong>the</strong> solid <strong>Earth</strong><br />
tides. This is done by solving <strong>the</strong> steady state Navier equation <strong>of</strong> motion, Equation<br />
(3.3.3) described in Section 3.3. The problem is to discretize <strong>the</strong> Navier equation by<br />
<strong>the</strong> numerical method <strong>of</strong> finite difference outlined in Chapter 4 with <strong>the</strong> main Equations<br />
(4.3.5) - (4.3.6) and <strong>the</strong> seven boundary conditions given in Section 4.4. Finally <strong>the</strong><br />
displacement is untangled by <strong>the</strong> Tikhonov regularization, see Section 5.2. In next<br />
chapter <strong>the</strong> models are analyzed and tested, but first this chapter starts by describing <strong>the</strong><br />
common model assumptions, <strong>the</strong>n homogeneous <strong>Earth</strong> model is introduced followed<br />
by <strong>the</strong> layered <strong>Earth</strong> model and in <strong>the</strong> end <strong>of</strong> this chapter <strong>the</strong> structure and development<br />
<strong>of</strong> <strong>the</strong> models are outlined<br />
6.1 Model Assumptions and Parameters<br />
The models are defined to be continuous, homogeneous, isotropic, infinite and an elastic<br />
medium, and from this <strong>the</strong> Navier equation <strong>of</strong> motion, Equation (3.3.3) is valid.<br />
The right side <strong>of</strong> <strong>the</strong> equation are <strong>the</strong> tidal accelerations given by <strong>the</strong> Equations (2.2.6)-<br />
(2.2.7) in Section 2.2.1. Fur<strong>the</strong>r <strong>the</strong> Moon is assumed to be <strong>the</strong> only attracting body,<br />
<strong>the</strong>reby neglecting <strong>the</strong> perturbations from <strong>the</strong> Sun and o<strong>the</strong>r planetary objects. The<br />
<strong>Earth</strong> is assumed to be spherical symmetric about <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and rotation<br />
symmetry is applied, hence <strong>the</strong> <strong>Earth</strong>-Moon system is stationary in relation to each<br />
o<strong>the</strong>r, and <strong>the</strong> <strong>Earth</strong>-Moon line is chosen going through <strong>the</strong> Equator. In Table 6.1.1,<br />
<strong>the</strong> universal constants used in <strong>the</strong> model is given.
46 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
Name Symbol Size Unit<br />
Mean radius <strong>of</strong> <strong>the</strong> <strong>Earth</strong> r⊕ 6.3710 · 10 6 m<br />
Orbital distance <strong>of</strong> <strong>the</strong> Moon R 3.8440 · 10 8 m<br />
Mass <strong>of</strong> <strong>the</strong> <strong>Earth</strong> M⊕ 5.9742 · 10 24 kg<br />
Mass <strong>of</strong> <strong>the</strong> Moon M 7.3483 · 10 22 kg<br />
Constant <strong>of</strong> gravitation G 6.6726 · 10 −11 m 3 kg −1 s −2<br />
Table 6.1.1: Universal constants with five significant numbers given by Woan [2003].<br />
6.2 The Homogeneous <strong>Earth</strong> Model<br />
The homogeneous <strong>Earth</strong> model is assumed to be a Poisson solid. For a Poisson solid<br />
<strong>the</strong> Lamé parameters µ and λ equals each o<strong>the</strong>r, which is close to be true for rocks, [Lay<br />
and Wallace, 1995]. The Lamé parameters are determined from <strong>the</strong> seismic velocities<br />
given in Section 3.3, and can be derived as<br />
µ = ρ β 2<br />
and λ = ρ α 2 − 2 ρ β 2 . (6.2.1)<br />
Poisson’s ratio σ and Young’s modulus are given in Section 3.3.1. The model is assuming<br />
a P-wave velocity, α, <strong>of</strong> 8000 ms −1 , which corresponds to <strong>the</strong> top <strong>of</strong> <strong>the</strong> mantle<br />
or at <strong>the</strong> core-mantle boundary, [Dziewonski and Anderson, 1981]. The ratio <strong>of</strong> <strong>the</strong><br />
seismic velocities in a Poisson solid is α/β = √ 3. The S-wave velocity, β, is calculated<br />
from this ratio.<br />
Since <strong>the</strong> <strong>Earth</strong> is homogeneous, <strong>the</strong> density <strong>of</strong> <strong>the</strong> <strong>Earth</strong> model is given by <strong>the</strong><br />
simple expression ρ = M⊕/(4πr 3 ⊕ ). M⊕ and r⊕ are <strong>the</strong> mass and <strong>the</strong> mean radius <strong>of</strong><br />
<strong>the</strong> <strong>Earth</strong>, respectively. The seismic velocities, density and <strong>the</strong> elastic parameters are<br />
given in Table 6.2.1. Young’s modulus, <strong>the</strong> bulk modulus and <strong>the</strong> Lamé parameters for<br />
rocks are in <strong>the</strong> magnitude <strong>of</strong> 20 − 120 GPa, [Fowler, 1990], which correspond to <strong>the</strong><br />
table values.<br />
Name Symbol Size Unit<br />
P-velocity α 8.0000 · 10 4 ms −1<br />
S-velocity β 4.6188 · 10 4 ms −1<br />
Density ρ 5.5153 · 10 3 kg m −3<br />
Shear modulus µ 1.1766 · 10 11 Pa<br />
Lamé Parameter no. 2 λ 1.1766 · 10 11 Pa<br />
Poisson ratio σ 0.2500 Pa<br />
Young’s modulus E 2.9419 · 10 11 Pa<br />
Table 6.2.1: Seismic velocities, density and elastic parameters <strong>of</strong> <strong>the</strong> homogeneous <strong>Earth</strong><br />
model, given with five significant digits. The P-wave velocity is an assumption and <strong>the</strong> bottom<br />
part <strong>of</strong> <strong>the</strong> table is calculated from this value and from <strong>the</strong> universal constants given by <strong>the</strong><br />
previous table, Table 6.1.1.
The Layered <strong>Earth</strong> Model 47<br />
6.3 The Layered <strong>Earth</strong> Model<br />
In this section, <strong>the</strong> homogeneous model in previous Section 6.2, is expanded to apply<br />
for a layered <strong>Earth</strong>, this results in some new assumptions and new parameters have to<br />
be introduced.<br />
6.3.1 Assumptions and Parameters<br />
The layered model uses <strong>the</strong> same universal constants as given in Table 6.1.1. The<br />
difference between <strong>the</strong> two models are that <strong>the</strong> layered <strong>Earth</strong> model, as <strong>the</strong> name suggest,<br />
is composed <strong>of</strong> four layers; inner core, outer core, mantle and crust. There are<br />
not implemented any o<strong>the</strong>r boundaries within <strong>the</strong> layers such as <strong>the</strong> transition zone,<br />
Moho or core-mantle boundary. The seismic velocities and <strong>the</strong> density in each layer<br />
are assumed to be constant through out <strong>the</strong> layer.<br />
The model uses data from <strong>the</strong> isotropic Preliminary Reference <strong>Earth</strong> Model (PREM)<br />
by Dziewonski and Anderson [1981]. The geophysical properties from PREM are averaged<br />
over each layer. In Figure 6.3.1, <strong>the</strong> seismic body wave velocities and <strong>the</strong><br />
densities through <strong>the</strong> <strong>Earth</strong> are depicted as functions <strong>of</strong> <strong>the</strong> radius. The different layers<br />
in <strong>the</strong> <strong>Earth</strong> are characterized by abrupt changes in <strong>the</strong>se values. The inner core is defined<br />
in <strong>the</strong> range <strong>of</strong> 0-1221 km. The outer core is given from 1221-3480 km, and <strong>the</strong><br />
materials within <strong>the</strong> outer core are fluids. This is apparent from <strong>the</strong> zero shear velocity,<br />
remember from Section 3.3 shear velocity can not exist in fluids, also <strong>the</strong> Poison ratio<br />
<strong>of</strong> 0.5 states a fluid material. The mantle is ranging from 3480-5701 km and <strong>the</strong> outer<br />
layer or <strong>the</strong> crust is defined from 5701-6371 km.<br />
Figure 6.3.1: Plot <strong>of</strong> <strong>the</strong> seismic body wave velocities and <strong>the</strong> density through <strong>the</strong> <strong>Earth</strong>, as a<br />
function <strong>of</strong> radius (m). Data is adapted from <strong>the</strong> isotropic PREM by Dziewonski and Anderson<br />
[1981]. The blue curve shows <strong>the</strong> P-wave velocity and green is <strong>the</strong> S-wave velocity both in m s .<br />
The density is represented by <strong>the</strong> red curve, given in kg<br />
3 . The different layers in <strong>the</strong> <strong>Earth</strong> are<br />
m<br />
characterized by abrupt changes in <strong>the</strong>se values.
48 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
Parameters Inner core Outer core Mantle Crust<br />
Radius range (km) r 0-1221 1221-3480 3480-5701 5701-6371<br />
P-velocity (km s −1 ) α 11.1835 9.4203 12.4495 4.6833<br />
S-velocity (km s −1 ) β 3.6128 0 6.7792 3.6508<br />
Density (10 3 kg m −3 ) ρ 12.9792 11.2955 4.9835 3.2160<br />
Shear modulus (10 11 Pa) µ 1.6941 0 1.6644 0.4284<br />
Lamé no. 2 (10 11 Pa) λ 16.2329 9.9706 7.7230 0.7053<br />
Young’s modulus (10 11 Pa) E 4.9222 0 4.6982 1.1234<br />
Poisson ratio σ 0.4528 0.5000 0.4114 0.31102<br />
Table 6.3.1: Physical properties <strong>of</strong> <strong>the</strong> different layers used in <strong>the</strong> analysis <strong>of</strong> <strong>the</strong> layered<br />
<strong>Earth</strong>. The seismic velocities, <strong>the</strong> density and <strong>the</strong> elastic parameters are averaged over each<br />
layer, with calculated from PREM.<br />
The blue curve in Figure 6.3.1 is <strong>the</strong> P-wave velocity given in m s , green is <strong>the</strong> Swave<br />
velocity also in m kg<br />
s , and <strong>the</strong> red curve is <strong>the</strong> density given in<br />
m3 . The radius is<br />
measured from <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong> to <strong>the</strong> surface in m. The values are written in<br />
Table 6.3.1, and plotted in Appendix C.<br />
6.4 Model Development<br />
The two model are constructed in a similar fashion only <strong>the</strong> constants differ from each<br />
o<strong>the</strong>r. The idea with <strong>the</strong> modeling is to make it as simple as possible. The equations<br />
are complex enough, without using any smart solving method. The complexity arises<br />
due to <strong>the</strong> chosen approach with <strong>the</strong> polar coordinate system, and <strong>the</strong> cross derivative<br />
terms urθ and vrθ. When solving cross derivatives in finite difference equations, it is<br />
not possible to use <strong>the</strong> simple method <strong>of</strong> dimension by dimension, which is a simple<br />
and widely used method to solve FDE’s. This is because <strong>the</strong> partial difference equation<br />
have no diagonal dominance as systems without cross derivative terms. In general<br />
<strong>the</strong>se terms make <strong>the</strong> modeling <strong>of</strong> <strong>the</strong> problem more difficult, [Li, 2004; Holmes,<br />
2007], <strong>the</strong>reby <strong>the</strong> model uses a simple iterative scheme. The model is developed in<br />
Matlab and <strong>the</strong> program code for <strong>the</strong> layered <strong>Earth</strong> model can be found in Appendix<br />
B. This section is not meant to go through <strong>the</strong> whole modeling process but just to point<br />
out and explain selected parts.<br />
6.4.1 Grid Points and Meshes<br />
The radial (u) and tangential (v) inner equations, from <strong>the</strong> Equations (4.3.5)-(4.3.6)<br />
in Section 4.3, are discretized for each direction separately. Each FDE is made with<br />
n = 66 grid points in <strong>the</strong> radial direction and m = 56 grid points in <strong>the</strong> co-latitudinal<br />
direction, this were <strong>the</strong> limit allowed by Matlab. The mesh spans a part <strong>of</strong> a circle<br />
sector, as described in Section 4.3, and a simple 4 × 4 grid is depicted in Figure 6.4.1,<br />
where <strong>the</strong> starts with (1,1) in <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>.<br />
The mesh sizes are respectively hr = 96.530 km and hθ = 0.028000 radians or<br />
1.6043 ◦ , this gives arcs <strong>of</strong> 178.34 km at <strong>the</strong> surface, 120.98 km at a radius <strong>of</strong> 4413 km
Model Development 49<br />
n<br />
(4, 1)<br />
(3, 1)<br />
(2, 1)<br />
(1, 1)<br />
(2, 2)<br />
(2, 3)<br />
(2, 4)<br />
(4, 2)<br />
(3, 2)<br />
n<br />
(3, 3)<br />
(3, 4)<br />
m<br />
(4, 3)<br />
(n, m) = (4, 4)<br />
Figure 6.4.1: The figure is displaying a 4 × 4 grid, where (1,1) starts at <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>,<br />
and <strong>the</strong> numbering is working out from <strong>the</strong> center counting from <strong>the</strong> Pole axis towards <strong>the</strong><br />
Equator axis, where <strong>the</strong> Equator is <strong>the</strong> grid point (n, m). n corresponds to <strong>the</strong> radius and m<br />
corresponds to <strong>the</strong> colatitude.<br />
(mantle), 64.88 km at a radius <strong>of</strong> 2313 km (outer core) and 10.00 km at a radius <strong>of</strong> 392<br />
km (inner core). Because <strong>of</strong> <strong>the</strong> uniform grid in <strong>the</strong> polar coordinate system, <strong>the</strong> grid<br />
points are allocated much closer in <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong> than at <strong>the</strong> surface. In stead<br />
<strong>of</strong> using <strong>the</strong> staggered grid to deal with <strong>the</strong> pole singularity, as described in Section<br />
4.4.2, a simple constrain <strong>of</strong> zero displacement is put on <strong>the</strong> center boundary condition.<br />
Due to <strong>the</strong> coupled surface boundary conditions, as described in Section 4.4.1,<br />
<strong>the</strong> radial and tangential linear system <strong>of</strong> equations have to be solved in one system.<br />
Figure 6.4.2 shows a sketch <strong>of</strong> <strong>the</strong> linear system <strong>of</strong> equations Ax = F. As it is seen<br />
both <strong>the</strong> radial and <strong>the</strong> tangential systems are evaluated toge<strong>the</strong>r. u and v are <strong>the</strong> radial<br />
and tangential displacements, respectively. Au and Av are <strong>the</strong> coefficients <strong>of</strong> <strong>the</strong> radial<br />
and tangential problem, given in Equation (4.3.5)-(4.3.6). Ac represent <strong>the</strong> boundary<br />
condition in <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, Aue and Aup are <strong>the</strong> radial boundary conditions for<br />
<strong>the</strong> equatorial and polar axes. Ave and Avp represent <strong>the</strong> boundary conditions for <strong>the</strong><br />
tangential problem. Last Ars and Ats are <strong>the</strong> radial and tangential boundary conditions<br />
at <strong>the</strong> surface. The right-hand side represents <strong>the</strong> tidal acceleration and <strong>the</strong> boundary<br />
constrains, where Fr = atides r and Fθ = atides θ , from <strong>the</strong> Equations (2.2.6)-(2.2.7). The<br />
boundary constrains are <strong>the</strong> constrain <strong>of</strong> no shear at <strong>the</strong> surface and <strong>the</strong> symmetry<br />
constrains <strong>of</strong> <strong>the</strong> polar and equatorial axes. The diagonal dotted lines in Au, Av, Ars<br />
and Ats shows <strong>the</strong> tridiagonal form <strong>of</strong> <strong>the</strong> matrices.<br />
6.4.2 The Coefficient Matrix<br />
The tricky part is to construct <strong>the</strong> total coefficient matrix for both <strong>the</strong> radial and tangential<br />
system. It is crucial to know <strong>the</strong> structure <strong>of</strong> <strong>the</strong> coefficient matrix, to keep track<br />
<strong>of</strong> <strong>the</strong> entries. The two systems are connected through a row ordering matrix, with <strong>the</strong>
50 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
Au<br />
0-matrix<br />
Ac<br />
Aue<br />
Aup<br />
Ave<br />
Avp<br />
Ars<br />
Ats<br />
0-matrix<br />
Av<br />
u<br />
v<br />
=<br />
Fr<br />
Fθ<br />
0 − vector<br />
Figure 6.4.2: Sketch <strong>of</strong> <strong>the</strong> linear system <strong>of</strong> equations, Ax = F. u and v are <strong>the</strong> radial and tangential<br />
displacement, respectively. Au and Av are <strong>the</strong> coefficients <strong>of</strong> <strong>the</strong> radial and tangential<br />
problem, given in <strong>the</strong> Equations (4.3.5) and (4.3.6). The coefficient matrices are dominated by<br />
tridiagonal lines, where <strong>the</strong> rest <strong>of</strong> <strong>the</strong> entries are zero. Ac represent <strong>the</strong> boundary condition<br />
in <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, Aue and Aup are <strong>the</strong> radial boundary conditions for <strong>the</strong> equatorial<br />
and <strong>the</strong> polar axes. Ave and Avp represents <strong>the</strong> boundary conditions for <strong>the</strong> tangential problem.<br />
Last Ars and Ats are <strong>the</strong> radial and tangential boundary conditions at <strong>the</strong> surface, constrained<br />
<strong>of</strong> both v and u. The right-hand side represents <strong>the</strong> tidal acceleration, where Fr = atides r and<br />
, and containing <strong>the</strong> boundary constrains.<br />
Fθ = a tides<br />
θ<br />
following relation<br />
k(i, j) = j + n(i − 1), for i = 1, 2, . . . , n, j = 1, 2, . . . , m.<br />
This is called natural row ordering. k is a one dimensional array counting from 1 in <strong>the</strong><br />
top right corner to m on <strong>the</strong> top left corner. The last number is given in <strong>the</strong> right bottom<br />
corner and is n · m. k(i, j) is a two dimensional array counting from (1, 1) to (n, m), and<br />
it is very useful in determine <strong>the</strong> location <strong>of</strong> a given value. The whole model is build<br />
on <strong>the</strong> row ordering matrix k(i, j). i and j are <strong>the</strong> same index as given in Section 4.3.1,<br />
where i represent <strong>the</strong> radial direction from <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong> to <strong>the</strong> surface, and j<br />
represent <strong>the</strong> co-latitudinal direction from <strong>the</strong> polar axis to <strong>the</strong> equatorial axis.<br />
The coupled coefficient matrix A is a sparse 7824 × 7392 matrix i.e. any nonzero<br />
element is squeezed out and <strong>the</strong> system is discretized with 57,833,333 grid points,<br />
where 54,641,660 are inner points and 3,193,344 grid points at <strong>the</strong> boundary. The data<br />
vector (or <strong>the</strong> tidal acceleration vector) has <strong>the</strong> dimension 7824 × 1, where Fr and Fθ<br />
both have <strong>the</strong> dimension 3696×1. The numerical rank <strong>of</strong> <strong>the</strong> full matrix A is calculated<br />
to 7338. A table <strong>of</strong> <strong>the</strong> system values are given in Table 6.4.1.<br />
A sketch <strong>of</strong> <strong>the</strong> coefficient matrix applied on <strong>the</strong> quarter <strong>Earth</strong>, is depicted in Figure<br />
6.4.3. It shows <strong>the</strong> <strong>Earth</strong> in polar coordinates (r, θ) from <strong>the</strong> center to <strong>the</strong> surface <strong>of</strong><br />
<strong>the</strong> <strong>Earth</strong> and from <strong>the</strong> North Pole to <strong>the</strong> Equator. The different parts <strong>of</strong> <strong>the</strong> coefficient<br />
matrix are applied on <strong>the</strong> sketch to visualize and explain <strong>the</strong> boundary conditions. To<br />
evaluate <strong>the</strong> final results, <strong>the</strong> resulting vector is divided in <strong>the</strong> radial and tangential
Model Development 51<br />
Model Parameters Symbol Size<br />
Radial mesh size (m) hr 9.6530 · 10 4<br />
Latitudinal mesh size (rad) hθ 0.0280<br />
Number <strong>of</strong> radial grid points n 66<br />
Number <strong>of</strong> latitudinal grid points m 56<br />
Number <strong>of</strong> grid points 3696<br />
Size <strong>of</strong> u n × m 66 × 56<br />
Size <strong>of</strong> v n × m 66 × 56<br />
Size <strong>of</strong> sparse coefficient matrix N × M 7824 × 7392<br />
Total number <strong>of</strong> grid points 57,835,008<br />
Number <strong>of</strong> inner grid points 54,641,660<br />
Number <strong>of</strong> boundary grid points 3,193,344<br />
Rank (numerical) rank(A) 7338<br />
Table 6.4.1: Summarizing <strong>the</strong> model parameters. The first part <strong>of</strong> <strong>the</strong> table contains <strong>the</strong> finite<br />
difference discretization parameters for <strong>the</strong> radial and <strong>the</strong> latitudinal problem, where <strong>the</strong><br />
number <strong>of</strong> grid points referred to <strong>the</strong> radial or <strong>the</strong> latitudinal problem. The second part are<br />
<strong>the</strong> parameters for <strong>the</strong> coupled system. The bottom part gives <strong>the</strong> matrix properties <strong>of</strong> <strong>the</strong><br />
coefficient matrix.<br />
The North Pole<br />
Pole axis<br />
Aup/Avp<br />
Ac<br />
Center <strong>of</strong> <strong>Earth</strong><br />
Au/Av<br />
The free surface<br />
Ars/Ats<br />
r<br />
Aue/Ave<br />
Equatorial axis<br />
θ<br />
The Equator<br />
Figure 6.4.3: Sketch <strong>of</strong> <strong>the</strong> coefficient matrix, applied on one quadrant. Ac represent <strong>the</strong><br />
boundary condition in <strong>the</strong> center <strong>of</strong> <strong>Earth</strong>, Aue and Aup are <strong>the</strong> radial boundary conditions<br />
for <strong>the</strong> equatorial and polar axes. Ave and Avp represent <strong>the</strong> boundary conditions for <strong>the</strong><br />
tangential problem. Last Ars and Ats are <strong>the</strong> radial and tangential boundary condition at <strong>the</strong><br />
surface, constrained <strong>of</strong> both v and u.<br />
components and <strong>the</strong> row ordering array is transfered back to only containing i’s and<br />
j’s. Finally <strong>the</strong> dimension <strong>of</strong> u and v becomes 66 × 56.
