Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

3.

06 Earth Tides
DC Agnew, University of California, La Jolla, CA, USA
ã 2015 Elsevier B.V. All rights reserved.

3.06.1 Introduction 151


3.06.1.1 An Overview 152
3.06.2 The Tidal Forces 152
3.06.2.1 The Tidal Potential (I) 153
3.06.2.2 Computing the Tides: Ephemeris-Based 154
3.06.2.3 Computing the Tides (II): Harmonic Decompositions 155
3.06.2.4 The Pole Tide 157
3.06.2.5 Radiational Tides 159
3.06.3 Tidal Response of the Solid Earth 159
3.06.3.1 Tidal Response of an SNREI Earth 159
3.06.3.1.1 Some combinations of Love numbers (I): gravity and tilt 160
3.06.3.1.2 Combinations of Love numbers (II): Displacement and strain tides 160
3.06.3.2 Response of a Rotating Earth 161
3.06.3.2.1 NDFW resonance 161
3.06.3.2.2 Coupling to other modes 163
3.06.3.2.3 Anelastic effects 164
3.06.3.3 Lateral Heterogeneities 165
3.06.4 Tidal Loading 165
3.06.4.1 Computing Loads I: Spherical Harmonic Sums 165
3.06.4.2 Computing Loads II: Integration Using Green Functions 166
3.06.4.3 Degree-One Effects in Ocean Loading 168
3.06.4.4 Ocean Tide Models 169
3.06.4.5 Computational Methods 169
3.06.5 Analyzing and Predicting Earth Tides 169
3.06.5.1 Tidal Analysis and Prediction 169
3.06.5.1.1 Predicting tides 171
3.06.6 Earth-Tide Instruments and Measurements 172
3.06.6.1 Local Distortion of the Tides 173
References 174

3.06.1 Introduction Accurate modeling of Earth tides is also easy because they
do not depend very strongly on Earth structure – which also
The Earth tides are the motions in the solid Earth, and the means that Earth-tide measurements, and accurate estimates of
changes in its gravitational potential, induced by the tidal the few parameters that describe the Earth’s response to the
forces from external bodies (these forces, acting on the rotating tides, do not provide much information about the Earth. This
Earth, also induce motions of its spin axis; see Chapter 3.09). was not always true; for example, in 1922, Jeffreys used tidal
Earth tides have three roles in geophysics: measurements of data to show that the average rigidity of the Earth was much
them can provide information about the Earth; models of less than that of the mantle, indicating that the core must be of
them can be used to remove tidal variations from measure- very low rigidity (Brush, 1996). But since then, seismology
ments; and the same models can be used to examine tidal has determined Earth structure in much more detail than
influence on some phenomenon. An example of the first role could be found using tides. Recently, Earth tides have become
would be measuring the nearly diurnal resonance in the gravity more important in geodesy, as the increasing precision of
tide to estimate the flattening of the core–mantle boundary measurements has required corrections for previously unde-
(Section 3.06.3.2.1); of the second, computing the expected tectable tidal effects. This chapter focuses on the theory needed
tidal displacements at a point so we can better estimate its to compute tidal effects as accurately as possible for under-
position with GPS; and of the third, finding the tidal stresses standing how phenomena are or are not influenced by tides.
to see if they trigger earthquakes. For the last two activities, we Section 3.06.6 summarizes, more briefly, measurement tech-
need to be able to model the Earth tides very accurately, some- niques and some results.
thing that is in fact not difficult. Notably, Earth tides are much The best short introduction to Earth tides remains that of
easier to model accurately than ocean tides, both because the Baker (1984). Melchior (1983) described the subject fully (and
Earth is more rigid than water and because the geometry of the with a very complete bibliography) but is now somewhat out
problem is much simpler. of date and should be used with caution by the newcomer to

Treatise on Geophysics, Second Edition http://dx.doi.org/10.1016/B978-0-444-53802-4.00058-0 151


152 Earth Tides

the field. The volume of articles edited by Wilhelm et al. (1997) responses sum to give the total tide caused by the nonrigidity of
is a better reflection of the current state of the subject, as are the the Earth; the final model, labeled ‘Site distortions,’ may be
quadrennial International Symposia on Earth Tides (e.g., used to describe how local departures from idealized models
Jentzsch et al., 2009). Harrison (1985) reprinted a number affect the result (Section 3.06.6.1). This nonrigid contribution
of important papers, with very thoughtful commentary; is summed with the tide from direct attraction to give the total
Cartwright (1999) is a history of tidal science (mostly the theoretical tide, at point T in the flowchart.
ocean tides) that also provides an interesting introduction to Mathematically, we can describe the processes shown on
some aspects of the field – which, as one of the older parts of this flow chart in terms of linear systems, something first
geophysics, has a terminology sometimes overly affected by its applied to tidal theory by Munk and Cartwright (1966). The
history. total signal y(t) is represented by a convolution integral:
ð
yðt Þ ¼ xT ðt  tÞwT ðtÞdt + nðt Þ [1]
3.06.1.1 An Overview
Figure 1 is a simple flowchart to indicate what goes into a tidal where xT(t) is the tidal forcing, n(t) is the noise (nontidal energy,
signal. We usually take the tidal forcing to be completely from whatever source), and t is a dummy variable. The function
known, but it is computed using a particular theory of gravity, w(t) is the impulse response of the system to the tidal forcing.
and Earth-tide measurements actually provided some of the Fourier transforming equation [1], and disregarding the noise,
best evidence available for general relativity as opposed to gives Y(f ) ¼ WT(f )XT(f ): W(f ) is the tidal admittance, which
some alternative theories (Warburton and Goodkind, 1976). turns out to be more useful than w(t), partly because of the
The large box labeled ‘Geophysics/Oceanography’ includes the bandlimited nature of XT and also because, with the exception
response of the Earth and ocean to the forcing, with the arrow of the NDFW resonance discussed in Section 3.06.3.2.1, W(f )
going around it to show that some tidal signals would be is a fairly smooth function of frequency. To predict the tides,
observed even if the Earth were oceanless and rigid. Finally, we assume W(f ) (perhaps guided by previous measurements);
measurements of Earth tides can detect other environmental to analyze them, we determine W(f ).
and tectonic signals. We describe the tidal forcing first, in some detail because
At this point, it is useful to introduce some terminology. the nature of this forcing governs the response and how tidal
The theoretical tides could be called the modeled tides, since measurements are analyzed. We next consider how the solid
they are computed from a set of models. The first model is the Earth responds to the tidal forcing and what effects this pro-
tidal forcing, or equilibrium tidal potential, produced by external duces. After this, we discuss the load tides, completing what we
bodies; this is computed from gravitational and astronomical need to know to produce the full theoretical tides. We conclude
theory and is the tide at point E in Figure 1. The next two with brief descriptions of analysis methods appropriate to
models are those that describe how the Earth and ocean Earth-tide data and instruments for measuring Earth tides.
respond to this forcing; in Figure 1, these are boxes inside the
large dashed box. The solid-Earth model gives what are called
the body tides, which are what would be observed on an ocean- 3.06.2 The Tidal Forces
less but otherwise realistic Earth. The ocean model (which
includes both the oceans and the elastic Earth) gives the load The tidal forces arise from the gravitational attraction of bodies
tides, which are changes in the solid Earth caused by the shift- external to the Earth. As noted earlier, computing them requires
ing mass of water associated with the ocean tides. These two only some gravitational potential theory and astronomy,

Direct attraction Non-tidal signals


Celestial
bodies Environment
and tectonics
Geophysics/Oceanography

Gravity Tidal E Solid Earth Environmental


+ data
theory forces and core

Site Tidal
+ + +
distortions signal
Ocean T
Earth loading
orbit/rotation

Total signal
Figure 1 Tidal flowchart. Entries in italics represent things we know (or think we know) to very high accuracy; entries in boldface (over the dashed
boxes) represent things we can learn about using tidal data. See text for details.
Earth Tides 153

1  n
with almost no geophysics. The extraordinarily high accuracy GMext X a
Vtot ¼ Pn ðcos aÞ [2]
of celestial mechanics makes it easy to describe the tidal forcing R n¼0 R
to much higher precision than can be measured: in this part
of the subject, the romance of the next decimal place has where P2 ðxÞ ¼ 12 ð3x2  1Þ and P3 ðxÞ ¼ 12 ð5x3  3xÞ.
exerted a somewhat excessive pull. The n ¼ 0 term is constant in space, so its gradient (the
Our formal derivation of the tidal forcing will use potential force) is zero, and it can be discarded. The n ¼ 1 term is
theory, but it is useful to start by considering the gravitational
forces exerted on one body (the Earth, in this case) by another. GMext GMext
a cos a ¼ x1 [3]
As usual in discussing gravitation, we work in terms of accel- R2 R2
erations. Put most simply, the tidal acceleration at a point where x1 is the Cartesian coordinate along the C–M axis. The
on or in the Earth is the difference between the acceleration gradient of this is a constant, corresponding to a constant force
caused by the attraction of the external body and the orbital along the direction to M; but this is just the orbital force at C,
acceleration – which is to say, the acceleration that the which we subtract to get the tidal force. Thus, the tidal poten-
Earth undergoes as a whole. This result is valid whatever the tial is eqn [2] with the two lowest terms removed:
nature of the orbit – it would hold just as well if the Earth
1  
were falling directly toward another body. For a spherically GMext X a n
Vtid ðt Þ ¼ Pn ½cos aðt Þ [4]
symmetrical Earth, the orbital acceleration is the acceleration Rðt Þ n¼2 Rðt Þ
caused by the attraction of the other body at the Earth’s center
of mass, making the tidal force the difference between the where we have made R and a, as they actually are, functions of
attraction at the center of mass and that at the point of time t – which makes V such a function as well.
observation. We can now put in some numbers appropriate to the Earth
Figure 2 shows the resulting force field. At the point directly and relevant external bodies to get a sense of the magnitudes of
under the attracting body (the subbody point), and at its anti- different terms. If r is the radius of the Earth, a/R ¼ 1/60 for the
pode, the tidal force is oppositely directed in space, though in Moon, so that the size of terms in the sum [4] decreases fairly
the same way (up) viewed from the Earth. It is in fact larger at rapidly with increasing n; in practice, we need only consider
the subbody point than at its antipode, though if the ratio a/R n ¼ 2 and n ¼ 3, and perhaps n ¼ 4 for the highest precision; the
is small (1/60 for the forces plotted in Figure 2), the difference n ¼ 4 tides are just detectable in very low-noise gravimeters.
is also small. These different values of n are referred to as the degree-n tides.
For the Sun, r/R ¼ 1/23 000, so the degree-2 solar tides
completely dominate.
3.06.2.1 The Tidal Potential (I) If we consider n ¼ 2, the magnitude of Vtid is proportional to
We now derive an expression for the tidal force – or rather, for GMext/R3. If we normalize this quantity to make the value for
the more useful tidal potential, following the development in the Moon equal to 1, the value for the Sun is 0.46, for Venus
Munk and Cartwright (1966). If Mext is the mass of the external 5  105, for Jupiter 6  106, and even less for all other
body, the gravitational potential, Vtot, from it at O is planets. So the lunisolar tides dominate and are probably the
only ones visible in actual measurements – though, as we will
GMext GMext 1 see, some expansions of the tidal potential include planetary
Vtot ¼ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
r R 2 tides.
1 + ða=RÞ  2ða=RÞcos a
At very high precision, we also need to consider another
using the cosine rule from trigonometry. The variables are as small effect: the acceleration of the Earth is exactly equal to the
shown in Figure 2: r is the distance of O from C, r the distance attraction of the external body at the center of mass only for a
from O to M, and a the angular distance between O and the spherically symmetrical Earth. For the real Earth, the C20 term
subbody point of M. We can write the square-root term as a in the gravitational potential makes the acceleration of the
sum of Legendre polynomials, using the generating-function Moon by the Earth (and hence the acceleration of the Earth
expression for these, which yields by the Moon) depend on more than just eqn [3]. The resulting

Earth
O
a r Moon
a
M
C
I

Figure 2 Tidal forcing. On the left is the geometry of the problem for computing the tidal force at a point O on the Earth, given an external body M.
The right plot shows the field of forces (accelerations) for the actual Earth–Moon separation; the scale of the largest arrow is 1.14 mm s2 for the
Moon, and 0.51 mm s2 for the Sun. The gray line shows the equipotential surface under tidal forcing, greatly exaggerated. The solid line is the
Earth’s equator; I is the inclination angle of the Moon.
154 Earth Tides

Earth-flattening tides (Dahlen, 1993; Wilhelm, 1983) are very geographic distribution of V/g on the Earth. The time depen-
small. dence of the tidal potential comes from time variations in x, y0 ,
We obtain further insight on the behavior of the tidal forces and f0 . The first two change relatively slowly because of the
if we use geographic coordinates, rather than angular distance orbital motion of M around the Earth; f0 varies approximately
from the subbody point. Suppose our observation point O is at daily as the Earth rotates beneath M. The individual terms in
colatitude y and east longitude f (which are fixed) and that the the sum over m in [6] thus separate the tidal potential of degree
subbody point of M is at colatitude y0 (t) and east longitude n into parts, called tidal species, that vary with frequencies
f0 (t). Then, we may apply the addition theorem for spherical around 0, 1, 2, . . . n times per day; for the largest tides
harmonics to get, instead of [4], (n ¼ 2), there are three such species. The diurnal tidal potential
1   varies once per day and with colatitude as sin y cos y: it is largest
GMext X a n 4p X n
Vtid ¼ * ðy0 ðt Þ, f0 ðt ÞÞYnm ðy, fÞ
Ynm at midlatitudes and vanishes at the equator and the poles. The
Rðt Þ n¼2 Rðt Þ 2n + 1 m¼n semidiurnal part (twice per day) varies as sin2 y and so is largest
[5] at the equator and vanishes at the poles. The long-period tide
varies as 3 cos2 y  1 and so is large at the pole and (with
where we have used the fully normalized complex spherical
reversed sign) at the equator. As we will see, these spatial
harmonics defined by
dependences do not carry over to those tides, such as strain
Ynm ðy, fÞ ¼ Nnm Pnm ðcos yÞeimf and tilt, that depend on horizontal gradients of the potential.
To proceed further, it is useful to separate the time-
where Nm
n is the normalizing factor dependent and space-dependent parts a bit more explicitly.
 1 We adopt the approach of Cartwright and Tayler (1971) who
2n + 1 ðn  mÞ! 2 produced what was for a long time the standard harmonic
Nnm ¼ ð1Þm
4p ðn + mÞ! expansion of the tidal potential. We can write [6] as
and Pmn is the associated Legendre polynomial of degree n and Vtid X 1
4p
order m (Table 1). ¼ Kn xn + 1 ½Yn0 ðy0 , f0 ÞYn0 ðy, fÞ
g n¼2
2n +1
As is conventional, we express the tidal potential as Vtid/g,
where g is the Earth’s gravitational acceleration; this combina- X
n
+ * ðy0 , f0 ÞYnm ðy, fÞ + Ynm
Ynm * ðy0 , f0 ÞYnm ðy, fÞ
tion has the dimension of length and can easily be interpreted m¼1
as the change in elevation of the geoid or of an equilibrium
X
1
4p
surface such as an ideal ocean (hence its name, the equilibrium ¼ Kn xn + 1 ½Yn0 ðy0 , f0 ÞYn0 ðy, fÞ
potential.) Part of the convention is to take g to have its value on n¼2
2n + 1
the Earth’s equatorial radius aeq; if we hold r fixed at that radius Xn
+ * ðy0 , f0 ÞYnm ðy, fÞ
2ℜ½Ynm
in [5], we get m¼1

Vtid Mext X
1
4p aeq n + 1 X
n
Now, define complex (and time-varying) coefficients Tnm(t) ¼
¼ aeq * ðy0 , f0 ÞYnm ðy, fÞ
Ynm
g ME n¼2 2n + 1 R m¼n am m
n (t) + ibn (t) such that
X1
4p n + 1 X
n " #
¼ Kn x * ðy0 , f0 ÞYnm ðy, fÞ
Ynm [6] Vtid X 1 X n

n¼2
2n + 1 m¼n
¼ℜ * ðt ÞYnm ðy, fÞ
Tnm [7]
g n¼2 m¼0

where the constant K includes all the physical quantities: XX


n¼1 n

  ¼ Nnm Pnm ðcos yÞ am


n ðt Þcos mf + bn ðt Þsin mf
m
[8]
Mext aeq n + 1 n¼2 m¼0
Kn ¼ aeq
ME R
Then, the Tnm coefficients are, for m equal to 0,
where ME is the mass of the Earth and R is the mean distance of
 1
the body; the quantity x ¼ R=R expresses the normalized vari- 4p 2
Tn0 ¼ Kn xn + 1 Pn0 ðcos y0 Þ [9]
ation in distance. For the Moon, K2 is 0.35837 m, and for the 2n + 1
Sun, 0.16458 m.
and, for m not equal to 0,
In both [5] and [6], we have been thinking of y and f as
giving the location of a particular place of observation; but if 8p 0
Tnm ¼ ð1Þm Kn xn + 1 Nnm Pnm ðy0 Þeif [10]
we consider them to be variables, the Ynm(y, f) describes the 2n + 1
from which we can find the real-valued, time-varying quanti-
Table 1 Associated Legendre functions ties am m
n (t) and bn (t), which we will use later in computing the
response of the Earth.
1  1 
P20 ðyÞ ¼ 3cos 2 y  1 P30 ðyÞ ¼ 5cos 3 y  3cosy
2 2
3 
P12(y) ¼ 3 sin y cos y P31 ðyÞ ¼ 5cos 2 y  1
2 3.06.2.2 Computing the Tides: Ephemeris-Based
P22(y) ¼ 3 sin2y P23(y) ¼ 15 sin2y cos y Equations [7]–[9] suggest a straightforward way to compute
P33(y) ¼ 15 3
sin y the tidal potential (and, as we will see, other theoretical tides).
First, use a description of the location of the Moon and Sun in
Earth Tides 155

