EnvPhys CH2 Solar System PDF
EnvPhys CH2 Solar System PDF
SCHAAP
TOPICS IN ENVIRON-
M E N TA L P H Y S I C S : T H E
SOLAR SYSTEM AND
EARTH
2 marcel g. schaap
Copyright © 2015, 2016, 2017, 2018 Marcel G. Schaap
All rights reserved. This book or any portion thereof may not be reproduced or used in any manner
whatsoever without the express written permission of the author or publisher except for the use of brief
quotations in a book review.
Selling class notes and/or other course materials to other students or to a third party for resale is not
permitted without the instructor’s express written consent. Providing student email addresses to a third
party is not permitted. Violations to this and other course rules are subject to the Code of Academic Integrity
and may result in course sanctions. Additionally, students who use D2L or UA email to sell or buy these
copyrighted materials are subject to Code of Conduct Violations for misuse of electronic resources provided
by The University of Arizona. This conduct may also constitute copyright infringement.
ISBN 0-9000000-0-0
Address information
books.dthetadt.com
Contents
3 Bibliography 61
A Critical Elements 65
B Albedo 69
1.1 Introduction
The previous chapter described how the sun emerged from a molecu-
lar cloud, how it produces its energy and discussed the constancy of
this energy source. This chapter will focus on the development and
structure of rest of the solar system and, in particular, Earth. As far
as we know, Earth is the only place where life evolved, though the ex-
istence of microbial life cannot be completely ruled out on Mars and
possibly in the subsurface of some of Jupiter’s or Saturn’s moons. As
we will see in this and other chapters, Earth’s position in the solar sys-
tem is nearly optimal: it is far enough from the sun’s scorching heat,
yet close enough to keep water liquid. In addition, the earth has a mag-
netic field that protects its atmosphere from erosion by the solar wind,
while the angular momentum of our moon serves a vital function in
keeping its rotation axis almost steady in space. Both are needed for
maintaining a stable climate that was suitable for the development of
life.
To get a better understanding of planet Earth, we will therefore
Video how this module fits
study the formation of the solar system, and -as a prelude to under- in the course:https://d2l.
standing the fundamentals of climate models later in this book- de- arizona.edu/content/enforced/
velop a simple model can predict the surface temperature of planets. 950156-867-2204-1ENVS420520001101/
02_Solar_System/HD_Intro_hd_720.mp4
• The Solar system formed out of the protoplanetary disk that sur-
rounded the proto-sun. The protoplanetary disk consisted of large
amounts of hydrogen and helium gas with a minor amount of dust
consisting of various kinds of ice, metals and silicates.
Video introduction to this chapter,
• There are four rocky planets in the inner solar system: Mercury, and origin of the solar system: https:
//d2l.arizona.edu/content/enforced/
Venus, Earth, and Mars. These formed relatively close to the sun 950156-867-2204-1ENVS420520001101/
where it was too hot to preserve much gas or ice. 02_Solar_System/M2S1_hd_720.mp4
10 marcel g. schaap
• Beyond the snow line where water ice remained stable, there was
enough gas to form the giants Jupiter and Saturn. The ice giants
Uranus and Neptune formed in the more distant and even colder
part of the solar system where there was abundant gas and ice.
• Smaller metallic, rocky and icy asteroids also formed among the
planets, particularly between the orbits of Mars and Jupiter. Comets
consist predominantly of ice and originated in the outer solar sys-
tem beyond the orbit of Neptune.
• Volatile elements include H, N, C, and the noble gases He, Ne, Ar,
Kr, Xe. Most high-density siderophile elements (Fe, Ni, Mn, and
some precious metals) sank to the core. Lithophile elements gen-
erally bind strongly with oxygen and form relatively light minerals
that are found in the crust and mantle, while chalcophile elements
prefer sulfur and form denser minerals located in ore bodies.
“ice” not only means water ice (H2 O), but also other volatile substances
such as carbon dioxide and -monoxide (CO2 and CO), methane (CH4 )
and ammonia (NH3 ), the melting points of which are listed in Table
1.1.
In a process called accretion, electrically charged dust grains merged
into clusters, which produced self-gravitating aggregates that after
many gentle collisions and mergers produced kilometer-sized planetes-
imals (Steinpilz et al., 2019). The composition of these planetesimals
depended on the distance from the proto-sun (Figure 1.2) and hence
the local temperature of the protoplanetary disk.
Silicates and metal Ice (H2O, CO, CO2, NH3, CH4) Gas
ble at correspondingly larger distances from the sun. Dust in the outer
solar system therefore was able to attract and retain large amounts gas
and ice, which facilitated a rapid coalescence of planetesimals (Cieza
et al., 2016).
The gas giants Jupiter and Saturn formed near the snow line by grav-
itational attraction of free hydrogen and helium gas onto their rocky
and icy cores. Farther out, where more ice was available the somewhat
smaller ice giants Uranus and Neptune formed. Although these plan-
ets predominantly contain hydrogen and helium, they have a much
higher content of water, ammonia and methane than the gas giants,
owing to their formation in the colder regions of the solar system. The
blueish color of Uranus and Neptune (Figure 1.4) is due to methane in
the atmospheres of these planets.
The planetary accretion process was halted after tens million years
because the increasingly bright sun generated solar winds strong enough
to remove most of the remaining gas from the solar system. Some of
the original planetesimals still remain at distances far beyond the orbit
of Neptune. In 2019, NASA’s New Horizons probe few by the as-
teroid MU69 (also named Arrokoth). This system is a contact-binary
(Figure 1.3), indicating a very gentle merger of two separate planetesi-
mals. The topography of the larger body suggests previous mergers of
smaller bodies.
pre-solar nebula. This indicates that most of Earth’s water was derived
from the planetesimals that formed Earth and was not derived from
comet impacts after the collision with Theia. Yet, it appears necessary
to invoke an impact of a Theia-sized body to explain the amount of
carbon in Earth’s mantle and crust (Li et al., 2016b; Budde et al., 2019)
as well as the almost identical isotopic compositions of the earth and
the moon. Additional evidence indicates that primordial water may
have survived the collision of proto-Earth and Theia (Lin et al., 2016;
Mercury
Williams et al., 2019). Some of these theories are contradictory and it is
clear that there are still many unknowns about formation of the early
solar system, including Earth and the origin of its water.
Venus
Things settled down a few hundred million years after the formation Moon
of the solar system with the planets located near their present day
orbits. Figure 1.4 and Table 1.2 provide an overview of the major fea-
tures of the solar system, while its structure is depicted in Figure 1.5. Mars
The distance of a planet from the sun is given in astronomical units Ceres
(AU), which is defined by the International Astronomical Union as
Vesta
149,597,870,700 meters exactly, or nearly 150 million kilometers. The
AU is closely related (but not quite identical) to Earth’s distance from
the center of the sun. Jupiter
the surface (quartz and feldspar minerals have densities of about 2650
kg/m3 ). These planets have sizable cores of iron and nickel. Mars has
a much lower density (3934 kg/m3 ) because it has a comparatively Neptune
much small iron.
As will be discussed in more detail in one of the next chapters,
Venus, Earth and Mars have atmospheres, while Mercury has none
due to its low gravity and exposure to an intense solar wind. Even Charon Pluto
though it has a mass and radius almost like Earth’s, Venus’ atmo-
sphere is extreme: it is more than 96% CO2 and contains sulfuric acid
Figure 1.4: Overview of the solar system.
clouds. The surface pressure of Venus is 92 bar, and its surface tem- The images are not to scale but scaled
perature is well over 400 °C. Liquid water (and life) cannot presently such to show some detail. Jupiter is
about ten times larger in diameter than
exist on Venus, though it is possible that more hospitable conditions Earth. All images indicate roughly how
were present several billion years ago (Way et al., 2016). the planet would look to the human eye,
except Venus which is shown in ultravi-
olet to exhibit some details in its other-
wise featureless cloud deck.
topics in environmental physics: the solar system and earth 15
Mars, too, has an atmosphere that is mostly CO2 but with a surface
pressure less than 1/100th of that of Earth. As a result, surface tem-
peratures only occasionally rise above the freezing point of water near
equator. Liquid water cannot exist in stable form due to the low sur-
face pressure, though there is evidence for ephemeral channels. Large
water ice and CO2 ice deposits exist in Mars’ subsurface, with perma-
nent icecaps at the poles that are visible through small telescopes (see
Figure 1.4).
While orbiting the sun, each planet also rotates around a rotation Solar system structure and
axis. For Mars and Earth a full rotation takes about a 24 hours (see general characteristics of plan-
ets, including albedo: https:
Table 1.2). Mercury and Venus rotate much slower: it takes Mercury
//d2l.arizona.edu/content/enforced/
nearly 59 days to rotate (i.e., a Mercury “day” lasts 59 Earth days). 950156-867-2204-1ENVS420520001101/
Venus rotates in 243 Earth days but orbits in 0.62 years or 226 Earth 02_Solar_System/M2S4_hd_720.mp4
Beyond Mars, we find the main asteroid belt (the white dots in the top
panel of Figure 1.5) which is an area of thousands of rocky and some-
times icy bodies that are essentially leftovers from the formation of the
solar system. The largest of these are Vesta and Ceres, which have radii
that are 263 and 473 km, respectively (Figure 1.4). Although these bod-
ies are much smaller than Earth, or even our moon (radius: 1737 km),
Vesta and Ceres had sufficient mass to gravitationally form roughly
spherical bodies Most other asteroids are smaller and less massive,
and consequently these bodies have irregular shapes. The total mass
of the asteroid belt is 0.0006 Earth masses or about two Ceres masses
(Krasinsky et al., 2002).