Chapter 7<br />
Model Considerations and Tests<br />
In last chapter <strong>the</strong> models were constructed and in this chapter <strong>the</strong> problems which<br />
have occurred in <strong>the</strong> modeling are described. This chapter stats out by analyzing <strong>the</strong><br />
linear system <strong>of</strong> equations, followed by a determination <strong>of</strong> <strong>the</strong> regularization parameter<br />
and description <strong>of</strong> <strong>the</strong> different test which have been made on <strong>the</strong> system.<br />
7.1 Model Considerations<br />
As revealed in <strong>the</strong> end <strong>of</strong> Chapter 4, it was not possible to get a reliable result by<br />
solving <strong>the</strong> linear system <strong>of</strong> equations with matrix inversion or even a damped least<br />
square solution.<br />
By using picard.m from Hansen [2008a], <strong>the</strong> singular values, σi, and <strong>the</strong> Fourier<br />
coefficients from Equation (5.2.7) are plotted. From Figure 7.1.1, it is seen, that <strong>the</strong><br />
ill-conditioned problem does not satisfy <strong>the</strong> discrete Picard condition. Following Section<br />
5.2.2, this states that <strong>the</strong> Tikhonov regularization can not give a reliable result for<br />
a discrete ill-posed problem. This could indicate, that <strong>the</strong> system is not a discrete illposed<br />
problem.<br />
As mentioned in Section 6.4.2, <strong>the</strong> numerical rank <strong>of</strong> <strong>the</strong> coefficient matrix A is 7338.<br />
The coefficient matrix has more rows (7824) than columns (7392), which makes it a<br />
thin matrix and an overdetermined system. There are more equations than unknowns,<br />
this means, <strong>the</strong> system is ill-conditioned and <strong>the</strong> coefficient matrix is both over- and<br />
underdetermined, [Aster et al., 2005].<br />
Regularization <strong>of</strong> <strong>the</strong> system is necessary, and as described in Chapter 5, choosing<br />
<strong>the</strong> type <strong>of</strong> regularization method is a cumbersome process and needs analysis <strong>of</strong> <strong>the</strong><br />
system in relation to errors and physical interpretation. The regularization choice fell<br />
on <strong>the</strong> Tikhonov regularization based on trial and error, this is described in fur<strong>the</strong>r<br />
details in Section 7.4.1.
54 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
10 15<br />
10 10<br />
10 5<br />
10 0<br />
10 −5<br />
10 −10<br />
10 −15<br />
10 −20<br />
σ<br />
i<br />
T<br />
|u F|<br />
i<br />
T<br />
|u F|/σi<br />
i<br />
10<br />
0 1000 2000 3000 4000 5000 6000 7000 8000<br />
−25<br />
i<br />
Figure 7.1.1: The Picard plot showing <strong>the</strong> singular values, σi, (blue line), |uT i | (green x) and<br />
<strong>the</strong>ir ratio (red o) , i are <strong>the</strong> entries.<br />
7.2 Analyzing <strong>the</strong> System<br />
To go into a deeper analysis <strong>of</strong> <strong>the</strong> systems ill-conditioning, <strong>the</strong> singular values <strong>of</strong> <strong>the</strong><br />
coefficient matrix A are analyzed. This is done separately for <strong>the</strong> homogeneous and<br />
<strong>the</strong> layered <strong>Earth</strong> models.<br />
7.2.1 Singular Value Spectrum for <strong>the</strong> Homogeneous Model<br />
The condition number <strong>of</strong> <strong>the</strong> homogeneous <strong>Earth</strong> system is calculated by Equation<br />
(5.2.4) to be 2.5130 · 10 23 . The large condition number <strong>of</strong> matrix A, states a very illconditioned<br />
system, that means it is very sensitive to perturbations in <strong>the</strong> data space,<br />
vector F.<br />
In Figure 7.2.1, <strong>the</strong> singular value spectra for <strong>the</strong> homogeneous model is depicted.<br />
Figure 7.2.1(a) shows <strong>the</strong> whole singular value spectrum. From this plot it is reasonable<br />
to think that <strong>the</strong> system is rank deficient, because <strong>of</strong> <strong>the</strong> large gap between<br />
<strong>the</strong> large and small singular values. Notice <strong>the</strong> large value <strong>of</strong> <strong>the</strong> singular values on<br />
5.69 · 10 6 . The o<strong>the</strong>r three figures, Figure 7.2.1(b)-(d) are zoom levels <strong>of</strong> <strong>the</strong> singular<br />
value spectrum. Figure 7.2.1(b) is from <strong>the</strong> beginning <strong>of</strong> <strong>the</strong> spectrum, and shows <strong>the</strong><br />
first cluster <strong>of</strong> large singular values, It is most <strong>of</strong> all shown because, this part looks<br />
much like a discrete ill-posed problem with <strong>the</strong> sequently decay <strong>of</strong> singular values<br />
without any significant gap. Figure 7.2.1(c) is a plot <strong>of</strong> <strong>the</strong> small singular values. The<br />
lower values start at i = 113 with a singular value <strong>of</strong> 33.85. This i states <strong>the</strong> singular<br />
value entries and is not to be compared with <strong>the</strong> FDE index representing <strong>the</strong> radius.<br />
The last plot, Figure 7.2.1(d) shows <strong>the</strong> end <strong>of</strong> <strong>the</strong> spectrum, where <strong>the</strong> singular values<br />
goes to zero, or very close to (10· −16 ) , which is most likely due to round-<strong>of</strong>f errors,<br />
this happens at i = 7339, this also agree with <strong>the</strong> rank <strong>of</strong> <strong>the</strong> system <strong>of</strong> 7338.
Analyzing <strong>the</strong> System 55<br />
Singular values<br />
Singular values<br />
x 106<br />
6<br />
5<br />
4<br />
3<br />
2<br />
1<br />
0<br />
0 1000 2000 3000 4000 5000 6000 7000 8000<br />
i<br />
1.8<br />
1.6<br />
1.4<br />
1.2<br />
1<br />
0.8<br />
0.6<br />
0.4<br />
0.2<br />
0<br />
(a)<br />
1000 2000 3000 4000 5000 6000 7000<br />
i<br />
(c)<br />
Singular values<br />
Singular values<br />
5.4<br />
5.35<br />
5.3<br />
5.25<br />
x 10 6<br />
5.2<br />
15 20 25 30 35<br />
i<br />
40 45 50<br />
x 10<br />
5<br />
−4<br />
4<br />
3<br />
2<br />
1<br />
0<br />
−1<br />
(b)<br />
7290 7300 7310 7320 7330 7340 7350<br />
i<br />
Figure 7.2.1: Singular spectra <strong>of</strong> <strong>the</strong> homogeneous model. (a) shows <strong>the</strong> whole singular<br />
spectrum. Notice <strong>the</strong> large gap between <strong>the</strong> high and small singular values. (b) shows <strong>the</strong><br />
spectrum from i = 14 to i = 52. This is <strong>the</strong> second cluster <strong>of</strong> high singular values. (c) shows<br />
a section <strong>of</strong> <strong>the</strong> low singular values, <strong>the</strong>y clearly goes towards zero. (d) here is an o<strong>the</strong>r break<br />
point, where <strong>the</strong> singular values goes from values <strong>of</strong> 10 −4 to values <strong>of</strong> 10 −16 .<br />
7.2.2 Weighting <strong>of</strong> <strong>the</strong> Surface Boundary Conditions<br />
The inner points are valued much more in <strong>the</strong> Tikhonov regularization, due to <strong>the</strong> larger<br />
amount <strong>of</strong> points. Remember <strong>the</strong> 54, 641, 660 inner points against <strong>the</strong> 3, 193, 344<br />
boundary points, where 827, 904 <strong>of</strong> <strong>the</strong> points are connected to <strong>the</strong> boundary. Multiplying<br />
a factor, f fBC, on <strong>the</strong> surface boundary conditions, assign a weight on <strong>the</strong><br />
surface boundary conditions. This factor is not to be confused with <strong>the</strong> filter factor<br />
applied by <strong>the</strong> Tikhonov regularization to <strong>the</strong> whole system. The boundary factor is<br />
chosen to f fBC = 10 −5 , on <strong>the</strong> basis <strong>of</strong> <strong>the</strong> error plot in Figure 7.2.2 and physical interpretation,<br />
surprisingly <strong>the</strong> best value was less than 1, i.e. <strong>the</strong> factor puts less weight<br />
on <strong>the</strong> surface boundary condition. The figure show <strong>the</strong> residual error against different<br />
factor values in a double logarithm representation.<br />
The result <strong>of</strong> this weighting is very clear when plotting <strong>the</strong> new singular spectrum, see<br />
<strong>the</strong> plot in Figure 7.2.3. Figure 7.2.3(a) shows <strong>the</strong> whole singular spectrum with <strong>the</strong><br />
applied boundary factor. Notice how <strong>the</strong> large singular values have gone from a factor<br />
<strong>of</strong> 10 6 to 56.94. Looking at <strong>the</strong> spectrum from i = 1050 to i = 1850 given in Figure<br />
7.2.3(c), it is not clear, that <strong>the</strong> spectrum is rank deficient, but zooming at i = 300 to<br />
(d)
56 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
|F−A*X|<br />
10 −4<br />
10 −5<br />
10 −6<br />
10 −7<br />
10 −8<br />
10 −9<br />
10 −15<br />
10 −10<br />
Filter factor on Surface BC<br />
10 −10<br />
10 −5<br />
10 0<br />
Surfece factor, ffBC<br />
Figure 7.2.2: Logarithmic error plot <strong>of</strong> different surface boundary factors. f fBC = 10 −5 is<br />
chosen after physical interpretation <strong>of</strong> <strong>the</strong> results and <strong>the</strong> errors <strong>of</strong> <strong>the</strong> system.<br />
Singular values<br />
Singular values<br />
60<br />
50<br />
40<br />
30<br />
20<br />
10<br />
0<br />
0 1000 2000 3000 4000 5000 6000 7000 8000<br />
i<br />
0.25<br />
0.2<br />
0.15<br />
0.1<br />
0.05<br />
0<br />
(a)<br />
1100 1200 1300 1400 1500 1600 1700 1800<br />
i<br />
(c)<br />
Singular values<br />
Singular values<br />
1.5<br />
1.4<br />
1.3<br />
1.2<br />
1.1<br />
1<br />
x 10<br />
14<br />
−4<br />
12<br />
10<br />
8<br />
6<br />
4<br />
2<br />
0<br />
10 5<br />
10 10<br />
350 400 450 500<br />
i<br />
(b)<br />
7150 7200 7250 7300 7350<br />
i<br />
Figure 7.2.3: Singular spectra for <strong>the</strong> homogeneous <strong>Earth</strong> model with surface boundary factor<br />
<strong>of</strong> f fBC = 10 −5 . (a) shows <strong>the</strong> whole spectrum for <strong>the</strong> new singular values. (b) is <strong>the</strong> spectrum<br />
from 300 to 550, <strong>the</strong> characteristic gap for a rank deficient problem is very clear. (c) shows <strong>the</strong><br />
spectrum from 1050 to 1850. (d) shows <strong>the</strong> end part <strong>of</strong> <strong>the</strong> spectrum, where <strong>the</strong> cluster <strong>of</strong> very<br />
small singular values begin. This spectrum goes from 7100 to 7392.<br />
(d)
Analyzing <strong>the</strong> System 57<br />
i = 550, 7.2.3(b) we see <strong>the</strong> characteristic gaps. The last subfigure, Figure 7.2.3(d)<br />
shows <strong>the</strong> spectrum going to zero at i = 7339 or very close to (2.6125 · 10 −16 ).<br />
As a consequence <strong>of</strong> lowering <strong>the</strong> large singular values, <strong>the</strong> condition number has<br />
decreased. The new condition number <strong>of</strong> <strong>the</strong> system is 1.2848 · 10 18 , which is a clear<br />
improvement <strong>of</strong> <strong>the</strong> earlier result <strong>of</strong> 2.5130·10 23 . The condition number is still to large<br />
to rely on <strong>the</strong> solution i.e. regularization is necessary!<br />
7.2.3 Singular Value Spectrum for <strong>the</strong> Layered Model<br />
The layered <strong>Earth</strong> model is analyzed like <strong>the</strong> homogeneous model in previous section,<br />
and <strong>the</strong> same factor f fBC = 10 −5 on <strong>the</strong> surface boundary conditions are applied. It has<br />
been chosen not to show <strong>the</strong> singular spectrum for <strong>the</strong> unfiltered boundary condition<br />
because it resembles <strong>the</strong> homogeneous plot, Figure 7.2.1 in Section 7.2.1.<br />
The singular spectrum for <strong>the</strong> layered model with <strong>the</strong> boundary factor is plotted in<br />
Figure 7.2.4. Figure 7.2.4(a) shows <strong>the</strong> whole spectrum from i = 1 to i = 77329. At<br />
first sight <strong>the</strong> large gap between <strong>the</strong> large and small singular values are <strong>the</strong>re no more,<br />
this will be discussed in more details in Section 9.1. The largest singular value is 66.14.<br />
In Figure 7.2.4(b) <strong>the</strong> spectrum ranges from i = 20 to i = 44. The singular values<br />
decay almost gradually. Figure 7.2.4(c) shows <strong>the</strong> singular spectrum from n = 88 to<br />
n = 148, and it is a sequence with a clear gap in <strong>the</strong> values. The last figure, Figure<br />
7.2.4(d), shows how <strong>the</strong> singular values go towards zero at i = 7339, just like in <strong>the</strong><br />
homogeneous model. The condition number has changed from 1.6340·10 23 to 1.7363·<br />
10 18 with applying <strong>the</strong> boundary condition factor, which is clear an improvement <strong>of</strong><br />
<strong>the</strong> system stability, but still a too large value to relay on <strong>the</strong> system. This means that<br />
regularization is still needed to solve <strong>the</strong> system.