celestial coordinates (an ephemeris); other planets can be expansion of the tidal potential. In this, we express the Tnm as a
included if we wish. Then, convert this celestial location to sum of sinusoids, whose frequencies are related to combina-
the geographic coordinates y0 and f0 of the subbody point tions of astronomical frequencies and whose amplitudes are
and the distance R, using standard transformations (Pétit and determined from the expressions in the ephemerides for R, y0 ,
Luzum, 2010, Chapter 1.05). Finally, use eqns [9] and [10] to and f0 . In such an expansion, we write the complex Tnm’s as
get Tnm(t). Once we have the Tnm, we can combine these with
X
Knm
the spatial factors in eqn [7] to get Vtid/g, either for a specific Tnm ðt Þ ¼ Aknm eið2pfknm t + ’knm Þ [11]
location or as a distribution over the whole Earth. Also, having k¼1
done the astronomy to produce the Tnm, we can vary the spatial
factors to find, not just the potential, but other observables, where, for each degree and order, we sum Knm sinusoids with
including tilt and strain. specified real amplitudes, frequencies, and phases A, f, and ’.
This direct computation has the advantage, compared with The individual sinusoids are called tidal harmonics (not the
the harmonic methods (discussed in the succeeding text), of same as the spherical harmonics of 3.1.)
being limited in accuracy only by the quality of the ephemeris. This method has the conceptual advantage of decoupling
If we take derivatives of eqn [10] with respect to R, y0 and f0 , we the tidal potential from the details of astronomy and the
find that relative errors of 104 in Vtid/g would be caused by practical advantage that a table of harmonic amplitudes and
errors of 7  105 rad (1400 ) in y0 and f0 and 3  105 in x. (We frequencies, once produced, is valid over a long time. Such an
pick this level of error because it usually exceeds the accuracy expansion also implicitly puts the description into the fre-
with which the tides can be measured, either because of noise quency domain, something as useful here as it is in other
or because of instrument calibration.) The errors in the angular parts of geophysics. We can use the same frequencies for any
quantities correspond to errors of about 400 m in the location tidal phenomenon, provided that it comes from a linear
of the subbody point, so our model of Earth rotation, and our response to the driving potential – which is essentially true
station location, needs to be good to this level – which requires for the Earth tides. So, while this expansion was first used for
1 s accuracy in the timing of the data. ocean tides (for which it remains the standard), it works just as
Two types of ephemerides are available: analytic ephemer- well for any type of Earth tides.
ides, which provide a closed-form algebraic description of the To get the flavor of this approach, and also introduce some
motion of the body, and the much more precise numerical terminology, we consider tides for a very simple situation: a
ephemerides, computed from numerical integration of the body moving at constant speed b in a circular orbit, the orbital
equations of motion, with parameters chosen to best fit astro- plane being inclined at an angle I to the Earth’s equator. The
nomical data. While numerical ephemerides are more accurate, angular distance from the ascending node (where the orbit
they are less convenient for most users, being available only as plane and the equatorial plane intersect) is bt. The rotation of
tables; analytic ephemerides are or can be made available as the Earth, at rate O, causes the terrestrial longitude of the
computer code. ascending node to be Ot; since the ascending node is fixed in
The first tidal-computation program based directly on an space, O corresponds to one revolution per sidereal day. We
ephemeris was that of Longman (1959), still in use for making further assume that at t ¼ 0 the body is at the ascending node
rough tidal corrections for gravity surveys. Longman’s pro- and longitude 0 is under it. Finally, we take just the real part of
gram, like some others, computed accelerations directly, mak- eqn [6] and do not worry about signs.
ing it less obvious that an ephemeris-based approach could be With these simplifications we consider first the diurnal
applied to all tidal computations. Munk and Cartwright (1966) degree-2 tides (n ¼ 2, m ¼ 1). After some tedious spherical trig-
applied the ephemeris method to find the tidal potential. onometry and algebra, we find that
 
Subsequent programs (Broucke et al., 1972; Harrison, 1971; 6p 1
V=g ¼ K2 sin Icos Isin Ot + sin Ið1 + cos IÞsin ðO  2bÞt
Merriam, 1992; Tamura, 1982) have used ever more precise 5 2
ephemerides, with the lunar part derived from subsets of 
1
Brown’s lunar theory. + sin Ið1  cos IÞsin ðO + 2bÞt
2
Numerical ephemerides have been used primarily to pro-
duce reference time series, rather than for general-purpose pro- This shows that the harmonic decomposition includes
grams, although the current IERS standards use such a method three harmonics, with arguments (of time) O, O  2b, and
for computing tidal potentials and displacements (with correc- O + 2b; their amplitudes depend on I, the inclination of the
tions described in Section 3.06.3.2.2). Most precise calculations orbital plane. If I were zero, there would be no diurnal tides at
(e.g., Hartmann and Wenzel, 1995) have relied on the numer- all. For our simple model, a reasonable value of I is 23.44 , the
ical ephemerides produced by JPL (Standish et al., 1992). The inclination of the Sun’s orbital plane and the mean inclination
resulting tidal series forms the basis for a harmonic expansion of of the Moon’s. These numbers produce the harmonics given in
the tidal potential, a standard method to which we now turn. Table 2, in which the frequencies are given in cycles per solar
day (cpd). Both the Moon and Sun produce a harmonic at 1
cycle per sidereal day. For the Moon, b corresponds to a period
3.06.2.3 Computing the Tides (II): Harmonic
of 27.32 days (the tropical month) and for the Sun 365.242
Decompositions
days (1 year), so the other harmonics are at 2 cycles per
Since the work of Thomson and Darwin in the 1870s and 1880s, month, or 2 cycles per year, from this. Note that there is
the commonest method of analyzing and predicting the tides, not a harmonic at 1 cycle per lunar (or solar) day – this is
and of expressing tidal behavior, has been through a harmonic not unexpected given the degree-2 nature of the tidal potential.
156 Earth Tides

Table 2 Tidal constituents (simple model)

Moon Sun
Argument
Freq. (cpd) Period Amp. (m) Freq. (cpd) Period Amp. (m)

Long-period tides
0.000000 1 0.217 0.000000 1 0.100
2b 0.073202 13d15h51m36s 0.066 0.005476 182d14h45m40s 0.030
Diurnal tides
O 1.002738 23h56m04.1s 0.254 1.002738 23h56m04.1s 0.117
O  2b 0.929536 25h49m09.6s 0.265 0.997262 24h03m57.2s 0.122
O + 2b 1.075940 22h18m21.9s 0.011 1.008214 23h48m16.1s 0.005
Semidiurnal tides
2O 2.005476 11h58m02.0s 0.055 2.005476 11h58m02.0s 0.025
2O  2b 1.932274 12h25m14.2s 0.640 2.000000 12h00m00.0s 0.294

It is convenient to have a shorthand way of referring to 1 1


cos o0 t + Acos ½ðo0 + om Þt  + Acos ½ðo0  om Þt 
these harmonics; unfortunately, the standard naming system, 2 2
now totally entrenched, was begun by Thomson for a few tides so we can retain a development purely in terms of sinusoids,
and then extended by Darwin in a somewhat ad hoc manner. but with three harmonics, one at the central frequency and two
The result is a series of conventional names that simply have to smaller ones (called satellite harmonics) separated from it by
be learned as is (though only the ones for the largest tides are one cycle in 18.61 years.
really important). For the Moon, the three harmonics have the An accurate ephemeris also includes the ellipticity of the
Darwin symbols K1, O1, and OO1; for the Sun, they are K1 orbits and all the periodic variations in I and in other orbital
(again, since this has the same frequency for any body), P1, parameters, which lead to many more harmonics; for a
and f1.For the semidiurnal (m ¼ 2 case), the result is detailed description, see Bartels (1957/1985). The first full
  harmonic expansion, including satellite harmonics, was by
24p
V=g ¼ K2 1  cos 2 I cos 2Ot Doodson (1921), done algebraically from an analytic ephem-
5
 eris; the result had 378 harmonics. Doodson needed a nomen-
1 1
+ ð1 + cos IÞ2 cos ð2O  2bÞt + ð1  cos IÞ2 cos ð2O + 2bÞt clature for these tides and introduced one that relies on the fact
2 2 that, as our simple ephemeris suggests, the frequency of any
harmonic is the sum of multiples of a few basic frequencies.
again giving have three harmonics, though for I equal to
For any (n, m), we can write the argument of the exponent in
23.44 , the third one is very small. Ignoring the last term, we
eqn [11] as
have two harmonics, also listed in Table 2. The Darwin symbol !
for the first argument is K2; again, this frequency is the same for X6 X
6

the Sun and the Moon, so these combine to make a lunisolar 2p fk t + fk ¼ Dlk 2p fl t + Dlk ’l
l¼1 l¼1
tide. The second argument gives the largest tides: for the Moon,
M2 (for the Moon) or S2 (for the Sun), at precisely 2 cycles per where the fl’s are the frequencies corresponding to various astro-
lunar (or solar) day, respectively. nomical periods and the ’l’s are the phases of these at some
Finally, the m ¼ 0, or long-period, case has suitable epoch. Table 3 gives a list. (Recent tabulations extend
p
this notation with up to five more arguments to describe the
V=g ¼ K2 1:5sin 2 I  1  1:5sin 2 Icos 2bt motions of the planets. As the tides from these are small, we
5
ignore them here.) The l ¼ 1 frequency is chosen to be one cycle
which gives one harmonic at zero frequency (the so-called per lunar day exactly, so for the tide, the Dl’s are 2,0,0,0,0,0. This
permanent tide) and another with an argument of 2b, making makes the solar tide, S2, have the Dl’s 2,2,2,0,0,0. In practice, all
tides with frequencies of two cycles per month (Mf, the fort- but the smallest tides have Dlk ranging from 5 to 5 for l > 1.
nightly tide, from the Moon) and two cycles per year (Ssa, the Doodson therefore added 5 to these numbers to make a com-
semiannual tide, from the Sun). pact code, so that M2 becomes 255555 and S2273555. This is
This simple model demonstrates one other important attri- called the Doodson number; the numbers without 5 added are
bute of the tides, namely, that they depend on the orbital sometimes called Cartwright–Tayler codes (Table 4).
inclination I. For the Sun, this inclination (called the obliquity Figure 3 shows the full spectrum of amplitude coefficients,
of the ecliptic) changes gradually by only 1.3  104 /year, but from the expansion of Hartmann and Wenzel (1995). The top
for the Moon, it varies by 5.13 from the mean, with a period panel shows all harmonics on a linear scale, making it clear
of 18.61 years. This produces a variation in amplitude in all the that only a few are large, and the separation into different
lunar tides, which is called the nodal modulation. The simple species around 0, 1, and 2 cycles per day: These are referred
expressions show that the resulting variation is 18% for O1 to as the long-period, diurnal, and semidiurnal tidal bands. The
and 3% for M2, though the M2 change can be larger in two lower panels show an expanded view of the diurnal and
amplitude. Such a modulated sinusoid can be written as semidiurnal bands, using a log scale of amplitude to include
cos o0t(1 + A cos omt), with o0  om; this is equal to the smaller harmonics. What is apparent from these is that
Earth Tides 157

Table 3 Fundamental tidal frequencies

l Symbol Frequency (cycles per day) Period What


h m s
1 t 0.9661368 24 50 28.3 Lunar day
2 s 0.0366011 27.3216d Moon’s longitude: tropical month
3 h 0.0027379 365.2422d Sun’s longitude: solar year
4 p 0.0003095 8.847y Lunar perigee
5 N0 0.0001471 18.613y Lunar node
6 ps 0.0000001 20 941y Solar perigee

‘Longitude’ refers to celestial longitude, measured along the ecliptic.

Table 4 Largest tidal harmonics, for n ¼ 2, sorted by size for each Tayler (1971) revisited the subject. Using eqn [6], they computed
species the potential from a more modern lunar ephemeris and then
applied special Fourier methods to analyze, numerically, the
Darwin
resulting series and get amplitudes for the various harmonics.
Amplitude (m) Doodson number Frequency (cpd) symbol
The result was a compendium of 505 harmonics, which (with
Long-period tides errors corrected by Cartwright and Edden, 1973) soon became
0.31459 055555 0.0000000 M0, S0 the standard under the usual name of the CTE representation. (A
0.06661 075555 0.0732022 Mf few small harmonics at the edges of each band, included by
0.03518 065455 0.0362916 Mm Doodson but omitted by Cartwright, are sometimes added to
0.03099 057555 0.0054758 Ssa make a CTED list with 524 harmonics.)
0.02793 055565 0.0001471 N More extensive computations of the tidal potential and its
0.02762 075565 0.0733493
harmonic decomposition have been driven by the very high
0.01275 085455 0.1094938 Mtm
precision available from the ephemerides and the desire for
0.00673 063655 0.0314347 MSm
0.00584 073555 0.0677264 MSf more precision for analyzing some tidal data (gravity tides
0.00529 085465 0.1096409 from superconducting gravimeters). Particular expansions are
Diurnal tides those of Bullesfeld (1985), Tamura (1987), Xi (1987),
0.36864 165555 1.0027379 K1 Hartmann and Wenzel (1995), and Roosbeek (1995). The
0.26223 145555 0.9295357 O1 latest is that of Kudryavtsev (2004), with 27 000 harmonics.
0.12199 163555 0.9972621 P1 Figure 4 shows the amplitude versus number of harmonics; to
0.05021 135655 0.8932441 Q1 get very high accuracy demands a very large number. But not
0.05003 165565 1.0028850 many are needed for a close approximation; the CTED expan-
0.04947 145545 0.9293886
sion is good to about 0.1% of the total tide.
0.02062 175455 1.0390296 J1
0.02061 155655 0.9664463 M1
0.01128 185555 1.0759401 OO1
0.00953 137455 0.8981010 r1 3.06.2.4 The Pole Tide
0.00947 135645 0.8930970
–0.00801 127555 0.8618093 s1 Both in our elementary discussion and in our mathematical
0.00741 155455 0.9658274 development of the tidal forcing, we treated the Earth’s rota-
–0.00730 165545 1.0025908 tion only as a source of motion of the subbody point. But
0.00723 185565 1.0760872 changes in this rotation also cause spatial variations in the
–0.00713 162556 0.9945243 p1 gravitational potential, and since these have the same effects
as the attraction of external bodies, they can also be regarded as
tides. The only significant one is the pole tide, which is caused
each tidal species is split into a set of bands, separated by one by changes in the direction of the Earth’s spin axis relative to a
cycle per month; these are referred to as groups: in each group, reference frame fixed in the Earth. The spin produces a centrif-
the first two digits of the Doodson number are the same. All ugal force, which depends on the angular distance between the
harmonics with the same first three digits of the Doodson spin axis and some location: so as the spin axis moves, this
number are in clusters separated by one cycle per year; these distance, and the centrifugal force, changes.
clusters are called constituents, though this name is also some- For a spin vector O, the potential at a location r is
times used for the individual harmonics. As a practical matter,
1 2 2

this is usually the finest frequency resolution attainable; on the V¼ jOj jrj  jOrj2 [12]
scale of this plot, finer frequency separations, such as the nodal 2
modulation, are visible only as a thickening of some of the We assume that the rotation vector is nearly along the 3-axis, so
lines. All this fine-scale structure poses a challenge to tidal that we have O ¼ Oðm1 x^1 + m2 x^2 + x^3 Þ, with m1 and m2 both
analysis methods (Section 3.06.5.1). much less than one; m1 and m2 describe the instantaneous
Since Doodson provided the tidal potential to more than rotation axis. If we put expression for this axis into eqn [12],
adequate accuracy for studying ocean tides, further developments and subtract V for m1 and m2 both zero, we get the potential
did not take place for the next 50 years, until Cartwright and height for the pole tide:
158 Earth Tides

Tidal potential: amplitude spectra

0.6 M2

0.5 All tidal species (degree 2)

Amplitude
0.4 P1,K1
0.3 S2
O1
0.2
0.1 Mf K2
0.0
0.0 0.5 1.0 1.5 2.0

100
O1 K1
Diurnal tides
P1
10−1 Q1 M1
2Q1 s1 r1 J1 OO1
10−2 X1 p1
log Amplitude

t1 q1 SO1 n1
10−3

10−4

10−5

10−6
0.80 0.85 0.90 0.95 1.00 1.05 1.10

100 M2 S2
Semidiurnal tides N2
10−1 K2
2N2 m2 n2 L2
T2
log Amplitude

−2
10 e2 λ2 h2
3N2 a2 b2
10−3 g2 δ2 2T2 x2

10−4

10−5

10−6
1.80 1.85 1.90 1.95 2.00 2.05
Frequency (cycles per day)

Figure 3 The spectrum of the tidal potential. Since all variations are purely sinusoidal, the spectrum is given by the amplitudes of the tidal harmonics,
taken from Hartmann and Wenzel (1995), though normalized according to the convention of Cartwright and Tayler (1971). The Darwin symbols are
shown for the larger harmonics (top) and all named diurnal and semidiurnal harmonics, except for a few that are shown only in Figure 5.