Jupiter is the main reason for the existence of the asteroid belt be-
tween it and Mars: Jupiter’s gravity caused so much disturbance in
this region that no true planet could ever form by prolonged accretion.
The many bodies that formed established an equilibrium between ac-
cretion and occasional collisions that would again break up the plan-
etesimals.
The bottom panel of Figure 1.5 shows an oblique view of the solar
system, with the asteroid belt as a flattened disk of tens of thousands
of bodies (only 200 are shown here). This disk indicates that the orig-
inal flattened structure of the protoplanetary disk (Figure 1.1) is still
16 marcel g. schaap
Because their lower mass limits the internal pressure (necessary for
hydrostatic equilibrium), no metallic hydrogen layer is expected to
exist on Uranus and Neptune. Instead, the cores is thought to be
surrounded by water, ammonia, and methane. The upper layers are
predominantly hydrogen, helium and methane gases. The methane
content of Uranus and Neptune is larger than that of Jupiter and Sat-
urn, which is a result of the colder conditions in the protoplanetary
disk where these ice giants formed (see also Table 1.1).
Mercury 0.055 0.387 0.2408 0.206 7.0 2439.7 58.646 0.034 3.7 ~0 5427 9093 0.106 0.068
Venus 0.815 0.723 0.6152 0.007 3.4 6,051.9 -243.013 2.64 8.87 9.2 5243 2605 0.65 0.77
MPa
Earth 1 1.000 1.0000 0.017 0 6,378.1 0.99727 23.44 9.81 101.3 5513 1361 0.367 0.306
kPa
Mars 0.107 1.523 1.8808 0.093 1.9 3396.1 1.02595 25.19 3.71 ~0.8 3934 586.8 0.17 0.25
Vesta 2.6×10-4 2.362 3.6294 0.099 7.1 262.7d 0.2226 ≈29l 0.25 0 3456 243.9 0.423 0.20j
Ceres 9.4×10-4 2.767 4.6040 0.116 10.6 473d 0.3781 ≈4m 0.29 0 2160 177.8 0.09 0.24i
Pluto 0.002 39.48 247.92 0.249 17.2 1151 6.3872 119.6 0.66 ~1 Pa 2050 0.8730 0.5-0.7 0.4-0.6
Moon 0.0123 383,399 km 27.32d 0.055 5.1 1737 27.32 1.67 1.62 10-7 3346 1361 0.12 0.11
Pa
Table 1.2: Selected properties of the planets in the solar system. Abbreviations: AU: Astronomical Unit (1 AU is 149,597,870.700 km, by definition); Ecc.: eccentricity; Incl.:
inclination of the orbit with respect to the ecliptic plane (Earth: 0 by definition); g: acceleration of gravity; Press: surface pressure (gas and ice giants have no surface); S(d):
planet-specific solar constant as a function of distance, d (Eq. 1.1). References and notes: a : NASA (2015b); b : pdf file aprx_pos_planets.pdf listed on NASA (2015a), for the
moon we list the distance to Earth in km; c : Sheppard (2015); d : mean; e : gas/ice planets have no solid surface and therefore no defined surface pressure; f : NASA (2016); l :
bond albedo from Li et al. (2016a); j : Li et al. (2013); k : bottom of photosphere; ’: Thomas et al. (1997); m : Schorghofer et al. (2016).
19
20 marcel g. schaap
Tectonic uplift
(mountains and
plateaus) Figure 1.9: Simplified depiction of the
Sediments Mid-Oceanic upper mantle, asthenosphere and the
Ridge oceanic and continental lithosphere (not
Ocean
to scale).
Oceanic crust Continental crust
Astenosphere
Mantle
Sedimentary
basins
ally more stable than oceanic crust. The stablest sections of conti-
nental lithosphere are known as cratons and are billions of years old.
Oceanic crust is much younger than continental crust (less than 200
million years) because it is constantly renewed at mid-oceanic ridges
(1.9) while it disappears by subduction into the mantle under the less
dense granitic continental crust (Chopin, 2003). Extension at mid-
oceanic ridges and sinking motion near subduction zones is in the
order of millimeters to centimeters per year. Continental and oceanic
crusts are usually covered with sediments that are erosion products
of primary crustal rocks or other sediments. Metamorphic rocks are
formed in the lithosphere where temperatures and pressures are high
enough to physically and chemically alter igneous and sedimentary
rock.
The ocean and atmosphere are the least dense of all and therefore
situated above the crust. As discussed in the following section, both
contain volatile elements with low melting points.
The formation of the solar system and the differentiation of Earth also
explain the abundance of elements in Earth’s crust. Figure 1.10 shows
that Earth’s elemental abundance exhibits both similarities and differ-
ences with the abundance of elements in the solar system that was
discussed in the previous module.
It is useful to use the Goldschmidt (1937) classification into volatile,
chalcophile, lithophile and siderophile elements as indicated by the back-
ground colors in Figure 1.10. The primary reason for this division is
due to the electron configuration of the elements which determines
their place in the periodical system, their chemical reactivity but also
Goldschmidt N Density
their density (Table 1.4). To normalize cosmic and crustal abundances Classification kg/m3
both series to the same scale, both were by their respective abundance Volatile 8 0.210
of silicon, one of the dominant elements in the crust. Data for Earth’s Lithophile 47 1.124
Chalcophile 17 3.960
inaccessible mantle and core are less reliable and are not shown in the Siderophile 13 7.869
figure.
Table 1.4: Goldschmidt classification
with mean elemental densities. The den-
1.5.1 Volatiles sity is given as an abundance-weighted
quantities (abundance as found for
Earth’s crust). N indicates the number
Volatile elements are noble gases (He, Ne, Ar, Kr, Xe), or form com- of elements in each class.
pounds such as H2 O, N2 , CO2 , all with low melting and boiling points
and therefore subject to degassing from the mantle and crust. Relative
to the cosmic abundance, volatile elements are depleted in the crust
but dominant in the oceans and atmosphere. Carbon is an exception.
Even though it is classified as a volatile, most of Earth’s is present as
organic matter (kerogen) and limestone in Earth’s crust due to biolog-
topics in environmental physics: the solar system and earth 23
1.5.2 Lithophiles
1.5.3 Chalcophiles
Chalcophile elements form strong bonds with sulfur, rather than the
chemically similar oxygen (both elements belong to group VI in the
periodical system). Chalcophiles tend to form ore bodies of insoluble
sulfides, which concentrates these minerals and facilitates mining op-
erations. Sulfide minerals tend to be more dense than lithophile ele-
ments (Table 1.4) and are generally located deeper in the crust or in
the upper mantle.
1.5.4 Siderophiles
Most of Earth’s iron and nickel settled into core. Siderophile elements
are able to dissolve in the liquid outer core or amalgamate into the
solid Fe − Ni core. Siderophile elements are therefore all depleted in
the crust, relative to the cosmos. For example, gold and platinum as
24 marcel g. schaap
well as some other elements are (very) rare in the crust, but they are
likely enriched in the core because of their siderophillic nature.
Iron itself, however, is not particularly depleted in the crust and
it is actually more abundant than some common lithophile elements
such as K, Ca and Ti. The relative abundance of Fe in the crust is
a consequence of iron-containing ultra-mafic volcanism that occurred
during the first few billion years. The ubiquitous iron ore bodies
presently found in Earth’s crust are actually a result of the evolution of
cyanobacteria that use oxygenic photosynthesis to produce O2 which
reacted with soluble ferric (Fe2+ ) into insoluble iron Fe2 O3 and Fe3 O4
(hematite and magnetite) that precipitated out in oceans as banded
iron formations.
Pm
Am
Tm
Mn
Ga
Na
Co
Cu
Rb
Nb
Rh
Ho
Re
Np
Ag
Sb
Cs
Eu
Au
Pa
Sc
As
Tb
Ac
Tc
La
Lu
Ta
Br
Pr
Cl
At
Fr
Al
Bi
In
Tl
Li
H
N
B
Y
F
Ir
I
He
Be
C
O
Ne
Mg
Si
S
Ar
Ca
Ti
Cr
Ni
Zn
Ge
Se
Kr
Sr
Zr
Mo
Ru
Pd
Cd
Sn
Xe
Ba
Ce
Nd
Sm
Gd
Dy
Er
Yb
Hf
W
Os
Pt
Hg
Pb
Rn
Ra
Th
U
Pu
Fe
Po
Te
Earth’s crust
Si=1 classification into volatile, chalcophile,
Volatile
Chalcophile
lithophile and siderophile is provides as
1e+00 background color. A white background
Lithophile
Siderophile indicates that the element is unstable
Unstable
and does not occur in significant quan-
1e−03 tities. The downward pointing trian-
gles under the element labels (top) in-
dicate that the U.S. Government consid-
1e−06 ers these elements (or their compounds)
critical for national security and econ-
omy (Appendix A)
1e−09
1 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85 90 95
Atomic number (Z)
The sun is the single energy source in the solar system. Some of the
Introduction and context of
sun’s energy is absorbed by the planets and converted into heat, giving a simple planetary surface
each a characteristic mean surface temperature. It is therefore interest- temperature model: https:
//d2l.arizona.edu/content/enforced/
ing to digest the data in Table 1.2 and to formulate a simple model
950156-867-2204-1ENVS420520001101/
that predicts the surface temperature of a planet using the solar en- 02_Solar_System/M2S3_hd_720.mp4
ergy received, the planet’s albedo, and the law of Stefan-Boltzmann
for black-body radiation. With such a model, we can quickly analyze
why Earth is so unique, but also immediately indicate why Earth re-
quires an atmosphere greenhouse effect to support liquid water and life.