58 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
Singular values<br />
Singular values<br />
70<br />
60<br />
50<br />
40<br />
30<br />
20<br />
10<br />
0<br />
0 1000 2000 3000 4000 5000 6000 7000 8000<br />
i<br />
22<br />
20<br />
18<br />
16<br />
14<br />
12<br />
(a)<br />
10<br />
90 100 110 120<br />
i<br />
130 140<br />
(c)<br />
Singular values<br />
Singular values<br />
51<br />
50.5<br />
50<br />
49.5<br />
49<br />
48.5<br />
8<br />
7<br />
6<br />
5<br />
4<br />
3<br />
2<br />
1<br />
0<br />
x 10 −5<br />
22 24 26 28 30 32 34 36 38 40 42<br />
i<br />
(b)<br />
7320 7325 7330 7335 7340 7345 7350<br />
i<br />
Figure 7.2.4: Singular value spectra for <strong>the</strong> layered <strong>Earth</strong> model with a filter factor on <strong>the</strong><br />
surface boundary conditions. (a) shows <strong>the</strong> whole spectrum, <strong>the</strong> first and largest value is<br />
66.14. (b) shows <strong>the</strong> singular values from i = 20 − 44, <strong>the</strong> values decay almost sequently. (c)<br />
<strong>the</strong> spectrum from i = 88 − 147, it shows large gaps between <strong>the</strong> values. (d) is a zoom <strong>of</strong> <strong>the</strong><br />
spectrum going toward zero at i = 7339.<br />
7.3 Choosing <strong>the</strong> Regularization Parameters<br />
In this section <strong>the</strong> regularization parameters for both <strong>the</strong> homogeneous and layered<br />
<strong>Earth</strong> model are found, and <strong>the</strong> Tikhonov regularization parameters are described. The<br />
regularization parameter is found with <strong>the</strong> L-curve method described in Section 5.2.2,<br />
where <strong>the</strong> residual norm is plotted against <strong>the</strong> solution norm. Choosing various values<br />
<strong>of</strong> ε ranging from 1 to 10 −20 <strong>the</strong> L-curve is generated with l_curve.m from Hansen<br />
[2008a].<br />
The Homogeneous <strong>Earth</strong><br />
The program calculated <strong>the</strong> best regularization parameter to be ε = 4.4802 · 10 −9 , <strong>the</strong><br />
plot can be seen in Figure 7.3.1. I should be mentioned, that <strong>the</strong> numbers in <strong>the</strong> figure<br />
are generated by <strong>the</strong> l_curve.m and do not necessarily show <strong>the</strong> regulation parameter,<br />
which is <strong>the</strong> case in this figure. The parameter is found by <strong>the</strong> program l_corner.m.<br />
For testing <strong>the</strong> regularization effects, regularization parameters deflecting from <strong>the</strong><br />
L-curve generated value, have been tested <strong>the</strong> model. The main characteristics were a<br />
(d)
Choosing <strong>the</strong> Regularization Parameters 59<br />
solution norm || x || 2<br />
10 5<br />
10 0<br />
10 −5<br />
10 −10<br />
10 −8<br />
2.5994e−12<br />
9.1764e−14<br />
4.4802e−09<br />
1.6738e−06<br />
10 −7<br />
10 −6<br />
10 −5<br />
residual norm || A x − b ||<br />
2<br />
4.7414e−05<br />
0.0013431<br />
0.038046<br />
Figure 7.3.1: The L-curve for <strong>the</strong> homogeneous model. The regularization parameter is ε =<br />
4.4802 · 10 −9 . The curve is generated from l_curve.m from Hansen [2008a], and shows <strong>the</strong><br />
residual norm against <strong>the</strong> solution norm.<br />
larger displacement when more regularization was applied, for ε = 10 −6 <strong>the</strong> displacement<br />
<strong>of</strong> <strong>the</strong> <strong>Earth</strong> was 130.7 mm and for ε = 10 −12 <strong>the</strong> maximum radial displacement<br />
was 143.1 mm. For 10 −13 < ε < 10 −5 <strong>the</strong> physical interpretation does not appear<br />
satisfied any more. In Appendix D.1 some early test can be found.<br />
The Layered <strong>Earth</strong><br />
The regularization parameter is determine like <strong>the</strong> homogeneous model, Section 7.3,<br />
by <strong>the</strong> L-curve method, described in Section 5.2.2. See <strong>the</strong> L-curve plot in Figure<br />
7.3.2, <strong>the</strong> regularization parameter is calculated to ɛ = 1.0157 · 10 −8 .<br />
solution norm || x || 2<br />
10 5<br />
10 0<br />
10 −5<br />
10 −10<br />
10 −10<br />
10 −15<br />
2.3496e−13<br />
10 −9<br />
10 −8<br />
6.6557e−12<br />
1.0157e−8<br />
10 −7<br />
residual norm || A x − b || 2<br />
10 −6<br />
4.2857e−06<br />
10 −5<br />
1.0778<br />
2.7595<br />
10 −4<br />
0.0001214<br />
Figure 7.3.2: The L-curve for <strong>the</strong> layered model, with corner at 1.0157 · 10 −8 , which is <strong>the</strong><br />
chosen regularization parameter.<br />
0.003439<br />
0.097416<br />
10 −4
60 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
7.3.1 Tikhonov Filter Factors<br />
The filter factors generated by <strong>the</strong> Tikhonov regularization, Equation 5.2.9 adds a filter<br />
factor corresponding to all <strong>the</strong> singular values, σi, and <strong>the</strong> regularization parameter<br />
ε. The filter factors can be generated by fil_fac.m from Hansen [2008a]. In Figure<br />
7.3.3, <strong>the</strong> filter factors are plotted for <strong>the</strong> homogeneous and layered models.<br />
For <strong>the</strong> homogeneous <strong>Earth</strong> model, <strong>the</strong> first 1402 entries are one, hereafter <strong>the</strong> filter<br />
factors begins decaying, i.e. <strong>the</strong> regularization is applied. Until <strong>the</strong> 7338th value <strong>the</strong><br />
entries are very close to zero, i.e. <strong>the</strong>se values are only smoo<strong>the</strong>n very little. At a value<br />
<strong>of</strong> i = 7338 a filter factor <strong>of</strong> 10 −8 is applied, it decays towards 9.91 · 10 −17 to <strong>the</strong> last<br />
singular value, i = 7392. Notice that <strong>the</strong> regularization begins at i = 7338 <strong>the</strong> same<br />
value as <strong>the</strong> rank <strong>of</strong> <strong>the</strong> system, and where <strong>the</strong> singular values goes towards zero, i.e.<br />
<strong>the</strong> filter factor, filter out <strong>the</strong> small singular values.<br />
The filter factor applied by <strong>the</strong> Tikhonov regularization to <strong>the</strong> layered model starts<br />
filtering at i = 910, and at i = 7338 <strong>the</strong> filter factor decreases rapidly towards 1.44 ·<br />
10 −17 , i.e. <strong>the</strong> values after i = 7338 are more smoo<strong>the</strong>d.<br />
Filter Factor<br />
10 0<br />
10 −5<br />
10 −10<br />
10 −15<br />
homo<br />
layered<br />
−20<br />
10<br />
0 1000 2000 3000 4000 5000 6000 7000 8000<br />
Index, n<br />
(a)<br />
Filter Factor<br />
10 −14<br />
10 −15<br />
−16<br />
10<br />
Filter Factor<br />
10 0<br />
10<br />
−1e−06<br />
homo<br />
layered<br />
7340 7350 7360 7370 7380 7390<br />
Index, n<br />
(c)<br />
7310 7315 7320 7325 7330 7335 7340<br />
Index, n<br />
homo<br />
layered<br />
Figure 7.3.3: Semilogarithmic plot <strong>of</strong> <strong>the</strong> filter factors for <strong>the</strong> homogeneous and layered models.<br />
(b)
Testing <strong>the</strong> System 61<br />
7.4 Testing <strong>the</strong> System<br />
This section aims to show some <strong>of</strong> <strong>the</strong> test, which have been applied on <strong>the</strong> system,<br />
besides <strong>the</strong> test <strong>of</strong> <strong>the</strong> regularization parameters, which have already been covered. The<br />
testing includes, boundary test <strong>of</strong> <strong>the</strong> resulting solution method, regularization steps,<br />
o<strong>the</strong>r solving methods and last <strong>the</strong> applied tidal acceleration is tested. In <strong>the</strong> following<br />
a physical meaningful result, should be understood as a result that fulfill <strong>the</strong> <strong>the</strong>oretical<br />
knowledge <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s tidal displacement field, described in Section 3.4. The tests<br />
are carried out for both models, but <strong>the</strong> results are trivial, <strong>the</strong>refore only <strong>the</strong> results for<br />
<strong>the</strong> layered <strong>Earth</strong> are shown.<br />
7.4.1 Test <strong>of</strong> Solving Methods<br />
When it was clear that <strong>the</strong> solution could not be solved with a simple inversion, X =<br />
A −1 F, because A was not invertible and badly scaled, <strong>the</strong> first test applied <strong>of</strong> <strong>the</strong> model<br />
was a regular Tikhonov solution , or damped least square solution, Equation 5.2.1.<br />
For approximately 10 −6 > ε > 10 −8 , <strong>the</strong>re were no physical meaningful solution,<br />
for values in between, <strong>the</strong> results were physical acceptable. The maximum radial<br />
displacements yield 69.0 ± 0.5 mm, <strong>the</strong> minimum displacement yield −18.5 ± 0.5 mm,<br />
and <strong>the</strong> maximum tangential displacement was 45.1 ± 0.5 mm. This method were<br />
less computational intensive and could <strong>the</strong>refore handle larger arrays, but <strong>the</strong> damped<br />
least square solution was rejected because <strong>of</strong> an infinitely large condition number. The<br />
largest achievable coefficient matrix was 10579 × 10082, with hr = 89.732 km and<br />
hθ = 1.2676 ◦ . For <strong>the</strong> best result, with ε = 10 −5 , <strong>the</strong> residual norm was F − AX =<br />
7.575 · 10 −6 .<br />
Though <strong>the</strong> Truncated Singular Value Decomposition (TSVD) is <strong>the</strong> most advisable<br />
approach <strong>of</strong> rank deficient problems, [Hansen, 1996; Aster et al., 2005], <strong>the</strong><br />
Tikhonov regularization gave better results for this study. The TSVD method is advised<br />
for rank-deficient problems, because it handles <strong>the</strong> numerical rank <strong>of</strong> <strong>the</strong> coefficient<br />
matrix. The singular values are cut <strong>of</strong> at <strong>the</strong> chosen truncation factor, normally<br />
this factor is <strong>the</strong> rank. The filter factors in TSVD decay slower than <strong>the</strong> Tikhonov filter<br />
factors, [Hansen, 1996] . The TSVD have been carried out for different truncations, but<br />
no satisfying results have been found. Also QR regularization with Householder and<br />
a damped singular value decomposition (DSVD) have been tested, and also without<br />
<strong>the</strong> fruitful results. In short, QR decomposition factorize <strong>the</strong> coefficient matrix into an<br />
orthogonal and a upper-triangular matrix, where householder is a computational tool<br />
to make a matrix transformation. The DSVD method is quite similar to TSVD, but <strong>the</strong><br />
DSVD does not cut away any singular values, [Aster et al., 2005]. Examples <strong>of</strong> <strong>the</strong><br />
tests can be found in Appendix D.2.<br />
7.4.2 The General Solution<br />
The general damped least square method has been tested, and been carried out by<br />
changing <strong>the</strong> roughening matrix L in Equation (5.2.1). L is changed from <strong>the</strong> identity<br />
matrix by implementing a finite difference approximation to <strong>the</strong> Laplacian i.e. second<br />
order derivatives in both directions. When solving <strong>the</strong> general Tikhonov, Matlab has
62 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
to make a general singular value decomposition, gsvd. This function needs to handle<br />
an even larger data set than <strong>the</strong> svd function, because it is storing two orthonormal<br />
matrices, a non singular matrix and two diagonal matrices, where <strong>the</strong> svd function<br />
only stores three matrices, described in Section 5.2.1.<br />
The maximum grid numbers achieved with this method were (40 × 40) corresponding<br />
to mesh sizes <strong>of</strong> hr = 155.39 km and hθ = 2.1951 ◦ . The results were<br />
physical meaningful, with a best fit <strong>of</strong> ε = 1.7453 · 10 −7 . The maximum radial displacements<br />
was 76.4365 mm and <strong>the</strong> minimum was −23.7960 mm, and <strong>the</strong> maximum<br />
tangential displacement was 47.5471 mm. These results gave <strong>the</strong> residual norm<br />
F − AX = 7.1012 · 10 −6 . It could have been possible to use <strong>the</strong> general form, but<br />
it has been chosen not use this method, because <strong>of</strong> <strong>the</strong> grid number limitations, and<br />
because <strong>the</strong> general solution did not elucidate better results.<br />
7.4.3 Testing <strong>the</strong> Weighting <strong>of</strong> Grid Numbers<br />
The layered <strong>Earth</strong> model has been tested for different grid numbers. In Figure 7.4.1,<br />
<strong>the</strong>re are shown four selected tests for both <strong>the</strong> radial and tangential displacement<br />
fields. In Figure 7.4.1(a)-(b) <strong>the</strong> grid numbers are (6 × 6) with mesh sizes <strong>of</strong> hr =<br />
1061 km and hθ = 15.0 ◦ , Figure 7.4.1(c)-(d) are a (16 × 16) grid with mesh sizes <strong>of</strong><br />
hr = 398 km and hθ = 5.6 ◦ , Figure 7.4.1(e)-(f) displays a (36 × 36) grid , with <strong>the</strong><br />
mesh sizes hr = 177 km and hθ = 2.5 ◦ and last in Figure 7.4.1(g)-(g) a (66 × 56) grid<br />
is shown, with mesh sizes <strong>of</strong> hr = 97 km and hθ = 1.6 ◦ . As it is shown from <strong>the</strong> figure,<br />
<strong>the</strong> model fails to give a physical meaningful result with only (6 × 6) grid numbers.<br />
The radial displacement field, Figure 7.4.1(a), becomes a purely expansive field, and<br />
<strong>the</strong> tangential field has developed a negative traction at <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. Figure<br />
7.4.1(e)-(e) gives a nicer picture, <strong>the</strong> radial displacement field has a compression at <strong>the</strong><br />
North Pole, though it only extend to about θ = 15 ◦ . The result from Figure 7.4.1(g)-(g)<br />
will be described in more details in Section 8.2.<br />
Comparing <strong>the</strong> three tests, displayed in Table 7.4.1, it is clear how <strong>the</strong> magnitude <strong>of</strong><br />
<strong>the</strong> displacements are dependent <strong>of</strong> <strong>the</strong> grid numbers. The magnitude increases as <strong>the</strong><br />
number <strong>of</strong> grid points increases, <strong>the</strong> growths goes very fast until about a grid number<br />
<strong>of</strong> 30 × 30, where <strong>the</strong> growth increases to a smaller degree, it could appear as <strong>the</strong><br />
growth <strong>of</strong> <strong>the</strong> displacement stabilizes. In Figure 7.4.2 a plot <strong>of</strong> <strong>the</strong> radial displacement<br />
growth is displayed. Notice also how <strong>the</strong> displacement for <strong>the</strong> grid 50 × 50 actually is<br />
larger than <strong>the</strong> 66 × 56 grid.
Testing <strong>the</strong> System 63<br />
R<br />
R<br />
R<br />
R<br />
R<br />
R<br />
R<br />
R<br />
(a)<br />
(c)<br />
(e)<br />
(g)<br />
θ<br />
θ<br />
θ<br />
θ<br />
0.035<br />
0.03<br />
0.025<br />
0.02<br />
0.015<br />
0.01<br />
0.005<br />
0<br />
−0.005<br />
−0.01<br />
−0.015<br />
0.06<br />
0.05<br />
0.04<br />
0.03<br />
0.02<br />
0.01<br />
0<br />
−0.01<br />
0.07<br />
0.06<br />
0.05<br />
0.04<br />
0.03<br />
0.02<br />
0.01<br />
0<br />
−0.01<br />
Radial displacement, u(r, θ) in m .<br />
0.07<br />
0.06<br />
0.05<br />
0.04<br />
0.03<br />
0.02<br />
0.01<br />
0<br />
−0.01<br />
−0.02<br />
Radial displacement, u(r, θ ) in m .<br />
Radial displacement, u(r, θ ) in m .<br />
Radial displacement, u(r, θ ) in m .<br />
Figure 7.4.1: Test <strong>of</strong> four different grid sizes given by n × m where n is <strong>the</strong> grid in <strong>the</strong> radial<br />
direction and m is <strong>the</strong> grid in <strong>the</strong> co-latitudinal direction. (a)-(b) are a 6 × 6 grid (c)-(d) are a<br />
16 × 16 grid, (e)-(f) are a 36 × 36 grid and (g)-(h) are a 66 × 56 grid.<br />
R<br />
R<br />
R<br />
R<br />
R<br />
R<br />
R<br />
R<br />
(b)<br />
(d)<br />
(f)<br />
(h)<br />
θ<br />
θ<br />
θ<br />
θ<br />
x 10 −3<br />
0<br />
16<br />
14<br />
12<br />
10<br />
8<br />
6<br />
4<br />
2<br />
0<br />
−2<br />
0.05<br />
0.04<br />
0.03<br />
0.02<br />
0.01<br />
−0.01<br />
0.05<br />
0.04<br />
0.03<br />
0.02<br />
0.01<br />
0<br />
−0.01<br />
0.05<br />
0.04<br />
0.03<br />
0.02<br />
0.01<br />
0<br />
Tangential displacement, v(r, θ ), in m .<br />
−0.01<br />
Tangential displacement, v(r, θ ), in m .<br />
Tangential displacement, v(r, θ ), in m .<br />
Tangential displacement, v(r, θ), in m .
64 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
Grid size<br />
(n, m)<br />
u(n, m)<br />
in mm<br />
u(n, 1)<br />
in mm<br />
v(n, m/2)<br />
in mm<br />
Cond(A)<br />
·10<br />
ε<br />
F − AX<br />
F<br />
18 ·10−8 %<br />
6 x 6 33.9553 2.3392 · 10−4 7.4574 0.1723 180.0938 81.1490<br />
16 x 16 64.1823 -12.7884 38.7450 5.5237 22.0989 53.3956<br />
36 x 36 73.3261 -17.8062 44.9401 1.9434 5.6186 36.4931<br />
66 x 65 74.3154 -26.2925 45.3977 1.6218 2.7670 28.1446<br />
Table 7.4.1: Table <strong>of</strong> <strong>the</strong> parameters in <strong>the</strong> grid size test. The columns are: grid size, where<br />
n is <strong>the</strong> radial direction and m is <strong>the</strong> co-latitudinal direction, u is <strong>the</strong> radial displacement<br />
given at <strong>the</strong> maximum magnitude (n, m), v is <strong>the</strong> maximum tangential displacement, ε is <strong>the</strong><br />
F − AX<br />
regularization parameter and is <strong>the</strong> error given by <strong>the</strong> residual norm over <strong>the</strong> data<br />
F<br />
norm and Cond(A) is <strong>the</strong> condition number <strong>of</strong> <strong>the</strong> system before regularization.<br />
Radial displacement displacement in mm<br />
80<br />
75<br />
70<br />
65<br />
60<br />
55<br />
50<br />
45<br />
40<br />
35<br />
30<br />
6x6<br />
16X16<br />
16X16<br />
21X21<br />
26x26<br />
31x31<br />
Number <strong>of</strong> grid points<br />
Figure 7.4.2: plot <strong>of</strong> <strong>the</strong> growth <strong>of</strong> <strong>the</strong> radial displacement due to increase <strong>of</strong> grid numbers.<br />
7.4.4 Testing <strong>the</strong> Boundary Conditions<br />
The boundary conditions are tested by applying five point sources <strong>of</strong> −10 −3 m s 2 in <strong>the</strong><br />
radial direction instead <strong>of</strong> <strong>the</strong> tidal acceleration, and <strong>the</strong>re is no tangential acceleration.<br />
The negative sign makes it attracting point sources, as <strong>the</strong> tidal acceleration. The five<br />
point sources are placed at <strong>the</strong> grid points (3, 3), (n/2, m/2), (n − 1, 3), (n − 1, m/2) and<br />
(n − 1, m − 3), where n and m are <strong>the</strong> number <strong>of</strong> grid points in <strong>the</strong> radial and latitudinal<br />
directions, counting from <strong>the</strong> center <strong>of</strong> <strong>Earth</strong> and from <strong>the</strong> pole respectively. n/2 and<br />
m/2 corresponds respectively to half <strong>of</strong> <strong>the</strong> radius and quarter <strong>of</strong> <strong>the</strong> circle, review <strong>the</strong><br />
grid in Figure 6.4.1.<br />
This results in <strong>the</strong> displacement field depicted in Figure 7.4.3. The plot shows that<br />
when a point source is applied, no points in <strong>the</strong> center <strong>of</strong> <strong>the</strong> homogeneous <strong>Earth</strong> are<br />
displaced. There is a little deflection in <strong>the</strong> central point, where <strong>the</strong> point source attack<br />
on <strong>the</strong> half radius and quarter circle. The displacement is in this point 170 mm. At <strong>the</strong><br />
surface <strong>of</strong> <strong>the</strong> homogeneous model <strong>the</strong> displacement is <strong>the</strong> greatest, with a maximum<br />
36X36<br />
51X51<br />
66X56
Testing <strong>the</strong> System 65<br />
R<br />
R<br />
Figure 7.4.3: Testing <strong>the</strong> surface conditions with point sources applied at <strong>the</strong> grid points (3, 3),<br />
(n/2, m/2), (n − 1, 3), (n − 1, m/2) and (n − 1, m − 3), where n and m are <strong>the</strong> number <strong>of</strong> grid<br />
−3 points. The point sources has a values <strong>of</strong> −10 m<br />
2 . The figure show <strong>the</strong> resultant displacement<br />
s<br />
field after <strong>the</strong> point sources are applied.<br />
displacement <strong>of</strong> 634 mm. The behavior <strong>of</strong> <strong>the</strong> point source reacts just like if somebody<br />
pulled at a single point. The displacement are largest at <strong>the</strong> attacking point and become<br />
smaller, but still positive, and only in a local area away from <strong>the</strong> pulling direction.<br />
It is remarkable to notice how <strong>the</strong> displacement is largest at <strong>the</strong> attacking point,<br />
and how it become smaller moving far<strong>the</strong>r away, to finally not being influenced by <strong>the</strong><br />
point source anymore.<br />
7.4.5 Testing <strong>the</strong> <strong>Tidal</strong> Acceleration<br />
The radial and tangential components <strong>of</strong> <strong>the</strong> tidal acceleration are plotted in Figure<br />
7.4.4(a)-(b), as functions <strong>of</strong> radius and colatitude on a quarter <strong>Earth</strong>. Note <strong>the</strong> different<br />
color scales. The radial component <strong>of</strong> <strong>the</strong> tidal acceleration is positive from <strong>the</strong> North<br />
Pole to θ = 36◦ and negative from this point on to <strong>the</strong> equator axis. The acceleration<br />
applied on <strong>the</strong> surface at Equator is Fr = −1.0999m s2 . The tangential component <strong>of</strong><br />
<strong>the</strong> tidal acceleration has <strong>the</strong> largest negative acceleration at θ = 45 ◦ with a value <strong>of</strong><br />
Fθ = −0.8224m s2 . These tidal component values are consistent with Torge [2001].<br />
Figure 7.4.4(c) shows <strong>the</strong> radial displacement field in m <strong>of</strong> <strong>the</strong> quarter <strong>Earth</strong>, where<br />
only a radial dependent tidal acceleration is applied. As excepted <strong>the</strong> <strong>Earth</strong> expands<br />
outwards over <strong>the</strong> whole volume.<br />
θ<br />
0. 6<br />
0. 5<br />
0.4<br />
0.3<br />
0.2<br />
0.1<br />
0<br />
−0.1<br />
−0.2<br />
Radial displacement in m, u(r, θ ) , in m.
66 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
R<br />
R<br />
(a)<br />
R<br />
θ<br />
x 10<br />
5<br />
−7<br />
0<br />
−5<br />
−10<br />
Radial comp. <strong>of</strong> tidal acc. in m/s 2 .<br />
R<br />
(c)<br />
Figure 7.4.4: (a) The radial component and (b) tangential component <strong>of</strong> <strong>the</strong> tidal accelerations<br />
applied on a quarter <strong>Earth</strong>. (c) The radial displacement field <strong>of</strong> <strong>the</strong> <strong>Earth</strong> due to a radial<br />
dependent tidal acceleration alone. Note <strong>the</strong> different color scales.<br />
R<br />
θ<br />
0.035<br />
0.03<br />
0.025<br />
0.02<br />
0.015<br />
0.01<br />
0.005<br />
0<br />
−0.005<br />
−0.01<br />
θ), in m .<br />
R<br />
(b)<br />
Radial displacement, u(r,<br />
θ<br />
x 10<br />
0<br />
−7<br />
−1<br />
−2<br />
−3<br />
−4<br />
−5<br />
−6<br />
−7<br />
−8<br />
Tangential comp. <strong>of</strong> tidal acc. in m/s 2<br />
.