Amplitude distribution of tidal harmonics


Diurnal harmonics Semidiurnal harmonics
100 100
Amp (m, CT normalization)

LS, n = 3
10−1 LS, n = 2
10−1

10−2 CTE cutoff 10−2 CTE cutoff


n=3 n=3
10−3 10−3

10−4 10−4

10−5 Flatten 10−5 Planet

10−6 10−6
Planet
10−7 10−7
100 101 102 103 100 101 102 103
Harmonic number Harmonic number
Figure 4 Distribution of harmonic amplitudes for the catalog of Hartmann and Wenzel (1995), normalized according to Cartwright and Tayler (1971).
The line ‘LS, n ¼ 2’ refers to lunisolar harmonics of degree two; those with large dots have Darwin symbols associated with them. Constituents
of degree 3, from other planets, and from Earth flattening are shown as separate distributions. The horizontal line shows the approximate cutoff level
of the Cartwright and Tayler (1971) list.
Earth Tides 159

V O2 ground temperature fairly directly. Ground-temperature


¼  ½2ðm1 r1 r3 + m2 r2 r3 Þ
g 2g changes in turn cause thermoelastic deformations with these
O2 a2 periods (Berger, 1975; Harrison and Herbst, 1977; Müller,
¼ sin ycos yðm1 cos f + m2 sin fÞ 1977). Air pressure changes driven by solar heating, usually
g
known as atmospheric tides (Chapman and Lindzen, 1970),
This is a degree-2 change in the potential, of the same form as load the Earth enough to cause deformations and changes in
for the diurnal tides. However, the periods involved are very gravity. The availability of better models of some atmospheric
different, since the largest pole tides come from the largest tides (Ray, 2001; Ray and Ponte, 2003; Ray and Poulose, 2005)
polar motions, at periods of 14 months (the Chandler wob- and their inclusion in ocean-tide models (Ray and Egbert,
ble) and 1 year. The maximum range of potential height is a 2004) has allowed their effects to be compared with gravity
few cm, small but not negligible; pole-tide signals have been observations (Boy et al., 2006b).
observed in very precise gravity measurements and also, not That some of these thermal tidal harmonics coincide with
strongly, in sea-level data (Desai, 2002; Haubrich and Munk, harmonics in the tidal potential poses a real difficulty for
1959; Miller and Wunsch, 1973; Trupin and Wahr, 1990). precise analysis of the latter, since if all we have is the sum of
Like the polar motion, the actual pole tide can only be found two harmonics with the same frequency, it is impossible to tell
from observations of the Earth’s orientation. These do not how much each part contributes. The only way to estimate this
actually give the position of V, but rather that of what is is by making additional assumptions, about how the response
called the celestial intermediate pole (CIP), whose coordi- to the potential or to temperature behaves at other frequencies.
nates are denoted by xp and yp (Pétit and Luzum, 2010). The Even then, it is likely that these estimates will have large sys-
position of V is given by tematic errors; this is why when estimating tidal responses, the
y_p x_p large K1 tide is less used than the smaller O1 tide.
m1 ¼ x p  x p  m2 ¼ yp + yp  [13]
O O
where x, y are mean values of the pole position (Pétit and
3.06.3 Tidal Response of the Solid Earth
Luzum, 2010, Section 7.1.4) and the third term of each expres-
sion corrects for the difference between the CIP and V
Having described the tidal forces, we next turn to the response
(Brzezı́nski and Capitaine, 1993; Gross, 1992); at periods of
of the solid Earth – which, as is conventional, we assume to be
a year and longer, this term is much smaller than the others.
oceanless, putting in the effect of the ocean tides at a later step.
We start with the usual approximation of a spherical Earth in
order to introduce a number of concepts, many of them ade-
3.06.2.5 Radiational Tides quate for all but the most precise modeling of the tides. We
The harmonic treatment used for the gravitational tides can then describe what effects a better approximation has, in
also be applied to the various phenomena associated with solar enough detail to enable computation; the relevant theory is
heating. As a first approximation, the input radiation can be beyond the scope of this treatment, though outlined in the
taken to be proportional to the cosine of the Sun’s elevation article by Dehant and Mathews.
during the day and zero at night; Munk and Cartwright (1966)
called this the radiational tide. The day–night asymmetry pro-
3.06.3.1 Tidal Response of an SNREI Earth
duces harmonics of degrees 1 and 2; these have been tabulated
by Cartwright and Tayler (1971) and are shown in Figure 5 as To a good approximation, we can model the tidal response of
crosses (for both degrees), along with the tidal potential har- an oceanless Earth by assuming an SNREI Earth model: that is,
monics shown as in Figure 3. The unit for the radiational tides one that is spherical, non-rotating, elastic, and isotropic. As in
is S, the solar constant: 1366 W m2. normal-mode seismology (from which this acronym comes),
Changes in solar irradiation drive changes in air tempera- this means that we assume the elastic properties to vary only
ture and pressure in very complicated ways but changes in with depth. In addition to these restrictions on the Earth

100 100
P1 K1
Log amplitude (m or S)

S2
10 –1 10–1 K2
T2
p1 S1
10–2 f1 10–2
y1 R2

10–3 10–3 2T2

10–4 10–4
0.990 1.000 1.010 1.990 2.000 2.010
Frequency (cpd) Frequency (cpd)
Figure 5 Radiational tides. The crosses show the amplitudes of the radiational tidal harmonics (degree 1 and 2) from Cartwright and Tayler (1971); the
amplitudes are for the solar constant S being taken to be 1.0. The lines are harmonics of the gravitational tides, as in Figure 3.
160 Earth Tides

model, we add one more about the tidal forcing: that it has a with the different combinations of a and b with the f depen-
much longer period than any normal modes of oscillation of dence showing that this tilt is phase-shifted relative to the
the Earth so that we can use a quasistatic theory, taking the potential, by 90∘ if we use a harmonic decomposition.
response to be an equilibrium one. Since the longest-period Tidal variations in gravity were for a long time the com-
normal modes for such an Earth have periods of less than an monest type of Earth-tide data. For a spherical Earth, the tidal
hour, this is a good approximation. potential is, for degree n,
It is simple to describe the response of an SNREI Earth to the  r n an + 1
tidal potential ( Jeffreys, 1976). Because of symmetry, only the Vn + kn Vn
a r
degree n is relevant. If the potential height at a point on the
surface is V(y, f)/g, the distortion of the Earth from tidal forces where the first term is the potential caused by the tidal forcing
produces an additional gravitational potential knV(y, f), a ver- (and for which we have absorbed all nonradial dependence into
tical (i.e., radial) displacement hnV(y, f)/g, and a horizontal Vn) and the second is the additional potential induced by the
displacement ln(r1V(y, f)/g), where r1 is the horizontal Earth’s deformation. The corresponding change in local gravita-
gradient operator on the sphere. So defined, kn, hn, and ln are tional acceleration is the radial derivative of the potential:
dimensionless; they are called Love numbers, after A. E. H. Love    an + 1   
@ r n n kn
(though the parameter ln was actually introduced by T. Shida). Vn + kn ¼ Vn  ðn + 1Þ [16]
@r a r r¼a a a
For a standard modern Earth model (PREM), h2 ¼ 0.6032,
k2 ¼ 0.2980, and l2 ¼ 0.0839. For comparison, the values for In addition to this change in gravity from the change in the
the much older Gutenberg–Bullen Earth model are 0.6114, potential, there is a change from the gravimeter being moved
0.3040, and 0.0832 – not very different. In this section, we up by an amount hnVn/g. The change in gravity is this displace-
adopt values for a and g that correspond to a spherical Earth: ment times the gradient of g, 2g/a, plus the displacement times
6.3707  106 m and 9.821 m s(2), respectively. o2, where o is the radian frequency of the tidal motion – that
is, the inertial acceleration. (We adopt the Earth-tide conven-
tion that a decrease in g is positive.) If we ignore this last part
(which is at most 1.5% of the gravity-gradient part), we get a
3.06.3.1.1 Some combinations of Love numbers (I): gravity total change of
and tilt        
Until there were data from space geodesy, neither the potential n n+1 2hn nVn n+1 2
Vn  kn + ¼ 1 k n + hn [17]
nor the displacements could be measured; what could be were a a a a n n
ocean tides, tilt, changes in gravity, and local deformation
The nVn/a term is the change in g that would be observed on a
(strain), each of which possessed its own expression in terms
rigid Earth (with h and k zero); the term which this is multi-
of Love numbers – which we now derive. Since the first three of
plied by, namely,
these would exist even on a rigid Earth, it is common to
 
describe them using the ratio between what they are on an 2hn n+1
dn ¼ 1 +  kn
elastic Earth (or on the real Earth) and what they would be on a n n
rigid Earth.
is called the gravimetric factor. For the PREM model,
The simplest case is that of the effective tide-raising poten-
d2 ¼ 1.1562: the gravity tides are only about 16% larger than
tial: that is, the one relevant to the ocean tide. The total tide-
they would be on a completely rigid Earth, so that most of the
raising potential height is (1 + kn)V/g, but the solid Earth (on
tidal gravity signal shows only that the Moon and Sun exist, but
which a tide gauge sits) goes up by hnV/g, so the effective tide-
does not provide any information about the Earth. The expres-
raising potential is (1 + kn  hn)V/g, sometimes written as gnV/g,
sion for the gravity tide is of course very similar to eqn [8]:
gn being called the diminishing factor. For the PREM model,
g2 ¼ 0.6948. Since tilt is just the change in slope of an equipo- X X

g n¼1 n
tential surface, again relative to the deforming solid Earth, it dg ¼ dn Nm P m ðcos yÞ am
n ðt Þcos mf + bn ðt Þsin mf
m
[18]
a n¼2 m¼0 n n
scales in the same way that the potential does: the tilt on an
SNREI Earth is gn times the tilt on a rigid Earth. The NS tilt is,
using eqn [8] and expressions for the derivatives of Legendre
functions, 3.06.3.1.2 Combinations of Love numbers (II):
X
n¼1 X
n Displacement and strain tides
gn @V 1
lN ¼ ¼ g N m ncos yPnm ðcos yÞ For a tidal potential of degree n, the displacements at the
ga @y asin y n¼2 n m¼0 n
surface of the Earth (r ¼ a) will be, by the definitions of the

ðn + mÞPn1
m
ðcos yÞ am
n ðt Þcos mf + bn ðt Þsin mf
m
[14] Love numbers ln and hn,

where the sign is chosen such that a positive tilt to the north hn V ln @V ln @V
ur ¼ uy ¼ uf ¼ [19]
would cause a plumb line to move in that direction. The east g g @y gsin y @f
tilt is
in spherical coordinates. Comparing these with [18], [14], and
gn @V 1 n¼1 X X n [15], we see that the vertical displacement is exactly propor-
lE ¼ ¼ gn mNnm Pnm ðcos yÞ tional to changes in gravity, with the scaling constant being
gasin y @f asin y n¼2 m¼0
m
hna/2dng ¼ 1.692  105 s2, and that the horizontal displace-
bn ðt Þcos mf  am
n ðt Þsin mf [15] ments are exactly proportional to tilts, with the scaling
Earth Tides 161

constant being lna/gn ¼ 7.692  105 m; we can thus use rotation and slightly inelastic behavior. Such models have
eqns [18], [14], and [15], suitably scaled, to find tidal three consequences for the tidal response:
displacements.
1. The ellipticity of the core–mantle boundary and the rota-
Taking the derivatives of [19], we find the tensor compo-
tion of the Earth combine to produce a free oscillation in
nents of the surface strain are
which the fluid core (restrained by pressure forces) and
  solid mantle precess around each other. This is known as
1 @2V
eyy ¼ hn V + ln 2 the nearly diurnal free wobble (NDFW) or free core
ga @ y
  nutation. Its frequency falls within the band of the diurnal
1 @V ln @ 2 V
eff ¼ hn V + ln cot y + tides, which causes a resonant response in the Love num-
ga @y sin y @ 2 f bers near one cycle per day. The diurnal tides also cause
 2 
ln @ V @V changes in the direction of the Earth’s spin axis (the astro-
eyf ¼  cot y
gasin y @y@f @f nomical precessions and nutations), and the NDFW affects
these as well.
We again use [8] for the tidal potential and get the following 2. Ellipticity and rotation couple the response to forcing of
expressions that give the formulas for the three components of degree n to spherical harmonics of other degrees and sphe-
surface strain for a particular n and m; to compute the total roidal to toroidal modes of deformation. As a result, the
strain, these should be summed over all n 2 and all m from Love numbers become slightly latitude-dependent, and
0 to n (though in practice, the strain tides with n > 3 or m ¼ 0 additional terms appear for horizontal displacement.
are unobservable): 3. The imperfect elasticity of the mantle (finite Q) modifies
Nnm the Love numbers in two ways: they become complex, with
eyy ¼ hn sin 2 y + ln n2 cos 2 y  n Pnm ðcos yÞ  2ln ðn  1Þ small imaginary parts; and they become weakly frequency-
asin 2 y
dependent because of anelastic dispersion.
ðn + mÞcos yPn1
m
ðcos yÞ + ln ðn + mÞðn + m  1ÞPn2
m
ðcosyÞ
m
The full theory for these effects, especially the first two, is
an ðt Þcos mf + bn ðt Þsin mf
m
quite complicated. Love (1911) provided some theory for the
Nnm effects of ellipticity and rotation, and Jeffreys and Vincente
eff ¼ hn sin 2 y + ln ncos 2 y  m2 Pnm ðcos yÞ
asin 2 y (1957) and Molodenskii (1961) for the NDFW, but the mod-

ern approach for these theories was first described by Wahr
ln ðn + mÞcos yPn1m
ðcos yÞ am
n ðt Þcos mf + bn ðt Þsin mf
m
(1981a,b); a simplified version is given by Neuberg et al.
mNnm ln
(1987) and Zürn (1997b). For more recent developments, see
eyf ¼ ðn  1Þcos yPnm ðcos yÞ  ðn + mÞPn1
m
ðcos yÞ 
asin 2 y the article by Dehant and Mathews in this volume and
m
Mathews et al. (1995a,b, 1997), Wang (1997), Dehant et al.
bn ðt Þcos mf  am
n ðt Þsin mf
(1999), Mathews et al. (2002), and Mathews and Guo (2005).
Note that the combination of the longitude factors with am n (t)
To obtain the full accuracy of these theories, particularly for
and bm n (t) means that eyy and eff are in phase with the poten-
the NDFW correction, requires the use of tabulated values of
tial, while eyf is not. the Love numbers for specific harmonics, but analytic approx-
One consequence of these expressions is that, for n equal to imations are also available. The next three sections outline
two, and m equal to either one or two, the areal strain, (1/2) these, using values from the IERS standards (Pétit and
(eyy + eff), is equal to V(h2  3l2)/ga: the areal strain, vertical Luzum, 2010) for the Love numbers and from Dehant et al.
displacement, potential, and gravity are all scaled versions of (1999) for the gravimetric factor.
each other. Close to the surface, the free-surface condition
means that deformation is nearly that of plane stress, so verti-
cal and volume strains are also proportional to areal strain and 3.06.3.2.1 NDFW resonance
are also just a scaled version of the potential. The most important result to come out of the combination of
If we combine these expressions for spatial variation improved theoretical development and observations has been
with the known amplitudes of the tidal forces, we can see that the frequency of the NDFW, fNFDW, is significantly differ-
how the rms amplitude of the body tides varies with lati- ent from that originally predicted. This frequency is usually
tude and direction (Figure 6). There are some complications expressed by the period of the NDFW in a space-fixed frame,
in the latitude dependence; for example, the EW semidiur- TNDFW ¼ (fNFDW  1.0027379)1, where fNFDW is the resonance
nal strain tides go to zero at 52.4 latitude. Note that while frequency for an Earth-fixed frame, in cycles per solar day.
the tilt tides are larger than strain tides, most of this signal TNDFW is in large part determined by the ellipticity of the
is from the direct attractions of the sun and moon; the core–mantle boundary (CMB), for a hydrostatic Earth theory
purely deformational part of the tilt is about the same size gives TNDFW ¼ 460 d. Both tidal and nutation data (Amoruso
as the strain. et al., 2012; Herring et al., 2002; Rosat and Lambert, 2009;
Zürn, 1997a) give values of about 430 d. This difference
implies that the ellipticity of the CMB departs from a hydro-
static value by about 5%, the equivalent of a 500 m variation
3.06.3.2 Response of a Rotating Earth
difference in radius. This departure is generally thought to
We now turn to models for tides on an oceanless and isotropic reflect distortion of the core–mantle boundary by mantle con-
Earth, still with properties that depend on depth only but add vection (e.g., Steinberger and Holme, 2008).
162 Earth Tides

RMS diurnal tides RMS semidiurnal tides

Vertical displacement (mm)


700 120 700 120
600 100 600 100

Gravity (nm s–2)


500 80 500 80
400 400
60 60
300 300
40 40
200 200
100 20 100 20
0 0 0 0

Horizontal displacement (mm)


45 45
40 30 40 30
35 35
25 25
30 30
Tilt (nrad)

25 20 25 20
20 15 20 15
15 15
10 10
10 10
5 5 5 5
0 0 0 0

14 14
Solid NS, dashed EW
Linear strain (10–9)

12 12
10 10
8 8
6 6
4 4
2 2
0 0
80 60 40 20 0 80 60 40 20 0
Latitude Latitude
Figure 6 RMS tides. The left plots show the rms tides in the diurnal band, and the right plots the rms in the semidiurnal band. The uppermost
frame shows gravity and vertical displacement (with scales for each), which have the same latitude dependence; the next horizontal displacement and
tilt; and the bottom linear strain. In all plots but the top, dashed is for EW measurements, solid for NS.

Table 5 Coefficients (real and imaginary parts) used in eqn [20] to find the frequency dependence of the Love numbers (including corrections for
ellipticity) in the diurnal tidal band

ℜ(Sz) ℑ(Sz) ℜ(S1) ℑ(S1) ℜ(S2) ℑ(S2)

d0 1.15802 0.0 2.871  103 0.0 4.732  105 0.0


k(0) 0.29954 1.412  103 7.811  104 3.721  105 9.121  105 2.971  106
h(0) 0.60671 2.420  103 1.582  103 7.651  105 1.810  104 6.309  106
l(0) 0.08496 7.395  104 2.217  104 9.672  106 5.486  106 2.998  107
d+ 1.270  104 0.0 2.364  105 0.0 1.564  106 0.0
k+ 8.040  104 2.370  106 2.090  106 1.030  107 1.820  107 6.500  109
h(2) 6.150  104 1.220  105 1.604  106 1.163  107 2.016  107 2.798  109
l(2) 1.933  104 3.819  106 5.047  107 1.643  108 6.664  109 5.090  1010
l(1) 1.210  103 1.360  107 3.169  106 1.665  107 2.727  107 8.603  109
lP 2.210  104 4.740  108 5.776  107 3.038  108 1.284  107 3.790  109