The central assumption in our model is that the amount of energy
received from the sun by a planet equals the amount of energy that
is re-radiated back into the universe as infrared black-body radiation.
This is also known as thermal equilibrium. A planet would warm up if
it receives more solar radiation than it emits as infrared (IR) back into
space. Conversely, the planet would cool down if its IR emission is
larger than the radiation it receives from the sun. In both cases, the
change in temperature would cause a change in IR emission and re-
sult in a new thermal equilibrium. We therefore have a simple balance:
energy absorbed by the planet equals the energy emitted as IR radia-
tion, which is controlled by the temperature of our planet. Figure 1.11
provides a schematic overview of the model.
n
Re
day night
L⊙
S(d) = (1.1)
4πd2
where L⊙ is the sun’s luminosity (3.828 × 1026 J/s). Planet-specific
values of the solar constant S(d) are listed in Table 1.2.
The amount of energy absorbed by a planet is proportional to its
solar constant times the area that intercepts the radiation (i.e. the day-
light side), less the amount that is reflected back into space without
ever being “used”. Albedo (A) is the fraction of radiation reflected
back into space.
There are two types of albedo: Geometric albedo (A g ) used for Figure 1.12: From the perspective of the
sun, spherical planets look like disks
imaging purposes, and Bond albedo (Ab ), which is useful for energy (shadows). The intercepted solar radi-
balance calculations, such as done here (also see Appendix B). Values ation is therefore proportional to πR2 ,
with R being the radius of the planet.
for the Bond albedo listed in Table 1.2 indicate that no planet is a per-
Shown are (left to right): Mercury,
fect reflector (Ab = 1), nor a perfect absorber (Ab = 0). Figure 1.11 Venus, Earth and Moon, and Mars.
represents incoming solar radiation and reflected solar radiation with
triangular arrows. The amount of energy absorbed by Earth is:
Where Re is the radius of Earth, and πR2e is the area that intercepts
the solar radiation (see also Figure 1.11). Note that this corresponds to
the area of a disk and not the area of a hemisphere because the planets
“look” like disks from the perspective of the sun (see Figure 1.4).
Let us now assume that the amount of absorbed solar radiation is
Video about the development of
converted into heat that gives our planet a surface temperature (Tsur f ). the temperature model. https:
Most solid or liquid materials objects emit black body radiation that is //d2l.arizona.edu/content/enforced/
proportional to the fourth power of their temperature. This energy is 950156-867-2204-1ENVS420520001101/
02_Solar_System/M2S5_hd_720.mp4
radiated into space as infrared (IR) radiation and allows the planet
to maintain thermal equilibrium by balancing the absorbed solar ra-
diation. We will assume that there is no storage of absorbed energy
(e.g. heat stored in oceans, rocks, soils or the atmosphere) and, impor-
tantly, we assume that the entire spherical surface of our planet has
one uniform temperature. The IR emission is therefore uniform over
the surface as indicated by the red barbed arrows in Figure 1.12.
We must, however, immediately realize the limitations of this model.
Especially the assumption of isothermal conditions is unrealistic be-
cause it ignores the usually lower temperatures at the night side of the
planet (Figure 1.11). There would also be a problem near the poles
where the rays of the sun will come in more obliquely, thus reducing
the intensity of solar radiation which results in less heating and less IR
emission. We also ignore the axis tilt of a planet, which would cause
opposite seasons on each hemisphere as the planet goes around its or-
bit. However, a uniform planetary temperature is an assumption we
are presently willing to make in order to keep our model simple.
topics in environmental physics: the solar system and earth 27
4
Ebb = ǫσsb Tsur f (1.3)
4 2
Eout = σsb Tsur f 4πRe (1.4)
Ein = Eout
(1.5)
4 4πR2
(1 − Ab )S(d)πR2e = σsb Tsur f e
The term πR2e on the left-hand side (lhs) and right-hand side (rhs) can-
cels and can be removed. Because we would like to know the surface
temperature, we must now solve for Tsur f . After some simple rear-
rangement of the symbols, we arrive the following expression:
1
(1 − A b ) S ( d )
4
Tcalc = Tsur f = (1.6)
4σsb
This model tells us that the surface temperature depends on the solar
constant (S(d)), which is different for each planet, see Table 1.2) and
the planet’s Bond albedo (Ab ) and no other variables (σsb is a physical
constant).
How useful is our model? This can be evaluated by comparing
Video about an evaluation of
predicted model outcomes with reality (observations). To this end we the temperature model. https:
make explicit that Tcalc = Tsur f (Eq. 1.6) is a model prediction, while //d2l.arizona.edu/content/enforced/
Tobs is the observed surface temperatures (i.e reality). Table 1.5 lists the 950156-867-2204-1ENVS420520001101/
02_Solar_System/M2S6_hd_720.mp4
values for eight planets and three other solar system bodies as well
as the disagreement between observations and calculations, Tdi f f =
Tobs − Tcalc . A positive value indicates that the model overestimates
the real temperature, a negative value implies an underestimate.
The temperature difference should be zero for a perfect model, but
by glancing at Table 1.5 it is clear that the model certainly is not per-
fect (e.g., Venus). On the other hand, the small differences for some
28 marcel g. schaap
Earth 1361 0.306 254 288 34 ets do not have a solid surface.
Mars 586.8 0.25 210 208 -2
Jup.a 50.28 0.343 110 163 53
Sat.a 14.97 0.342 81 133 52
Uran.a 3.696 0.3 58 78 20
Nep.a 1.505 0.29 47 73 26
Ceres 177.8 0.24 156 168 13
Vesta 243.9 0.20 171 177 6
Pluto 0.8730 0.5 37 44 7
bodies indicate that our simple model works quite well in some cases
(Mercury, Pluto, Ceres, Mars). So, the model has at least some merit.
It is evident that the model works poorly for larger bodies. We
already mentioned Venus where the observations are 510 K hotter than
modeled, while Earth is 34 K warmer. Jupiter and Saturn are also
substantially warmer by about 50 K. As will be discussed further in
a next chapter, both Venus and Earth have atmospheres that exhibit
greenhouse effects (which is extreme for Venus), allowing the trapping
of heat. The situation is different for Jupiter and Saturn: even after 4.6
billion years these planets have still not reached thermal equilibrium
state and have large internal sources of heat. This make these planets
warmer than calculated from a solar energy supply alone.
Even though our surface temperature model is not very accurate (it
is much too simple, really), in some cases it is useful because the differ-
ence with the observations tells us that we have overlooked something,
namely a planet’s greenhouse effect or internal heat production. The
discrepancy between the model outcomes and observations can there-
fore give us hints where the model can be improved.
The model is also useful in a qualitative sense. It predicts that an
increase in S will lead to an increase in temperature while an increase
in albedo (Ab ) will lead to a temperature decrease. Both are demon-
strated with an experiment on a scale model in Appendix C. This
aspect of the model is relevant when studying Earth because more
clouds and larger ice-caps lead to more reflection of sunlight (increase
in albedo) and therefore lower temperatures. Conversely, shrinking ice
caps, snow fields, and sea ice tend to decrease the albedo and lead to
more absorption of solar radiation. In the following chapters we will
discuss refinements to the model, such as the greenhouse effect, global
circulation in atmosphere and oceans.
The usefulness of the model is also demonstrated by a major recent
topics in environmental physics: the solar system and earth 29
No planet in the solar system is quite like another, but it appears that
Earth is truly unique in the sense that life evolved there. Several factors
are important. First of all Earth has an enormous amount of liquid wa-
ter with an active hydrological cycle. The water is not only necessary
for life, but also allows for efficient, mixing and transport of nutrients
and dissolved gases. Water allows for weathering of rocks and (facili-
tated by oceanic micro-organisms) deposition of excess carbon-dioxide
as limestone.
Earth has liquid water due to its distance from the sun and an at-
mosphere with a pressure that is sufficiently high. As a result of a
moderate greenhouse effect, its surface temperatures are neither too
cold such that all water freezes (Mars), nor too hot (Venus) such that all
water eventually vaporizes. Without a greenhouse effect, Earth’s water
would probably frozen solid too, because our temperature model in-
dicated that Earth’s mean temperature “should” be 254 K (-18 °C). The
atmospheric pressure is important because it raises the boiling point of
liquids. Mars’ surface pressure (about 0.8 kPa, or less than 1 millibar)
is too low to permit liquid water at its surface.