Chapter 8<br />
The <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Earth</strong><br />
This chapter aims to describe <strong>the</strong> resulting deformation <strong>of</strong> <strong>the</strong> homogeneous and layered<br />
<strong>Earth</strong> Models due to <strong>the</strong> tidal acceleration. Also a analytic solution <strong>of</strong> <strong>the</strong> displacement<br />
at <strong>the</strong> surface is elaborated, and compared to <strong>the</strong> numerical results.<br />
8.1 <strong>Deformation</strong> in <strong>the</strong> Homogeneous <strong>Earth</strong> Model<br />
In this section, <strong>the</strong> final results from <strong>the</strong> modeling is given. The results are given as <strong>the</strong><br />
displacement <strong>of</strong> <strong>the</strong> homogeneous <strong>Earth</strong> due to <strong>the</strong> solid tides. There are no particular<br />
difference in <strong>the</strong> results for <strong>the</strong> homogeneous <strong>Earth</strong> models with and without boundary<br />
factor. The boundary factor f fBC only changed <strong>the</strong> displacement by 10 −4 mm, it is<br />
chosen not to show <strong>the</strong> results for <strong>the</strong> model without <strong>the</strong> boundary factor, because <strong>of</strong><br />
triviality.<br />
The results for <strong>the</strong> homogeneous <strong>Earth</strong> model are plotted in Figure 8.1.1, and a representation<br />
<strong>of</strong> <strong>the</strong> displacements are summarized in Table 8.1.1. The figure shows <strong>the</strong><br />
displacements <strong>of</strong> <strong>the</strong> <strong>Earth</strong> in polar coordinates, displayed on a quarter sphere, with <strong>the</strong><br />
center <strong>of</strong> <strong>the</strong> <strong>Earth</strong> at <strong>the</strong> bottom left corner, <strong>the</strong> North Pole at <strong>the</strong> top left corner and<br />
<strong>the</strong> Equator at <strong>the</strong> bottom right corner. The radius are ranging from <strong>the</strong> center to <strong>the</strong><br />
surface <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. . The colatitude from <strong>the</strong> Pole to <strong>the</strong> Equator. The displacement<br />
is visualized by a color scale ranging from -0.05 m to about 0.13 m. The colours in<br />
<strong>the</strong> colour bar are not exactly alike in all figures, but <strong>the</strong>re is emphasis on that green<br />
symbolizing zero displacement, blue negative displacement and <strong>the</strong> yellow-reddish<br />
colour symbolizing positive displacement. Notice also <strong>the</strong> contour lines showing <strong>the</strong><br />
displacement values.<br />
The table columns are given as follows: 1.) grid points (i, j) where i and j represents<br />
<strong>the</strong> radius and colatitude, respectively, 2.) <strong>the</strong> radius coordinate, r, in kilometer<br />
and 3.) <strong>the</strong> colatitudinal coordinate, θ in degrees corresponding to <strong>the</strong> grid. 4) is <strong>the</strong><br />
radial displacement, v, 5.) <strong>the</strong> tangential displacement u, 6.) <strong>the</strong> total displacement, u,<br />
all <strong>the</strong> displacements are given in millimeters.<br />
The three figures, displayed in Figure 8.1.1, are <strong>the</strong> radial displacement field, Figure
68 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
8.1.1(a), <strong>the</strong> tangential displacement field, Figure 8.1.1(b), and <strong>the</strong> total displacement,<br />
Figure 8.1.1(c). The total displacement field is <strong>the</strong> squared sum <strong>of</strong> <strong>the</strong> radial and tangential<br />
displacement fields.<br />
The displacement is negative at <strong>the</strong> North Pole with a value <strong>of</strong> u = −5 mm, i.e.<br />
<strong>the</strong> point is compressed towards <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. Going along <strong>the</strong> surface from<br />
<strong>the</strong> Pole to approximately θ = 15 ◦ , <strong>the</strong> displacement becomes positive, and continuing<br />
towards <strong>the</strong> Equator <strong>the</strong> radial displacement gets larger until reaching its maximum<br />
value at <strong>the</strong> Equator, here <strong>the</strong> displacement is u = 131 mm expanding towards <strong>the</strong><br />
Moon. In <strong>the</strong> direction from <strong>the</strong> surface <strong>of</strong> <strong>the</strong> <strong>Earth</strong> to <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong> <strong>the</strong><br />
displacement looses strength until it becomes zero at <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. At <strong>the</strong><br />
equatorial axis <strong>the</strong> displacement is u = 125 mm in a radius <strong>of</strong> 4313 km, u = 86 mm at<br />
2353 km and u = 21 mm at 392 km. Notice <strong>the</strong> contour lines being perpendicular to<br />
<strong>the</strong> polar and <strong>the</strong> equatorial axes, but not at <strong>the</strong> surface boundary.<br />
The tangential displacement field is characterized by being zero at <strong>the</strong> polar and<br />
<strong>the</strong> equatorial axes with maximum at <strong>the</strong> surface at θ = 45 ◦ . The model parameters are<br />
not completely zero, but this is probably a numerical error. The maximum tangential<br />
displacement is v = 41 mm, which is a tractive movement in <strong>the</strong> tangential plane. At<br />
θ = 45 ◦ and a radius <strong>of</strong> 4313 km <strong>the</strong> tangential displacement is v = 39 mm, at 2353<br />
km a point is stretched v = 11 mm and at 393 km from <strong>the</strong> center <strong>the</strong> displacement<br />
is v = 0.075 mm. The contour lines are not completely tangent to <strong>the</strong> surface. The<br />
total displacement field has a maximum displacement at <strong>the</strong> Equator, θ = 90 ◦ <strong>of</strong> u =<br />
131 mm.
<strong>Deformation</strong> in <strong>the</strong> Homogeneous <strong>Earth</strong> Model 69<br />
(a) Radial displacement<br />
(b) Tangential displacement<br />
(c) Total displacement<br />
Figure 8.1.1: The displacement (in meter) for <strong>the</strong> homogeneous <strong>Earth</strong> model with surface<br />
boundary factor and Tikhonov regularization with ɛ = 4.4802 · 10 −9 . (a) shows <strong>the</strong> radial displacement<br />
field, (b) shows <strong>the</strong> tangential displacement field and (c) shows <strong>the</strong> total displacement<br />
field. The displacements are given for a quarter sphere, where <strong>the</strong> bottom left corner is<br />
<strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, top left corner is <strong>the</strong> North Pole, and <strong>the</strong> bottom right corner is <strong>the</strong><br />
Equator.
70 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
Grid points r (km) θ ( ◦ ) u (mm) v (mm) u (mm)<br />
(1,1) 0 0 −6.6723 · 10 −15 1.2442 · 10 −13 1.2460 · 10 −13<br />
(5,1) 392 0 19.2574 2.4379 · 10 −14 19.2574<br />
(5,28) 392 45 20.5049 0.075 20.5051<br />
(5,55) 392 90 20.7787 3.9377 · 10 −14 20.7787<br />
(25,1) 2353 0 36.562 −5.9404 · 10 −14 36.562<br />
(25,10) 2353 15 43.319 6.4133 43.7912<br />
(25,28) 2353 45 71.0693 10.86 71.8942<br />
(25,45) 2353 72 84.794 5.2619 84.9571<br />
(25,56) 2353 90 86.1656 4.9518 · 10 −14 86.1656<br />
(45,1) 4313 0 -0.4317 −2.7243 · 10 −12 14.4317<br />
(45,10) 4313 15 25.0779 18.6913 31.2773<br />
(45,25) 4313 39 61.6813 38.0125 81.1367<br />
(45,28) 4313 45 81.5458 38.7285 90.2752<br />
(45,45) 4313 72 119.0393 23.4878 121.3344<br />
(45,56) 4313 90 125.4658 1.6143 · 10 −13 125.4658<br />
(66,1) 6371 0 -4.5777 −6.0341 · 10 −9 4.5777<br />
(66,10) 6371 15 -1.5318 13.5711 14.3078<br />
(66,20) 6371 31 35.4273 31.3569 47.3112<br />
(66,28) 6371 45 67.2545 41.2706 78.9078<br />
(66,40) 6371 72 121.2756 30.4703 125.0449<br />
(66,50) 6371 80 128.4876 18.3878 129.7967<br />
(66,56) 6371 90 131.2487 5.5113 · 10 −9 131.2487<br />
Table 8.1.1: Displacement table for <strong>the</strong> homogeneous <strong>Earth</strong> model. u, v, u are <strong>the</strong> radial,<br />
lateral and total displacement respectively. The displacements are given in millimeters for<br />
different locations within <strong>the</strong> <strong>Earth</strong>, displaced as grid points (i, j) and <strong>the</strong> corresponding radius<br />
in meter and colatitude in degrees.<br />
8.2 <strong>Deformation</strong> in <strong>the</strong> Layered <strong>Earth</strong> Model<br />
This section describes <strong>the</strong> resulting tidal deformation <strong>of</strong> <strong>the</strong> layered <strong>Earth</strong> model. In<br />
Figure 8.2.1, <strong>the</strong> radial, tangential and total displacements are displayed for <strong>the</strong> layered<br />
model with boundary factor. The blue arcs represent <strong>the</strong> boundaries between <strong>the</strong> four<br />
layers; inner core, outer core, mantle and crust. In Table 8.2.1 a representation <strong>of</strong> <strong>the</strong><br />
model results are shown. The overall picture <strong>of</strong> <strong>the</strong> displacements are <strong>the</strong> same as<br />
described in Section 8.1 for <strong>the</strong> homogeneous model, but <strong>the</strong> size <strong>of</strong> <strong>the</strong> displacements<br />
are quite different.<br />
For <strong>the</strong> layered <strong>Earth</strong> model, <strong>the</strong> radial displacement, Figure 8.2.1(a), has a maximum<br />
expansion at <strong>the</strong> Equator with a displacement <strong>of</strong> u = 74 mm, and a maximum<br />
compressional displacement <strong>of</strong> u = −26 mm at <strong>the</strong> North Pole. The displacements are<br />
changing signs, or going from a compression to an expansion, around a colatitude <strong>of</strong><br />
θ = 28 ◦ . Moving in a line from <strong>the</strong> Equator towards <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, <strong>the</strong> displacement<br />
decays, and at a radius <strong>of</strong> 4313 km, which is in <strong>the</strong> mantle, <strong>the</strong> displacement<br />
become u = 68 mm, when moving fur<strong>the</strong>r towards <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, to a radius<br />
<strong>of</strong> 2353 km (<strong>the</strong> o<strong>the</strong>r core), <strong>the</strong> displacement is u = 44 mm, for finally having a radial
<strong>Deformation</strong> in <strong>the</strong> Layered <strong>Earth</strong> Model 71<br />
displacement value <strong>of</strong> u = 9 mm, at 392 km (<strong>the</strong> inner core) from <strong>the</strong> center.<br />
The tangential displacement field, Figure 8.2.1(b), has a maximum traction at θ =<br />
45 ◦ with a displacement <strong>of</strong> about v = 45 mm. In <strong>the</strong> mantle, at a radius <strong>of</strong> 4313, km<br />
<strong>the</strong> tangential displacement is v = 33 mm, in <strong>the</strong> outer core (2353 km) <strong>the</strong> value is<br />
v = 2 mm and in <strong>the</strong> inner core ( 393 km) <strong>the</strong> displacement is v = 0.0011 mm, all<br />
<strong>the</strong>se values are for θ = 45 ◦ . The total displacement field, 8.2.1(c) has a maximum at<br />
θ = 74 ◦ , with a value <strong>of</strong> u = 77 mm.<br />
Grid points u (km) θ ( ◦ ) u (mm) v (mm) u (mm)<br />
(1,1) 0 0 -3.1101 · 10 −14 −3.02321 · 10 −14 4.3373 · 10 −14<br />
(5,1) 39 0 8.3851 −3.6523 · 10 −15 8.3851<br />
(5,28) 393 45 8.4905 0.0113 8.4905<br />
(5,55) 393 90 8.5009 −6.3158 · 10 −14 8.5009<br />
(25,1) 2353 0 17.5184 −7.1280 · 10 −12 17.5184<br />
(25,10) 2353 15 27.2964 0.6479 27.3041<br />
(25,28) 2353 45 41.3682 1.6261 41.4002<br />
(25,45) 2353 72 43.5761 2.4750 43.6463<br />
(25,56) 2353 90 43.7002 8.1170 · 10 −12 43.7002<br />
(45,1) 4313 0 -10.0585 −3.4946 · 10 −11 10.0585<br />
(45,10) 4313 15 3.8140 17.4486 17.8606<br />
(45,25) 4313 39 42.3912 32.5096 53.4218<br />
(45,28) 4313 45 48.9089 31.7561 58.3141<br />
(45,45) 4313 72 66.8801 10.053 67.6315<br />
(45,56) 4313 90 67.8094 2.6334 · 10 −14 67.8094<br />
(66,1) 6371 0 -26.2925 −3.5447 · 10 −9 26.2925<br />
(66,10) 6371 15 -21.1048 13.0612 24.8195<br />
(66,20) 6371 31 3.4461 33.7783 33.9537<br />
(66,28) 6371 45 28.2988 45.3977 53.4955<br />
(66,40) 6371 72 68.8496 33.9149 76.7495<br />
(66,50) 6371 80 73.2741 20.1883 76.0044<br />
(66,56) 6371 90 74.3154 −7.2879 · 10 −11 74.3154<br />
Table 8.2.1: Displacement table for <strong>the</strong> layered <strong>Earth</strong> model. u, v, u are <strong>the</strong> radial, lateral<br />
and total displacement respectively. The displacements are given in millimeters for different<br />
locations within <strong>the</strong> <strong>Earth</strong>, displaced as grid points (i, j) and <strong>the</strong> corresponding radius in meter<br />
and colatitude in degrees.
72 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
(a) Radial displacement<br />
(b) Tangential displacement<br />
(c) Total displacement<br />
Figure 8.2.1: The displacement fields (in meter) for <strong>the</strong> layered <strong>Earth</strong> model with a surface<br />
boundary factor <strong>of</strong> f fBC = 10 −5 and Tikhonov regularization with ɛ = 2.8137 · 10 −8 . (a)<br />
shows <strong>the</strong> radial displacement, (b) shows <strong>the</strong> tangential displacement and (c) shows <strong>the</strong> total<br />
displacement. The displacements are given for a quarter sphere, where <strong>the</strong> bottom left corner<br />
is <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, top left corner is <strong>the</strong> North Pole, and <strong>the</strong> bottom right corner is <strong>the</strong><br />
Equator.
Analytic Solution 73<br />
8.3 Analytic Solution<br />
In Section 3.4 <strong>the</strong> displacements were given by a function <strong>of</strong> <strong>the</strong> tidal potential and <strong>the</strong><br />
Love numbers. This analytic solution for <strong>the</strong> surface displacement was given by <strong>the</strong><br />
Equations (3.4.1) and (3.4.2), and <strong>the</strong> results <strong>of</strong> <strong>the</strong> calculated values are displayed in<br />
Table 8.3.1.<br />
The analytic solution resembles <strong>the</strong> general picture <strong>of</strong> <strong>the</strong> displacement fields from<br />
<strong>the</strong> following sections. The radial displacement field is expansive and compressive,<br />
and <strong>the</strong> tangential field is tractive. The maximum radial expansion is about a factor<br />
<strong>of</strong> 2.5 larger than <strong>the</strong> maximum tractive tangential displacement. The magnitude <strong>of</strong><br />
<strong>the</strong> displacement components on <strong>the</strong> o<strong>the</strong>r hand differentiate from <strong>the</strong> analytic and numerical<br />
models. The maximum radial displacement is calculated to u = 214 mm and<br />
<strong>the</strong> maximum compression is u = −107 mm. The change from compressive to expansive<br />
displacement happens at θ = 35 ◦ . The tangential displacement has a maximum<br />
traction <strong>of</strong> v = 86 mm at θ = 45 ◦ .<br />
Grid points r(km) θ( ◦ ) u (mm) v (mm)<br />
(66,1) 6371 0 -107.0426 0.0000<br />
(66,10) 6371 15 -86.2891 42.1091<br />
(66,20) 6371 31 -21.4086 75.7376<br />
(66,28) 6371 45 48.9362 85.5599<br />
(66,40) 6371 72 183.32025 50.3345<br />
(66,50) 6371 80 204.7475 28.8777<br />
(66,56) 6371 90 214.0852 0.0000<br />
Table 8.3.1: Values <strong>of</strong> <strong>the</strong> analytic solution given by <strong>the</strong> Equations (3.4.1) and (3.4.2).<br />
8.3.1 Comparison Between Results<br />
For <strong>the</strong> homogeneous <strong>Earth</strong> model <strong>the</strong> radial displacement field is about a factor 1.6<br />
times less compressive than <strong>the</strong> analytic model, and about 23 times less expansive than<br />
<strong>the</strong> analytic model. The tangential displacement is between 1.6 to to 3.1 times smaller<br />
than <strong>the</strong> analytic results, with a smaller difference at <strong>the</strong> Equator.<br />
The displacement field for layered <strong>Earth</strong> model, does also show a smaller deformation<br />
compared to <strong>the</strong> analytic results. For <strong>the</strong> radial displacement field it is about<br />
2.6 to 4.0 less expansive than <strong>the</strong> analytic solution, and for <strong>the</strong> tangential displacement<br />
field it is about a factor 1.4-3.2 too small.<br />
The surface displacements for <strong>the</strong> three results are depicted in Figure 8.3.1, where<br />
Figure 8.3.1(a) are showing <strong>the</strong> radial displacement fields and 8.3.1(b) are showing <strong>the</strong><br />
tangential displacement fields. The displacements are given as a function <strong>of</strong> colatitude,<br />
where θ = 0 ◦ is at <strong>the</strong> Pole and θ = 90 ◦ is at <strong>the</strong> Equator. The figure clearly shows <strong>the</strong><br />
large difference in <strong>the</strong> results, especially in <strong>the</strong> radial displacement fails.
74 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
(a)<br />
(b)<br />
Figure 8.3.1: Comparing <strong>the</strong> radial and tangential displacements at <strong>the</strong> surface as a function<br />
<strong>of</strong> <strong>the</strong> colatitude in degrees.
Chapter 9<br />
Discussion<br />
In this chapter <strong>the</strong> finite difference discretization, <strong>the</strong> developed models and <strong>the</strong> final<br />
results are discussed. The discussion is divided into four parts: First <strong>the</strong> results<br />
and errors are discussed, second <strong>the</strong> reliability <strong>of</strong> <strong>the</strong> model, third some applications<br />
<strong>of</strong> tidal deformation are outlined and fourth suggestions to future developments and<br />
improvements <strong>of</strong> <strong>the</strong> numerical modeling are discussed.<br />
9.1 Results and Errors<br />
In <strong>the</strong> Chapters 4, 6 and 7 an <strong>Earth</strong> model is developed with <strong>the</strong> numerical methods <strong>of</strong><br />
finite difference equations and treated with <strong>the</strong> Tikhonov regularization to stabilize <strong>the</strong><br />
solution. In <strong>the</strong> Chapter 8 <strong>the</strong> final results are given.<br />
9.1.1 The Displacement Fields<br />
Looking at <strong>the</strong> overall picture <strong>of</strong> <strong>the</strong> radial and tangential deformation fields, <strong>the</strong> numerical<br />
results <strong>of</strong> <strong>the</strong> displacement models are at <strong>the</strong> surface consistent with <strong>the</strong> normal<br />
physical interpretation. The radial displacement field expands with a maximum at <strong>the</strong><br />
Equator and contracts with a maximum at <strong>the</strong> Pole. The tangential displacement field<br />
shows only tractive displacements, with a maximum at θ = 45 ◦ , and <strong>the</strong>re is no displacement<br />
at <strong>the</strong> polar and <strong>the</strong> equatorial axes.<br />
Compared to <strong>the</strong> analytic results, <strong>the</strong> numerical models did not accomplish to generate<br />
sufficiently large displacements. The analytic result was dependent <strong>of</strong> <strong>the</strong> averaged<br />
Love numbers, but <strong>the</strong> magnitude <strong>of</strong> <strong>the</strong> analytic result is verified by, among<br />
o<strong>the</strong>rs Melchior [1978] and Xing et al. [2007]. Xing et al. [2007] have made a 3dimensional<br />
finite element analysis and found a total displacement <strong>of</strong> 260 mm at <strong>the</strong><br />
side closest to <strong>the</strong> Moon. The <strong>Earth</strong> model was build much like <strong>the</strong> layered model in<br />
this <strong>the</strong>sis, with four homogeneous layers, but <strong>the</strong> outer core was not assumed to be<br />
completely fluid and <strong>the</strong> outer layer included both <strong>the</strong> crust and <strong>the</strong> transition zone.<br />
The most crucial result is <strong>the</strong> lack <strong>of</strong> compression in <strong>the</strong> homogeneous <strong>Earth</strong><br />
model, it was 23 times too small! In general <strong>the</strong> compressions are fare too small in <strong>the</strong><br />
numerical results compared to <strong>the</strong> analytic result, and extending to an insufficient area.