The resonant behavior of the Love numbers from the where L(f ) is the frequency-dependent Love number (of what-
NDFW is confined to the diurnal band; within that band, it ever type) for frequency f in cycles per solar day. The expansion
can be approximated by an expansion in terms of the resonant in use for the IERS standards includes three resonances: the
frequencies: Chandler wobble, the NDFW, and the free inner core nutation
(FICN); to a good approximation (better than 1%), the last can
X2
Sk be ignored. Table 5 gives the values of the S’s according to the
Lð f Þ ¼ S z + [20] IERS standards (and to Dehant et al., 1999 for the gravimetric
k¼1
f  fk
Earth Tides 163

factors), scaled for f in cycles per solar day; in these units, the the result will be a ratio that depends on latitude: we say that
resonance frequencies are the Love number has become latitude-dependent.
Such a generalization raises issues of normalization; unlike
f1 ¼ 2:60812  103  1:365  104 i the spherical case, the potential and the response may be
f2 ¼ 1:0050624 + 2:5  105 i evaluated on different surfaces. This has been a source of
some confusion. The normalization of Mathews et al.
and the FICN frequency (not used in eqn [20]) is 1.00176124 (1995b) is the one generally used for displacements: it uses
+ 7.82  104i. Dehant et al. (1999) used purely real-valued the response in displacement on the surface of the ellipsoid but
frequencies, with f1 ¼  2.492  103 and f2 ¼ 1.0050623, as takes these to be relative to the potential evaluated on a sphere
well as purely real values of the S’s. with the Earth’s equatorial radius. Because of the inclusion of
Figure 7 shows the NDFW resonance, for a signal (areal such effects as the inertial and Coriolis forces, the gravimetric
strain) relatively sensitive to it. Unfortunately, the tidal har- factor is no longer the combination of potential and displace-
monics do not sample the resonance very well; and because the ment Love numbers, but an independent ratio, defined as the
largest effect is for the small c1 harmonic, which is also affected ratio of changes in gravity on the ellipsoid to the direct attrac-
by radiational tides, the best estimates of fNDFW have come tion at the same point. Both quantities are evaluated along the
from nutation rather than from tidal data. normal to the ellipsoid, as a good approximation to the local
One consequence of the NDFW resonance is that we cannot vertical. Wahr (1981a) used the radius vector instead, produc-
use equations of the form [18] to compute the theoretical ing a much larger apparent latitude effect.
diurnal tides, since the factor for them varies with frequency. The standard expression for the gravimetric factor is given
If we construct the am m
n (t) and bn (t) using [11], it is easy to adjust by Dehant et al. (1999):
the harmonic amplitudes and phases appropriately. Alterna-
tively, if we find am m
n (t) and bn (t) using an ephemeris, we can Ynm+ 2 Ym
dðyÞ ¼ d0 + d + + d n2 [21]
compute the diurnal tides assuming a frequency-independent Ynm Ynm
factor and then apply corrections for the few harmonics that
n ’s, d ¼ 0 except for m
n  2, so for
By the definition of the Ym
are both large and affected by the resonance; Mathews et al.
the n ¼ 2 tides, we have for the diurnal tides
(1997) and Pétit and Luzum (2010) described such a proce-
dure for displacements and for the induced potential. pffiffiffi
3
dðyÞ ¼ d0 + d + pffiffiffi 7cos 2 y  3
2 2
3.06.3.2.2 Coupling to other modes
and for the semidiurnal tides,
The other effect of rotation and ellipticity is to couple the
spheroidal deformation of degree n, driven by the potential, pffiffiffi
3
to spheroidal deformations of degree n  2 and toroidal defor- dðyÞ ¼ d0 + d + 7cos 2 y  1
2
mations of degree n  1. Thus, the response to the degree-2 part
of the tidal potential contains a small component of degrees The expression for the induced potential (Wahr, 1981a) is
0 and 4. If we generalize the Love numbers, describing the similar, namely, that the potential is obtained by replacing
response as a ratio between the response and the potential, the term Ynm(y, f) in equation [7] with

Diurnal resonance in areal strain


2.4 y1

2.2
Tidal lines shown as crosses,
Relative response for areal strain

with size proportional to log(amp)


2.0

1.8

1.6

1.4

1.2 f1
O1 J1
1.0 P1

0.8 K1

0.6
0.95 1.00 1.05 1.10
Frequency (cpd)
Figure 7 NDFW response for the combination of Love numbers that gives the areal strain, normalized to 1 for the O1 tide. The crosses show the
locations of tidal harmonics, with size of symbol proportional to the logarithm of the amplitude of the harmonic.
164 Earth Tides

a n + 1 a n + 3 Table 6 gives the generalized Love numbers for selected tides,


e e
k0 Ynm ðy, fÞ + k + Ynm+ 2 ðy, fÞ [22]
r r including ellipticity, rotation, the NDFW, and anelasticity
(which we discuss later). The values for the diurnal tides are
which recovers the conventional Love number for k+ ¼ 0.
from exact computations rather than the resonance approxima-
The expressions for displacements are more complicated,
tions given earlier. It is evident that the latitude dependence
not only because this is a vector quantity but also because
ranges from small to extremely small, the latter applying to the
the horizontal displacements include spheroidal–toroidal
gravimetric factor, which varies by only 4  104 from the equa-
coupling, which affects neither the vertical, the potential, nor
tor to 60 N. For the displacements, Mathews et al. (1997)
the gravity. For the degree-2 tides, the effect of coupling to the
showed that the various coupling effects are at most 1 mm; the
degree-4 deformation is allowed for by defining
latitude dependence of h(y) changes the predicted displace-
1 ments by 0.4 mm out of 300.
hðyÞ ¼ hð0Þ + hð2Þ 3cos 2  1
2
1
lðyÞ ¼ lð0Þ + lð2Þ 3cos 2 y  1 [23] 3.06.3.2.3 Anelastic effects
2
All modifications to the Love numbers discussed so far apply to
Then, to get the vertical displacement, replace the term
an Earth model that is perfectly elastic. However, the materials
Ynm(y, f) in equation [7] with
of the real Earth are slightly dissipative (anelastic), with a finite
dm0 hP Q. Measurements of the Q of Earth tides were long of interest
hðyÞY2m ðy, fÞ + [24] because of their possible relevance to the problem of tidal
N20
evolution of the Earth–Moon system (Cartwright, 1999);
where the dm0 is the Kronecker delta, since hP (usually called h0 though it is now clear that almost all of the dissipation of
in the literature) only applies for m ¼ 0; the N02 factor arises from tidal energy occurs in the oceans (Ray et al., 2001), anelastic
the way in which hP was defined by Mathews et al. (1995b). effects on tides remain of interest because tidal data (along
The displacement in the y^ (north) direction is obtained by with the Chandler wobble) provide the only information on
replacing the term Ynm(y, f) in eqn [7] with Q at frequencies below about 103 Hz.
@Y2m ðy, fÞ ml1 cos y m dm1 lP if Over the seismic band (approximately 103 to 1 Hz), Q
lðyÞ  Y2 ðy, fÞ + e [25] appears to be approximately independent of frequency.
@y sin y N21
A general model for frequency dependence is
where again the lP (usually called l0 ) applies only for the par-  a
ticular value of m ¼ 1, for which it applies a correction such that f
Q ¼ Q0 [27]
there is no net rotation of the Earth. And finally, to get the f0
displacement in the f^ (east) direction, we replace the Ynm(y, f) where f0 and Q0 are reference values. In general, in a dissipative
term in eqn [7] with material, the elastic modulus m will in general also be a function
  of frequency, with m(f ) and Q(f ) connected by the Kramers–
mlðyÞ m @Y m ðy, fÞ dm0 sin ylP
i Y2 ðy, fÞ + l1 cos y 2 + [26] Kronig relation (Dahlen and Tromp, 1998, Chapter 1.07). (We
sin y @y N20
use m because this usually denotes the shear modulus; in pure
where again the lP term applies, for m ¼ 0, a no-net-rotation compression, Q is very high and the Earth can be treated as
correction. The multiplication by i means that when this is applied elastic.) This frequency dependence is usually termed anelastic
to [7] and the real part taken, the time dependence will be bm n (t) dispersion. For the frequency dependence of Q given by eqn [27],
cos mf  am m m
n (t)sin mf instead of an (t)cos mf + bn (t)sin mf. and a small, the modulus varies as

Table 6 Love numbers for an Earth that includes ellipticity, rotation, anelasticity, and the NDFW resonance

Ssa Mf O1 P1 K1 c1 M2 M3

d0 1.15884 1.15767 1.15424 1.14915 1.13489 1.26977 1.16172 1.07338


d+ 0.00013 0.00013 0.00008 0.00010 0.00057 0.00388 0.00010 0.00006
d 0.00119 0.00118
ℜ[k(0)] 0.3059 0.3017 0.2975 0.2869 0.2575 0.5262 0.3010 0.093
ℑ[k(0)] 0.0032 0.0021 0.0014 0.0007 0.0012 0.0021 0.0013
k+ 0.0009 0.0009 0.0008 0.0008 0.0007 0.0011 0.0006
ℜ[h(0)] 0.6182 0.6109 0.6028 0.5817 0.5236 1.0569 0.6078 0.2920
ℑ[h(0)] 0.0054 0.0037 0.0006 0.0006 0.0006 0.0020 0.0022
ℜ[l(0)] 0.0886 0.0864 0.0846 0.0853 0.0870 0.0710 0.0847 0.0015
ℑ[l(0)] 0.0016 0.0011 0.0006 0.0006 0.0007 0.0001 0.0006
h(2) 0.0006 0.0006 0.0006 0.0006 0.0007 0.0001 0.0006
l(2) 0.0002 0.0002 0.0002 0.0002 0.0002 0.0002 0.0002
l(1) 0.0012 0.0012 0.0011 0.0019 0.0024
lP –0.0002 –0.0002 –0.0003 –0.0001

Values for the gravimetric factors are from Dehant et al. (1999) and for the other Love numbers from the IERS standards (Pétit and Luzum, 2010).
Earth Tides 165

   a   a 
1 2 f0 f0 Egbert, 2012); using the IERS value of fm gave a better fit for a
mð f Þ ¼ m0 1 + 1 +i [28]
Q0 ap f f between 0.15 and 0.25.
so there is a slight variation in the modulus with frequency,
and the modulus becomes complex, introducing a phase lag
3.06.3.3 Lateral Heterogeneities
into its response to sinusoidal forcing. In the limit as a
approaches zero (constant Q), the real part has a logarithmic Relative to an SNREI Earth, the Earth’s ellipticity can be viewed
frequency dependence. Including a power-law variation [27] at as a lateral heterogeneity with very long wavelength. Since the
frequencies below a constant-Q seismic band, the frequency Earth is heterogeneous on scales shorter than this, a natural
dependence becomes question is what effect these lateral heterogeneities will have on
  a  the Earth tides. Because of the range of spatial scales for lateral
1 2 f0 heterogeneities, this is a very broad question, ranging from
mð f Þ ¼ m0 1 + ln +i f > fm [29]
Q0 p f differences over thousands of kilometers to variations with
      a   a   scales of less than a meter close to the measurements. The last
1 2 fm fm fm
mð f Þ ¼ m0 1 + a ln +1 +i f < fm are usually referred to as site effects and are discussed sepa-
Q0 ap f0 f f
rately in Section 3.06.6.1.
One approach (Fu and Sun, 2007; Molodensky, 1977) is to
where fm is the frequency of transition between the two Q
express the lateral heterogeneities as spherical harmonics and use
models.
perturbation theory to compute their effects. Other methods
Adding anelasticity to an Earth model has three effects on
include numerical procedures such as spectral elements
the computed Love numbers:
(Metivier et al., 2005, 2006) or finite-volume methods
1. Anelastic dispersion means that the elastic constants of an (Latychev et al., 2009). All of these have been applied to find
Earth model found from seismology must be adjusted the effect of lateral heterogeneities on gravity and displacement.
slightly to be appropriate for tidal frequencies. As an exam- Given realistic lateral heterogeneities in density, elastic moduli,
ple, Dehant et al. (1999) found that for an elastic Earth and prestress, the perturbations to the semidiurnal tide are less
model, the gravimetric factor d0 is 1.16030 for the M2 tide; than 1 mm for the radial displacements for radial motion and
an anelastic model gives 1.16172; for h(0), the correspond- 2 nm s2 for gravity. Comparing with Figure 6 shows that this is
ing values are 0.60175 and 0.61042. no more than 5 103 in displacement and 103 in gravity.
2. Dispersion also means that the Love numbers vary within While gravity measurements have a very high precision, in com-
the tidal bands. For the semidiurnal and diurnal tides, the paring them with theory, the limiting factor is their accuracy: as
effect is small, especially compared to the NDFW reso- described in Section 3.06.6, gravimeter calibrations are at best
nance; in the long-period bands, it is significant as f good to only 103. It may be easier to observe perturbations in
approaches zero. The usual formulation for this (Pétit and displacements – and widespread GPS measurements provide
Luzum, 2010) is based on a slightly different form of [29], many more estimates of these than are available from tidal
from Smith and Dahlen (1981) and Wahr and Bergen gravimeters.
(1986); the Love numbers vary in the long-period band as
  a   a 
ap fm fm
Lð f Þ ¼ A  B cot 1 +i [30] 3.06.4 Tidal Loading
2 f f
where A and B are constants for each Love number. For the A major barrier to using Earth tides to find out about the solid
IERS standards, a ¼ 0.15 and fm ¼ 432 cpd (a period of 200 s). Earth is the observations also include the effects of the ocean
A and B are 0.29525 and 5.796  104 for k0, 0.5998 and tides – which, depending on what is being studied, may be
9.96  104 for h(0), and 0.0831 and 3.01  104 for l(0). signal or noise. Even on a rigid Earth, the redistribution of mass
3. As eqn [30] shows, the Love numbers also become in the ocean tides would create signals from the attraction of
complex-valued, introducing small phase lags into the the water; on the real Earth, this redistribution also causes the
tides. There are additional causes for this; in particular, Earth to deform, which creates additional changes. These
the NDFW frequency can have a small imaginary part induced signals are called the load tides, which combine with
because of dissipative core–mantle coupling, and this will the body tide to make up the total tide (Figure 1).
produce complex-valued Love numbers even in an elastic
Earth. Complex-valued Love numbers can be used in exten-
sions of eqns [7] and [8]; for example, if the elastic Love 3.06.4.1 Computing Loads I: Spherical Harmonic Sums
number combination introduces no phase shift (as for
To compute the load, we start with a description of the ocean
gravity) in phase, the real part is multiplied by [am n (t)
tides, almost always as a complex-valued function H(y0 , f0 ),
cos mf + bm
n (t)sin mf] and the imaginary part by [bn (t)
m
giving the amplitude and phase of a particular constituent over
cos mf  amn (t)sin mf].
the ocean; we discuss such ocean-tide models in more detail in
The most recent examination of tidal data for anelastic Section 3.06.4.4. The loads can then be computed in two ways:
effects (Benjamin et al., 2006) combined data from diurnal using a sum of spherical harmonics or as a convolution of the
tides (in the potential, as measured by satellites), the Chandler tide height with a Green function.
wobble, and the 19-year nodal tide. They find a good fit for a In the first approach, we expand the tidal elevation in
between 0.2 and 0.3 with fm ¼ 26.7 cpd (see also Ray and spherical harmonics:
166 Earth Tides

X
1 X
n 1 X
Grw a X n
4p 0
Hðy0 , f0 Þ ¼ Hnm Ynm ðy0 , f0 Þ [31] uz ðy, fÞ ¼ h Hnm Ynm ðy, fÞ
n¼0 m¼n g n¼0 m¼n 2n + 1 n
where Ynm are as in the section on tidal forcing and Hnm would
be found from rw X1 X n
3h0n
¼ Hnm Ynm ðy, fÞ [37]
rE n¼0 m¼n 2n + 1
ðp ð
2p ð
Hnm ¼ siny0 dy0 dfHðy0 , f0 ÞYnm
* Hðy0 , f0 ÞYnm
* dO [32] where rE is the mean density of the Earth. A similar expression
0 0 O applies for the induced potential, with kn0 replacing hn0 ;
for the effective tide-raising potential, sometimes called the
where we use O for the surface of the sphere. Note that there ‘self-attraction loading’ or SAL (Ray, 1998), we would use
will be significant high-order spherical harmonic terms in Hnm, 1 + kn0  hn0 .
because the tidal height goes to zero over land: any function Many terms are needed for a sum in eqn [37] to converge,
with a step behavior will decay only gradually with increasing but such a sum provides the response over the whole Earth. Ray
degree. and Sanchez (1989) used this method to compute radial dis-
The mass distribution H causes a gravitational potential on placement over the whole Earth, with n ¼ 256, and special
the surface of the Earth, which we call VL. This potential is methods to speed the computation of the Hnm coefficients in
given by the integral over the surface of the potential function eqn [32]. In any method that sums harmonics, there is always
times H; the potential function is proportional to r1, where r room for concern about the effects of Gibbs’ phenomenon
is the linear distance from the location (y, f) to the mass at (Hewitt and Hewitt, 1979) near discontinuities, but no such
(y0 , f0 ), making the integral effect was observed in the displacements computed near coast-
ð lines. Mitrovica et al. (1994) independently developed the
Hðy0 , f0 Þ
V L ðy, fÞ ¼ Grw a2 dO [33] same method, extended it to the more complicated case of
r horizontal displacements, and were able to make calculations
O
where rw is the density of seawater and G and a are as in with n ¼ 2048.
Section 3.06.2.1. We can write the r1 in terms of angular Given a global ocean-tide model and a need to find loads
distance D: over the entire surface, this summation technique requires
much less computation than the convolution methods to be
1 1 1X 1
discussed in the next section. For gravity, the contributions
¼ ¼ Pn ðcos DÞ
r 2asin ðD=2Þ a n¼0 from the Earth’s response are, from eqn [17], (n + 1)kn0 VLn/a
1X 1 X n
4p from the induced potential and 2hn0 VLn/a from the displace-
¼ Ynm ðy0 , f0 ÞYnm
* ðy, fÞ [34] ment, making the sum
a n¼0 m¼n 2n + 1
where we have again used the addition theorem [5]. Combin- 3grw X1 X n
2h0n  ðn + 1Þk0n
ing the last expression in eqn [34] with the spherical harmonic dgðy, fÞ ¼ Hnm Ynm ðy, fÞ [38]
arE n¼0 m¼n 2n + 1
expansion [31] and the expression for the potential [33] gives
the potential in terms of spherical harmonics: While this sum might appear to converge more slowly than
eqn [37] because of the n + 1 multiplying kn0 , the convergence is
X
1 X
n
V L ¼ Grw a
4p
Hnm Ynm ðy, fÞ [35] similar because for large n, nkn0 approaches a constant value,
n¼0 m¼n
2n +1 which we term k10. All three load Love numbers have such
We have found the potential produced by the load because this asymptotic limits for large n:
potential is used, like the tidal potential, in the specification of
As n ) 1 h0n ) h01 nk0n ) k01 nl0n ) l01 [39]
the Earth’s response to the load. Specifically, we define the
load Love numbers kn0 , hn0 , and ln0 such that, for a potential VL of so that the sum [38] converges reasonably well. A similar sum
degree n, we have can be used to get the gravity from the direct attraction of the
water:
VnL h 0 r1 VnL
uzn ¼ h0n u n ¼ ln Vn ¼ k0n VnL [36]
g g 3grw X1 X n
1
Hnm Ynm ðy, fÞ
where uzn is the vertical displacement (also of degree n), uhn is the arE n¼0 m¼n 4n + 2
horizontal displacement, and Vn is the additional potential
produced by the deformation of the Earth. These load Love (Agnew, 1983; Merriam, 1980).
numbers, like the Love numbers for the tidal potential, are However, summation over harmonics is not well suited to
found by integrating the differential equations for the defor- quantities that involve spatial derivatives, such as tilt or strain.
mation of the Earth, but with a different boundary condition at To find the load tides for these, we need instead to employ
the surface: a normal stress from the load, rather than zero convolution methods, which we now turn to.
stress. For a spherical Earth, these load numbers depend only
on the degree n of the spherical harmonic.
3.06.4.2 Computing Loads II: Integration Using Green
To compute the loads, we combine the definition of the
Functions
load Love numbers [36] with the expression [35], using which-
ever combination is appropriate for some observable. For If we only want the loads at a few places, the most efficient
example, for vertical displacement uz, this procedure gives approach is to multiply the tide model by a Green function,
Earth Tides 167