Earth’s magnetic field is also crucial as it protects our atmosphere
from the solar wind. Mars, on the other hand, lost its magnetic field,
most of its atmosphere and most of its liquid water early in its his-
tory. The water that remained froze solid in its subsurface or its ice-
caps. The erosion of its atmosphere continues into the present as doc-
umented recently by NASA’s Maven orbiter (Jakosky et al., 2015) and
further analysis by Fedorova et al. (2020). Although after Earth, Mars
is the most hospitable planet for human settlement, it is unlikely that
we will be able to create Earth-like conditions (“Terraform”) on Mars
- even if we had the technology to do it. Recent evidence indicates
that there is insufficient CO2 in Mars’ icecaps to create a sufficiently
strong greenhouse effect for plants or humans to survive at the surface
(Jakosky and Edwards, 2018).
Based on mass and size, Venus is the most Earth-like planet in the
solar system (Table 1.2), yet its atmospheric conditions are character-
30 marcel g. schaap
2.1 Introduction
It would be very convenient if Earth and other bodies in the solar sys-
tem orbited the sun with constant circular motions without any mutual
effects and without any change. Unfortunately, this is not the case. We
have already seen that Jupiter’s gravity had a dominant role in shaping
of the structure of the early solar system by affecting the orbits of other
planets. This continues to this day: all planets exert small gravitational
forces continuously change the orbital characteristics of, especially, the
less massive bodies in the solar system. The result of this is that all
planets follow quasi-periodically changing elliptical (non-circular) or-
bits around the sun (Figure 2.1).
Planets also spin around their respective rotation axes, which are
tilted (oblique) with respect to the plane of their orbits, thus caus-
ing the seasons that are so apparent on Earth. Slow orbital and ax-
ial changes can change the solar intensity (insolation), particularly at
higher northern and southern latitudes which has the potential to
cause cold and warm climatic periods (glacials and interglacials, re-
spectively).
The interplay of orbits, axial tilt and their changes is complicated,
and we will discuss these one by one. To this end, we will first discuss
the most important characteristics that are important for seasonality on
Earth. With seasonality we mean the variation of solar energy input
throughout the year which drives the annual pattern in weather, ecol-
ogy, and agricultural production. We will see that there are global ef-
fects and latitude-specific effects. We will then study the slow changes
in orbital parameters and discuss their significance for long-term quasi
periodic changes in our global climate.
tance from the sun, which is controlled by time and the eccentricity
of Earth’s orbit.
If you could go high above the North Pole you would observe that
the earth is orbiting the sun in a counter clockwise (CCW) direction
(Figure 2.1) and that the orbit is not circular, but slightly elongated.
You would further notice that the earth is also spinning in a CCW
fashion around its rotation axis. This axis connects north and south pole
and is perpendicular to the equatorial plane that separates the northern
and southern hemispheres (abbreviated as NH and SH, respectively).
It would also be clear that Earth’s rotation axis is tilted with respect
to the plane of the orbit. The elliptical orbit combined with the axis
tilt brings about seasons on Earth. The following sections will describe
the most important concepts in more detail.
Ecliptic plane
Earth orbit Eq
ua
tor
ne
la
lp
ia
or
t
ua
eq
rotation axis
The ecliptic plane is the plane in which Earth orbits around the sun
(Figure 2.4). All other planets circumscribe the sun in planes that make
a small angle (inclination) with the ecliptic plane (see the “Inclination”
column in Table 1.2). The ecliptic plane is therefore a convenient, but
geocentric reference plane. As seen from Earth, the sun would appear
to travel past the stars that are fixed on an infinitely far celestial sphere.
From the perspective of the sun, Earth would appear to travel past the
same stars half a year later.
Because the orbital planes of the other planets have only small an-
gles with the ecliptic plane (see again Table 1.2) they travel roughly
along similar paths on the celestial sphere. Their speeds are different
and consistent with their orbital periods thus causing an ever-changing
configuration of planets along the “ecliptic”. Traditionally, star group-
ings along the ecliptic were divided into 12 constellations (signs) of the
zodiac which, however, no longer align with the 13 constellations that
are presently recognized. The term “ecliptic plane” finds its origin in
34 marcel g. schaap
the fact that solar or lunar eclipses occur when the moon crosses the
ecliptic while in line with Earth and Sun.
semi minor
and eccentricity (e = 0.5), r: the distance
between Earth and the sun, A: Aphelion,
axis (b)
Earth P: perihelion. The red arrows indicate
relative velocities.
Sun r
(focus)
center θ
A P
semi major axis (a)
ae
a(1+e) a(1-e)
2.2.2 Eccentricity
Eccentricity is an index of non-circularity of an elliptical orbit, which is
characterized by a major (longest) axis (a) and a perpendicular minor Video about elliptical orbits. https:
(shortest) axis (b, see also Figure 2.2). Eccentricity, e, is defined as: //d2l.arizona.edu/content/enforced/
950156-867-2204-1ENVS420520001101/
r 02_Solar_System/M2S8_hd_720.mp4
b2
e= 1− (2.1)
a2
A zero eccentricity indicates a perfectly circular orbit while larger val-
ues cause an increased “stretching” of the ellipse. Note in Figure 2.2
that the sun is not in the center of the ellipse (the intersection of the
major and minor axes), but located in the focus which is offset from
the center by a factor ae. There is a second focus on the other side of
the center that is empty. Eccentricity therefore determines how much
the focus is displaced from the center. Table 1.2 provides the eccentric-
ity of the planets orbits as well as their characteristic “distance” from
the sun, which is the semi-major axis (i.e., one-half of the major axis).
The orbit’s elliptical shape and the off-center position of the sun
cause the Sun-Earth distance to vary throughout the year according to:
a 1 − e2
r= (2.2)
1 + e cos θ
topics in environmental physics: the solar system and earth 35
Where θ is the angle of Sun-Earth (focus) with the major axis (Figure
2.2). The closest point is reached around January 3rd and is called the
perihelion (P in Figure 2.2). Around July 4th the Sun-Earth distance is
largest and is called the aphelion (A in Figure 2.2).
The fact that Earth has an elliptical orbit around the sun means that
the Sun-Earth distance varies throughout the year, according to Equa-
tion 2.2. On January 3, when θ equals 0° and Earth is at perihelion, the
Sun-Earth (center-center) distance is therefore 147.1 million km, while
on July 4’th the distance is 152.1 million km (when Earth is at aphelion
and θ equals 180°, Figure 2.2).
• In equal periods of time, planets carve out equal area slices from
the elliptical “pie” (Fig. 2.3).
Gm1 m2
Fg = (2.3)
r2
Where Fg is the mutual gravitational force (in Newtons) of two bodies
with mass m1 and m2 separated by distance r. The force of gravity
becomes the centripetal force that holds body 2 (say, Earth) in orbit
around body 1 (the sun). The centripetal force depends on the square
velocity v and distance r:
mv2
Fc = (2.4)
r
When we set Fg = Fc , we obtain:
m2 v2 Gm1 m2
= (2.5)
r r2
and thus:
Gm1
v2 = (2.6)
r
36 marcel g. schaap
This simply states that the square of the orbital speed of body 2 de-
pends on the mass of the body 1 and the distance between the centers
of mass of both bodies. This is only true when m1 ≫ m2 . When this
is not the case m1 and m2 orbit a common center of mass known as
the barycenter. The Earth-Moon barycenter is inside the Earth, 4,671
km from its center. The Pluto-Charon barycenter is outside Pluto and
about 1000 above Pluto’s surface.
Equation 2.6 tells us that the square of orbital speed increases pro-
portionally to the mass of the parent body and decreases inversely
proportionally with their distance. Planets further from the sun (or
satellites further from Earth) move with slower speeds.
The orbital period of a body is the time it takes to complete a full
orbit. Assuming a circular orbit with a circumference 2πr, the period
P is:
2πr
P= (2.7)
v
Kepler’s third law can be found by squaring Eq. 2.7 and combining
this with Eq. 2.6:
4π 2 r3
P2 = (2.8)
Gm1
Although this equation was derived for a circular orbit, it remains
valid when the orbit is elliptical, in which case the semi-major axis
(a) should be used instead of r. This results in P2 ∝ a3 as stated in
4π 2
Kepler’s third law with a constant of proportionality equal to Gm . A
1
more practical version of this equation is:
s
a3
P = 2π (2.9)
Gm1
This implies that the orbital period only depends on the semi-major
axis and not on the eccentricity (e) or semi-minor axis (b). This conclu-
sion is important later when we discuss changes in Earth’s orbit.
Kepler’s first law (elliptical orbits) can be derived by allowing ki- 2
plies that the planet will move faster during perihelion and slower
during aphelion. When a planet orbits the sun in an elliptic orbit, nei-
ther the distance nor its speed are constant. However, the area of the
“pie-slices” carved out in equal time are constant (see Figure 2.3).