76 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
The compression <strong>of</strong> <strong>the</strong> radial displacement would have been expected to expand from<br />
<strong>the</strong> North Pole down to <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and extending to θ = 38 ◦ as <strong>the</strong> analytic<br />
result showed. The tangential field should, at <strong>the</strong> maximum displacement, have been<br />
about two and a half times <strong>the</strong> strength <strong>of</strong> <strong>the</strong> maximum radial field. The numerical<br />
result give a three times smaller tangential displacement field for <strong>the</strong> homogeneous<br />
<strong>Earth</strong> model and one and a half smaller in <strong>the</strong> layered <strong>Earth</strong> model. Again <strong>the</strong> layered<br />
model show better results though <strong>the</strong> homogeneous model has a larger displacement<br />
It is not surprising that <strong>the</strong> homogeneous <strong>Earth</strong> model is more deformed at <strong>the</strong> crustal<br />
layer than <strong>the</strong> layered <strong>Earth</strong> model, because <strong>of</strong> <strong>the</strong> smaller shear modulus. Nei<strong>the</strong>r<br />
is it surprising that <strong>the</strong> layered model shows better results compared to <strong>the</strong> analytic<br />
solution, it is in fact more a reassemble to <strong>the</strong> real <strong>Earth</strong>.<br />
The total displacement field fails to give a conceiving picture <strong>of</strong> <strong>the</strong> reel <strong>Earth</strong><br />
picture, due to <strong>the</strong> small compressions in <strong>the</strong> radial displacement fields. The total<br />
displacement should have had a maximum displacement at <strong>the</strong> Equator.<br />
From <strong>the</strong>se results with less fluctuations in <strong>the</strong> values from <strong>the</strong> Pole to <strong>the</strong> Equator<br />
and <strong>the</strong> smaller magnitude <strong>of</strong> <strong>the</strong> displacements it could seem to reflect a too large<br />
damping <strong>of</strong> <strong>the</strong> system due to <strong>the</strong> regularization, but it was showed in Section 7.3 that<br />
<strong>the</strong> regularization parameter did not have that great <strong>of</strong> a weight on <strong>the</strong> system to be <strong>the</strong><br />
reason for <strong>the</strong> bad scaled results.<br />
When regarding <strong>the</strong> inner structures <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, it is uncertain how large <strong>the</strong> real<br />
displacement fields are, except that <strong>the</strong>y becomes smaller moving down <strong>the</strong> <strong>Earth</strong> until<br />
reaching <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, where <strong>the</strong>re is no displacement at all. This is in fact<br />
due to <strong>the</strong> symmetry <strong>of</strong> <strong>the</strong> spherical <strong>Earth</strong>. This is indeed accomplished in <strong>the</strong> two<br />
models.<br />
The surface boundaries were tested (Figure 7.4.4) by applying five attacking point<br />
sources, and <strong>the</strong>y seemed to be working. The model behaved as if a head <strong>of</strong> a pin was<br />
affixed on a piece <strong>of</strong> leader and somebody pulled <strong>the</strong> pin from <strong>the</strong> opposite side. The<br />
leader would be displaced in <strong>the</strong> direction <strong>of</strong> <strong>the</strong> pull, depending on <strong>the</strong> force <strong>of</strong> <strong>the</strong><br />
pulling. The surrounding areas would also be displaced, but not in <strong>the</strong> same degree.<br />
The radial contour lines are perpendicular to <strong>the</strong> polar and <strong>the</strong> equatorial axes, due<br />
to <strong>the</strong> constrains <strong>of</strong> symmetry. The contours are not perpendicular to <strong>the</strong> surface <strong>of</strong> <strong>the</strong><br />
<strong>Earth</strong>, this is due to <strong>the</strong> large contribution <strong>of</strong> <strong>the</strong> partial derivatives with respect to θ,<br />
from Equations (3.3.7) and (3.3.8). This can be shown by <strong>the</strong> simple calculation <strong>of</strong> <strong>the</strong><br />
grid point (n, 40), with <strong>the</strong> discretization given by <strong>the</strong> Equations (4.4.1) and (4.4.2),<br />
where n is <strong>the</strong> last grid point in <strong>the</strong> radial direction, i.e. <strong>the</strong> surface n = a, we get<br />
1<br />
a<br />
∂u<br />
∂r = ua,40 − ua−1,40<br />
hr<br />
∂u<br />
∂θ<br />
9.1.2 Regularization<br />
1 ua,39 − ua,41<br />
=<br />
a 2hθ<br />
= −7.50 · 10 −09<br />
= −3.06 · 10 −09 .<br />
It is essential to investigate how good <strong>the</strong> regularized solution xreg is compared to <strong>the</strong><br />
result <strong>of</strong> <strong>the</strong> true solution. The bias put on <strong>the</strong> system by <strong>the</strong> Tikhonov regularization,<br />
means that <strong>the</strong> solution would never be <strong>the</strong> true solution for a non zero regularization<br />
parameter, ε, even with noise free data, [Aster et al., 2005]. It is not possible to
Results and Errors 77<br />
compare with <strong>the</strong> true solution <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s displacement field, because such solution<br />
does not exist. The knowledge <strong>of</strong> <strong>the</strong> inner structures <strong>of</strong> <strong>the</strong> <strong>Earth</strong> are only based on<br />
<strong>the</strong>oretical assumptions, though it is possible with seismology and airborne satellites<br />
to give a pretty good picture <strong>of</strong> <strong>of</strong> <strong>the</strong> displacement field within <strong>the</strong> <strong>Earth</strong>.<br />
It is possible to examine <strong>the</strong> stability and errors <strong>of</strong> <strong>the</strong> models. In Table 9.1.1, <strong>the</strong><br />
condition numbers and regularization parameters are summarized with and without <strong>the</strong><br />
boundary factor.<br />
When <strong>the</strong> boundary condition factor is applied, it approves <strong>the</strong> solution, especially<br />
<strong>the</strong> stability <strong>of</strong> <strong>the</strong> system i.e. <strong>the</strong> condition number, has been enhanced by this step.<br />
The stability has been approved, but it would be false to call <strong>the</strong> system stable, <strong>the</strong><br />
condition number is still much lager than one. The difference on ε shows that <strong>the</strong><br />
layered <strong>Earth</strong> model is less regularized compared to <strong>the</strong> homogeneous model, but <strong>the</strong><br />
regularization parameter are relative high, i.e. a large damping is embedded on <strong>the</strong><br />
system.<br />
The boundary factor did not effect <strong>the</strong> results to give an observable change in <strong>the</strong><br />
displacement fields, but changed <strong>the</strong> condition number remarkable. Because <strong>of</strong> <strong>the</strong><br />
smaller number <strong>of</strong> grid points at <strong>the</strong> surface compared to inner points, a streng<strong>the</strong>ning,<br />
i.e. a greater weight <strong>of</strong> <strong>the</strong> surface boundary conditions were expected, but this was<br />
not <strong>the</strong> case.<br />
Cond(A) ε Cond(A) f fBC ε f fBC<br />
Homogeneous 2.1635 · 10 23 4.4802 · 10 −9 2.5534 · 10 18 2.2531 · 10 −9<br />
Layered 1.7301 · 10 23 2.8137 · 10 −8 1.7363 · 10 18 1.10157 · 10 −8<br />
Table 9.1.1: Comparison <strong>of</strong> <strong>the</strong> condition number and <strong>the</strong> regularization parameters for <strong>the</strong><br />
homogeneous <strong>Earth</strong> model and <strong>the</strong> layered <strong>Earth</strong> model. The values are given both with and<br />
without <strong>the</strong> surface boundary factor applied.<br />
The error <strong>of</strong> <strong>the</strong> system is calculated with <strong>the</strong> residual norm divided by <strong>the</strong> data norm<br />
F − AX<br />
F − AX<br />
and <strong>the</strong> solution norm . The results are shown in Table 9.1.2,<br />
F<br />
X<br />
For <strong>the</strong> residual norm it does only shows a difference in <strong>the</strong> fifth decimal. In general<br />
<strong>the</strong> layered model is more accurate than <strong>the</strong> homogeneous model. The residual norm<br />
and data norm ratio, show a 10% error for <strong>the</strong> layered model and a 28% error for <strong>the</strong><br />
homogeneous model. These values are still too large for being acceptable.<br />
Homogeneous<br />
F − AX<br />
8.6236 · 10<br />
F − AX<br />
X<br />
F − AX<br />
F<br />
−6 1.90152 · 10−6 28.1462%<br />
Layered 8.6234 · 10−6 5.4212 · 10−6 10.2352%<br />
Table 9.1.2: Errors <strong>of</strong> <strong>the</strong> homogeneous and layered <strong>Earth</strong> models given by <strong>the</strong> residual norm,<br />
<strong>the</strong> residual norm over <strong>the</strong> solution norm and <strong>the</strong> residual norm over <strong>the</strong> data norm.
78 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
9.1.3 The Discretization<br />
The <strong>Earth</strong> was discretized into 66 × 56 grid points for both <strong>the</strong> radial and latitudinal<br />
systems, which gave mesh sizes <strong>of</strong> 96.53 km × 1.6043 ◦ , spanning a part <strong>of</strong> a circle<br />
sector. Even after <strong>the</strong> important values were packed in a temporary directory, and <strong>the</strong><br />
unimportant values were deleted on <strong>the</strong> go, this were <strong>the</strong> finest discretization, possible<br />
by Matlab due to limited memory. Closer grid would have been preferable, as <strong>the</strong><br />
regularization parameter, ε, gets smaller as <strong>the</strong> number <strong>of</strong> grid points increases, as<br />
described in Section 7.4.3, Table 7.4.1. The computational limit was due to <strong>the</strong> use<br />
<strong>of</strong> <strong>the</strong> Matlab function svd, <strong>the</strong> singular value decomposition. The function stores<br />
<strong>the</strong> unitary matrices Up and Vp, containing <strong>the</strong> left and right singular value vectors,<br />
and Matlab needs a lot <strong>of</strong> memory for this step. The server used did not have enough<br />
memory to allow an increase <strong>of</strong> <strong>the</strong> grid numbers.<br />
From Figure 7.4.1 it can be seen how <strong>the</strong> compression <strong>of</strong> <strong>the</strong> radial displacement<br />
field is dependent <strong>of</strong> <strong>the</strong> grid sizes. For very small grid <strong>the</strong> displacement field fails<br />
to give any compression at all. It is also clear from <strong>the</strong> table, Table 7.4.1 and Figure<br />
7.4.2 that <strong>the</strong> magnitude <strong>of</strong> <strong>the</strong> displacement grow as <strong>the</strong> grid points increases. It is<br />
hard to say what would have happened with a finer grid, and if <strong>the</strong> magnitude <strong>of</strong> <strong>the</strong><br />
displacements have reached a stable level, as described in Section 4.2. Though it is<br />
hard to believe <strong>the</strong> displacements would have grown sufficient to match <strong>the</strong> analytic<br />
results. The displacement field with 66×56 grid points is smaller than <strong>the</strong> displacement<br />
field with 51 × 51 grid points, in Aster et al. [2005] it is briefly mentioned, that square<br />
systems could give a better stability. Even though this is true, it would not have given<br />
<strong>the</strong> extra needed displacement.<br />
The body force applied by finite difference equations, only attacks at a single point<br />
not a volume element, as in reality. It is possible that a normalization <strong>of</strong> <strong>the</strong> tidal<br />
acceleration by adding a volume element should have been applied. This enhances <strong>the</strong><br />
importance <strong>of</strong> <strong>the</strong> mesh sizes, and it could be causing <strong>the</strong> errors, but as it is seen, <strong>the</strong><br />
mesh sizes do not change <strong>the</strong> displacement to a large degree, so this is probable not<br />
<strong>the</strong> case. Durran [1999] states, that stability <strong>of</strong> pole singularity problems requires very<br />
small step sizes.<br />
9.1.4 Ill-conditioning<br />
Determination <strong>of</strong> <strong>the</strong> systems ill-conditioning by <strong>the</strong> singular value spectra is a cumbersome<br />
process, and <strong>the</strong> literature is very sparse. Never<strong>the</strong>less has <strong>the</strong>re in this study<br />
been a great focus on <strong>the</strong> process. It is <strong>the</strong> belief, that a great deal <strong>of</strong> information can<br />
be extracted from <strong>the</strong>se singular value spectra.<br />
Ill-conditioning <strong>of</strong> this system is not due to measurement errors, but is caused<br />
because <strong>the</strong> system is underdetermined, i.e. <strong>the</strong>re is too many solutions. The question<br />
whe<strong>the</strong>r <strong>the</strong> singular value spectrum for <strong>the</strong> layered <strong>Earth</strong> model shows a rank-deficient<br />
problem, can be discussed. The definition <strong>of</strong> rank-deficient problems and discrete illposed<br />
problems are very weak. Compared to Hansen [1996], also stated in Section 5.1,<br />
<strong>the</strong>re has to be a well-defined gap between <strong>the</strong> large and small values, what precisely<br />
a well defined means, is not clear, nei<strong>the</strong>r is it stated how little a small singular values<br />
is. Rojas [1996] makes a slightly different definition, that is, rank-deficient problems
Results and Errors 79<br />
(a) Discrete ill-posed (b) Rank-deficient<br />
Figure 9.1.1: Example <strong>of</strong> discrete ill-posed and rank-deficient problem by Rojas [1996]. The<br />
singular values are given for (a) a discrete ill-posed problem (b) a rank deficient problem.<br />
have a clear gap between large and small singular values and <strong>the</strong>re is a (usually) small<br />
cluster <strong>of</strong> small singular value. Compared to a plot from Rojas [1996] <strong>of</strong> a discrete<br />
ill-posed problem and rank-deficient problem, see Figure 9.1.1, it still looks more like<br />
a rank-deficient problem than a discrete ill-posed problem. The definition <strong>of</strong> small is<br />
still uncertain, but it is believed to be values close to zero.<br />
Actually in this case it does not make <strong>the</strong> greatest difference what type <strong>of</strong> illconditioning<br />
<strong>the</strong> system had. Different solving methods have been tested, and <strong>the</strong><br />
choice <strong>of</strong> regularization method fell on Tikhonov based on trial an error. As mentioned<br />
in Section 7.4.1, it is usual to solve rank deficient problems by <strong>the</strong> normal or<br />
truncated singular value decompositions, because <strong>the</strong>y are handling <strong>the</strong> rank <strong>of</strong> <strong>the</strong><br />
coefficient matrix. But even though <strong>the</strong> Tikhonov regularization do not involve <strong>the</strong><br />
numerical rank, it can be used with great success. There is <strong>of</strong> course <strong>the</strong> third option<br />
that <strong>the</strong> system is nei<strong>the</strong>r rank deficient nor discrete ill-posed. Hansen [1996] states<br />
such solutions have to be solved as accurate as possible, without regularization, and he<br />
suggest iterative refinement techniques.<br />
Comparing <strong>the</strong> filter factor plot in Figure 7.3.3, and <strong>the</strong> singular value spectra,<br />
Figure 7.2.3 and 7.2.4 shows how <strong>the</strong> singular values much larger than zero have filter<br />
factors close to or exactly zero. From <strong>the</strong> figure it can also be seen how <strong>the</strong> singular<br />
values much smaller <strong>the</strong>n ε, are very close to zero. Though Tikhonov regularization do<br />
not involve numerical rank directly, it clearly puts a larger filter on <strong>the</strong> singular values<br />
larger than <strong>the</strong> rank, (Rank(A) = 7338).<br />
9.1.5 Computer Errors<br />
When dealing with errors in numerical methods, one can not ignore <strong>the</strong> round-<strong>of</strong>f<br />
errors in numerical differentiation, and <strong>of</strong> course <strong>the</strong> machine precision. The errors<br />
in numerical differentiation have already been dealt with in Section 4.2, where <strong>the</strong><br />
truncation errors was introduced. Because <strong>the</strong> computer is only able to use a finite<br />
number <strong>of</strong> digits in its calculations, <strong>the</strong> computational errors actually gets worse, when
80 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
<strong>the</strong> mesh sizes, hr and hθ decrease and it could be a fatal error with values very close<br />
to zero, [Holmes, 2007]. The large mesh sizes in this study, allows us to disregard this<br />
problem. The machines precision varies depending on <strong>the</strong> computer, system and <strong>the</strong><br />
program used, it is a full-study alone.<br />
9.2 The Reliability <strong>of</strong> <strong>the</strong> Model<br />
Throughout <strong>the</strong> <strong>the</strong>sis, <strong>the</strong>re have been given a lot <strong>of</strong> physical assumptions for <strong>the</strong><br />
purpose <strong>of</strong> creating a simple physical system in which ma<strong>the</strong>matics could be applied.<br />
Some assumptions do not play a major role in <strong>the</strong> physical expression <strong>of</strong> <strong>the</strong> <strong>Earth</strong>,<br />
when o<strong>the</strong>rs changes <strong>the</strong> <strong>Earth</strong> remarkable.<br />
9.2.1 Elasticity Assumptions<br />
In order to apply <strong>the</strong> Navier equation <strong>of</strong> motion, in Section 3.3, <strong>the</strong> <strong>Earth</strong> was assumed<br />
to be continues, homogeneous, isotropic, infinite and elastic. The assumptions <strong>of</strong> continuity<br />
and infinity <strong>of</strong> <strong>the</strong> medium greatly simplify <strong>the</strong> ma<strong>the</strong>matic framework, and are<br />
valid in <strong>the</strong> large scale we are looking at. Inhomogeneity is a local change in physical<br />
parameters within a larger area, [Fowler, 1990]. The homogeneity assumption is not<br />
true for <strong>the</strong> <strong>Earth</strong>, but when looking at <strong>the</strong> rough global picture <strong>the</strong> assumption is valid<br />
to some degree, and will not cause any trouble. Studies concerning tides or elasticity<br />
<strong>the</strong>ory in an inhomogeneous <strong>Earth</strong> can be found in Chakravorty [1972] and Fu and<br />
Sun [2007]. The assumption <strong>of</strong> an isotropic media, i.e. <strong>the</strong> stress-strain relationship<br />
is independent <strong>of</strong> <strong>the</strong> orientation <strong>of</strong> <strong>the</strong> material, is true for most <strong>Earth</strong> material, but<br />
<strong>the</strong> upper mantle have layers which are anisotropic and also <strong>the</strong> inner core indicates<br />
anisotropy, [Lay and Wallace, 1995].<br />
The <strong>Earth</strong> is not completely elastic, as already mentioned in Chapter 3, in fact<br />
<strong>the</strong> mantle <strong>of</strong> <strong>the</strong> <strong>Earth</strong> can be seen as viscoelastic. Though considering <strong>the</strong> tides, <strong>the</strong><br />
mantle could be regarded as a solid body, but may react as a fluid depending on <strong>the</strong><br />
timescale applied, [Wang, 1999]. Because <strong>the</strong> outer core is fluid it is deformed more<br />
than <strong>the</strong> mantle by <strong>the</strong> external gravitational potential, [Hinderer et al., 1987].<br />
9.2.2 The Real Tides<br />
In Chapter 2, <strong>the</strong> <strong>Earth</strong> tides was described by a completely solid <strong>Earth</strong>. The real <strong>Earth</strong><br />
has a solid crust with thin layers <strong>of</strong> ocean bounded by continents. The ocean tides<br />
plays a significant role in <strong>the</strong> crustal deformation. The oceans have a large effect and<br />
can be calculated by ocean models, see for example Farrell [1972], which describes<br />
<strong>the</strong> ocean loading by <strong>the</strong> elastic Greens function, or Khan and C.C.Tscherning [2001],<br />
which studied <strong>the</strong> semi-diurnal ocean tide loading using GPS data. In reality <strong>the</strong> solid<br />
<strong>Earth</strong> tides are dominated by <strong>the</strong> compressive and expansive radial component <strong>of</strong> <strong>the</strong><br />
tidal force. The large oceans are dominated by <strong>the</strong> tangential components <strong>of</strong> <strong>the</strong> tidal<br />
forces. The tidal bulges are small compared with <strong>the</strong> radius <strong>of</strong> <strong>Earth</strong>, but it still raises<br />
a huge amount <strong>of</strong> water. When <strong>the</strong> continents are added, <strong>the</strong> ocean reflect on <strong>the</strong><br />
shorelines setting up currents and standing waves. Also <strong>the</strong> coastal topography plays
Application <strong>of</strong> Tides 81<br />
an significant role in intensifying water level. This is only due to <strong>the</strong> tidal bulge,<br />
[Jentzsch, 1997].<br />
The <strong>Earth</strong> is not spherical symmetric and it rotes, this means in reality <strong>the</strong> tidal<br />
force is more complicated than described. The tidal displacement is actually larger at<br />
<strong>the</strong> side closest to <strong>the</strong> Moon, than on <strong>the</strong> side turning away, [de Pater and Lissauer,<br />
2001]. The tidal forces causes a precession <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s rotation axis by applying a<br />
torque attempting to pull <strong>the</strong> equatorial bulge into <strong>the</strong> plane <strong>of</strong> <strong>the</strong> ecliptic. The <strong>Earth</strong>’s<br />
nutation is a result <strong>of</strong> <strong>the</strong> diurnal frequency band. The precession and nutation are true<br />
even for a symmetric sphere, [Broche and Schuh, 1998]. Also <strong>the</strong> tidal deformation<br />
is influenced by <strong>the</strong> rotation <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. On short times scales less than a month,<br />
<strong>the</strong> core is decoupled from <strong>the</strong> mantle rotation, and this causes <strong>the</strong> love number k<br />
to decrease by approximately 8%, [Wang, 1997]. Atmospheric tides, due to diurnal<br />
heating <strong>of</strong> <strong>the</strong> atmosphere induces surface loading, also has to be considered, [Petrov<br />
and Boy, 2003].<br />
9.3 Application <strong>of</strong> Tides<br />
The solid <strong>Earth</strong> tides influence <strong>the</strong> entire universe and <strong>the</strong> tides can tell us about <strong>Earth</strong>’s<br />
interior, earthquake occurrence and estimation <strong>of</strong> rock parameters in bore holes <strong>of</strong> connected<br />
aquifer, [Kümpel, 1997]. The global tidal response is not possible to recover<br />
from tilt and strain tide observations, <strong>the</strong>y can only determine local effects. Seismological<br />
investigation is by far <strong>the</strong> best method to extract information about <strong>the</strong> <strong>Earth</strong>’s<br />
interior, but also studies <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s magnetic field and gravity anomalies can reveal<br />
information <strong>of</strong> <strong>the</strong> interior. The nearly-diurnal free wobble-resonance, caused by a<br />
resonance between <strong>the</strong> mantle and <strong>the</strong> outer core, short described in Section 3.4.1, is<br />
unrecoverable from seismic interpretation but can be recovered from VLBI data and<br />
gravity models from for example superconducting gravimeters, [Zürn, 1997].<br />
The <strong>Earth</strong>’s response to <strong>the</strong> tides have been used to study o<strong>the</strong>r stress induced<br />
phenomenas in <strong>the</strong> <strong>Earth</strong> such as earthquake triggering and volcanic eruptions. The<br />
<strong>Earth</strong>’s response to tectonic stress are much lager than <strong>the</strong> tidal induced stress level,<br />
but <strong>the</strong> stress rates, i.e. stress in a given period, are comparable, [Heaton, 1982; Steacy<br />
et al., 2005]. <strong>Earth</strong>quake triggering due to <strong>the</strong> solid <strong>Earth</strong> tides have among o<strong>the</strong>rs<br />
been studied by Schuster [1897]; Heaton [1982]; Emter [1997]; Tananka et al. [2002,<br />
2004]; Beeler and Lockner [2003]; Cochran et al. [2004]; Métivier et al. [2009].<br />
Ano<strong>the</strong>r approach to <strong>the</strong> deformation <strong>of</strong> <strong>the</strong> <strong>Earth</strong> is by looking at <strong>the</strong> elastic energy.<br />
Getino and Ferrandez [1991]; Barkin and Ferrandiz [2002]; Barkin et al. [2006]<br />
have studied <strong>the</strong> tension <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s elastic mantle as a perturbation <strong>of</strong> a combined<br />
rotational and tidal forces from <strong>the</strong> Moon and <strong>the</strong> Sun. The strain components are <strong>the</strong><br />
partial divided <strong>of</strong> <strong>the</strong> displacement.<br />
Even though tidal deformation have been studied in many years, no comprehensive<br />
studies <strong>of</strong> <strong>the</strong> global displacement field have been deduced. There have been few<br />
investigations <strong>of</strong> tidal deformation throughout <strong>the</strong> whole spherical <strong>Earth</strong>. <strong>Tidal</strong> deformations<br />
is normally carried out for <strong>the</strong> surface <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, and if <strong>the</strong> mantle or core<br />
are investigated it is with <strong>the</strong> emphasis to determine <strong>the</strong> displacement at <strong>the</strong> surface,<br />
see for example Métivier et al. [2007]. This is because <strong>the</strong> known <strong>the</strong>ory about <strong>the</strong> de-
82 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
formation <strong>of</strong> <strong>the</strong> deeper structures may only be deviating little from <strong>the</strong> true <strong>Earth</strong>, due<br />
to <strong>the</strong> small displacements in depths, also <strong>the</strong> crust is most relevant to our lives. Xing<br />
et al. [2007] have studied <strong>the</strong> global deformation, but <strong>the</strong>y are only concerned about<br />
<strong>the</strong> crustal layer, not <strong>the</strong> inner <strong>Earth</strong>. There have been several local investigations <strong>of</strong><br />
<strong>the</strong> mantle and crust and a lot <strong>of</strong> those are concerning <strong>the</strong> studies <strong>of</strong> tidal triggering<br />
<strong>of</strong> earthquakes. Varga and Grafarend [1996] have made a study <strong>of</strong> <strong>the</strong> stress tensor<br />
components in <strong>the</strong> mantle due to perturbations from <strong>the</strong> Moon and <strong>the</strong> Sun, <strong>the</strong> emphasis<br />
were to determine <strong>the</strong> tidal influences <strong>of</strong> mantle viscoelasticity and <strong>the</strong> lateral<br />
heterogeneities for a rotation <strong>Earth</strong>. They find, that <strong>the</strong> mantle viscoelasticity causes<br />
a amplitude decrease and phase delay <strong>of</strong> <strong>the</strong> <strong>Earth</strong>’s response function. Takeuchi and<br />
Saito [1962] gives <strong>the</strong> lateral and radial displacements in <strong>the</strong> Mantle and a homogeneous<br />
core for various rigidity values <strong>of</strong> <strong>the</strong> core by spherical harmonics. <strong>Tidal</strong> bulges<br />
influences GPS and is important in <strong>the</strong> satellite calibration and VLBI measurements,<br />
[Torge, 2001].<br />
Tides influence nutation and angular rotation and <strong>the</strong>refore knowledge <strong>of</strong> tides<br />
is important in astronomy to make precise angular measurements, [Melchior, 1978].<br />
The particle accelerator in Cern also take <strong>the</strong> tides into account, for making precise<br />
operations [Takao and Shimada, 2000].<br />
9.4 Model Improvements<br />
In <strong>the</strong> previous discussion, it has been described how studies <strong>of</strong> <strong>the</strong> tidal deformation<br />
<strong>of</strong> <strong>the</strong> <strong>Earth</strong>, can be carried out in much more details, than have been done in this<br />
<strong>the</strong>sis. It have never been <strong>the</strong> intention, with this one year study, to develop a model<br />
that could match those who have been in development for years. In future perspectives<br />
a lot <strong>of</strong> improvements can be carried out, but keeping it in a realistic perspective and<br />
without loosing <strong>the</strong> fundamental approaches, it is chosen only to suggest <strong>the</strong> most apparent<br />
improvements.<br />
If <strong>the</strong> work could have been done all over again, it would have been preferable to<br />
use a programming language such as Fortran instead <strong>of</strong> Matlab. A lot <strong>of</strong> <strong>the</strong> trouble<br />
which has been encountered doing <strong>the</strong> study, would probable have been avoided by<br />
taken <strong>the</strong> trouble <strong>of</strong> writing <strong>the</strong> code in Fortran, because Fortran would speed up <strong>the</strong><br />
calculations and most important <strong>the</strong> memory problem would not have bene as big an<br />
issue. The model is programmed with for loops which are not very efficient within<br />
Matlab, instead <strong>the</strong> arrays should have been written with inner and outer matrices,<br />
[Acklam, 2003].<br />
One <strong>of</strong> <strong>the</strong> most obvious improvements is <strong>the</strong> finite difference discretization. First<br />
<strong>of</strong> all a higher order <strong>of</strong> accuracy in <strong>the</strong> finite difference formula could have been applied.<br />
In this study a second order centered finite difference formula is used. For an<br />
example, <strong>the</strong> third order accurate for ur can easily be written as<br />
ur(r, θ) = 2u(r + hr, θ) + 3u(r, θ) − 6u(r − hr, θ) + u(r − 2hr, θ)<br />
6hr<br />
.