which gives the response to a point load, and integrate over the which may again have the asymptotic part nln0 removed and
area that is loaded by the tides. That is, we work in the spatial replaced by an analytic expression
domain rather than, as in Section 3.06.4.1, in the wave number
al01 X
1
1 @Pn ðcos DÞ a X 1 @Pn ðcosDÞ
domain; there is a strict analogy with Fourier theory, in which Gh ðDÞ ¼ + l0  l0
convolution of two functions is the same as multiplying ME n¼0 n @D ME n¼0 n 1 @D
their Fourier transforms. Convolution methods have other al01 cos ðD=2Þ½1 + 2sin D=2
¼
advantages, such as the ability to combine different ocean-tide ME 2sin ðD=2Þ½1 + sin D=2
models easily, include more detail close to the point of obser- a X1 1 @Pn ðcos DÞ
vation, and handle any of the Earth-tide observables. The stan- + nl0  l0 [45]
ME n¼0 n 1 n @D
dard reference on the procedure remains the classic paper of
Farrell (1972); Jentzsch (1997) is a more recent summary. which shows the same dependence on D for small distances.
More formally, we find the integral over the sphere (in For gravity, there are two parts to the loading: the direct
practice over the oceans) attraction of the water mass (often called the Newtonian part)
and the change caused by elastic deformation of the Earth. The
ðp ð
2p
first part can be found analytically by using the inverse square
rsiny0 dy0 r df0 GL ðy, f, y0 , f0 Þrw Hðy0 , f0 Þ [40] law and computing the vertical part of the attraction. If the
0 0 elevation of our point of observation is ea, with e small, this
where GL is the Green function for an effect (of whatever type) Green function is
" #
at (y, f) from a point mass (d-function) at (y0 , f0 ); g e + 2sin 2 D=2
rwgHr2 sin y dy df is the applied force. Ggn ðDÞ ¼  [46]
ME ð4ð1 + eÞsin 2 D=2 + e2 Þ3 =2
The Green functions are found, not directly, but by
forming sums of combinations of the load Love numbers.
The elastic part of the Green function follows from the har-
The first step is to find the potential from a point mass. Take
monic expression [38]:
H ¼ rwa2d(y0 , f0 ) (where d is the Dirac delta function). Substi-
tute this into eqn [33], using the sum in Pn(cos D) in eqn [34]. g X1
g 0
This gives the potential as Gge ¼ ð2h0n  ðn + 1Þk0n ÞPn ðcos DÞ ¼ 2 hn  h01
ME n¼0 ME
ð  
X1
ga X 1
h01 1
V L ðy, fÞ ¼ Grw a H y0 , f0 Pn ðcosDÞdO ¼ Pn ðcosDÞ [41] +  k01  1  log sin D=2 + sin 2 D=2
n¼0
M E n¼0 sin D=2 2sinD=2
O
   #
which shows that the degree-n part of the potential is VLn ¼ ga/ X
1 k01
0 0 0
+ 2 hn  h1  ðn + 1Þ kn  Pn ðcosDÞ
ME, independent of n. So, to compute vertical displacement, n¼1
n
we would apply this potential to the load Love number hn0 , [47]
getting the displacement
1
which shows, again, a D singularity for D small.Likewise, the
a X 1
VL a X 1
Green function for the tide-raising potential makes use of the
u ¼
z
h0n n ¼ h0 Pn ðcos DÞ ¼ Gz ðDÞ [42]
ME n¼0 g ME n¼0 n combination 1  kn0 + hn0 :

which is thus the loading Green function for vertical displace- a X1


Gpot ¼ 1 + k0n  h0n Pn ðcosDÞ
ment. Some insight into the behavior of this function can be ME n¼0

obtained by using the asymptotic value of hn0 to write a 1  h01 X1
¼ + k0  h0n  h01 Pn ðcos DÞ [48]
2ME sin D=2 n¼0 n
a X1
a X 1
Gz ðDÞ ¼ h01 Pn ðcos DÞ + h0n  h01 Pn ðcosDÞ
ME n¼0 ME n¼0
and the Green function for tilt uses the same combination of
ah01 a X1
¼ + h0n  h01 Pn ðcos DÞ [43] Love numbers, but with the derivative of the Pn’s:
2ME sinD=2 ME n¼0
1 X 1 @Pn ðcos DÞ
where we have made use of [34]. The new sum will con- Gt ¼ 1 + k0n  h0n
ME n¼0 @D
verge much more rapidly; in practice, it needs to include
only enough terms for hn0 to have approached h10 to ade- 1  h01 cos ðD=2Þ X 1 @Pn ðcosDÞ
¼  k0n  h0n  h01 [49]
quate precision. For D small, the sum approaches zero, so 2
4ME sin D=2 n¼0
@D
the analytic part shows that, for loads nearby, Gz varies as
D1. This is the vertical displacement seen for a point load which has a D2 singularity for D small. The tilt is thus much
on an elastic half-space, in what is called the Boussinesq more sensitive to local loads than the other observables we
solution; in the limit of short distance, the loading problem have so far discussed.
reduces to this, which provides a useful check on numerical The remaining Green function is that for strain, specifically
computations. for the strain in the direction of the load
For the horizontal displacement, the Green function is
1 @uh uz
eDD ¼ +
ga X1 0
ln @VnL a X 1
@Pn ðcos DÞ a @D a
uh ¼ ¼ l0 ¼ Gh ðDÞ [44]
ME n¼0 g @D ME n¼0 n @D from which the Green function is, from [42] and [44],
168 Earth Tides

1 X 1
1 X 1
@ 2 Pn ðcosDÞ stress, to describe the deformation of the Earth by wind stress
GDD ¼ h0n Pn ðcos DÞ + l0
ME n¼0 ME n¼0 n @D2 or ocean currents, including tidal currents; see Merriam (1985,
1986) and Wilhelm (1986).
h01 1 X1
¼ + h0  h01 Pn ðcos DÞ
2ME sin D=2 ME n¼0 n
3.06.4.3 Degree-One Effects in Ocean Loading
l01 X
1
1 @ 2 Pn ðcos DÞ 1 X 1 @ 2 Pn ðcos DÞ
+ + l0n  l01
ME n¼0 n @D 2 M E n¼0 @D2 Satellite-based measurements introduce a complication into
ocean-loading computations, one most easily explained by
1 X
h01 1
¼ + h0  h01 Pn ðcos DÞ supposing the Earth to be rigid. All the load Love numbers
2ME sin D=2 ME n¼0 n
would then be zero and so would the predicted displacements
l01 1 + sinD=2 + sin 2 D=2 1 X
1 @ 2 Pn ðcos DÞ from tidal loading. But observed displacements would be non-
+ + l0  l0
ME 4sin 2 D=2½1 + sinD=2 ME n¼0 n 1 @D2 zero because the origin of a satellite reference frame is the
[50] center of mass of everything below the satellite orbit: solid
Earth, ocean, and atmosphere. Ocean tides on a rigid Earth
2
which again shows a near-field singularity of D . This behav- would shift this combined center of mass (CM) relative to the
ior also holds for the strain perpendicular to the direction to center of mass of the solid Earth (CE), to which the observation
the load; since this is given by points would be tied. This varying displacement between CM
and CE is usually termed motion of the geocenter.
uz cos D uh In the spherical harmonic expansion of tidal height
+ [51]
a sin D a (eqn [31]), the only terms that correspond to a change in the
there is no need to compute a separate Green function for it. The center of mass of the ocean (and hence the displacement
Green function for linear strain that would be used in [40] is between CE and CM) are those with n equal to one. Because
only the degree-one load Love numbers (eqn [36]), h10 , k10 , and
 
Gz ðDÞ Gh ðDÞ l10 , combine with the n ¼ 1 terms in the load potential
GL ðD, zÞ ¼ GDD ðDÞcos 2 z + + cot D sin 2 z [52] (eqn [35]), only these numbers change when computing load-
a a
ing relative to CM instead of to CE. Before space geodesy, all
where z is the azimuth of the load relative to the direction of Earth-tide observations were made entirely on the solid Earth,
extension. Areal strain has a complicated dependence on dis- and degree-one load Love numbers were computed assuming
tance, because it is zero for a point load on a half-space, except CE to be fixed.
right at the load. Farrell (1972) discussed the difference between CE and CM
All of the Green functions are computed by finding the load loading briefly; most subsequent treatments assumed that this
Love numbers for a range of n and forming the various sums. could be completely represented by a rigid-body displacement
Farrell (1972) formed the sums up to n ¼ 10 000; the Love of all points. However, because the Earth deforms, a motion of
numbers can be computed at values of n spaced logarithmi- CM relative to CE can involve more general motions. Blewitt
cally and interpolated to intermediate values. Several numeri- (2003) showed that adding a constant to all three degree-one
cal methods to accelerate the convergence of the sums are load Love numbers produces a motion that leaves the Earth
described by Farrell (1972), Francis and Dehant (1987), and spherical, but with patterns of vertical and horizontal displace-
Guo et al. (2004). The Green functions tabulated by Farrell ment that depend on the constant added. The particular con-
(1972) (with the addition of the potential function by Farrell stant to be added to convert from CE to CM is given by
(1973)) are still widely used; Jentzsch (1997) and Guo et al. considering the induced potential. If CE is fixed, the additional
(2004) tabulated the functions for the PREM model. potential caused by the Earth’s deformation is zero, which we
0
Kamigaichi (1998) had discussed the variations in the strain write as k1:CM ¼0.
and tilt Green functions at shallow depths, forming sums up to If we take CM as fixed, a degree-one load does not change
0
n ¼ 4  106; his results show a smooth transition between the the external potential at all, so we must have k1:CM ¼  1 to
surface functions for an SNREI Earth model and the Boussi- cancel the direct effect of the mass causing the load. To com-
nesq solution for a half-space, but at a nonzero depth. The pute loads in the CM frame, we thus need to subtract one from
response at depth lacks the singularity present in the strain and the degree-one load Love numbers before using them, whether
tilt Green functions at the surface. Examinations of the extent in harmonic sums such as eqn [37] or in Green functions such
to which local structure, particularly lateral variations, affects as eqn [42]. It remains true that it is incorrect to compute CM
computed load tides have not been plentiful – perhaps mostly loads for phenomena referenced to the geocenter, such as tidal
because the data most sensitive to such effects, strain, and tilt, gravity measured on the Earth’s surface: though actually
are affected by other local distortions (Section 3.06.6.1). As eqns [47]–[49] are all unaffected by constant shifts to the
with the Love numbers for the body tides, the load Love degree-one load Love numbers.
numbers will be affected by rotation, ellipticity, anisotropy, In fact, as Blewitt (2003) pointed out, the CE point cannot
and anelasticity. The first two produce, again, a resonant be determined from surface observations; but it can be approx-
response from the NDFW, though only for loads of degree imated by choosing the load Love numbers to produce dis-
two and order one (Wahr and Sasao, 1981); Pagiatakis placements whose average is minimized over the Earth’s
(1990) has examined effects from the others. surface, a choice Blewitt terms the CF frame. For the Love
It is also possible to define load Love numbers, and Green numbers given by Farrell (1972), the degree-one load Love
functions derived for them, for transverse rather than normal numbers for that frame differ from those for the CE frame by
Earth Tides 169

amounts that cause only a 2% difference in displacements, so 3.06.4.5 Computational Methods


we can take the CF frame as equivalent to the CE frame.
Essentially, all load programs perform the convolution [40]
Fu et al. (2012) analyzed a large suite of GPS data to
directly, either over the grid of ocean cells (perhaps more finely
compare the effects of using the CM or CE loads in processing
divided near the load) or over a radial grid. Two that are
GPS data with predetermined orbits and clocks. If the same
generally available are GOTIC (Matsumoto et al., 2001) and
load computation (CE or CM) was used both in finding the
SPOTL (Agnew, 1996, 2012). Bos and Baker (2005) and Penna
orbits and in reprocessing the data, the residual tides were not
et al. (2008) had compared the results from different programs
large; but if different load computations were used, these
for gravity and displacement, finding variations of a few per-
tides became significant. Current orbit products are all set
cent because of different computational assumptions and dif-
in the CM frame, so loads computed using these Love num-
ferent versions of the coastline. Most global ocean-tide models
bers should be used for all modeling of GPS station
do not represent coastlines very accurately, so some local
displacements.
refinement is often needed. Fortunately, this can be provided
using the global coastal representations made available by
Wessel and Smith (1996) – except in the Antarctic, where this
3.06.4.4 Ocean Tide Models coastline is (in places) the ice shelves rather than the grounding
line (Bindschadler et al., 2011; Rignot et al., 2011).
To compute loads, we also need a description of the ocean
Figure 8 shows the computed loads for a region (Northwest
tides. Finding the global pattern of the ocean tides is a difficult
Europe) with large and complex local tides. The vertical dis-
task that has been pursued for some time (Cartwright, 1977,
placement and gravity loads have roughly similar forms, but
1999). The intractability of the relevant equations, the com-
the tilt and linear strain have a very different pattern, being
plexity of the geometry, and the sensitivity of the results to
much more concentrated near the coast, as might be expected
details in the models long precluded numerical solutions, one
from the different near-field behavior of their Green functions.
difficulty being that the oceans have barotropic modes of
oscillation with periods close to the diurnal and semidiurnal
tidal bands. At the same time, it was very difficult to measure
tides in deep water. All this meant that until relatively recently, 3.06.5 Analyzing and Predicting Earth Tides
there were no good tidal models for computing loads.
From the Earth-tide standpoint, what is important is that As noted in ‘Abstract,’ the tidal forces can be described to
increasing computational power has rendered numerical solu- extremely high precision and accuracy, and the body tides
tions possible for realistic geometries and that satellite altime- and tidal loading can often be modeled to an accuracy that
try has provided global coverage for tidal data. Starting in the exceeds that of tidal measurements. Such measurements do
late 1990s, these developments made it possible to construct however provide a check on these models and in some cases
global tidal models, whose data are mostly (often entirely) allow them to be improved, so we briefly describe how tidal
derived from altimetry and computed on a relatively coarse parameters are extracted from the data and how the data are
mesh (0.125 –0.5 ). Comparisons between different models obtained. This is important whether we aim to measure the
(Andersen et al., 1995; Desai et al., 1997; Le Provost et al., tides, use modeled tides as a calibration signal, or predict the
1995; Shum et al., 1997) showed variations in the open ocean tides to high accuracy to check the quality of ongoing
of less than 1 cm. With more altimetry data and more comput- measurements.
ing power, additional models (Egbert and Erofeeva, 2002;
Lyard et al., 2006; Matsumoto et al., 2000) have become avail-
3.06.5.1 Tidal Analysis and Prediction
able, and loads computed from them usually differ by only
small amounts. As noted in ‘Abstract,’ the analysis of time series for tidal
But there is still room for improvement. The polar regions response is just a special case of finding the transfer function,
are poorly covered by altimetry data, and much effort has gone or admittance, of a linear system, a concept first introduced
into improving tidal models there (King and Padman, 2005; into tidal analysis by Munk and Cartwright (1966). Because the
Padman and Erofeeva, 2004; Padman et al., 2002). More gen- tides are very bandlimited, we can find the tidal admittance,
erally, global models often lack the resolution (in grid size and WT(f ), only for frequencies at which XT(f ) contains significant
representation of bathymetry) to accurately model tides in energy. For ocean tides, it is most meaningful to take xT(t) to be
marginal seas, many of which have large tides because of the local value of the tide-raising potential or for some analyses
local resonances. There is thus still a need for local tidal the tide computed for a nearby site (Cartwright et al., 1969). In
models, to be used when Earth-tide data are collected nearby. Earth tide studies, it may be more convenient to take as refer-
Such models use a finer mesh and rely more on local tide- ence the tides expected for an oceanless, but otherwise realistic,
gauge data, than global models do. Ray et al. (2011) reviewed Earth model, so that any departure of W(f ) from unity will
the state of local tidal modeling, which in many places remains then reflect the effect of ocean loads or the inadequacy of the
imperfect. model.
Most tidal models (global or local) are given for particular The theory described in Section 3.06.3.2 shows that, except
tidal constituents, usually at least one diurnal and one semidi- for the NDFW resonance, W(f ) for an oceanless Earth varies
urnal. Unless a local resonance is present, the loads for other only very slightly with frequency. The ocean tides show more
harmonics can be found by scaling using the ratios of the variability, but only in limited areas do they have resonances
amplitudes in the equilibrium tide (Le Provost et al., 1991). within the tidal bands, so that in general, the ocean load also
170 Earth Tides

Gravity Tide-raising potential


10
58 58 10

60
20
40
56 56
20 60 10
54 54

10
40 40
52 100 52 60
120

10
80
50 50 80
140
20

48 48

-10 -5 0 5 10 -10 -5 0 5 10
Vertical displacement Horizontal displacement

58 5 5 58
6
56 20 56
10 4
5 2 2
54 54 6
10 4
20
10
52 40
52
8 10
50 50 6

4
30
50

48 48

-10 -5 0 5 10 -10 -5 0 5 10
Max tilt Max linear strain
30
58 58
100

3
56 56 30
30

54 54 30
30

30

52 52
300

10
300 30 3
50 50
00
30

300 1
10
10

48 48
0

100

-10 -5 0 5 10 -10 -5 0 5 10
Figure 8 Loads for the M2 tide, computed by the Green function method for a 2004 empirical global model (GOT4.7, developed using the methods of Ray
et al., 1999), combined with a detailed model (done by Egbert and Erofeeva (2002) in 2010) of the tides on the European Shelf. Only the amplitude
of the complex-valued quantities is shown; for tilt, displacement, and strain, the value is taken along the azimuth that maximizes the amplitude.