The difference in velocity near perihelion and aphelion causes the Figure 2.3: Kepler’s second law ex-
plained in graphical form. Because the
time between the fall solstice and spring solstice to be shorter than the distance to the sun and the planet’s or-
time between the spring solstice and fall solstice (by about two days). bital velocity changes, the colored “pie”
section near the perihelion (P) has the
This is the main reason why February has 28 (or 29) days while July
same area as the narrow colored area
and August have 31 days. near the aphelion (A). The time it takes
to go from 1 to 2 and from 3 to 4 is the
same. The red arrows indicate relative
velocities.
topics in environmental physics: the solar system and earth 37
NH summer NH spring
equinox NH winter Figure 2.4: Solstices, aphelion and peri-
solstice (June 21)
solstice (Dec 21)
helion dates in Earth’s orbit.
Perihelion (Jan 3)
Aphelion (July 4)
l plane
Equatoria
23.44 °
NH fall
Earth axis tilt
Ecliptic plane equinox
It may surprise the reader that Earth’s rotation period is not ex-
actly 24 hours, but listed in Table 1.2 as 0.9927 days instead (23 hours,
56 minutes and 4.1 seconds). This is called sidereal rotation, which is
the time needed for an exact 360 degree rotation. This rotation time
is determined by measuring the time it takes between two meridian
crossings (i.e., stars being exactly south).
The sidereal rotation of Earth is subject to slight quasi-random fluc-
tuations due to changes on Earth (melting glaciers, ocean and wind
currents, tectonics) and because of the occurrence of tides in Earth’s
oceans (Appendix D). A smaller effect is caused by tides in the oceans
which gradually slow down the rotation of our planet, but increase the
semi-major axis of the moons orbit around Earth.
An Earth day (or solar or synodic day ) is 24 hours (on average) be-
cause for most daily life we use the sun as a reference for our time
(not the stars). The nearly four minute difference is due to Earth’s or-
bit around the sun: while a full (360 degree) rotation it takes 23 hours,
Figure 2.5: Star tracks around the north
56 minutes and 4.1 seconds, it takes an extra 4 minutes if we take the celestial pole in a long-exposure photo-
sun as our reference star. This apparent discrepancy is caused by the graph. Credit: LCGS Russ.
fact that during a day Earth also moves in its orbit (Top left diagram
in Figure 2.6), which means that it must rotate about 361 degrees for
a solar day, rather than the 360 degrees necessary for a sidereal day.
38 marcel g. schaap
2.3 Seasons
The axis tilt and the elliptical orbit explain the occurrence of seasons on
Earth. Both hemispheres always have opposing seasons (if it is winter
in one, it is summer in the other), but there are also strong latitudinal
differences in the intensity of the seasons as will be discussed next.
topics in environmental physics: the solar system and earth 39
2.3.2 Declination
Because Earth’s axis is tilted by 23.44° the sun will make an angle (or
declination) with the equatorial plane that varies between 23.44° and
-23.44° as the earth goes around its orbit (Figures 2.4, and 2.6). There
are four distinct dates in the year:
• The northern solstice occurs around June 21 when the solar decli- Video about solar declination with
Stellarium demonstration. https:
nation reaches 23.44°. At this date the sun is above the latitude of //d2l.arizona.edu/content/enforced/
Cancer (the tropic of Cancer) and it is summer in the northern hemi- 950156-867-2204-1ENVS420520001101/
02_Solar_System/M2S11_hd_720.mp4
sphere, but winter in the southern hemisphere. Days are longer
than 12 hours in the NH but shorter than 12 hours in the SH.
The tropics of Cancer (Capricorn) are named such because the north-
ern (southern) solstice occurred when the sun was in the constella-
tion of Cancer (Capricorn). However, that was 2500 years ago, due to
axis precession (discussed later) the northern solstice occurs in Gem-
ini, while the southern solstice occurs in Sagittarius. Note that the
northern (southern) solstice is close in time to -but does not coincide
with- the aphelion (perihelion) of Earth. This distinction is important
when studying orbital changes and ice ages.
The angle of the sun with the equatorial plane varies between -23.44
and 23.44° irrespective of the location on Earth. The effect of the solar
40 marcel g. schaap
t + 10
δ (t) = −23.44 cos(360 ) (2.10)
365
where t is the time in (possibly fractional) days since midnight on Jan-
uary 1 (=0.0); t+10 merely approximates the number of days since the
NH winter solstice, which occurs around December 21. This equa-
tion assumes that Earth’s orbit is perfectly circular and that one orbit
is exactly 365 days. As long as it is not used for applications that
need high-precision values (e.g. navigation) this equation is practical
because Earth’s orbital eccentricity is small.
Solar declination is a global value, but because we live on a sphere
we are commonly interested in the angle that the sun makes with the
local horizon. This is called the elevation angle and is useful because
it provides us with the angle by which the sun’s rays will strike the
ground. To further pinpoint the position in the local sky we need az-
imuth which provides the angle of the sun with the geographic north.
For any given NH latitude north of the tropic of Cancer the daily
noontime solar elevation angle is given by:
June 21
Figure 2.7: Simulated view of the noon-
time sky in Tucson, Arizona (+32 de-
grees NH latitude) for June 21 and De-
cember 21 (NH summer and winter sol-
stice). The top diagram shows the blue
sky that prevents use from seeing the
stars. The second and third diagrams
show the sky with atmospheric light
scattering taken out, showing the back-
ground stars with an equatorial grid.
Solar declination is +23 and -23 de-
grees in June and December 21, respec-
June 21 tively. The yellow lines indicate the
path taken by the sun on those days,
+40
showing that sunrise/sunset is not in
the east/west. The bottom two dia-
Betelgeuse grams show the same dates and times,
+30 but now with an azimuthal grid, indi-
cating that the elevation angle changes
during the day, reaching a maximum
+20 at noon. These images were generated
+10 0 -10 -20 -30 -40 by the open-source package Stellarium
(http://stellarium.org/).
December 21
+40
+30
+20
+10 0 -10 -20 -30 -40
+20
+10
+20
+10
0
42 marcel g. schaap
aphhelion
solstice
solstice
Lat.
90
Day length (hours)
18 60
30
0
12 −30
−60
−90
6
0
0 50 100 150 200 250 300 350
Day of Year (doy)
As discussed there are three factors that control the seasons on Earth:
the axis tilt, latitude, and eccentricity of Earth’s orbit. The solar dec-
lination in combination with latitude has a much stronger local effect
on the insolation than eccentricity of Earth’s orbit, which determines
variability of the total energy input to the planet.
It probably does not need much explanation that at noontime the
Video about orbit, axis tilt,
sun is always near the zenith at low latitudes (i.e. tropical latitudes be- latitude and insolation https:
tween the tropics of Cancer and Capricorn) and always rather low in //d2l.arizona.edu/content/enforced/
the sky at far northern or southern latitudes. The amount of solar en- 950156-867-2204-1ENVS420520001101/
02_Solar_System/M2S10_hd_720.mp4
ergy that reaches the surface of the earth is proportional to the sine of
the elevation angle. When the sun’s elevation angle is low, the sunlight
is spread out over a larger area, as indicated by the yellow bars striking
Earth in Figure 2.9. This diagram ignores significant and variable ef-
fects in Earth’s atmosphere, such as reflection (clouds), scattering and
topics in environmental physics: the solar system and earth 43
Ca Ca
Night
pri
cor Sun pri
cor
n n
Day Day Night
~1.017 AU ~0.983 AU
SH Winter not to scale SH Summer
absorption of light.
At the equator, the variation is relatively small as shown in the right-
hand panel of Figure 2.10. For example, at the equinoxes when the
declination is 0° the noon-time elevation angle is 90° and the incom-
ing energy is S (the solar constant). During the northern solstice the
amount of energy received is equal to sin(90 − 23.44) × S = 0.92 × S,
thus only 8% less. However, at 60 degree latitude the incoming energy
at noon is sin(90 − 60 + 23.44) × S = 0.80 × S at the summer solstice,
but only 0.11 × S during the southern solstice (an 85% variation, see
Eq. 2.11). In addition, there are larger variations in day length the
farther north or south you go from the equator (see Figure 2.8).
Latitudes outside the tropics do not only see less insolation because
the incoming sunlight is spread out over a larger area, but also experi-
ence a larger annual variability. Seasonality therefore tends to increase
at more extreme northern and southern latitudes. The situation is, of
course, most extreme at the poles: here we essentially have days and
nights that last up to half a year (Figure 2.8).
A surprising fact is that during the summer solstice, the total daily Video about solar power dis-
tribution on Earth and climate
incoming energy is greatest of all latitudes (Figure 2.10). The 24-hour differences between northern and
day length at the pole (2.8) compensates for the rather low elevation southern hemispheres. https:
angle. Conversely, near the winter solstice no solar energy is available //d2l.arizona.edu/content/enforced/
950156-867-2204-1ENVS420520001101/
whatsoever, and temperatures at the surface and in the atmosphere 02_Solar_System/M2S12_hd_720.mp4
can reach very low values, especially at higher altitudes near the south
pole. These conclusions are not just trivia as they are highly relevant
in explaining the melting of arctic ice, or the dynamics of antarctic
stratospheric ozone.