Model Improvements 83<br />
This led to an o<strong>the</strong>r FDE improvement, instead <strong>of</strong> using <strong>the</strong> centered finite difference<br />
a bunch <strong>of</strong> o<strong>the</strong>r methods could be used to discretized <strong>the</strong> Navier equation. Different<br />
approximations have different truncation errors, Li [2004] suggest <strong>the</strong>se simple<br />
methods to be favorable for handling hyperbolic problems: Crank-Nicholson scheme,<br />
Lax-Wendr<strong>of</strong>f method, Leap-frog method, Up-wind scheme, Lax-Freidrichs method<br />
and <strong>the</strong> Beam-Warming method.<br />
The general limitation with finite FDE concerns <strong>the</strong> use <strong>of</strong> grid points, and by this<br />
approach <strong>the</strong> model do not care about what lying in between <strong>the</strong> grid points. Finite<br />
element methods on <strong>the</strong> o<strong>the</strong>r hand is concerned about <strong>the</strong> defined elements instead <strong>of</strong><br />
<strong>the</strong> grid points.<br />
O<strong>the</strong>r improvements to <strong>the</strong> FDE concerning <strong>the</strong> boundary conditions. To avoid <strong>the</strong><br />
pole singularity a staggered grid can be used for r = 0. This was described in Section<br />
4.4.2 and solves <strong>the</strong> singularity problem by avoiding grid points directly at <strong>the</strong> pole.<br />
Also <strong>the</strong> clustering <strong>of</strong> grid points at <strong>the</strong> center can be avoided, [Mosheni and Colonius,<br />
2000; Vasilyev et al., 2004]. At <strong>the</strong> free surface, <strong>the</strong> traction free conditions could have<br />
been handled with <strong>the</strong> ghost point method, where fictive grid points are added to keep<br />
<strong>the</strong> second order accuracy at <strong>the</strong> boundaries, [Pérez-Ruiz et al., 2007]. Finally <strong>the</strong> grid<br />
sizes have to be mentioned in this context, but it have been discussed earlier.<br />
Next step in improving <strong>the</strong> model would be to incorporate some assumptions handling<br />
<strong>the</strong> fluid core <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and going from two dimensional problem to a three<br />
dimensional problem with <strong>the</strong> <strong>Earth</strong> rotating about its own axis.
Chapter 10<br />
Conclusion<br />
The study <strong>of</strong> tides is an important field in geophysics, <strong>the</strong>re are many factors to deal<br />
with and many approaches to <strong>the</strong> problem. The behavior <strong>of</strong> <strong>the</strong> solid <strong>Earth</strong> tides is well<br />
understood, because it is determined by <strong>the</strong> known orbital motions <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and <strong>the</strong><br />
Moon, and also by <strong>the</strong> Sun and o<strong>the</strong>r external objects. Combining this knowledge<br />
and with <strong>the</strong> elasticity <strong>of</strong> <strong>Earth</strong>, given by <strong>the</strong> Love numbers, <strong>the</strong> deformation due to <strong>the</strong><br />
solid <strong>Earth</strong> tides can be found. Most studies are calculating an analytic solution, where<br />
<strong>the</strong> leading approach is to expand <strong>the</strong> tidal potential into spherical harmonics and using<br />
Love numbers to determine <strong>the</strong> rigidity <strong>of</strong> <strong>the</strong> inner structures. The <strong>Earth</strong>’s responds to<br />
<strong>the</strong> tidal force, in a particular point, can be deduced by Green’s function. The tides are<br />
<strong>of</strong>ten studied in terms <strong>of</strong> <strong>the</strong> different tidal periods, and <strong>the</strong> <strong>Earth</strong>’s deformation due to<br />
<strong>the</strong> tidal force is <strong>of</strong>ten limited to <strong>the</strong> surface or <strong>the</strong> crust <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. Also a coupling<br />
<strong>of</strong> advanced <strong>Earth</strong>, ocean and atmospheric models are widely used. The <strong>Earth</strong> is a<br />
complex body, and a good knowledge <strong>of</strong> <strong>the</strong> driving force is not enough to observe <strong>the</strong><br />
<strong>Earth</strong>’s deformation, because <strong>of</strong> <strong>the</strong> many o<strong>the</strong>r factors which also deform <strong>the</strong> <strong>Earth</strong>.<br />
The main purpose <strong>of</strong> this <strong>the</strong>sis, has been to determine <strong>the</strong> tidal deformation <strong>of</strong> <strong>the</strong><br />
solid <strong>Earth</strong> by using a steady state version <strong>of</strong> <strong>the</strong> Navier equation <strong>of</strong> motion, verified<br />
by <strong>the</strong> slow deformation processes. The approach was meant to be as simple as possible<br />
and <strong>the</strong> equations were discretized with <strong>the</strong> numerical method <strong>of</strong> finite difference<br />
equations.<br />
There has been developed two <strong>Earth</strong> models. The first model, called <strong>the</strong> homogeneous<br />
<strong>Earth</strong> model, was assumed to be a pure Poisson solid, where <strong>the</strong> seismic<br />
velocities and <strong>the</strong> elastic parameters were constant throughout <strong>the</strong> <strong>Earth</strong>. In <strong>the</strong> second<br />
model, called <strong>the</strong> layered model, <strong>the</strong> homogeneous model was expanded from being<br />
a complete Poisson solid to including four layers representing an inner core, an outer<br />
core,a mantle and a crust. The properties <strong>of</strong> each layer were taken from <strong>the</strong> isotropic<br />
Preliminary Reference Model and averaged over each layer. This resulted in a more<br />
realistic <strong>Earth</strong> model with a fluid outer core.<br />
A dead end was reached, in <strong>the</strong> numerical process <strong>of</strong> solving <strong>the</strong> linear system <strong>of</strong><br />
equations, when <strong>the</strong> system, to a first approach, proved to be unsolvable. The system<br />
was rank deficient, and both over- and underdetermined. Especially because <strong>of</strong> too
86 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
many best fitted solutions (underdetermined) <strong>the</strong> best physical result was hard to find.<br />
To overcome <strong>the</strong> obstacle a Tikhonov regularization was applied.<br />
The displacement field induced by <strong>the</strong> solid <strong>Earth</strong> tides, were presented for <strong>the</strong><br />
radial, <strong>the</strong> tangential and <strong>the</strong> total displacement in a two dimensional quarter <strong>Earth</strong>.<br />
In <strong>the</strong> view <strong>of</strong> physical interpretations <strong>the</strong> displacement fields have gone from not<br />
showing anything meaningful at all, to (after regularization) give a solution which in<br />
broad terms provided a physical meaningful result.<br />
The magnitude <strong>of</strong> <strong>the</strong> displacement fields were too small. For <strong>the</strong> homogeneous<br />
model, <strong>the</strong> maximum expansive displacement was calculated to 131 mm, <strong>the</strong> maximum<br />
compressive displacement was −5 mm. For <strong>the</strong> layered model <strong>the</strong>se displacements<br />
were 74 mm and −26 mm. These results compared to <strong>the</strong> analytic result were<br />
2.6 to 4 times too small for <strong>the</strong> layered model and for <strong>the</strong> homogeneous model <strong>the</strong><br />
expansive radial displacement field was 1.6 times too small, and <strong>the</strong> compressive part<br />
<strong>of</strong> <strong>the</strong> radial displacement field was 23 times too small. The displacement became<br />
smaller moving fur<strong>the</strong>r towards <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>, which <strong>of</strong> course is due to <strong>the</strong><br />
tidal force losing strength until it cancels out in <strong>the</strong> center <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. The tangential<br />
displacement fields were also determined. For <strong>the</strong> layered <strong>Earth</strong> model <strong>the</strong> maximum<br />
traction was 45 mm, and <strong>the</strong> results for <strong>the</strong> homogeneous model gave a maximum traction<br />
<strong>of</strong> 41 mm. For <strong>the</strong> layered <strong>Earth</strong> model this deviated with a traction with a factor<br />
<strong>of</strong> 1.4 to 3.2 times too small. For <strong>the</strong> homogeneous model <strong>the</strong> tractions where a factor<br />
1.6 to 3.1 times too small compared to <strong>the</strong> analytic results.<br />
Except for <strong>the</strong> radial displacement field that did not provide enough compression,<br />
probably due to <strong>the</strong> rough grid, <strong>the</strong> modeled displacement fields gave a physical interpretable<br />
pictures <strong>of</strong> <strong>the</strong> tidal deformation <strong>of</strong> <strong>the</strong> solid <strong>Earth</strong>.
Bibliography<br />
Acklam, P. J. (2003). M ATLAB array manipulation tips and tricks. http://home.<br />
online.no/~pjacklam.<br />
Agnew, D. (1986). Strainmeters and tiltmeters. Reviews <strong>of</strong> Geophysics, 24:579–624.<br />
Anandakrishnan, S. and Alley, R. B. (1997). <strong>Tidal</strong> forcing <strong>of</strong> basal seismicity <strong>of</strong> ice<br />
stream C, West Antarcitica. In American Geophysical Union.<br />
Aster, R., Borchers, B., and Thurber, C. (2005). Parameter Estimation and Inverse<br />
Problems. Academic Press.<br />
Barkin, Y. V. and Ferrandiz, J. M. (2002). <strong>Tidal</strong> elastic energy in <strong>the</strong> planetary systems<br />
and its dynamic role. Astronomical ans Astrophysical Transactions, 23(4):369–384.<br />
Barkin, Y. V., Ferrandiz, J. M., Ferrandez, M. G., and Navarro, J. F. (2006). The elastic<br />
energy <strong>of</strong> rotational and lunisolar tides and <strong>the</strong>ir role in <strong>the</strong> earth seismic activity.<br />
Astronomical and Astrophysical Transactions, 26:163–198.<br />
Beeler, N. M. and Lockner, D. A. (2003). Why earthquakes correlate weekly with <strong>the</strong><br />
solid earth tides: Effects <strong>of</strong> periodic stress on <strong>the</strong> rate and probability <strong>of</strong> earthquake<br />
occurrence. Journal <strong>of</strong> Geophysical Reseach, 108:ESE 8–1 – ESE 8–17.<br />
Boudin, F., Bernard, P., Longuevergne, L., Florsch, N., Larmat, C., Courteille, C.,<br />
Blum, P., Vincent, T., and Kammentaler, M. (2008). A silica long base tiltmeter<br />
with high stability and resolution. Review <strong>of</strong> Scientific Instruments, 79:034502.<br />
Broche, P. and Schuh, H. (1998). Tides and <strong>Earth</strong> rotation. Surveys in Geophysics,<br />
19:417–430.<br />
Chakravorty, J. G. (1972). <strong>Deformation</strong> and stresses in a non-homogeneous earthmodel<br />
with a rigid core. Pure and Applied Geophysics PAGEOPH, 95:59–66.<br />
Cochran, E. S., Vidale, J. E., and Tananka, S. (2004). <strong>Earth</strong> tides can trigger shallow<br />
thrust fault earthquakes. Science, 306:1164–1166.<br />
de Pater, I. and Lissauer, J. J. (2001). Planetary Sciences. Cambridge.
88 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
Durran, D. (1999). Numerical methods for wave equations in geophysical fluid dynamics.<br />
Springer.<br />
Dziewonski, A. and Anderson, D. (1981). Preliminary reference <strong>Earth</strong> model (PREM).<br />
Phys. <strong>Earth</strong> Planet. Inter, 25:297–356.<br />
Emter, D. (1997). <strong>Tidal</strong> Triggering <strong>of</strong> <strong>Earth</strong>quakes and Volcanic Events. In Lecture<br />
Notes in <strong>Earth</strong> Sciences, <strong>Tidal</strong> Phenomena. Springer.<br />
Farrell, W. (1972). <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Earth</strong> by surface loads. Reviews <strong>of</strong> Geophysics<br />
<strong>of</strong> Space Physics, 10:767–797.<br />
Fowler, C. (1990). The <strong>Solid</strong> <strong>Earth</strong>. Cambridge Univ. Press, New York.<br />
Fu, G. and Sun, W. (2007). Effects <strong>of</strong> lateral inhomogeneity in a spherical <strong>Earth</strong> on<br />
gravity <strong>Earth</strong> tides. Jounal <strong>of</strong> Geophysical Research, 112:B06409.<br />
Getino, J. and Ferrandez, M. G. (1991). A Hamiltonian Theory for an Elastic <strong>Earth</strong>:<br />
Elastic Energy <strong>of</strong> <strong>Deformation</strong>. Celestial Mechanics and Dynamical Astronomy.,<br />
51:17–43.<br />
Hansen, P. C. (1996). Rank-Deficient and Discrete Ill-Posed Problems. Polyteknisk<br />
Forlag.<br />
Hansen, P. C. (2008a). regtools - a matlab package. http://www.mathworks.com/<br />
matlabcentral/fileexchange/52. Matlab Central.<br />
Hansen, P. C. (2008b). Regularization Tool. A Matlab Package for Analysis and<br />
Solution <strong>of</strong> Discrete Ill-Posed Problems, Informatics and Ma<strong>the</strong>matical Modelling,<br />
Technical University <strong>of</strong> Copenhagen, DK-2800 Lyngby, Denmark.<br />
Heaton, T. H. (1975). <strong>Tidal</strong> triggering <strong>of</strong> earthquakes. Geophysical Journal R. astr.<br />
Soc., 43:307–326.<br />
Heaton, T. H. (1982). <strong>Tidal</strong> triggering <strong>of</strong> earthquakes. Bull. Seismol. Soc. Am.,<br />
72:2181–2200.<br />
Hinderer, J., Legros, H., and Amalvict, M. (1987). <strong>Tidal</strong> motions within <strong>the</strong> <strong>Earth</strong>’s<br />
fluid core: resonance process and possible variations. Phys. <strong>Earth</strong> Planet. Inter,<br />
49:213–221.<br />
Holmes, M. H. (2007). Introduction to Numerical Methods in Differential Equations.<br />
Springer.<br />
Hughes, W. F. and Gaylord, E. W. (1964). Basic Equations <strong>of</strong> Engineering Science.<br />
Schaum’s Outlinne Series. McGraw-Hill Book Campany.<br />
Hurford, T. A. and Greenberg, R. (2002). Tides on a Compressible Sphere: Sensitivity<br />
<strong>of</strong> <strong>the</strong> H2 Love Number. In Lunar and Planetary Science Conference XXXlll.<br />
Jeffreys, H. and Vicente, R. O. (1957). The <strong>the</strong>ory <strong>of</strong> nutation and <strong>the</strong> variation <strong>of</strong><br />
latitude. Royal Astronomical Society, 117:142–161.