varies smoothly with frequency (Garrett and Munk, 1971). noise will bias the amplitudes of the coefficients to be larger
Even the local resonances in certain bays and gulfs have a low than the true values.
enough Q that the response is smoothly varying over (say) the If spectral analysis is to be used, a much better method is
entire semidiurnal band (Ku et al., 1985). So, the more closely the cross spectral technique described by Munk and Cartwright
spaced two frequencies are, the closer the corresponding values (1966). This method has poorer frequency resolution than
of W(f ) will be, an assumption Munk and Cartwright (1966) others to be discussed but makes the fewest assumptions
dubbed the ‘credo of smoothness.’ about the form of W(f ) and also provides estimates of the
A naive way to find the tidal response is to take the Fourier noise as a function of frequency. This is useful because many
transform of the data (using a fast Fourier transform) and use methods assume the noise to be the same at all frequencies,
the amplitudes and phases of the result. This is a poor choice and it may not be; in particular, the noise is sometimes
for two reasons. One problem is that the frequencies computed observed to rise sharply in the tidal bands, a phenomenon
by the usual definition of the discrete Fourier transform usually called ‘tidal cusping’ (Colosi and Munk, 2006; Munk et al.,
do not coincide with the frequencies of the tidal harmonics – 1965; Ponchaut et al., 2001). The cross spectral method does
especially if the length of the transform, N, is chosen to work however require large amounts of data to perform reliably. The
well with a fast Fourier transform algorithm. In addition, any procedure is described in full by Munk and Cartwright (1966);
Earth Tides 171

it depends on finding the cross spectrum between a noise-free Pawlowicz et al., 2002; Tamura et al., 1991; Van Camp and
reference series and the data, using a slow Fourier transform to Vauterin, 2005; Wenzel, 1996).
make the Fourier frequencies match the tidal frequencies rela- A quite different approach to tidal analysis is the response
tively well, windowing to reduce bias from spectral leakage, method, introduced by Munk and Cartwright (1966). This does
and averaging to get a statistically consistent estimate. not use an expansion of the tidal potential into harmonics, but
By far the commonest approach to tidal analysis is least- rather treats it as a time series to be fit to the data, using a set of
squares fitting of a set of sinusoids with known frequencies, weights to express the admittance. Lambert (1974) and
namely, the frequencies of the largest tidal constituents. That is, Merriam (2000) had applied this method to Earth tides, and
we aim to minimize the sum of squares of residuals: it has been used for estimating tides from satellite altimetry
" #2 (Smith et al., 1997).
XN X
L
The basic approach is to find the tides as a weighted sum
yn  ðAl cos ð2pfl tn Þ + Bl sin ð2pfl tn ÞÞ [53]
n¼0 l¼1
over the time variations of each spherical harmonic (not har-
monics in time):
which expresses the fitting of L sine–cosine pairs with frequen-
X
1 X
n X
Lnm
m

cies fl to the N data yn, the f’s being fixed to the tidal harmonic y ðt Þ ¼ nl an ðt  lDÞ + ibn ðt  lDÞ
wm m
[54]
frequencies and the A’s and B’s being solved for. n¼2 m¼nl¼Lnm
The usual assumption behind a least-squares analysis is that
the residual after fitting the sinusoids will be statistically indepen- where the am m
n (t) and bn (t) are the time-varying functions that
dent random variables; but this is valid only if the noise spectrum sum to give the potential in eqn [8]. The complex-valued
is white, which is usually not so. One departure from whiteness is weights wm nl are called response weights; their Fourier transform
the presence of increased long-period noise outside the tidal gives the admittance W(f ). So, for example, a single complex
bands; this can be removed by filtering the data before analyzing weight for each n and m (i.e., setting Lnm ¼ 0) amounts to
it. A more difficult problem is the tidal cusping just referred to. If assuming a constant W for each degree and order – though
the noise spectrum rises to a higher level inside the tidal bands, even one complex weight can express both amplitude and
perhaps very much higher around the frequencies of the radia- phase response. Including more weights, with time lags (the
tional tides, this needs to be allowed for in fitting the tides and in sum over l), allows the admittance to vary smoothly with
finding the errors in the final tidal parameters. In particular, the frequency across each tidal band. The lag interval is usually
relative error for a harmonic of amplitude A analyzed over a total chosen to be 2 days, which makes the admittance smooth over
time span T is approximately 2P(f )/A2T, where P(f ) is the noise frequencies of greater than 0.5 cpd; note that the weights can
power spectral density at the frequency of that harmonic include the potential at future times because the admittance is
(Cartwright and Amin, 1986; Munk and Hasselmann, 1964). If being fit over only a narrow frequency band.
excess energy in the tidal bands is not allowed for, the errors can
be underestimated by significant amounts.
3.06.5.1.1 Predicting tides
The main problem with using eqn [53] directly for tidal
All tidal predictions, other than those based on the response
analysis comes from the fine-scale frequency structure of the
method, use a harmonic expansion similar to eqn [11]:
tidal forcing, particularly the nodal modulations. Leaving such
variations out of eqn [53], and only solving for a few large X
K

harmonics, will be inaccurate. But the simplest way of includ- xð t Þ ¼ Ak cos 2pfk ðt  t0 Þ + f0k ðt0 Þ + fk [55]
k¼1
ing nodal and other modulations, namely, by including the
satellite harmonics in eqn [53], is not possible because the where the Ak’s and fk’s are amplitudes and phases (the har-
solution will be unstable unless we have 19 years of data. monic constants) for whatever is being predicted. The fk’s are the
This instability is general and applies whenever we try to frequencies of the different harmonics, and the f0k ’s are the
solve for the amplitudes of harmonics separated in frequency phases of these at a reference time t0.
by less than 1/T, where T is the record length (Munk and Any user of tidal constants should be aware of two pitfalls
Hasselmann, 1964). This problem is not restricted to the relating to the conventions for phase. One is a sign convention:
nodal modulation; for example, with only a month of data, whether positive phases represent lags (time delays), as is true in
we cannot get reliable results for the P1 and K1 lines, since they much of the older literature, or leads. The other is the reference
are separated by only 0.15 cycles per month. time used. The local phase is one choice, in which zero phase (for
All least-squares tidal analysis thus has to include assump- each harmonic) is at a time at which the potential from that
tions about tidal harmonics closely spaced in frequency – harmonic is locally a maximum. For ocean tides, this phase is
which comes to an implicit assumption about the smoothness usually denoted as k (with positive phases for lags). For Earth
of the admittance. Usually, the admittance is assumed to be tides, local phase is convenient because on an SNREI Earth, it is
constant over frequency ranges of width 1/T around the main zero for gravity, NS tilt, vertical and NS displacement, and areal
constituents, summing all harmonics within each such range strain. The other choice is the Greenwich phase G, in which the
to form (slowly varying) sinusoidal functions to replace the phase is taken to be zero (for each harmonic) at a time at which
sines and cosines of eqn [53]. Of course, if we then wish to its potential would be a maximum at 0 longitude. If given for a
assign the resulting amplitude to a particular harmonic (say number of places, this phase provides a ‘snapshot’ of the distri-
M2), we need to correct the amplitudes found by the ratio of bution of the tides at a particular instant; this phase is the norm
this sinusoidal function to the single harmonic. All this adds in ocean-tide models. Since the time between maximum at
complexity to the existing analysis programs (Foreman, 2004; Greenwich and maximum at a local place depends only on the
172 Earth Tides

spherical harmonic order m and on the Earth rotating 360 every of these used metallic springs, with changes in gravity causing
24 h, the relationship between k and G is simple and depends changes in spring length. This mode of operation produced
only on the tidal species number m and the longitude fW; the significant phase lags and hysteresis, but such measurements
frequency of the harmonic is not involved. The relationship is were able to show the widespread effects of load tides (Llubes
conventionally written as and Mazzega, 1997) and the effect of the NDFW on tidal data
(e.g., Neuberg et al., 1987). Effects caused by spring rheology
G ¼ k  mfW
are much reduced by applying feedback, which was first
where for both phases, a lag is taken to be positive, and done in the instrument of LaCoste and Romberg (Harrison
longitude fW to be positive going west. and LaCoste, 1978); these instruments, supplied with elect-
The primary complications in predicting the tides come, ronic feedback, remain useful for tidal measurements. The
once again, from the various long-term modulations, notably lowest noise and most stable calibration are provided by the
the nodal modulations discussed in Section 3.06.2.3. Classical superconducting gravimeter (Goodkind, 1999), in which a
prediction methods, which used only a few constituents to superconducting sphere is suspended in a magnetic field at
minimize computation, applied nodal corrections to the Ak’s liquid-helium temperatures. This gives little drift and very
and fk’s of these few harmonics to produce new values that low noise at long periods. At the periods of the semidiurnal
would be valid for (say) each year, a complication, since the tides, the noise on superconducting gravimeters ranges from
corrections themselves change with time. 130 to 140 dB (relative to 1 m2 s4), with the noise being
A more computationally intensive but conceptually simpler about 5 dB higher in the diurnal band (Rosat et al., 2004);
approach uses a large number of harmonics in the sum [55], these levels are about 5 dB below those of spring gravimeters
including all satellite harmonics, thus automatically producing (Cummins et al., 1991). Because the noise is so low, small tidal
the modulations. The amplitudes and phases of a few har- signals can be measured with great precision; recent examples
monics are interpolated to give those of all harmonics, again include the detection of loading from small nonlinear ocean
on the assumption that the admittance is smooth, for example, tides (Boy et al., 2004; Merriam, 1995) and the discrimination
through a spline interpolation of the real and imaginary parts between the loads from equilibrium and dynamic models
of W(f ) (Le Provost et al., 1991). for the long-period ocean tides (Boy et al., 2006a; Iwano
et al., 2005). The superconducting gravimeter also can provide
accurate measurements of the NDFW resonance (Florsch et al.,
1994; Sato et al., 1994, 2004; Zürn et al., 1986).
3.06.6 Earth-Tide Instruments and Measurements Comparing the size of the load tides in Figure 8 with the
rms body tides in Figure 6 shows that even the largest loads are
Because the Earth tides are so small, building instruments to but a few percent of the total gravity tide. So, to get accurate
detect them has long been a challenge, though digital measurements of the load tides, the gravimeter must be cali-
recording and improvements in transducers have made tidal brated with extreme accuracy, at least to a part in 103. An even
measurements easier than they used to be. higher level of accuracy is needed if tidal gravity measurements
The earliest measurements were of tidal tilts, and over the are to discriminate between the predictions of different Earth
years, a wide variety of tiltmeters have been designed; Agnew models. Calibration to this level has proven to be difficult. One
(1986) described many of them, and the designs have changed method, usable only with spring gravimeters, is to move them
little since then. They generally fall into two classes: small to different locations, making gravity measurements over a
instruments that sense the motion of a pendulum or of a wide range and interpolating to the small range of the tides.
bubble in a fluid and larger systems that measure the motion This is not feasible for superconducting gravimeters; they can
of a fluid in a long pipe. The former are usually referred to as be calibrated by measuring the response to known moving
short-base tiltmeters and are now generally installed in bore- masses (Achilli et al., 1995), something not generally possible
holes in order to obtain adequate thermal stability and rela- with spring gravimeters because of their complicated mass
tively low rates of drift. The latter, called long-base systems, are distribution. A generally applicable method is to place the
usually installed in tunnels a few tens of meters long or longer instrument on a platform that can be moved vertically by
(e.g., d’Oreye and Zürn, 2005), though a very few instruments, known amounts to produce small accelerations of the same
several hundred meters in length, have been installed near the size as the tides (Richter et al., 1995). Currently, the common-
ground surface. A similar division exists in strainmeters est calibration method is to operate an absolute gravimeter
(Agnew, 1986): there are very short-base systems, many next to a tidal system for several days and find the scale factor
installed in boreholes, longer instruments installed in tunnels, by direct comparison (Francis et al., 1998; Hinderer et al.,
and a few very long instruments, using laser light rather than 1991; Tamura et al., 2005). The calibrations of absolute gravi-
physical length standards, some at the surface and others meters are known to much higher accuracy than is required, so
underground. One other class of instrument sensitive to tidal the main source of error is noise in the measurements. Francis
deformations is the ring-laser gyroscope (Schreiber et al., and van Dam (2002) confirmed that this method can provide
2003), which detects changes in the orientation of the instru- calibrations to 103. Baker and Bos (2003) summarized tidal
ment relative to the Earth’s spin axis: these involve both tilts measurements from both spring and superconducting gravi-
and horizontal displacements. meters; from the greater scatter for the in-phase components of
Historically, the second type of Earth tide to be detected was the residual tide, they conclude that systematic errors in the
changes in gravity, and many such measurements have been calibration remain, with errors up to 4  103 in some cases.
made with a variety of types of gravimeters (Torge, 1989). Most These authors, and Boy et al. (2003), found too much scatter in
Earth Tides 173

the results to distinguish between different Earth models or to models specifically designed for this region predicting the
detect any latitude dependence. loads better than the older global models.
The newest procedures for measuring Earth tides are the The substantial growth in continuous GPS measurements
techniques of space geodesy. Since the induced potential has meant that measurements of displacement tides are avail-
affects satellite orbits, modeling of these provides constraints able from many locations, with some regions (Japan, Europe,
on k2; a particularly notable result was that of Ray et al. (2001), and the Western United States) having a much higher density
who were able to measure the phase lag of the solid-Earth of stations than has ever been available for any other type of
component of the M2 tide as 0.204  0.047 . Positioning Earth-tide measurement. As the available data spans have
techniques such VLBI and GPS are sensitive to all displace- increased, the errors in the tidal estimates have shrunk to a
ments, including tides. VLBI data provided some of the first few tenths of a millimeter. Ito et al. (2009) estimated tidal
observations of the body-tide displacements and now have displacements using GPS data from the Japanese GEONET
sufficient precision to be sensitive to load tides as well (Haas array; Ito and Simons (2011) and Yuan and Chao (2012)
and Schuh, 1998; Petrov and Ma, 2003; Sovers, 1994); the had done the same with GPS data from continuous GPS sta-
currently available series can provide the amplitudes of tidal tions in the Western United States. These authors found that,
constituents to better than 1 mm. However, VLBI data are after removing the body and load tides, there are spatially
available only at a few places, whose number is not increasing. coherent patterns in the residual tidal displacements; Figure 9
Continuous GPS data, in contrast, are available from a shows ones derived from Yuan and Chao (2012). Since these
large, and rapidly increasing, number of locations. The tidal residuals are not always largest near the coast, it seems unlikely
displacements of GPS stations need to be accurately modeled, that they are caused by inadequate modeling of the load tides;
since any inaccuracy (especially in the vertical) will bias GPS how to explain them in terms of the laterally heterogeneous
estimates of zenith delay and hence of water vapor (Dach and models described in Section 3.06.3.3 remains an open ques-
Dietrich, 2000; Dragert et al., 2000); in addition, for the stan- tion, as does their relationship to the kind of systematic depar-
dard 1-day processing, unmodeled tidal displacements may tures from theoretical tides documented by Langbein (2010)
produce, through aliasing, spurious long-period signals for long-base strain data.
(Penna and Stewart, 2003). All GPS processing software there-
fore includes models for the body-tide displacements (includ-
ing the pole tide) and often the load displacements as well.
Tidal motions can be measured with GPS in two ways. One 3.06.6.1 Local Distortion of the Tides
is to process data over short time spans (say, every hour) to Heterogeneities on a much smaller scale than those discussed
produce a time series that can then be analyzed for tidal in Section 3.06.3.3 can also affect Earth-tide measurements, in
motions; this is usually called a ‘kinematic’ method. The ways that make interpretation much more difficult. King and
other procedure is to include the complex amplitude of some Bilham (1973) and Baker and Lennon (1973) introduced
of the larger tidal constituents (in all three directions) as this concept into Earth-tide studies by suggesting that much
unknowns in the GPS solution, solving for any unmodeled of the scatter in tidal tilts measurements was caused by strain
tides. For a single-day solution, the errors will be large because tides, which were coupled into tilts by the presence of a free
of the close frequency spacing, but keeping the full set of surface – for example, the tunnel in which such instruments
covariances will retain information about what combinations were often housed. This cavity effect (Harrison, 1976) has
are well constrained. As daily solutions are combined (usually indeed turned out to be quite important; any strain and tilt
with a Kalman filter estimator), the different harmonics dec- measurements not made with surface-mounted long-base
orrelate, and the tidal estimates converge to a solution. instruments require a cavity, which creates a very strong inho-
Baker et al. (1995) and Hatanaka et al. (2001) took the first mogeneity that produces local deformations, notably rotations
approach for two local areas, finding good agreement between and strains that modify the tidal strains and tilts expected
observed and predicted loads. Khan and Scherneck (2003) also for an SNREI Earth. Similar effects are caused by other inhomo-
used this method, finding that the best approach was to esti- geneities such as topography (an irregular free surface),
mate zenith delays over the same (hourly) time span as the fractures, or changes in rock properties.
displacement was found for. Schenewerk et al. (2001) used the We can describe these local effects using coupling matrices
second method for a global set of stations, taking data from (Berger and Beaumont, 1976; King et al., 1976). We denote the
every third day over 3 years; they found generally good agree- actual displacement field in the Earth by u and divide it into
ment with predicted loads except at some high-latitude sites, two parts: u0, which is the displacement that would occur with
probably because of using an older ocean-tide model; Allinson no cavity (or other inhomogeneity), and the difference
et al. (2004) applied this method to tidal loading in Britain. du ¼ u  u0. We can perform the same decomposition on the
King (2006) compared the two methods for the GPS site at the strain tensor E and the local rotation vector O:
South Pole (for which only load tides are present), using about
5 years of data. He found that the second method gave better E ¼ E0 + dE O ¼ O0 + dO
results in the vertical; the two methods had comparable errors
in the horizontal, with the precision being somewhat less than The additional deformations dE and dO will depend on the
1 mm. The K1 and K2 tides give poor results because their details of the irregularity (i.e., the shape of the cavity or the
frequency is very close to the repeat time of the GPS constella- topography) and on E0, but not on O0 because it is a rigid
tion and its first harmonic. King et al. (2005) used GPS to rotation. If we suppose E0 to be homogeneous strain, then dE
validate different ocean-tide models around Antarctica, with and dO at any position are linear functions of E0 and can be
174 Earth Tides

East M2 residuals North M2 residuals

48

46

44

42

40

38

36

34
1 mm 1 mm
32
-125 -120 -115 -110 -105 -125 -120 -115 -110 -105
Figure 9 Residual M2 tides for east and north displacements, represented as phasors relative to the local potential (lags counterclockwise).
These come from the GPS results of Yuan and Chao (2012), after subtracting loads from both the GOT4.7 global model and local models for the eastern
Pacific, the Gulf of California, Hudson’s Bay, and the Arctic Ocean. If there are three or fewer results in a 1∘ square, they are plotted individually
as black arrows; otherwise, a red arrow shows the average value. Blue circles are station locations.