While insolation dynamics are driven by latitude, Earth’s axial tilt
and orbital eccentricity, these are not the only factors that determine
seasonality at any given terrestrial location: altitude, proximity to large
water bodies, and land cover also play a significant role in reducing or
enhancing the effect of isolation-driven seasonality. For example, one
would expect that the southern hemisphere experiences more season-
44 marcel g. schaap
ality because the southern solstice roughly aligns with the perihelion
(occurring on December 21st and Jan 4’th, respectively); the northern
solstice aligns with the aphelion. However, measurements indicate
that temperature fluctuations are less extreme than that of the north-
ern hemisphere (bottom panel in Figure 2.11). The explanation for this
contradiction is threefold:
• Due to Kepler’s second law Earth actually moves faster in its orbit
near its perihelion and slower near its aphelion causing the period
with low global insolation (during NH summer and SH winter) to
last longer than the period with high global insolation (NH winter
and SH summer).
• The albedo of snow and ice on Antarctica is rather high, which al-
lows the SH reflect solar radiation more efficiently and which limits
absorption of solar radiation during the SH summer (compare the
polar regions in the top and middle panels of Figure 2.11 and see
also equation 1.6).
What we have seen so far is that the sun is crucial in keeping the
earth warm. Although our simplistic radiation balance model (Eq.
1.6) is not very good, it provides a conceptual understanding that the
Earth is kept warm by solar insolation. We can also make it plausible
that variation in insolation is caused by Earth’s axis tilt, latitude and
orbital eccentricity as the main drivers of seasons.
On a human time scale it is possible to assume that eccentricity Introductory video about relation
between orbit, axis and climate. https:
and axial tilt are constant. However, these parameters vary consider- //d2l.arizona.edu/content/enforced/
ably over periods of thousands of years, along with the celestial di- 950156-867-2204-1ENVS420520001101/
rection with of Earth’s rotation axis as well as orientation of the orbit 02_Solar_System/M2S13_hd_720.mp4
itself. Variation in these features over time will cause significant sea-
sonal variation in insolation and is one reason for the occurrence of ice
ages in the past few millions of years. These orbital changes are com-
monly known as Milankovitch cycles, after the Serbian scientist who, in
ow )
(n ky
the early 20th century, identified the complex gravitational interaction
)
67 27
01 (+
among Earth and other celestial bodies that result in slow changes
0. 27
)
20 3
ky
(- 50
00
7
0
0.
0.
in the orbit, axis direction and tilt. Modern computation techniques A e=0 P
have superseded Milankovitch’ original hand-calculations and in the
following we will use work by Berger and Loutre (1991) and Berger
(1992) to discuss four major orbital factors that control the intensity
and timing of insolation.
The duration of the year does not change when eccentricity varies. Table 2.1: Sidereal, Tropic and Anomal-
The orbital period for an elliptic orbit around the sun is defined by Eq. istic years in days. One day is defined as
2.9 (Kepler’s third law). Eccentricity does not appear in this equation 86,400 seconds.
0
,00
22 and 22,000.
ner (see Top right-hand diagram in Figure 2.6). Therefore, the time
between perihelion passages (the anomalistic year Table 2.1) is a lit-
tle longer than the sidereal year. In other words: Earth needs to play
“catch up” with the perihelion (because the motion is in the same di-
rection), which takes almost 5 minutes per year. Figure 2.13 shows the
relative orbital positions in the year 2000, 12000 and 22000, with Earth
positioned at the aphelion in 2000 but lagging behind in 12,000 and
22,000.
The significance of apsidal precession is that the perihelion moves
with respect to the sidereal year. The effect of apsidal precession is that
perihelion (or aphelion) would occur 5 minutes later each sidereal year.
This implies that the timing of the maximum in global insolation (at
perihelion) changes slowly through the years. See below for a stronger
and opposite effect due to precession of Earth’s rotation axis.
Even so, the 2 degree variation does have a significant effect on Earth’s
climate: a larger value means that the axis is tilted more towards or
away from the sun at the solstices. An increased tilt therefore means
more insolation extremes during the summer solstices. Conversely, a
22.1°
decreased tilt means that there is less variation in insolation between 24.5°
winter and summer solstices. Seasonality in solar energy input will
therefore be increased at high angles and reduced at low tilt angles.
As the tilt of rotation axis changes, so does the angle of the equatorial
plane with the ecliptic (since the axis is perpendicular to the equator).
As a result, the tropics of Cancer and Capricorn will shift in latitude
Eq
ua
according to the axis tilt, but the equator itself will not be displaced. tor
Rotation (1 day)
CW Axis precession
26,000 y
(Table 2.6): the sidereal year, the anomalistic year, and the tropical
year. Of these, the sidereal year has no effect on the timing of Earth’s
seasons (it is simply a full, 360 degree orbit), but the other two do.
The anomalistic year relates to apsidal precession and therefore the
timing of the perihelion of Earth’s orbit. The tropical year depends
on axial precession. Because axis tilt has a much stronger effect on
Earth’s seasonality than when perihelion and aphelion occur, we use
the tropical year for Earth’s civilian calendar. If we used the sidereal
year or anomalistic year the solstices would slowly occur at a later or
earlier date. Because the tropical year is shorter than the sidereal year
the solstices and equinoxes move relative to the fixed stars and are the
main reason why the zodiac no longer lines up with the astrological
signs. Figure 2.15 illustrates the effect of axis precession at the aphelion
in the context of Earth’s orbit around the sun.
Orbit precession
Figure 2.16: Orbit and Axial precession
combined. The earth orbits CCW around
4000 5000 the sun in one sidereal year, while the
6000 earth orbit itself also precesses CCW in
3000 112 ky. Axis precession makes the NH
-14,000 winter solstice move (precess) CW along
-13,000 the orbit (26 ky) NH winter solstice hap-
2000 -12,000 pens somewhat earlier in each year, at a
-11,000 different distance from the sun. The ec-
-10,000 centricity of this diagram is exaggerated
-9000 for display purposes. The position of the
-8000 summer solstice relative to the perihelion is
ω measured by the angle ω for between the
1000 -7000
-6000 year -14,000 and + 6000 CE.
-5000
-4000
-3000
0
-2000
-1000
Axis precession
change your time around the race track, but there is a difference with
relation to each other.
Figure 2.16 illustrates how the summer solstice precesses around the
ellipse. Because the perihelion is presently occurring around January
3, there is a large angle (ω) between the summer solstice and the per-
ihelion. NH insolation is more intense when ω = 0 (northern solstice
occurs at perihelion) and less intense when the northern solstice occurs
at aphelion (note this was the case seven centuries ago, Figure 2.16).
When we take the cosine of ω we get a variable that varies between -1
and 1 when the solstice occurs at aphelion or perihelion, respectively
(see Figure 2.21, second panel).
Eccentricity can now be used as a scale factor that modulates the per-
ihelion and NH solstice alignment as e cos ω. When e is low, the effect
of perihelion and NH solstice alignment is negligible (0). In this case
it doesn’t really matter where the NH summer solstice occurs in the
orbit, because the orbit is almost circular (see Figure 2.12). However,
when e is large, e cos ω reaches a maximum value when NH solstice
occurs around the perihelion and a minimum (negative) value when
the SH solstice occurs at the perihelion.
With the above we therefore managed to “collapse” three effects
(eccentricity, orbit and axis precession) into one compound variable:
e cos ω. This is shown in the third panel of Figure 2.21, which now has
the periodical patterns of eccentricity and precession interfering with
each other.
All that remains now is to incorporate the effects of axial tilt (also
known as obliquity and displayed in the fourth panel of Figure 2.21),
which can further enhance e cos ω when obliquity is high (more pro-
nounced differences between seasons) or mute its effect when obliq-
uity is low (less extreme seasons across the globe).
We do not discuss the complex calculation of July insolation at 65
N (see Berger and Loutre, 1991 and Berger, 1992), but it appears in
the fifth panel of Figure 2.21. July’s insolation during the past 500 ky
varied by about 12% around a mean value of 440 W/m2 . July and Jan-
uary insolation near the tropics varied somewhat less by 10% around
a mean value of 455 W/m2 (not visible in the figure). These numbers
are much smaller than the solar constant (1361 W/m2 ) because they
factor in the sun’s elevation angle (which varies throughout the day)
and also the fact that the sun doesn’t shine at night.
Also shown here in blue is the January insolation at 65 South, which
-as might be expected- has an opposite effect (or phase): when July in-
sulation at 65 N is at a maximum, January insolation at 65 S (summer)
is at a minimum. This poses a direct problem: the figure indicates
that Milankovitch stimulation of ice ages should alternate between the
hemispheres. According to these data, a cold period (glacial) in the
52 marcel g. schaap
18 ky BP Present
Figure 2.17: Ice extent at 18 ky BP (left)
and in the present (right). Land ice is
white and sea ice is gray. Light green
colors denote coastal plains, while light
blue indicate shallow oceans. The sea
level is significantly lower during ice
ages because part of the ocean water
is “stored” as land ice. Adapted from:
http://earthobservatory.nasa.gov/
Features/BorealMigration/boreal_
migration2.php
Glacial ice
Sea ice
elemental isotopes, and dust was stored in annual layers (Figure 2.19).