BIBLIOGRAPHY 89<br />
Jentzsch, G. (1997). <strong>Earth</strong> tides and ocean tidal loading. In <strong>Tidal</strong> Phenomena. Lecture<br />
Notes in <strong>Earth</strong> Sciences, volume 66, pages 145–171. Springer.<br />
Khan, S. and C.C.Tscherning (2001). Determination <strong>of</strong> semi-diurnal ocean tide loading<br />
constituents using GPS in Alaska. Gephysical Review Letter, 28(11):2249–2252.<br />
Knudsen, J. and Hjorth, P. (2000). Elements <strong>of</strong> Newtonian Mechanics: Including<br />
Nonlinear Dynamics. Springer.<br />
Kreyszig, E. (1999). Advanced Engineering Ma<strong>the</strong>matics. Wiley.<br />
Kulessa, B., Hubbard, B., Brown, G., and Becker, J. (2003). <strong>Earth</strong> tide forcing <strong>of</strong><br />
glacier drainage. Geophys. Res. Lett, 30(1):1011.<br />
Kümpel, H. (1997). Tides in Water-Saturated Rock. In <strong>Tidal</strong> Phenomena, Lecture<br />
Notes in <strong>Earth</strong> Sciences, pages 277–293. Springer.<br />
Lammlein, D. R. (1977). Lunar Seismicity and Tectonics. Physics and <strong>Earth</strong> and<br />
Planetary Interiors, 14(14):224–273.<br />
Laplace, P.-S. (1832). Méchanique Céleste, volume 4 vols. (Translated by N.<br />
Bowditch), Boston.<br />
Latycheva, K., Mitrovicaa, J. X., Ishiib, M., Chana, N.-H., and Davisc, J. L. (2009).<br />
Body tides on a 3-D elastic earth: Toward a tidal tomography. <strong>Earth</strong> and Planetary<br />
Science Letters, 277:86–90.<br />
Lautrup, B. (2005). Physics <strong>of</strong> Continuous Matter: Exotic and Everyday Phenomena<br />
in <strong>the</strong> Macroscopic World. Institute <strong>of</strong> Physics Publishing.<br />
Lay, T. and Wallace, T. C. (1995). Modern Global Seismology, volume 58 (International<br />
Geophysics). Academic Press.<br />
Li, Z. (2004). MA584: Numerical Solutions <strong>of</strong> FDEs - Finite Difference Methods.<br />
Lecture notes. Lecture notes from <strong>the</strong> course: MA584: Numerical Solution <strong>of</strong><br />
Differential Equations at North Carolina State University: http://www4.ncsu.edu/<br />
~zhilin/.<br />
Longman, I. (1962). A Grenn’s function for determing <strong>the</strong> deformation f <strong>the</strong> <strong>Earth</strong><br />
under suface mass loads: 1. <strong>the</strong>ory. Journal <strong>of</strong> Geophysical Reseach, 67:845–850.<br />
Love, A. (1911). Some problems <strong>of</strong> geodynamics. Dover Publications, New York.<br />
Love, A. (1944). The Ma<strong>the</strong>matical Theory <strong>of</strong> Elasticity. Dover Publications, New<br />
York, 4th edition. First Edition (1892).<br />
Melchior, P. (1978). The tides <strong>of</strong> <strong>the</strong> planet earth. Oxford: Pergamon Press.<br />
Métivier, L., de Viron, O., d, C. P. C., phane Renault, S., Diament, M., and Patau,<br />
G. (2009). Evidence <strong>of</strong> earthquake triggering by <strong>the</strong> solid earth tides. <strong>Earth</strong> and<br />
Planetary Science Letters, 278:370–375.
90 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
Métivier, L., Greff-Lefftz, M., and Diament, M. (2007). Mantle lateral variations and<br />
elastogravitational deformations - II. Possible effests <strong>of</strong> a superplume on body tide.<br />
Geophysical Journal International, 168:897–903.<br />
Mosheni, K. and Colonius, T. (2000). Numerical treatment <strong>of</strong> polar coordinate suingularities.<br />
Journal <strong>of</strong> Computational Physics., 157:787–795.<br />
Munk, W. and Dzieciuch, M. (2001). Millennial Climate Variability: Is <strong>the</strong>re a <strong>Tidal</strong><br />
Connection? Journal <strong>of</strong> Climate, 15:370–385.<br />
Newton, I. (1833). Philosophiae naturalis principia ma<strong>the</strong>matica. (Translation <strong>of</strong><br />
1833), volume 1. [Cambridge Mass.] Harvard University Press. (3. ed. (1726),<br />
with variant readings, assembled and ed. by Alexandre Koyré ed.), (Translation <strong>of</strong><br />
1833).<br />
Parker, R. (1994). Geophysical inverse <strong>the</strong>ory. Princeton University Press.<br />
Pérez-Ruiz, J., Luzón, F., and Garcia-Jerez, A. (2007). Scattering <strong>of</strong> elastic waves in<br />
cracked media using a finite defference method. Stud. Geophys. Geod., 51:59–88.<br />
Petrov, L. and Boy, J. (2003). Study <strong>of</strong> <strong>the</strong> atmospheric pressure loading signal in<br />
VLBI observations. Arxiv preprint physics/0311096.<br />
Rinehart, J. (1972). 18.6-Year <strong>Earth</strong> Tide Regulates Geyser Activity. Science,<br />
177(4046):346–347.<br />
Rojas, M. (1996). Regularization <strong>of</strong> large–scale ill–conditioned least squares problems.<br />
Schuster, W. A. (1897). On lunar and solar periodicities <strong>of</strong> earthquakes. Proc. R. Soc.<br />
London, 61:455–456.<br />
Seeber, G. (2003). Satellite Geodesy. Walter de Gruyter.<br />
Shearer, P. M. (1999). Introduction to Seismology. Cambridge University Press.<br />
Sottili, G., Martino, S., Palladino, D. M., Paciello, A., and Bozzano, F. (2007). Effects<br />
<strong>of</strong> tidal stresses on volcanic activity at Mount Etna, Italy. Geophysical Research<br />
Letters, 34:L01311.<br />
Spiegel, M. R. and Liu, J. (1999). Ma<strong>the</strong>matical Handbook <strong>of</strong> Formulas and Tables.<br />
Schaum’s Outline series. McGraw-Hill, second edition.<br />
Stacey, F. D. and Davis, P. M. (2008). Physics <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. Cambridge, 4 edition.<br />
Steacy, S., Gomberg, J., and Cocco, M. (2005). Introduction to special section: Stress<br />
transfer, earthquake triggering, and time-dependent. Journal o<strong>of</strong> Geophysical Research,<br />
110:B05S01.
BIBLIOGRAPHY 91<br />
Surfaces and Contact Mechanics (2005). Surfaces and contact mechanics, part two,<br />
page 2. (lecture 13). http://frictioncenter.engr.siu.edu/course/file13.<br />
html. Webpage from a graduate-level course at Center for Advanced Friction Studies<br />
at Sou<strong>the</strong>rn Illinois University.<br />
Takao, M. and Shimada, T. (2000). Long Term Variation <strong>of</strong> <strong>the</strong> Circumference <strong>of</strong> The<br />
Spring-8 Storage Ring. In Proceedings <strong>of</strong> EPAC.<br />
Takeuchi, H. and Saito, M. (1962). Statistical <strong>Deformation</strong> and Free Oscillations <strong>of</strong> a<br />
Model <strong>Earth</strong>. Journal <strong>of</strong> Geophysical Reseach, 67:1141–1154.<br />
Tananka, S., Ohtake, M., and Sato, H. (2002). Evidence for tidal triggering <strong>of</strong> earthquakes<br />
as revealed from statistical analysis <strong>of</strong> global data. Journal <strong>of</strong> Geophysical<br />
Reseach, 107(B10):1–1 – 1–10.<br />
Tananka, S., Ohtake, M., and Sato, H. (2004). <strong>Tidal</strong> triggering <strong>of</strong> earthquakes in japan<br />
related to <strong>the</strong> regional tectonic stress. <strong>Earth</strong> Planets Space, 56:511–515.<br />
Torge, W. (2001). Geodesy. Walter de Gruyter, 3rd edition.<br />
Trefe<strong>the</strong>n, L. N. and Bau, D. (1997). Numerical Linear Algebra. SIAM.<br />
Tsantes, E. (1972). Note on <strong>the</strong> Tides. Am. J. Phys., 42:330–333.<br />
Turcotte, D. L. and Schubert, G. (1982). Geodynamics Application <strong>of</strong> Continuum<br />
Physics to Geological Problems. John Wiley and Sons, Inc.<br />
Varga, P. (1974). Dependence <strong>of</strong> <strong>the</strong> Love numbers upon <strong>the</strong> inner structure <strong>of</strong> <strong>the</strong><br />
<strong>Earth</strong> and comparison <strong>of</strong> <strong>the</strong>oretical models with results <strong>of</strong> measurements. Pure and<br />
Applied Geophysics, 112(5):777–785.<br />
Varga, P. and Grafarend, E. (1996). Distribution <strong>of</strong> <strong>the</strong> lunisolar tidal elastic stress<br />
tensor components within <strong>the</strong> <strong>Earth</strong>’s mantle. Physics <strong>of</strong> <strong>the</strong> <strong>Earth</strong> and Planetary<br />
Interiors, 93:285–297.<br />
Vasilyev, O. V., Ogi, T., and Morinishi, Y. (2004). Fully conservative finite difference<br />
scheme in cylindrical coordinates for incompressible flow simulations. Jounal <strong>of</strong><br />
Computational Physics, 197:686–710.<br />
Vogel, C. R. (1996). Non-convergence <strong>of</strong> <strong>the</strong> L-curve regularization parameter selection<br />
method. In Inverse Problems, pages 535–547.<br />
Wang, H. (1999). Surface vertical displacemnets, potential pertubations and gravity<br />
changes <strong>of</strong> a viscoelastic earth model induced by internal point dislocations. Geophysical<br />
Journal International, 137:429–440.<br />
Wang, R. (1997). <strong>Tidal</strong> Response <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong>. In <strong>Tidal</strong> Phenomena, Lecture<br />
Notes in <strong>Earth</strong> Sciences, pages 27–57. Springer.<br />
Wegener, A. (1966). The Origin <strong>of</strong> Continents and Oceans. Courier Dover Publications,<br />
4th edition. Translated by John Biram.
92 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
Wenzel, H.-G. (1997). Tide-Generating Potential for <strong>the</strong> <strong>Earth</strong>. In <strong>Tidal</strong> Phenomena,<br />
Lecture Notes in <strong>Earth</strong> Sciences, pages 9–26. Springer.<br />
wikimedia.org (2006). File:tidal-forces-calculated-right.png. http://commons.<br />
wikimedia.org/wiki/File:<strong>Tidal</strong>-forces-calculated-right.png.<br />
Woan, G. (2003). The Cambridge handbook <strong>of</strong> Physical Formulas. Cambridge University<br />
Press, 2003 edition.<br />
Xing, H. L., J.Zhang, and Yin, C. (2007). A Finite Element Aanalysis <strong>of</strong> <strong>Tidal</strong> <strong>Deformation</strong><br />
<strong>of</strong> <strong>the</strong> Entire <strong>Earth</strong> with a Discontinous Outer Layer. Geophysical Journal<br />
International, 170:961–970.<br />
Xu, J., Sun, H., and Ducarme, B. (2004). A global experimental model for gravity<br />
tides <strong>of</strong> <strong>the</strong> <strong>Earth</strong>. Journal <strong>of</strong> Geodynamics, 38(3-5):293–306.<br />
Zürn, W. (1997). The Nearly-Diurnal Free Wobble-Resonance. In <strong>Tidal</strong> Response <strong>of</strong><br />
<strong>the</strong> <strong>Solid</strong> <strong>Earth</strong> <strong>Tidal</strong> Phenomena. Springer.
Appendices
Appendix A<br />
Hyperbolic Function<br />
Considering a general partial difference equation in <strong>the</strong> domain Ω <strong>of</strong> <strong>the</strong> form<br />
a(r, θ)urr + 2b(r, θ)urθ + c(r, θ)uθθ + d(r, θ)ur + e(r, θ)uθ + g(r, θ)u(r, θ) (A.0.1)<br />
To verify that <strong>the</strong> Navier equation <strong>of</strong> motion, Equation (3.3.3), is a hyperbolic function<br />
has to be valid.<br />
b 2 − ac > 0 for all (r, θ) (A.0.2)<br />
In <strong>the</strong> figure below <strong>the</strong> determinant, b 2 − ac, has been plotted, and as <strong>the</strong> figure shows,<br />
<strong>the</strong> determinant is at all points larger than zero. This prove, that <strong>the</strong> partial difference<br />
equation is a hyperbolic function.
Appendix B<br />
Matlab Codes<br />
In this appendix a selection <strong>of</strong> <strong>the</strong> Matlab codes developed doing this <strong>the</strong>sis are displayed.<br />
B.1 FDE_premiso.m<br />
FDE_premiso.m is <strong>the</strong> main code for <strong>the</strong> layered <strong>Earth</strong> model. It is chosen not to show<br />
<strong>the</strong> code for <strong>the</strong> homogeneous <strong>Earth</strong> model, because <strong>of</strong> triviality.<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
%<br />
% This program s o l v e s r a d i a l u and t a n g e n t i a l v d i s p l a c e m e n t from<br />
% t h e N a v i e r e q u a t i o n <strong>of</strong> Motion .<br />
% In t h e domain : r = [ R_0 , a ] , t h e t a = [ t h e t a _ 0 , <strong>the</strong>ta_m ] .<br />
% R_0=0 , a =3671000 , t h e t a _ 0 =0 , <strong>the</strong>ta_m=p i / 2 ,<br />
% with seven boundary c o n d i t i o n s r=R_0 , two a t r=a and<br />
% two a t t h e t a _ 0 =0 and two a t <strong>the</strong>ta_m=p i / 2<br />
%<br />
% Master t h e s i s by S t i n e k i l d e g a a r d P o u l s e n .<br />
%<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
t i c<br />
c l e a r a l l , c l o s e a l l<br />
opengl neverselect<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
% CONSTANTS ( a l l i n SI − u n i t s )<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
g l o b a l rho G Mm rm<br />
%F _ t i d e s<br />
G = 6.6725985 e −11; % G r a v i t a t i o n a l c o n s t . m^ 3 / ( kg * s ^ 2)<br />
Mm = 7.3483 e22 ; % Mass <strong>of</strong> t h e Moon<br />
Me = 5.9742 e24 ;<br />
rm = 3.84400 e8 ; % O r b i t a l d i s t a n c e <strong>of</strong> t h e Moon<br />
a = 6 . 3 7 1 e6 ;<br />
% E l a s t i c p a r a m e t e r s d e r i v e d from t h e i s o t r o p i c PREM:<br />
l o a d . . / premiso / meanvalues_alpha . dat<br />
l o a d . . / premiso / meanvalues_beta . dat
96 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
l o a d . . / premiso / meanvalues_rho . dat<br />
lam = meanvalues_alpha ( 4 ) ^2 * meanvalues_rho ( 4 ) −2 * meanvalues_beta ( 4 ) ^ 2 ;<br />
mu = meanvalues_rho ( 4 ) * meanvalues_beta ( 4 ) ^ 2 ;<br />
sigma=lam / ( 2 * ( lam+mu ) ) ; % u n i t l e s s<br />
Emodulus = mu * (3 * lam+2 * mu ) / ( lam+mu ) ; % kg / (m * s ^ −2)<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
% S e t t i n g up t h e m a t r i c e s<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
% −−−−−−−−−−−−−R a d i a l d i r e c t i o n −−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
% R a d i a l C o n s t a n t s<br />
r0 = 0 ; % Radius <strong>of</strong> t h e e a r t h<br />
dr=a / 6 5 ; % loop s t e p<br />
istart = 1 ; % loop s t a r t<br />
radius = ( r0 : dr : a ) ' ; % G e n e r a t i n g t h e f i n a l r v e c t o r − s i z e : hr x 1<br />
iend = l e n g t h ( radius ) ; % # g r i d p o i n t s i n r and end <strong>of</strong> l oop<br />
hr = ( a−r0 ) / iend ; % Mesh s i z e i n r − d i r e c t i o n<br />
% −−−−−−−−−−−−The l a t i t u d i n a l d i r e c t i o n :−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
% L a t i t u d i n a l C o n s t a n t s<br />
<strong>the</strong>ta0 = 0 ; <strong>the</strong>taend = p i / 2 ; % l a t i t u d i n a l d i r e c t i o n<br />
d<strong>the</strong>ta =( <strong>the</strong>taend−<strong>the</strong>ta0 ) / 5 5 ; % loop s t e p s i z e<br />
jstart =1; % s t a r t <strong>of</strong> loop<br />
<strong>the</strong>ta = ( <strong>the</strong>ta0 : d<strong>the</strong>ta : <strong>the</strong>taend ) ' ; % G e n e r a t i n g l a t . S i z e : ( h t h e t a +1) x 1<br />
jend= l e n g t h ( <strong>the</strong>ta ) ; % # g r i d p o i n t i n t h e t a and end <strong>of</strong> loop<br />
h<strong>the</strong>ta =( <strong>the</strong>taend−<strong>the</strong>ta0 ) / jend ; % Mesh s i z e i r a d i a n s<br />
%−−−−−−−−−−−Making t h e row o r d e r i n g matrix −−−−−−−−−−−−−−−−−−−−<br />
k= z e r o s ( iend , jend ) ;<br />
f o r i=istart : iend<br />
f o r j=jstart : jend<br />
k ( i , j ) = j + jend * ( i −1) ;<br />
end<br />
end<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
% The C o e f f i c i e n t m a t r i x A and t h e d a t a k e r n e l F ( t i d a l acc . )<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
M = iend * jend ; % Dimension<br />
%−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
% BOUNDARY CONDITIONS<br />
%−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
%−−−−− BC a t c e n t e r <strong>of</strong> t h e E a r t h ( Ac and f c )−−−−−−−−−−−−−−−−<br />
Ac= s p a r s e ( jend , 2 * M ) ;<br />
fc = s p a r s e ( jend , 1 ) ' ;<br />
i =1;<br />
f o r j=jstart : jend<br />
Ac ( j , k ( i , j ) ) = 1 ;<br />
fc ( j ) = 0 ;<br />
end<br />
%−−−BC a t e q u a t o r f o r t h e r a d i a l component ( Aue and f u e )−−−−<br />
Aue = s p a r s e ( iend , 2 * M ) ;<br />
fue = s p a r s e ( iend , 1 ) ' ;<br />
f o r i=istart : iend<br />
Aue ( i , k ( i , jend −1) ) = 1 ;<br />
Aue ( i , k ( i , jend ) ) = −1;<br />
fve ( i ) = 0 ;<br />
end
FDE_premiso.m 97<br />
%−−−BC a t t h e p o l e f o r t h e r a d i a l komponent ( Aup and fup )−−−−<br />
Aup = s p a r s e ( iend , 2 * M ) ;<br />
fup = s p a r s e ( iend , 1 ) ' ;<br />
j=jend ;<br />
f o r i=istart : iend<br />