related to it by a fourth-order strain-strain coupling tensor CE usually known to make results from such a model more than a
and a third-order strain-rotation coupling tensor CO: rough guide. The measurements by Baker (1980), on a pier in
the middle of a tunnel, which showed a 5% change in tilts
dE ¼ CE E0 dO ¼ CO E0 [56] measured 0.5 m apart, imply that such modeling has to be
extremely detailed.
In general, strain-tilt coupling involves both CE and CO
One possible way to reduce the importance of coupling is to
(Agnew, 1986). Near the surface of the Earth, there are only
arrange the geometry to minimize it. Measuring strain along a
three independent components to E0; this fact, and the sym-
tunnel is one example; another is measuring tilt in a borehole:
metries of the strain tensors, means that both CE and CO have
since a horizontal strain does not cause rotation of the side of
nine independent components.
the borehole, attaching the tiltmeter to this should eliminate
An analytic solution for CE and CO exists for an ellipsoidal
the cavity effect. An array of such tiltmeters installed to use
cavity (Eshelby, 1957); Harrison (1976) had described results
observed tides to map crustal inhomogeneities (Levine et al.,
for the case where one axis of the ellipsoid is vertical: the
1989; Meertens et al., 1989) still produced widely scattered
coupled strains and tilts are largest when measured in the direc-
results. A subsequent test using closely spaced instruments in
tion along which the cavity is smallest, whether the external
nominally uniform geology (Kohl and Levine, 1995) also
strain is in that direction or not. Strain measured along a long
showed significant differences between boreholes and in one
narrow cavity is not amplified very much. The limit of this is an
borehole even between two different positions. These results
infinite horizontal circular tunnel, for which the strains along
suggest that any short-base measurement of tilt may be signif-
the tunnel (or shear strains across it) are unaltered, but both
icantly affected by local inhomogeneities and so will not be
vertical and horizontal strains are amplified, as has been
able to be accurately compared with theoretical tidal models.
observed in cross tunnel measurements of strain tides (Beavan
Sometimes, the aim is to use the tides to determine the
et al., 1979; Itsueli et al., 1975; Takemoto, 1981). Finite-element
effects of the inhomogeneity, so that other signals seen can
modeling (Berger and Beaumont, 1976) shows that slight depar-
be corrected. The most notable case is borehole strainmeters, in
tures from circularity do not much alter the strain from wall to
which the instrument, hole, and grout cause the strain inside
wall, but in a square tunnel, the strains concentrate near the
the instrument wall to be significantly different from that in
corners; strains along a finite tunnel are undistorted if they are
the far field. A simplified model that assumes axial symmetry
measured more than one tunnel diameter away from an end.
(Gladwin and Hart, 1985) provides two coupling coefficients,
Tiltmeters mounted on tunnel walls, or placed on a ledge or
one for areal and one for shear strain. This can be checked
near a crack, will be affected by strain-coupled tilts.
using the tides (Hart et al., 1996), which also allow estimation
Inhomogeneities (including cavities) create distortion but
of a full coupling tensor, along the lines of equation [56];
do not add noise. They thus mean that precise comparison
Roeloffs (2010) used a version of this in calibrating a later
between data and models is difficult at best, with the exception
generation of borehole strainmeters.
of frequency-dependent effects such as the NDFW. There are
semianalytic methods for approximating topographic effects
(Cayol and Cornet, 1997, 1998; Meertens and Wahr, 1986),
and while it is always possible to build a finite-element model
References
of any inhomogeneity (Emter and Zürn, 1985; Sato and Achilli V, Baldi P, Casula G, et al. (1995) A calibration system for superconducting
Harrison, 1990), not enough detail of the elastic constants is gravimeters. Bulletin Géodésique 69: 73–80.
Earth Tides 175

Agnew DC (1983) Conservation of mass in tidal loading computations. Geophysical Cartwright DE and Edden AC (1973) Corrected tables of tidal harmonics. Geophysical
Journal of the Royal Astronomical Society 72: 321–325. Journal of the Royal Astronomical Society 33: 253–264.
Agnew DC (1986) Strainmeters and tiltmeters. Reviews of Geophysics 24: 579–624. Cartwright D, Munk W, and Zetler B (1969) Pelagic tidal measurements: A suggested
Agnew DC (1996) SPOTL: Some programs for ocean-tide loading. SIO Reference Series procedure for analysis. EOS. Transactions of the American Geophysical Union
96-8. Scripps Institution of Oceanography. 50: 472–477.
Agnew DC (2012) SPOTL: Some programs for ocean-tide loading. SIO Technical Cartwright DE and Tayler RJ (1971) New computations of the tide-generating potential.
Report. Scripps Institution of Oceanography. http://escholarship.org/uc/item/ Geophysical Journal of the Royal Astronomical Society 23: 45–74.
954322pg. Cayol V and Cornet FH (1997) 3D mixed boundary elements for elastic deformation field
Allinson CR, Clarke PJ, Edwards SJ, King MA, Baker TF, and Cruddace PR (2004) analysis. International Journal of Rock Mechanics and Mining Science and
Stability of direct GPS estimates of ocean tide loading. Geophysical Research Letters Geomechanics Abstracts 34: 275–287.
31(15): L15603. http://dx.doi.org/10.1029/2004GL020588. Cayol V and Cornet FH (1998) Effect of topography on the interpretation of the
Amoruso A, Botta V, and Crescentini L (2012) Free core resonance parameters from deformation field of prominent volcanoes: Application to Etna. Geophysical
strain data: sensitivity analysis and results from the Gran Sasso (Italy) Research Letters 25: 1979–1982.
extensometers. Geophysical Journal International 189: 923–936. http://dx.doi.org/ Chapman S and Lindzen RS (1970) Atmospheric Tides: Gravitational and Thermal.
10.1111/j.1365-246X.2012.05440.x. New York: Gordon and Breach.
Andersen OB, Woodworth PL, and Flather RA (1995) Intercomparison of recent ocean Colosi JA and Munk WH (2006) Tales of the venerable Honolulu tide gauge. Journal of
tide models. Journal of Geophysical Research 100: 25,261–25,282. Physical Oceanography 36: 967–996.
Baker TF (1980) Tidal tilt at Llanrwst, north Wales: Tidal loading and earth structure. Cummins P, Wahr JM, Agnew DC, and Tamura Y (1991) Constraining core undertones
Geophysical Journal of the Royal Astronomical Society 62: 269–290. using stacked IDA gravity records. Geophysical Journal International 106: 189–198.
Baker TF (1984) Tidal deformations of the Earth. Science Progress Oxford 69: 197–233. Dach R and Dietrich R (2000) Influence of the ocean loading effect on GPS derived
Baker TF and Bos MS (2003) Validating Earth and ocean tide models using tidal gravity precipitable water vapor. Geophysical Research Letters 27: 2953–2956. http://dx.
measurements. Geophysical Journal International 152: 468–485. doi.org/10.1029/1999GL010970.
Baker TF, Curtis DJ, and Dodson AH (1995) Ocean tide loading and GPS. GPS World Dahlen FA (1993) Effect of the Earth’s ellipticity on the lunar potential. Geophysical
6(5): 54–59. Journal International 113: 250–251.
Baker TF and Lennon GW (1973) Tidal tilt anomalies. Nature 243: 75–76. Dahlen FA and Tromp J (1998) Theoretical Global Seismology. Princeton, NJ: Princeton
Bartels J (1957/1985) Tidal forces (English translation). In: Harrison JC (ed.) Earth University Press.
Tides, pp. 25–63. New York: Van Nostrand Reinhold. Dehant V, Defraigne P, and Wahr JM (1999) Tides for a convective Earth. Journal of
Beavan J, Bilham R, Emter D, and King G (1979) Observations of strain Geophysical Research 104: 1035–1058.
enhancement across a fissure. Veroeff Deutsche Geodaetische Kommission, Reihe B Desai SD (2002) Observing the pole tide with satellite altimetry. Journal of Geophysical
231: 47–58. Research 107(C11): 3186. http://dx.doi.org/10.1029/2001JC001224.
Benjamin D, Wahr JM, Ray RD, Egbert GD, and Desai SD (2006) Constraints on mantle Desai SD, Wahr JM, and Chao Y (1997) Error analysis of empirical ocean tide models
anelasticity from geodetic observations, and implications for the J2 anomaly. estimated from Topex/Poseidon altimetry. Journal of Geophysical Research
Geophysical Journal International 165: 3–16. http://dx.doi.org/10.1111/j.1365- 102: 25,157–25,172.
246X.2006.02915.x. Doodson AT (1921) The harmonic development of the tide generating potential.
Berger J (1975) A note on thermoelastic strains and tilts. Journal of Geophysical Proceedings of the Royal Society of London, Series A 100: 305–329.
Research 80: 274–277. d’Oreye NF and Zürn W (2005) Very high resolution long-baseline water-tube tiltmeter
Berger J and Beaumont C (1976) An analysis of tidal strain observations from the to record small signals from Earth free oscillations up to secular tilts. Review of
United States of America: II. The inhomogeneous tide. Bulletin of Seismological Scientific Instruments 76: 024501.
Society of America 66: 1821–1846. Dragert H, James TS, and Lambert A (2000) Ocean loading corrections for continuous
Bindschadler R, Choi H, Wichlacz A, et al. (2011) Getting around Antarctica: New GPS: A case study at the Canadian coastal site Holberg. Geophysical Research
high-resolution mappings of the grounded and freely-floating boundaries of the Letters 27: 2045–2048.
Antarctic ice sheet created for the International Polar Year. Cryosphere Discussion Egbert GD and Erofeeva SY (2002) Efficient inverse modeling of barotropic ocean tides.
5: 183–227. http://dx.doi.org/10.5194/tcd-5-183-2011. Journal of Atmospheric and Oceanic Technology 19: 183–204.
Blewitt G (2003) Self-consistency in reference frames, geocenter definition, and surface Emter D and Zürn W (1985) Observations of local elastic effects on earth tide tilts and
loading of the solid Earth. Journal of Geophysical Research 108: 2103. http://dx.doi. strains. In: Harrison JC (ed.) Earth Tides, pp. 309–327. New York: Van Nostrand
org/10.1029/2002JB002082. Reinhold.
Bos MS and Baker TF (2005) An estimate of the errors in gravity ocean tide loading Eshelby JD (1957) The determination of the elastic field of an ellipsoidal inclusion and
computations. Journal of Geodesy 79: 50–63. related problems. Proceedings of the Royal Society of London, Series A
Boy J-P, Llubes M, Hinderer J, and Florsch N (2003) A comparison of tidal ocean 241: 376–396.
loading models using superconducting gravimeter data. Journal of Geophysical Farrell WE (1972) Deformation of the earth by surface loads. Reviews of Geophysics
Research 108. http://dx.doi.org/10.1029/2002JB002050, ETG 6-1–6-17. 10: 761–797.
Boy J-P, Llubes M, Ray R, Hinderer J, and Florsch N (2006a) Validation of long-period Farrell WE (1973) Earth tides, ocean tides, and tidal loading. Philosophical Transactions
oceanic tidal models with superconducting gravimeters. Journal of Geodynamics of the Royal Society A 272: 253–259.
41: 112–118. Florsch N, Chambat F, Hinderer J, and Legros H (1994) A simple method to retrieve the
Boy J-P, Llubes M, Ray R, et al. (2004) Non-linear oceanic tides observed by complex eigenfrequency of the Earth’s nearly diurnal free wobble: Application to the
superconducting gravimeters in Europe. Journal of Geodynamics 38: 391–405. Strasbourg superconducting gravimeter data. Geophysical Journal International
Boy J-P, Ray R, and Hinderer J (2006b) Diurnal atmospheric tide and induced gravity 116: 53–63.
variations. Journal of Geodynamics 41: 253–258. Foreman MGG (2004) Manual for tidal heights analysis and prediction. Pacific Marine
Broucke RA, Zürn W, and Slichter LB (1972) Lunar tidal acceleration on a rigid Earth. Science Report 77-10. Institute of Ocean Sciences, Patricia Bay, Sidney, B.C.
American Geophysical Union, the Geophysical Monograph 16: 319–324. Francis O and Dehant V (1987) Recomputation of the Green’s functions for
Brush SG (1996) Nebulous Earth: The Origin of the Solar System and the Core of the tidal loading estimations. Bulletin d’Information des Marées Terrestres
Earth from Laplace to Jeffreys. Cambridge: Cambridge University Press. 100: 6962–6986.
Brzezı́nski A and Capitaine N (1993) The use of the precise observations of the Celestial Francis O, Niebauer TM, Sasagawa G, Klopping F, and Gschwind J (1998) Calibration
Ephemeris Pole in the analysis of geophysical excitation of earth rotation. Journal of of a superconducting gravimeter by comparison with an absolute gravimeter FG5 in
Geophysical Research 98: 6667–6675. http://dx.doi.org/10.1029/92JB02874. Boulder. Geophysical Research Letters 25: 1075–1078.
Bullesfeld F-J (1985) Ein Beitrag Zur Harmonischen Darstellung Des Francis O and van Dam T (2002) Evaluation of the precision of using absolute
Gezeitenerzeugenden Potentials. Deutsche Geodaetische Kommission, Reihe C gravimeters to calibrate superconducting gravimeters. Metrologia 39: 485–488.
31: 3–103. Fu Y, Freymueller J, and vanDam T (2012) The effect of using inconsistent ocean tidal
Cartwright DE (1977) Oceanic tides. Reports on Progress in Physics 40: 665–708. loading models on GPS coordinate solutions. Journal of Geodesy 86: 409–421.
Cartwright DE (1999) Tides: A Scientific History. New York: Cambridge University http://dx.doi.org/10.1007/s00190-011-0528-1.
Press. Fu GY and Sun WK (2007) Effects of lateral inhomogeneity in a spherical Earth on
Cartwright DE and Amin M (1986) The variances of tidal harmonics. Deutsche gravity Earth tides. Journal of Geophysical Research 112. http://dx.doi.org/
Hydrographische Zeitschrift 39: 235–253. 10.1029/2006JB004512.
176 Earth Tides