From these cores it is possible to reconstruct the dramatic shifts be-
tween the warm and cold periods. For example, the temperature dif-
ference between the deepest lows (e.g. about 20 ky BP) and the highest
maximums is more than 10 centigrade (see black curve in the bottom
left panel in Figure 2.21). We refer the reader to NICL (2017) for an ex-
cellent overview of ice-core sampling, storage and the procedure used
topics in environmental physics: the solar system and earth 53
1.0
to get data from the collected ice.
Temperature swings on land and in the atmosphere were responsi- 0.8
cold ocean water can dissolve more CO2 , as shown in Figure 2.18 for an
0.4
atmospheric CO2 concentration of 280 ppm (parts per million). When
conditions cool, CO2 will be transferred from the atmosphere to the 0.2
and
One of the oxygen atoms in CaCO3 is therefore derived from the ocean
water.
When it is colder foraminifera will therefore build an increased
amount of 18 O into their calcareous skeletons, allowing the 18 O en-
richment history to become part of the geological record. Careful ex-
54 marcel g. schaap
The data in the bottom panel 2.21 are invaluable in comparing the
present-day climate with more typical values in the late Quaternary.
There is something odd, however, when you compare the pattern of
insolation with the actual response in temperature, CO2 and d18 O: the
pattern of insolation at 65 NH has a dominant period of about 20 ky
(there are about 4 to 5 full sine waves per 100 ky). Earth’s temperature,
CO2 and 18 O responses, however, appear to have dominant period of
about 100 ky. It therefore appears that -even though there is stimulus
in insolation with a period 20 ky- Earth’s climate, ice sheets and ocean
systems respond with a time constant that is about 5 times slower.
To investigate the differences in orbital stimuli and the earth system Video about a comparison be-
response, it is interesting to look at the data patterns in an alternative tween cycles in model predictions
and those in observed data. https:
way. The panels on the left-hand side of Figure 2.21 all display time
//d2l.arizona.edu/content/enforced/
series, which means they give the pattern of a variable or phenomenon 950156-867-2204-1ENVS420520001101/
over time. Such patterns can also be thought to consist of periodic 02_Solar_System/M2S22_hd_720.mp4
65N insolation
2 CO2
300 d18O
0
250
−2
200
CO2 (ppmv)
Rel. T (oC)
−4 150
−6 100
−8 50
−10 0
−500 −400 −300 −200 −100 0 10 20 40 100 200 400
time before present (1950, in ky) period (ky)
Figure 2.21: Orbital drivers of insolation and earth system response. The left panel shows (from top to bottom) the last 500 thousand years of: eccentricity, cos ω, e cos ω,
obliquity (axis tilt), insolation at 65 N and 65 S (black and blue); and combined in the bottom diagram: relative temperature (black), CO2 concentration (red), and d18 O (blue).
The right-hand panel shows the spectra of the items in the left-hand panel, normalized to the strength of the strongest peak. Vertical gray lines in these diagrams correspond
to periods of 23, 29, 37 and 103 thousand years. The right-hand panels break down the time-based signals of the left-hand panels into their constituent periodic cycles. Peaks
thus show dominant periods, valleys absence of cycles. The right-hand essentially contain the same information as the left-hand panels, except that they did not preserve the
relative phase of the cycles. One striking effect is the absence of the 103 ky peak in the insolation spectrum which appears as a dominant peak in temperature, CO2 and d18 O.
See text for further explanation. Data credits d18 O: Lisiecki and Raymo (2005), CO2 : Lüthi et al. (2008), deuterium: Jouzel et al. (2007), orbital parameters: Berger and Loutre
(1991) and Berger (1992)
55
56 marcel g. schaap
-2 1 4 7 10 mm/y
The slow response of especially the N. American crust is a major
factor in transforming the 19-41 ky cycles in insolation into the much
longer 100 ky cycle. It simply takes time for the crust to react to a
building ice-sheet and, similarly, it takes time for the crust to rebound
up once an ice-sheet has melted away.
58 marcel g. schaap
Abe-Ouchi, A., Saito, F., Kawamura, K., Raymo, M. E., Okuno, J., Takahashi, K., Blatter, H., 2013. Insolation-driven 100,000-year glacial
cycles and hysteresis of ice-sheet volume. Nature 500 (7461), 190–193.
URL https://doi.org/10.1038/nature12374
Altwegg, K., Balsiger, H., Bar-Nun, A., Berthelier, J.-J., Bieler, A., Bochsler, P., Briois, C., Calmonte, U., Combi, M., De Keyser, J., et al.,
2015. 67p/churyumov-gerasimenko, a jupiter family comet with a high d/h ratio. Science 347 (6220), 1261952.
Berger, A., 1992. Orbital variations and insolation database. IGBP PAGES/World Data Center-A for Paleoclimatology Data Contribution
Series 92 (007).
URL https://www.ncdc.noaa.gov/paleo-search/study/5776
Berger, A., Loutre, M.-F., 1991. Insolation values for the climate of the last 10 million years. Quaternary Science Reviews 10 (4), 297–317.
URL https://doi.org/10.1016/0277-3791(91)90033-q
Budde, G., Burkhardt, C., Kleine, T., 2019. Molybdenum isotopic evidence for the late accretion of outer solar system material to earth.
Nature Astronomy.
URL https://doi.org/10.1038/s41550-019-0779-y
Chapront, J., Chapront-Touzé, M., Francou, G., 2002. A new determination of lunar orbital parameters, precession constant and tidal
acceleration from llr measurements. Astronomy & Astrophysics 387 (2), 700–709.
URL https://doi.org/10.1051/0004-6361:20020420
Chopin, C., 2003. Ultrahigh-pressure metamorphism: tracing continental crust into the mantle. Earth and Planetary Science Letters
212 (1-2), 1–14.
URL https://doi.org/10.1016/s0012-821x(03)00261-9
Cieza, L. A., Casassus, S., Tobin, J., Bos, S. P., Williams, J. P., Perez, S., Zhu, Z., Caceres, C., Canovas, H., Dunham, M. M., et al., 2016.
Imaging the water snow-line during a protostellar outburst. Nature 535 (7611), 258–261.
URL https://doi.org/10.1038/nature18612
EOC, 2016.
URL http://hpiers.obspm.fr/eop-pc/index.php
Fedorova, A. A., Montmessin, F., Korablev, O., Luginin, M., Trokhimovskiy, A., Belyaev, D. A., Ignatiev, N. I., Lefèvre, F., Alday, J., Irwin,
P. G. J., Olsen, K. S., Bertaux, J.-L., Millour, E., Määttänen, A., Shakun, A., Grigoriev, A. V., Patrakeev, A., Korsa, S., Kokonkov, N.,
Baggio, L., Forget, F., Wilson, C. F., Jan. 2020. Stormy water on mars: The distribution and saturation of atmospheric water during
the dusty season. Science, eaay9522.
URL http://science.sciencemag.org/content/early/2020/01/08/science.aay9522.abstract
Geruo, A., Wahr, J., Zhong, S., 2013. Computations of the viscoelastic response of a 3-d compressible earth to surface loading: an
application to glacial isostatic adjustment in antarctica and canada. Geophysical Journal International 192 (2), 557–572.
Gillon, M., Triaud, A. H. M. J., Demory, B.-O., Jehin, E., Agol, E., Deck, K. M., Lederer, S. M., de Wit, J., Burdanov, A., Ingalls, J. G.,
Bolmont, E., Leconte, J., Raymond, S. N., Selsis, F., Turbet, M., Barkaoui, K., Burgasser, A., Burleigh, M. R., Carey, S. J., Chaushev,
A., Copperwheat, C. M., Delrez, L., Fernandes, C. S., Holdsworth, D. L., Kotze, E. J., Van Grootel, V., Almleaky, Y., Benkhaldoun, Z.,
Magain, P., Queloz, D., Feb. 2017. Seven temperate terrestrial planets around the nearby ultracool dwarf star trappist-1. Nature 542,
456.
URL https://doi.org/10.1038/nature21360
62 marcel g. schaap
Goldschmidt, V. M., 1937. The principles of distribution of chemical elements in minerals and rocks. the seventh hugo müller lecture,
delivered before the chemical society on march 17th, 1937. Journal of the Chemical Society (Resumed), 655–673.
URL https://doi.org/10.1039/jr9370000655
Gomes, R., Levison, H. F., Tsiganis, K., Morbidelli, A., 2005. Origin of the cataclysmic late heavy bombardment period of the terrestrial
planets. Nature 435 (7041), 466–469.
Hallis, L. J., Huss, G. R., Nagashima, K., Taylor, G. J., Halldórsson, S. A., Hilton, D. R., Mottl, M. J., Meech, K. J., 2015. Evidence for
primordial water in earth’s deep mantle. Science 350 (6262), 795–797.
URL https://doi.org/10.1126/science.aac4834
Imbrie, J., Imbrie, K. P., 1986. Ice ages: solving the mystery. Harvard University Press.