Aup ( i , k ( i , 1 ) ) = −1;<br />
Aup ( i , k ( i , 2 ) ) = 1 ;<br />
fup ( i ) = 0 ;<br />
end<br />
%−−BC a t E q u a t o r and t h e P ole f o r l a t . component ( Aue , Aup and fuep )−−<br />
Aue = s p a r s e ( iend , 2 * M ) ;<br />
Aup = s p a r s e ( iend , 2 * M ) ;<br />
fuep = s p a r s e ( iend , 1 ) ' ;<br />
f o r i=istart : iend<br />
Aue ( i , k ( i , jend )+M ) = 1 ;<br />
Aup ( i , k ( i , 1 ) +M ) = 1 ;<br />
fuep ( i ) = 0 ;<br />
end<br />
%−−BC a t t h e s u r f a c e <strong>of</strong> t h e E a r t h − r a d i a l d i r e c t i o n ( Ars , f r s )−−−−−<br />
taukonst=Emodulus / ( ( 1 + sigma ) * (1 −2 * sigma ) ) ;<br />
Ars = s p a r s e ( jend , 2 * M ) ;<br />
frs = s p a r s e ( jend , 1 ) ' ;<br />
% With no g h o s t p o i n t<br />
i=iend ;<br />
f o r j =2: jend<br />
Ars ( j , k ( i −1 ,j ) ) = −taukonst * (1 − sigma ) / hr ; % Au_ ( a −1 , j )<br />
Ars ( j , k ( i , j ) ) =taukonst * (1 − sigma ) / hr . . .<br />
+ 2 * taukonst * sigma / a ; % Au_aj<br />
Ars ( j , k ( i , j −1)+M ) = −taukonst * sigma / ( h<strong>the</strong>ta * a ) ; % Av_ ( a , j −1)<br />
Ars ( j , k ( i , j )+M ) = taukonst * sigma / ( h<strong>the</strong>ta * a ) . . .<br />
+ taukonst * sigma * c o t ( <strong>the</strong>ta ( j ) ) / a ; % Av_ ( a , j )<br />
frs ( j ) = 0 ; % F ( a )<br />
end<br />
Ars ( 1 , k ( iend , 1 ) )=Ars ( 2 , k ( iend , 2 ) ) ;<br />
Ars ( 1 , k ( iend , 1 ) +M )=Ars ( 2 , k ( iend , 1 ) +M ) ;<br />
Ars ( 1 , k ( iend , 2 ) +M )=Ars ( 2 , k ( iend , 2 ) +M ) ;<br />
Ars ( 1 , k ( iend −1 ,1) )=Ars ( 2 , k ( iend −1 ,2) ) ;<br />
frs ( 1 ) = 0 ;<br />
%−−BC a t t h e s u r f a c e <strong>of</strong> t h e E a r t h − ( r , t h e t a ) d i r e c t i o n ( Ats , f t s )−−−−−<br />
Ats = s p a r s e ( jend , 2 * M ) ;<br />
fts = s p a r s e ( jend , 1 ) ' ;<br />
i=iend ;<br />
f o r j =2: jend<br />
Ats ( j , k ( i , j ) ) = mu / ( a * h<strong>the</strong>ta ) ; % Au_ ( a , j )<br />
Ats ( j , k ( i , j −1) ) = −mu / ( a * h<strong>the</strong>ta ) ; % Au_ ( a , j −1)<br />
Ats ( j , k ( i , j )+M ) = mu / hr . . .<br />
− mu / a ; % Av_ ( a , j )<br />
Ats ( j , k ( i −1 ,j )+M ) = −mu / hr ; % Av_ ( a −1 , j )<br />
fts ( j ) = 0 ; % F ( a )<br />
end<br />
Ats ( 1 , k ( iend , 1 ) )=−Ats ( 2 , k ( iend , 2 ) ) ;<br />
Ats ( 1 , k ( iend , 2 ) )=Ats ( 2 , k ( iend , 2 ) ) ;<br />
Ats ( 1 , k ( iend , 1 ) +M )=Ats ( 2 , k ( iend , 2 ) +M ) ;<br />
Ats ( 1 , k ( iend −1 ,1)+M )=−Ats ( 2 , k ( iend , 2 ) +M ) ;<br />
fts ( 1 ) = 0 ;<br />
%−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
% The main Coef . m a t r i x Au and fu ( r a d i a l d i r e c t i o n )
98 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
%−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
fu = z e r o s ( M , 1 ) ' ; % Making t h e s i z e <strong>of</strong> t h e d a t a k e r n e l<br />
%−−−−−−−−−−−−−−−− The t i d a l a c c e l e r a t i o n ( r a d i a l ) −−−−−−−−−−−−−−<br />
f o r i =1: iend<br />
f o r j =1: jend<br />
fu ( k ( i , j ) ) = ftides_r ( radius ( i ) , <strong>the</strong>ta ( j ) ) ; % c a l l i n g −Fu ( r , t h e t a )<br />
end<br />
end<br />
% P r e p a r i n g t h e Av m a t r i x<br />
Au = s p a r s e ( M , M ) ; % Making t h e m a t r i x<br />
f o r i =2: iend−1<br />
f o r j =2: jend−1<br />
% −−−−−−−−−−−−−−− The main FDE e q u a t i o n s :−−−−−−−−−−−−−−−−−−−−−−−−<br />
Au ( k ( i , j ) , k ( i , j ) ) = − 2 * mean_beta ( i ) ^2 . / ( radius ( i ) . ^ 2 * . . .<br />
h<strong>the</strong>ta ^2)− 2 * mean_alpha ( i ) ^2 . / ( hr ^ 2) − ( mean_alpha ( i ) ^2 − . . .<br />
mean_beta ( i ) ^ 2) . / ( radius ( i ) . ^ 2 ) ; % Au_ij<br />
%−−−−−−−−−−−−−−−− The r a d i a l d i r e c t i o n : −−−−−−−−−−−−−−−−−−−−−−−−−<br />
Au ( k ( i , j ) , k ( i −1 ,j ) ) = mean_alpha ( i ) . ^ 2 * (2 * radius ( i ) + hr ) . . .<br />
. / (2 * radius ( i ) * hr . ^ 2 ) ; % Au_ ( i −1 , j )<br />
Au ( k ( i , j ) , k ( i+1 ,j ) ) = mean_alpha ( i ) . ^ 2 * (2 * radius ( i ) − hr ) . . .<br />
. / (2 * radius ( i ) * hr . ^ 2 ) ; % Au_ ( i +1 , j )<br />
%−−−−−−−−−−−−−−−− The l a t i t u d i n a l d i r e c t i o n :−−−−−−−−−−−−−−−−−−−−−<br />
Au ( k ( i , j ) , k ( i , j −1) ) = (− ( mean_alpha ( i ) . ^ 2 − mean_beta ( i ) . ^ 2 ) . . .<br />
* h<strong>the</strong>ta + 2 * mean_beta ( i ) ^ 2) . . .<br />
. / (2 * radius ( i ) . ^ 2 * h<strong>the</strong>ta ^ 2) ; % Au_ ( i , j −1)<br />
Av ( k ( i , j ) , k ( i , j+1) ) = ( ( mean_alpha ( i ) . ^ 2 − mean_beta ( i ) . ^ 2 ) . . .<br />
* h<strong>the</strong>ta + 2 * mean_beta ( i ) ^ 2) . . .<br />
. / (2 * radius ( i ) . ^ 2 * h<strong>the</strong>ta ^ 2) ; % Au_ ( i , j +1)<br />
% −−−−−−−−−−The mixed t e r m s:−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
Au ( k ( i , j ) , k ( i+1 ,j+1) ) = ( mean_alpha ( i ) .^2 − mean_beta ( i ) . ^ 2 ) . / . . .<br />
(4 * radius ( i ) * hr * h<strong>the</strong>ta ) ; %Au_ ( i +1 , j −1)<br />
Au ( k ( i , j ) , k ( i+1 ,j −1) ) = −( mean_alpha ( i ) .^2 − mean_beta ( i ) . ^ 2 ) . / . . .<br />
(4 * radius ( i ) * hr * h<strong>the</strong>ta ) ;%Au_ ( i +1 , j −1)<br />
Au ( k ( i , j ) , k ( i −1 ,j+1) ) = −( mean_alpha ( i ) .^2 − mean_beta ( i ) . ^ 2 ) . / . . .<br />
(4 * radius ( i ) * hr * h<strong>the</strong>ta ) ;%Au_ ( i −1 , j +1)<br />
Au ( k ( i , j ) , k ( i −1 ,j −1) ) = ( mean_alpha ( i ) .^2 − mean_beta ( i ) . ^ 2 ) . / . . .<br />
(4 * radius ( i ) * hr * h<strong>the</strong>ta ) ; %Au_ ( i −1 , j −1)<br />
end<br />
end<br />
%−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
% The main Coef . m a t r i x Av and fv ( t a n g e n t i a l d i r e c t i o n )<br />
%−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
% P r e p a r i n g t h e Av m a t r i x<br />
Av = s p a r s e ( M , M ) ; % Making t h e m a t r i x<br />
fv = s p a r s e ( M , 1 ) ' ; % Making t h e s i z e <strong>of</strong> t h e d a t a k e r n e l<br />
%−−−−−−−−−−−−−−−− The t i d a l a c c e l e r a t i o n ( a n g u l a r ) −−−−−−−−−−−−−−<br />
f o r i =1: iend<br />
f o r j =1: jend<br />
fv ( k ( i , j ) ) = ftides_<strong>the</strong>ta ( radius ( i ) , <strong>the</strong>ta ( j ) ) ; % t i d a l acc . −Fv<br />
end<br />
end<br />
f o r i =2: iend−1<br />
f o r j =2: jend−1
FDE_premiso.m 99<br />
end<br />
end<br />
% −−−−−−−−−−−−The main FDE e q u a t i o n s −−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
Av ( k ( i , j ) , k ( i , j ) ) = − 2 * mean_beta ( i ) . ^ 2 / ( hr ^ 2) . . .<br />
− 2 * mean_alpha ( i ) . ^ 2 . / ( radius ( i ) . ^ 2 * h<strong>the</strong>ta ^ 2) ;% v _ i j<br />
%−−−−−−−−−−−−The r a d i a l d i r e c t i o n −−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
Av ( k ( i , j ) , k ( i −1 ,j ) ) = mean_beta ( i ) . ^ 2 * (2 * radius ( i ) − hr ) . . .<br />
. / (2 * radius ( i ) * hr ^2) ; % v_ ( i −1 , j )<br />
Av ( k ( i , j ) , k ( i+1 ,j ) ) = mean_beta ( i ) . ^ 2 * (2 * radius ( i ) + hr ) . . .<br />
. / (2 * radius ( i ) * hr ^2) ; % v_ ( i +1 , j )<br />
%−−−−−−−−−−−−−The l a t i t u d i n a l d i r e c t i o n :−−−−−−−−−−−−−−−−−−−−−−−<br />
Av ( k ( i , j ) , k ( i , j −1) ) = (2 * mean_alpha ( i ) . ^ 2 + ( mean_alpha ( i ) . ^ 2 . . .<br />
− mean_beta ( i ) . ^ 2 ) * h<strong>the</strong>ta ) . . .<br />
. / (2 * radius ( i ) . ^ 2 * h<strong>the</strong>ta ^2) ;<br />
Av ( k ( i , j ) , k ( i , j+1) ) = (2 * mean_alpha ( i ) . ^ 2 − ( mean_alpha ( i ) ^2 . . .<br />
− mean_beta ( i ) . ^ 2 ) * h<strong>the</strong>ta ) . . .<br />
. / (2 * radius ( i ) . ^ 2 * h<strong>the</strong>ta ^ 2) ;<br />
% −−−−−−−−−−The mixed t e r m s:−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
Av ( k ( i , j ) , k ( i+1 ,j+1) ) = ( mean_alpha ( i ) . ^ 2 − mean_beta ( i ) . ^ 2 ) . / . . .<br />
(4 * radius ( i ) * hr * h<strong>the</strong>ta ) ;<br />
Av ( k ( i , j ) , k ( i+1 ,j −1) ) = −( mean_alpha ( i ) . ^ 2 − mean_beta ( i ) . ^ 2 ) . / . . .<br />
(4 * radius ( i ) * hr * h<strong>the</strong>ta ) ;<br />
Av ( k ( i , j ) , k ( i −1 ,j+1) ) = − ( mean_alpha ( i ) . ^ 2 − mean_beta ( i ) . ^ 2 ) . / . . .<br />
(4 * radius ( i ) * hr * h<strong>the</strong>ta ) ;<br />
Av ( k ( i , j ) , k ( i −1 ,j −1) ) = ( mean_alpha ( i ) . ^ 2 − mean_beta ( i ) . ^ 2 ) . / . . .<br />
(4 * radius ( i ) * hr * h<strong>the</strong>ta ) ;<br />
%−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
% Assembling t h e c o e f f i c i e n t m a t r i x and f o r c e v e c t o r<br />
%−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−<br />
zero= z e r o s ( s i z e ( Au ) ) ;<br />
A1 = [ Au zero ] ;<br />
A2 = [ zero Av ] ;<br />
ff=1e −5; %BC f a c t o r<br />
% F i n a l Coef . M a t r i x<br />
A = [ A1 ; A2 ; Ac ; Aue ; Aup ; Ave ; Avp ; ff * Ars ; ff * Ats ] ;<br />
% F i n a l f o r c e v e c t o r<br />
F = [ fu , fv , fc , fue , fup , fvep , fvep , ff * frs , ff * fts ] ' ;<br />
A= s p a r s e ( A ) ;<br />
F= s p a r s e ( F ) ;<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
% SOLVING THE SYSTEM OF EQUATIONS<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
% Tikhonov r e g u l a r i z a t i o n f u n c t i o n from P . Hansen<br />
[ reg_corner , rho , eta , reg_param ]= l_curve ( U , s , F , ' Tikh ' ) ;<br />
epsilon=reg_corner ;<br />
[ U , s , V ] = csvd ( A ) ; % Compact SVD<br />
[ X , rho , eta ] = tikhonov ( U , s , V , F , lambda ) ;<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
% Transform back t o ( i , j ) form t o p l o t t h e s o l u t i o n<br />
% Making t h e f i n a l c o e f i c i e n t m a t r i x<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
Ur=X ( 1 : M , 1 ) ;<br />
U<strong>the</strong>ta=X ( M +1:2 * M , 1 ) ;<br />
ur= s p a r s e ( iend , jend ) ; % R a d i a l d i s p l a c e m e n t<br />
u<strong>the</strong>ta= s p a r s e ( iend , jend ) ; % Lat . d i s p l a c e m e n t
100 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
fr= s p a r s e ( iend , jend ) ; % R a d i a l t i d a l f o r c e<br />
f<strong>the</strong>ta= s p a r s e ( iend , jend ) ; % Lat . t i d a l f o r c e<br />
f o r l =1: M<br />
i = f l o o r ( ( l −1) / jend ) + 1 ;<br />
j = mod ( l −1 , jend ) +1;<br />
ur ( i , j ) = Ur ( l ) ;<br />
u<strong>the</strong>ta ( i , j )=U<strong>the</strong>ta ( l ) ;<br />
fr ( i , j )= fu ( l ) ;<br />
f<strong>the</strong>ta ( i , j )=fv ( l ) ;<br />
end<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
% Analyze and V i s u a l i z e t h e r e s u l t .<br />
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%<br />
NORMcalc=norm ( F−A * X ) / norm ( F ) ;<br />
R = ( 0 : dr : a ) ' ;<br />
T= p i / 2 −<strong>the</strong>ta ' ;<br />
XX=R * cos ( T ) ;<br />
YY=R * s i n ( T ) ;<br />
% R a d i a l d i s p l a c e m e n t<br />
f i g u r e ( 1 )<br />
a x i s equal tight<br />
p c o l o r ( XX , YY , ur ) , s h a d i n g flat<br />
hcb = c o l o r b a r ( ' WestOutside ' ) ;<br />
s e t ( g e t ( hcb , ' Y l a b e l ' ) , ' S t r i n g ' , ' R a d i a l d i s p l a c e m e n t i n m. ' )<br />
hold on<br />
ContourHandle= c o n t o u r ( XX , YY , ur , 2 0 , ' k−− ' ) ;<br />
c l a b e l ( ContourHandle , ' F o n t S i z e ' , 1 2 , ' Color ' , ' r ' , ' R o t a t i o n ' , 0 )<br />
hold on<br />
a x i s <strong>of</strong>f<br />
%[ px , py ] = g r a d i e n t ( ur , 0 . 2 , 0 . 2 ) ;<br />
%q u i v e r (XX,YY, px , py )<br />
% T a n g e n t i a l d i s p l a c e m e n t<br />
f i g u r e ( 2 )<br />
a x i s equal tight<br />
p c o l o r ( XX , YY , u<strong>the</strong>ta ) , s h a d i n g flat<br />
x l a b e l ( 'R (m) ' , ' F o n t S i z e ' , 1 2 )<br />
y l a b e l ( 'R (m) ' , ' F o n t S i z e ' , 1 2 )<br />
%t i t l e ( ' u_ \ t h e t a , A l l c o n s t a n t s \ lambda =0.5 e − 5 ' )<br />
hold on<br />
ContourHandle= c o n t o u r ( XX , YY , u<strong>the</strong>ta , 1 0 , ' k−− ' ) ;<br />
c l a b e l ( ContourHandle , ' F o n t S i z e ' , 1 2 , ' Color ' , ' r ' , ' R o t a t i o n ' , 0 )<br />
hcb = c o l o r b a r ( ' WestOutside ' ) ;<br />
s e t ( g e t ( hcb , ' Y l a b e l ' ) , ' S t r i n g ' , ' T a n g e n t i a l d i s p l a c e m e n t i n m. ' )<br />
hold on<br />
a x i s <strong>of</strong>f<br />
%hold on<br />
%[ qx , qy ] = g r a d i e n t ( u t h e t a , 0 . 2 , 0 . 2 ) ;<br />
%q u i v e r (XX,YY, qx , qy )<br />
t o c
Matlab Functions 101<br />
B.2 Matlab Functions<br />
This section contains Matlab functions called from <strong>the</strong> main program FDE_premiso.m.<br />
B.2.1 ftides_r.m<br />
The radial tidal acceleration.<br />
f u n c t i o n f = ftides_r ( radius , <strong>the</strong>ta )<br />
g l o b a l G Mm rm<br />
f = − G * Mm * radius . / ( rm ^ 3) * (3 * ( cos ( p i /2 − <strong>the</strong>ta ) ' ) . ^ 2 − 1) ;<br />
B.2.2 ftides_<strong>the</strong>ta.m<br />
The horizontal tidal acceleration.<br />
f u n c t i o n f = ftides_<strong>the</strong>ta ( radius , <strong>the</strong>ta )<br />
g l o b a l G Mm rm<br />
f = − 3 / 2 * G * Mm * radius . / ( rm ^ 3) * s i n ( 2 * ( p i /2 − <strong>the</strong>ta ) ' ) ;<br />
B.2.3 mean_alpha.m<br />
The seismic P-wave velocity for <strong>the</strong> four layers.<br />
f u n c t i o n f = mean_alpha ( radius )<br />
l o a d premiso / meanvalues_alpha . dat<br />
i f ( radius >= 0) && ( radius < 1221491)<br />
f = meanvalues_alpha ( 1 ) ; % i n n e r c o r e<br />
e l s e i f ( radius >= 1221491) && ( radius < 3479958)<br />
f = meanvalues_alpha ( 2 ) ; % o u t e r c o r e<br />
e l s e i f ( radius >= 3479958) && ( radius < 5701000)<br />
f = meanvalues_alpha ( 3 ) ; % m a n t l e<br />
e l s e<br />
f = meanvalues_alpha ( 4 ) ; % Outer l a y e r<br />
end
102 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
B.2.4 mean_beta.m<br />
The seismic S-wave velocity for <strong>the</strong> four layers.<br />
f u n c t i o n f = mean_beta ( radius )<br />
l o a d premiso / meanvalues_beta . dat<br />
i f ( radius >= 0) && ( radius < 1221491)<br />
f = meanvalues_beta ( 1 ) ;<br />
e l s e i f ( radius >= 1221491) && ( radius < 3479958)<br />
f = meanvalues_beta ( 2 ) ;<br />
e l s e i f ( radius >= 3479958) && ( radius < 5701000)<br />
f = meanvalues_beta ( 3 ) ;<br />
e l s e<br />
f = meanvalues_beta ( 4 ) ;<br />
end
Appendix C<br />
Averaging <strong>of</strong> PREM Layers<br />
This appendix shows how <strong>the</strong> P- and S-wave velocities and <strong>the</strong> density is averaged in<br />
each layer <strong>of</strong> <strong>the</strong> layered <strong>Earth</strong> model. The data is adapted from PREM, [Dziewonski<br />
and Anderson, 1981]. Blue is <strong>the</strong> P-wave velocity given in m s , green is <strong>the</strong> S-wave<br />
velocity also in m kg<br />
s , and <strong>the</strong> red curve is <strong>the</strong> density given in<br />
m3 . The radius is counting<br />
from <strong>the</strong> center <strong>of</strong> <strong>Earth</strong> to <strong>the</strong> surface.<br />
(a) Inner core (b) Outer core<br />
(c) Mantle (d) Outer layer
Appendix D<br />
Model Tests<br />
In this chapter some <strong>of</strong> <strong>the</strong> results from <strong>the</strong> testing <strong>of</strong> <strong>the</strong> model is shown.<br />
D.1 Tikhonov Tests<br />
Test <strong>of</strong> <strong>the</strong> Tikhonov regularization with λ = 10 −1 − λ = 10 −5 , where lambda is <strong>the</strong><br />
regularization parameter. A point source is applied at (25,25) in <strong>the</strong> radial direction.<br />
The grid numbers are 50 × 50.
Tikhonov Tests 105<br />
(e) λ = 10 −1 (f) λ = 10 −2<br />
(g) λ = 10 −3 (h) λ = 10 −4<br />
(i) λ = 10 −5
106 <strong>Tidal</strong> <strong>Deformation</strong> <strong>of</strong> <strong>the</strong> <strong>Solid</strong> <strong>Earth</strong><br />
D.2 Regularization Tests<br />
Surface plots <strong>of</strong> different regularization methods, only <strong>the</strong> best are showed. The grid<br />
size is 50 × 50. A point source is applied at (25,25) in <strong>the</strong> radial direction. Here λ is<br />
<strong>the</strong> regularization parameter.<br />
50<br />
45<br />
40<br />
35<br />
30<br />
25<br />
20<br />
15<br />
10<br />
5<br />
QR m. householder<br />
10 20 30 40 50<br />
(j) QR med householder,<br />
10−14 |F−A·X|<br />
|F|<br />
(l) TSVD, k = 3000, |F−A·X|<br />
|F|<br />
0.14<br />
0.12<br />
0.1<br />
0.08<br />
0.06<br />
0.04<br />
0.02<br />
0<br />
= 5.407 ·<br />
(n) TSVD, k = rank(A) = 4853, |F−A·X|<br />
|F|<br />
1.94<br />
(k) DSVD λ = 10 −2 , |F−A·X|<br />
|F|<br />
= 0.6115 (m) Tikhonov, lambda = 1e − 4,<br />
0.0341<br />
=<br />
= 0.409<br />
|F−A·X|<br />
|F|<br />
(o) TSVD, k = rank(A) = 5000, |F−A·X|<br />
|F|<br />
2.29 · 10−6 =<br />
=