Garrett C and Munk WH (1971) The age of the tide and the Q of the oceans. Deep-Sea King M (2006) Kinematic and static GPS techniques for estimating tidal displacements
Research 18: 493–503. with application to Antarctica. Journal of Geodynamics 41: 77–86.
Gladwin MT and Hart R (1985) Design parameters for borehole strain instrumentation. King GCP and Bilham R (1973) Tidal tilt measurement in Europe. Nature 243: 74–75.
Pure and Applied Geophysics 123: 59–80. King MA and Padman L (2005) Accuracy assessment of ocean tide models around
Goodkind JM (1999) The superconducting gravimeter. Review of Scientific Instruments Antarctica. Geophysical Research Letters 32: L23608. http://dx.doi.org/
70: 4131–4152. 10.1029/2005GL023901.
Gross R (1992) Correspondence between theory and observations of polar motion. King MA, Penna NT, Clarke PJ, and King EC (2005) Validation of ocean tide models
Geophysical Journal International 109: 162–170. around Antarctica using onshore GPS and gravity data. Journal of Geophysical
Guo J, Li Y, Huang Y, Deng H, Xu S, and Ning J (2004) Green’s function of the Research 110: B08401. http://dx.doi.org/10.1029/2004JB003390.
deformation of the Earth as a result of atmospheric loading. Geophysical King GCP, Zürn W, Evans R, and Emter D (1976) Site correction for long-period
Journal International 159: 53–68. http://dx.doi.org/10.1111/j.1365- seismometers, tiltmeters, and strainmeters. Geophysical Journal of the Royal
246X.2004.02410.x. Astronomical Society 44: 405–411.
Haas R and Schuh H (1998) Ocean loading observed by geodetic VLBI. In: Ducarme B Kohl ML and Levine J (1995) Measurement and interpretation of tidal tilts in a small
and Paquet P (eds.) Proceedings of the 13th International Symposium on Earth array. Journal of Geophysical Research 100: 3929–3941.
Tides, pp. 111–120. Brussels: Observatoire Royal de Belgique. Ku LF, Greenberg DA, Garrett C, and Dobson FW (1985) The nodal modulation of the
Harrison JC (1971) New programs for the computation of Earth tides. Internal Technical M2 tide in the Bay of Fundy and Gulf of Maine. Science 230: 69–71.
Report. CIRES, University of Colorado. Kudryavtsev SM (2004) Improved harmonic development of the Earth tide-generating
Harrison JC (1976) Cavity and topographic effects in tilt and strain measurement. potential. Journal of Geodesy 77: 829–838.
Journal of Geophysical Research 81: 319–328. Lambert A (1974) Earth tide analysis and prediction by the response method. Journal of
Harrison JC (1985) Earth Tides. New York: Van Nostrand Reinhold. Geophysical Research 79: 4952–4960.
Harrison JC and Herbst K (1977) Thermoelastic strains and tilts revisited. Geophysical Langbein J (2010) The effect of error in theoretical earth tide on calibration of borehole
Research Letters 4: 535–537. strainmeters. Geophysical Research Letters 37: L21303. http://dx.doi.org/
Harrison JC and LaCoste LJB (1978) The measurement of surface gravity. In: Mueller II 10.1029/2010GL044454.
(ed.) Applications of Geodesy to Geodynamics: Ninth GEOP Conference, Latychev K, Mitrovica J, Ishii M, Chan N, and Davis J (2009) Body tides on a 3-D elastic
pp. 239–243. Columbus, OH: Ohio State University, Department of Geodetic earth: Toward a tidal tomography. Earth and Planetary Science Letters 277: 86–90.
Science, rep. 280. http://dx.doi.org/10.1016/j.epsl.2008.10.008.
Hart RHG, Gladwin MT, Gwyther RL, Agnew DC, and Wyatt FK (1996) Tidal calibration Le Provost C, Bennett AF, and Cartwright DE (1995) Ocean tides for and from Topex/
of borehole strainmeters: Removing the effects of local inhomogeneity. Journal of Poseidon. Science 267: 639–642.
Geophysical Research 101: 25553–25571. Le Provost C, Lyard F, and Molines J (1991) Improving ocean tide predictions by using
Hartmann T and Wenzel H-G (1995) The HW95 tidal potential catalogue. Geophysical additional semidiurnal constituents from spline interpolation in the frequency
Research Letters 22: 3553–3556. domain. Geophysical Research Letters 18: 845–848.
Hatanaka Y, Sengoku A, Sato T, Johnson JM, Rocken C, and Meertens C (2001) Levine J, Meertens C, and Busby R (1989) Tilt observations using borehole tiltmeters:
Detection of tidal loading signals from GPS permanent array of GSI Japan. Journal 1. Analysis of tidal and secular tilt. Journal of Geophysical Research 94: 574–586.
of the Geodetic Society of Japan 47: 187–192. Llubes M and Mazzega P (1997) Testing recent global ocean tide models with loading
Haubrich RA and Munk WH (1959) The pole tide. Journal of Geophysical Research gravimetric data. Progress in Oceanography 40: 369–383.
64: 2373–2388. Longman IM (1959) Formulas for computing the tidal accelerations due to the Moon
Herring TA, Mathews PM, and Buffett BA (2002) Modeling of nutation-precession: Very and the Sun. Journal of Geophysical Research 64: 2351–2355.
long baseline interferometry results. Journal of Geophysical Research 107(B4): Love AEH (1911) Some Problems of Geodynamics. Cambridge: Cambridge University
2069. http://dx.doi.org/10.1029/2001JB000165. Press.
Hewitt E and Hewitt R (1979) The Gibbs–Wilbraham phenomenon: An episode in Lyard F, Lefevre F, Letellier T, and Francis O (2006) Modelling the global ocean tides:
Fourier analysis. Archive for History of Exact Sciences 21: 129–169. Modern insights from FES2004. Ocean Dynamics 56: 394–415.
Hinderer J, Florsch N, Maekinen J, Legros H, and Faller JE (1991) On the calibration of Mathews PM, Buffett BA, and Shapiro II (1995a) Love numbers for diurnal tides:
a superconducting gravimeter using absolute gravity measurements. Geophysical Relation to wobble admittances and resonance expansion. Journal of Geophysical
Journal International 106: 491–497. Research 100: 9935–9948.
Ito T, Okubo M, and Sagiya T (2009) High resolution mapping of Earth tide response Mathews PM, Buffett BA, and Shapiro II (1995b) Love numbers for a rotating spheroidal
based on GPS data in Japan. Journal of Geodynamics 48: 253–259. http://dx.doi. Earth: New definitions and numerical values. Geophysical Research Letters
org/10.1016/j.jog.2009.09.012. 22: 579–582. http://dx.doi.org/10.1029/95GL00161.
Ito T and Simons M (2011) Probing asthenospheric density, temperature, and elastic Mathews PM, Dehant V, and Gipson JM (1997) Tidal station displacements. Journal of
moduli below the western United States. Science 332: 947–951. http://dx.doi.org/ Geophysical Research 102: 20,469–20,477.
10.1126/science.1202584. Mathews PM and Guo JY (2005) Viscoelectromagnetic coupling in precession-nutation
Itsueli UJ, Bilham R, Goulty NR, and King GCP (1975) Tidal strain enhancement observed theory. Journal of Geophysical Research 110(B2): B02402. http://dx.doi.org/
across a tunnel. Geophysical Journal of the Royal Astronomical Society 42: 555–564. 10.1029/2003JB002915.
Iwano S, Fukuda Y, Sato T, Tamura Y, Matsumoto K, and Shibuya K (2005) Mathews PM, Herring TA, and Buffett BA (2002) Modeling of nutation and precession:
Long-period tidal factors at Antarctica Syowa Station determined from 10 years of New nutation series for nonrigid Earth and insights into the Earth’s interior. Journal
superconducting gravimeter data. Journal of Geophysical Research 110: B10403. of Geophysical Research 107(B4): 2068. http://dx.doi.org/10.1029/2001JB000390.
http://dx.doi.org/10.1029/2004JB003551. Matsumoto K, Sato T, Takanezawa T, and Ooe M (2001) GOTIC2: A program for
Jeffreys H (1976) The Earth: Its Origin, History and Physical Constitution. Cambridge: computation of oceanic tidal loading effect. Journal of the Geodetic Society of Japan
Cambridge University Press. 47: 243–248.
Jeffreys H and Vincente RO (1957) The theory of nutation and the variation of latitude. Matsumoto K, Takanezawa T, and Ooe M (2000) Ocean tide models developed by
Monthly Notices of the Royal Astronomical Society 117: 142–161. assimilating Topex/Poseidon altimeter data into hydrodynamical model: A global
Jentzsch G (1997) Earth tides and Ocean tidal loading. In: Wilhelm H, Zürn W, and model and a regional model around Japan. Journal of Oceanography 56: 567–581.
Wenzel HG (eds.) Tidal Phenomena, pp. 145–171. Berlin: Springer-Verlag. Meertens C, Levine J, and Busby R (1989) Tilt observations using borehole tiltmeters:
Jentzsch G, Jahr T, and Kroner C (2009) Special issue on “New challenges in Earth’s 2. Analysis of data from Yellowstone National Park. Journal of Geophysical
dynamics”, Proceedings of the 16th International Symposium on Earth Tides-Joined Research 94: 587–602.
Meeting of Sub-commissions 3.1 on Earth Rotation and Earth Tides, 3.2 on Crustal Meertens CM and Wahr JM (1986) Topographic effect on tilt, strain, and displacement
Deformation, 3.3 on Geophysical fluids, and the Global Geodynamics Project measurements. Journal of Geophysical Research 91: 14,057–14,062.
(GGP), held in Jena, Germany, 1–5 September 2008, Journal of Geodynamics Melchior P (1983) The Tides of the Planet Earth. New York: Pergamon.
48: 107–109. http://dx.doi.org/10.1016/j.jog.2009.09.041. Merriam JB (1980) The series computation of the gravitational perturbation due to an
Kamigaichi O (1998) Green functions for the earth at borehole installation sensor depths ocean tide. Physics of the Earth and Planetary Interiors 23: 81–86.
for surface point load. Papers in Meteorology and Geophysics 48: 89–100. Merriam JB (1985) Toroidal Love numbers and transverse stress at the Earth’s surface.
Khan SA and Scherneck H-G (2003) The M2 ocean tide loading wave in Alaska: Vertical Journal of Geophysical Research 90: 7795–7802.
and horizontal displacements, modelled and observed. Journal of Geodesy Merriam JB (1986) Transverse stress Green’s functions. Journal of Geophysical
77: 117–127. http://dx.doi.org/10.1007/s00190-003-0312-y. Research 91: 13903–13913.
Earth Tides 177

Merriam JB (1992) An ephemeris for gravity tide predictions at the nanogal level. Ray RD, Egbert GD, and Erofeeva SY (2011) Tide predictions in shelf and coastal waters:
Geophysical Journal International 108: 415–422. status and prospects. In: Vignudelli S, Kostianoy AG, Cipollini P, and Benveniste J
Merriam JB (1995) Nonlinear tides observed with the superconducting gravimeter. (eds.) Coastal Altimetry, pp. 191–216. New York: Springer-Verlag.
Geophysical Journal International 123: 529–540. Ray RD and Ponte RM (2003) Barometric tides from ECMWF operational analyses.
Merriam JB (2000) The response method applied to the analysis of superconducting Annales Geophysicae 21: 1897–1910.
gravimeter data. Physics of the Earth and Planetary Interiors 121: 289–299. Ray RD and Poulose S (2005) Terdiurnal surface-pressure oscillations over the
Metivier L, Greff-Lefftz M, and Diament M (2005) A new approach to computing continental United States. Monthly Weather Review 133: 2526–2534.
accurate gravity time variations for a realistic earth model with lateral Ray RD and Sanchez BV (1989) Radial deformation of the earth by oceanic tidal loading.
heterogeneities. Geophysical Journal International 162: 570–574. http://dx.doi.org/ NASA Tech. Mem. TM-100743, Goddard Space Flight Center, Greenbelt, MD.
10.1111/j.1365-246X.2005.02692.x. Richter B, Wilmes H, and Nowak I (1995) The Frankfurt calibration system for relative
Metivier L, Greff-Lefftz M, and Diament M (2006) Mantle lateral variations and gravimeters. Metrologia 32: 217–223.
elastogravitational deformations – I. Numerical modelling. Geophysical Journal Rignot E, Mouginot J, and Scheuchl B (2011) Antarctic grounding line mapping from
International 167: 1060–1076. http://dx.doi.org/10.1111/j.1365- differential satellite radar interferometry. Geophysical Research Letters 38: L10504.
246X.2006.03159.x. http://dx.doi.org/10.1029/2011GL047109.
Miller SP and Wunsch C (1973) The pole tide. Nature Physical Science 246: 98–102. Roeloffs E (2010) Tidal calibration of PBO borehole strainmeters: The roles of vertical
Mitrovica JX, Davis JL, and Shapiro II (1994) A spectral formalism for computing and shear coupling. Journal of Geophysical Research 115: B06405. http://dx.doi.
three-dimensional deformations due to surface loads: 1. Theory. Journal of org/10.1029/2009JB006407.
Geophysical Research 99: 7057–7074. Roosbeek F (1995) RATGP95: A harmonic development of the tide-generating potential
Molodenskii MS (1961) The theory of nutations and diurnal earth tides. Belgian Royal using an analytical method. Geophysical Journal International 126: 197–204.
Observatory Communications Series Geophysics Series 58: 25–56. Rosat S, Hinderer J, Crossley D, and Boy JP (2004) Performance of superconducting
Molodensky SM (1977) The influence of horizontal inhomogeneities in the gravimeters from long-period seismology to tides. Journal of Geodynamics
mantle on the amplitude of tidal oscillations. Izvestiya Physics of the Solid Earth 38: 461–476.
13: 77–80. Rosat S and Lambert SB (2009) Free core nutation resonance parameters from VLBI and
Müller G (1977) Thermoelastic deformations of a half-space: A Green’s function superconducting gravimeter data. Astronomy and Astrophysics 503: 287–291.
approach. Journal of Geophysics 43: 761–769. http://dx.doi.org/10.1051/0004-6361/200811489.
Munk WH and Cartwright DE (1966) Tidal spectroscopy and prediction. Philosophical Sato T and Harrison JC (1990) Local effects on tidal strain measurements at Esashi,
Transactions of the Royal Society A 259: 533–581. Japan. Geophysical Journal International 102: 513–526.
Munk WH and Hasselmann K (1964) Super-resolution of tides. In: Studies on Sato T, Tamura Y, Higashi T, et al. (1994) Resonance parameters of the free core
Oceanography; A Collection of Papers Dedicated to Koji Hidaka in Commemoration nutation measured from three superconducting gravimeters in Japan. Journal of
of His Sixtieth Birthday, pp. 339–344. Tokyo: University of Tokyo Press. Geomagnetism and Geoelectricity 46: 571–586.
Munk WH, Zetler BD, and Groves GW (1965) Tidal cusps. Geophysical Journal Sato T, Tamura Y, Matsumoto K, Imanishi Y, and McQueen H (2004) Parameters of the
10: 211–219. fluid core resonance inferred from superconducting gravimeter data. Journal of
Neuberg J, Hinderer J, and Zürn W (1987) Stacking gravity tide observations in central Geodynamics 38: 375–389. http://dx.doi.org/10.1016/j.jog.2004.07.016.
Europe for the retrieval of the complex eigenfrequency of the nearly diurnal free Schenewerk MS, Marshall J, and Dillinger W (2001) Vertical ocean-loading
wobble. Geophysical Journal of the Royal Astronomical Society 91: 853–868. deformations derived from a global GPS network. Journal of the Geodetic Society of
Padman L and Erofeeva S (2004) A barotropic inverse tidal model for the Arctic Ocean. Japan 47: 237–242.
Geophysical Research Letters 31: L02303. http://dx.doi.org/ Schreiber KU, Klugel T, and Stedman GE (2003) Earth tide and tilt detection by a ring
10.1029/2003GL01900. laser gyroscope. Journal of Geophysical Research 108. http://dx.doi.org/
Padman L, Fricker HA, Coleman R, Howard S, and Erofeeva L (2002) A new tide model 10.1029/2001JB000569.
for the Antarctic ice shelves and seas. Annals of Glaciology 34: 247–254. Shum CK, Woodworth PL, Andersen OB, et al. (1997) Accuracy assessment of recent
Pagiatakis SD (1990) The response of a realistic Earth to ocean tide loading. ocean tide models. Journal of Geophysical Research 102: 25,173–25,194.
Geophysical Journal International 103: 541–560. Smith A, Ambrosius B, Wakker K, Woodworth P, and Vassie J (1997) Comparison
Pawlowicz R, Beardsley B, and Lentz S (2002) Classical tidal harmonic analysis between the harmonic and response methods of tidal analysis using Topex/
including error estimates in MATLAB using T_TIDE. Computers and Geosciences Poseidon altimetry. Journal of Geodesy 71: 695–703. http://dx.doi.org/10.1007/
28: 929–937. s001900050137.
Penna N, Bos M, Baker T, and Scherneck HG (2008) Assessing the accuracy of Smith ML and Dahlen FA (1981) The period and Q of the Chandler wobble. Geophysical
predicted ocean tide loading displacement values. Journal of Geodesy 82: 893–907. Journal of the Royal Astronomical Society 64: 223–281.
http://dx.doi.org/10.1007/s00190-008-0220-2. Sovers OJ (1994) Vertical ocean loading amplitudes from VLBI measurements.
Penna NT and Stewart MP (2003) Aliased tidal signatures in continuous GPS height Geophysical Research Letters 21: 357–360.
time series. Geophysical Research Letters 30. http://dx.doi.org/ Standish EM, Newhall XX, Williams JG, and Yeomans DK (1992) Orbital ephemerides of
10.1029/2003GL018828, SDE 1–1. the Sun, Moon and planets. In: Seidelmann PK (ed.) Explanatory Supplement to the
Pétit G and Luzum B (2010), IERS Conventions (2010), IERS Technical Note 36. Verlag Astronomical Almanac, pp. 279–323. Sausalito, CA: University Science Books.
des Bundesamts für Kartographie und Geodäsie, Frankfurt am Main. Steinberger B and Holme R (2008) Mantle flow models with core–mantle boundary
Petrov L and Ma C (2003) Study of harmonic site position variations determined by very constraints and chemical heterogeneities in the lowermost mantle. Journal of
long baseline interferometry. Journal of Geophysical Research 108. http://dx.doi. Geophysical Research 113: B05403. http://dx.doi.org/10.1029/2007JB005080.
org/10.1029/2002JB001801, ETG 5–1. Takemoto S (1981) Effects of local inhomogeneities on tidal strain measurements.
Ponchaut F, Lyard F, and LeProvost C (2001) An analysis of the tidal signal in the Bulletin of the Disaster Prevention Research Institute, Kyoto University
WOCE sea level dataset. Journal of Atmospheric and Oceanic Technology 31: 211–237.
18: 77–91. Tamura Y (1982) A computer program for calculating the tide generating force.
Ray RD (1998) Ocean self-attraction and loading in numerical tidal models. Marine Publications of the International Latitude Observatory, Mizusawa 16: 1–20.
Geodesy 21: 181–192. Tamura Y (1987) A harmonic development of the tide-generating potential. Bulletin
Ray RD (2001) Resonant third-degree diurnal tides in the seas off Western Europe. d’Information des Marées Terrestres 99: 6813–6855.
Journal of Physical Oceanography 31: 3581–3586. Tamura Y, Sato T, Fukuda Y, and Higashi T (2005) Scale factor calibration of a
Ray RD, Bills BG, and Chao BF (1999) Lunar and solar torques on the oceanic tides. superconducting gravimeter at Esashi Station, Japan, using absolute gravity
Journal of Geophysical Research 104: 17653–17659. measurements. Journal of Geodesy 78: 481–488.
Ray RD, Eanes RJ, and Lemoine FG (2001) Constraints on energy dissipation in the Tamura Y, Sato T, Ooe M, and Ishiguro M (1991) A procedure for tidal analysis with a
Earth’s body tide from satellite tracking and altimetry. Geophysical Journal Bayesian information criterion. Geophysical Journal International 104: 507–516.
International 144: 471–480. Torge W (1989) Gravimetry. Berlin: Walter de Gruyter Verlag.
Ray RD and Egbert GD (2004) The global S1 tide. Journal of Physical Oceanography Trupin A and Wahr J (1990) Spectroscopic analysis of global tide gauge sea level data.
34: 1922–1935. Geophysical Journal International 100: 441–453.
Ray RD and Egbert GD (2012) Fortnightly Earth rotation, ocean tides and mantle Van Camp M and Vauterin P (2005) Tsoft: Graphical and interactive software for
anelasticity. Geophysical Journal International 189: 400–413. http://dx.doi.org/ the analysis of time series and Earth tides. Computers and Geosciences
10.1111/j.1365-246X.2012.05351.x. 31: 631–640.
178 Earth Tides

Wahr JM (1981a) Body tides on an elliptical, rotating, elastic and oceanless Earth. Wilhelm H (1983) Earth’s flattening effect on the tidal forcing field. Journal of
Geophysical Journal of the Royal Astronomical Society 64: 677–703. Geophysics 52: 131–135.
Wahr JM (1981b) A normal mode expansion for the forced response of a rotating Earth. Wilhelm H (1986) Spheroidal and torsional stress coefficients. Journal of Geophysics
Geophysical Journal of the Royal Astronomical Society 64: 651–675. 55: 423–432.
Wahr JM and Bergen Z (1986) The effects of mantle and anelasticity on Wilhelm H, Zürn W, and Wenzel HG (1997) Tidal Phenomena. Berlin: Springer-Verlag.
nutations, earth tides, and tidal variations in rotation rate. Geophysical Journal Xi Q (1987) A new complete development of the tide-generating potential for the epoch
87: 633–668. J2000.0. Bulletin d’Information des Marées Terrestres 99: 6766–6812.
Wahr JM and Sasao T (1981) A diurnal resonance in the ocean tide and in the Earth’s Yuan L and Chao B (2012) Analysis of tidal signals in surface displacement measured
load response due to the resonant free “core nutation” Geophysical Journal of the by a dense continuous GPS array. Earth and Planetary Science Letters
Royal Astronomical Society 64: 747–765. 355–356: 255–261. http://dx.doi.org/10.1016/j.epsl.2012.08.035.
Wang R (1997) Tidal response of the solid Earth. In: Wilhelm H, Zürn W, and Wenzel HG Zürn W (1997a) Earth tide observations and interpretation. In: Wilhelm H, Zürn W, and
(eds.) Tidal Phenomena, pp. 27–57. Berlin: Springer-Verlag. Wenzel HG (eds.) Tidal Phenomena, pp. 77–94. Berlin: Springer-Verlag.
Warburton RJ and Goodkind JM (1976) Search for evidence of a preferred reference Zürn W (1997b) The nearly-diurnal free wobble-resonance. In: Wilhelm H, Zürn W, and
frame. Astrophysical Journal 208: 881–886. Wenzel HG (eds.) Tidal Phenomena, pp. 95–109. Berlin: Springer-Verlag.
Wenzel H-G (1996) The nanogal software: Earth tide data processing package ETERNA Zürn W, Rydelek PA, and Richter B (1986) The core-resonance effect in the record from
3.3. Bulletin d’Information des Marées Terrestres 124: 9425–9439. the superconducting gravimeter at Bad Homburg. In: Viera R (ed.) Proceedings of
Wessel P and Smith WHF (1996) A global, self-consistent, hierarchical, high-resolution the 10th International Symposium on Earth Tides, pp. 141–147. Madrid: Consejo
shoreline database. Journal of Geophysical Research 101: 8741–8743. Superior de Investigaciones Cientificas.

You might also like