Jacobson, S. A., Rubie, D. C., Hernlund, J., Morbidelli, A., Nakajima, M., 2017. Formation, stratification, and mixing of the cores of earth
and venus. Earth and Planetary Science Letters 474, 375–386.
URL https://doi.org/10.1016/j.epsl.2017.06.023
Jakosky, B. M., Edwards, C. S., 2018. Inventory of co2 available for terraforming mars. Nature Astronomy 2 (8), 634.
URL https://doi.org/10.1038/s41550-018-0529-6
Jakosky, B. M., Lin, R. P., Grebowsky, J. M., Luhmann, J. G., Mitchell, D. F., Beutelschies, G., Priser, T., Acuna, M., Andersson, L., Baird,
D., et al., 2015. The mars atmosphere and volatile evolution (maven) mission. Space Science Reviews 195 (1-4), 3–48.
URL https://doi.org/10.1007/s11214-015-0139-x
Jouzel, J., Masson-Delmotte, V., Cattani, O., Dreyfus, G., Falourd, S., Hoffmann, G., Minster, B., Nouet, J., Barnola, J., Chappellaz,
J., et al., 2007. Epica dome c ice core 800kyr deuterium data and temperature estimates. IGBP PAGES/World Data Center for
Paleoclimatology data contribution series 91, 2007.
Krasinsky, G. A., Pitjeva, E. V., Vasilyev, M. V., Yagudina, E. I., 2002. Hidden mass in the asteroid belt. Icarus 158 (1), 98–105.
URL https://doi.org/10.1006/icar.2002.6837
Li, J.-Y., Le Corre, L., Schröder, S. E., Reddy, V., Denevi, B. W., Buratti, B. J., Mottola, S., Hoffmann, M., Gutierrez-Marques, P., Nathues,
A., et al., 2013. Global photometric properties of asteroid (4) vesta observed with dawn framing camera. Icarus 226 (2), 1252–1274.
URL https://doi.org/10.1016/j.icarus.2013.08.011
Li, J.-Y., Reddy, V., Nathues, A., Le Corre, L., Izawa, M. R. M., Cloutis, E. A., Sykes, M. V., Carsenty, U., Castillo-Rogez, J. C., Hoffmann,
M., et al., 2016a. Surface albedo and spectral variability of ceres. The Astrophysical Journal Letters 817 (2), L22.
Li, Y., Dasgupta, R., Tsuno, K., Monteleone, B., Shimizu, N., 9 2016b. Carbon and sulfur budget of the silicate earth explained by
accretion of differentiated planetary embryos. Nature Geosci advance online publication, letter.
URL http://dx.doi.org/10.1038/ngeo2801
Lin, Y., Tronche, E. J., Steenstra, E. S., van Westrenen, W., Nov. 2016. Evidence for an early wet moon from experimental crystallization
of the lunar magma ocean. Nature Geoscience 10, 14.
URL https://doi.org/10.1038/ngeo2845
Lisiecki, L. E., 2010. Links between eccentricity forcing and the 100,000-year glacial cycle. Nature geoscience 3 (5), 349.
URL https://doi.org/10.1038/ngeo828
Lisiecki, L. E., Raymo, M. E., 2005. A pliocene-pleistocene stack of 57 globally distributed benthic δ18o records. Paleoceanography 20 (1).
URL https://doi.org/10.1029/2004pa001071
Lüthi, D., Le Floch, M., Bereiter, B., Blunier, T., Barnola, J.-M., Siegenthaler, U., Raynaud, D., Jouzel, J., Fischer, H., Kawamura, K., et al.,
2008. High-resolution carbon dioxide concentration record 650,000–800,000 years before present. Nature 453 (7193), 379–382.
URL https://doi.org/10.1038/nature06949
Mann, A., 2018. Bashing holes in the tale of earth’s troubled youth. Nature 553 (7689).
URL https://doi.org/10.1038/d41586-018-01074-6
Meyers, S. R., Malinverno, A., 2018. Proterozoic milankovitch cycles and the history of the solar system. Proceedings of the National
Academy of Sciences.
URL http://www.pnas.org/content/early/2018/05/30/1717689115
Morbidelli, A., Levison, H. F., Tsiganis, K., Gomes, R., 2005. Chaotic capture of jupiter’s trojan asteroids in the early solar system.
Nature 435 (7041), 462–465.
URL https://doi.org/10.1038/nature03540
topics in environmental physics: the solar system and earth 63
Morlighem, M. E., Rignot, E., Mouginot, J., Seroussi, H., Larour, E., 2015. Icebridge bedmachine greenland.
URL https://doi.org/10.5067/AD7B0HQNSJ29
NASA, 2015a. Keplerian elements for approximate positions of the major planets.
URL http://ssd.jpl.nasa.gov/?planet_phys_par
Nimmo, F., 2002. Why does venus lack a magnetic field? Geology 30 (11), 987–990.
URL https://doi.org/10.1130/0091-7613(2002)030<0987:wdvlam>2.0.co;2
NIST, 2016.
URL http://physics.nist.gov/cuu/Units/second.html
Paillard, D., 2015. Quaternary glaciations: from observations to theories. Quaternary Science Reviews 107, 11–24.
URL https://doi.org/10.1016/j.quascirev.2014.10.002
Paulson, A., Zhong, S., Wahr, J., 2007. Inference of mantle viscosity from grace and relative sea level data. Geophysical Journal Interna-
tional 171 (2), 497–508.
URL https://doi.org/10.1111/j.1365-246x.2007.03556.x
Petit, J.-R., Jouzel, J., Raynaud, D., Barkov, N. I., Barnola, J.-M., Basile, I., Bender, M., Chappellaz, J., Davis, M., Delaygue, G., et al., 1999.
Climate and atmospheric history of the past 420,000 years from the vostok ice core, antarctica. Nature 399 (6735), 429–436.
URL https://doi.org/10.1038/20859
Rufu, R., Aharonson, O., Perets, H. B., 1 2017. A multiple-impact origin for the moon. Nature Geosci advance online publication, article.
URL http://dx.doi.org/10.1038/ngeo2866
Schorghofer, N., Mazarico, E., Platz, T., Preusker, F., Schröder, S. E., Raymond, C. A., Russell, C. T., 2016. The permanently shadowed
regions of dwarf planet ceres. Geophysical Research Letters 43 (13), 6783–6789.
URL https://doi.org/10.1002/2016gl069368
Steinpilz, T., Joeris, K., Jungmann, F., Wolf, D., Brendel, L., Teiser, J., Shinbrot, T., Wurm, G., 2019. Electrical charging overcomes the
bouncing barrier in planet formation. Nature Physics.
URL https://doi.org/10.1038/s41567-019-0728-9
Stevenson, D. J., 1987. Origin of the moon-the collision hypothesis. Annual review of earth and planetary sciences 15, 271–315.
URL https://doi.org/10.1146/annurev.earth.15.1.271
Thomas, P. C., Binzel, R. P., Gaffey, M. J., Zellner, B. H., Storrs, A. D., Wells, E., 1997. Vesta: Spin pole, size, and shape from hst images.
Icarus 128 (1), 88–94.
URL https://doi.org/10.1006/icar.1997.5736
Touma, J., Wisdom, J., Feb. 1993. The chaotic obliquity of mars. Science 259 (5099), 1294.
URL http://science.sciencemag.org/content/259/5099/1294.abstract
Tsiganis, K., Gomes, R., Morbidelli, A., Levison, H. F., 2005. Origin of the orbital architecture of the giant planets of the solar system.
Nature 435 (7041), 459–461.
URL https://doi.org/10.1038/nature03539
64 marcel g. schaap
Wade, J., Wood, B. J., 2016. The oxidation state and mass of the moon-forming impactor. Earth and Planetary Science Letters 442,
186–193.
URL https://doi.org/10.1016/j.epsl.2016.02.053
Way, M. J., Del Genio, A. D., Kiang, N. Y., Sohl, L. E., Grinspoon, D. H., Aleinov, I., Kelley, M., Clune, T., 2016. Was venus the first
habitable world of our solar system? Geophysical Research Letters 43 (16), 8376–8383, 2016GL069790.
URL http://dx.doi.org/10.1002/2016GL069790
Willacy, K., Alexander, C., Ali-Dib, M., Ceccarelli, C., Charnley, S. B., Doronin, M., Ellinger, Y., Gast, P., Gibb, E., Milam, S. N., et al.,
2015. The composition of the protosolar disk and the formation conditions for comets. Space Science Reviews 197 (1-4), 151–190.
URL https://doi.org/10.1007/978-94-024-1103-4_8
Williams, C. D., Mukhopadhyay, S., Rudolph, M. L., Romanowicz, B., Jul. 2019. Primitive helium is sourced from seismically slow
regions in the lowermost mantle. Geochem. Geophys. Geosyst. 0 (0).
URL https://doi.org/10.1029/2019GC008437
A
Critical Elements
column). Helium is the only element that can be used to attain tem-
perature conditions that approach zero kelvin and is therefore essen-
tial for super-conducting conditions necessary in many research and
medical applications (e.g. MRI imaging). The helium concentration
in the atmosphere, but some natural gas reservoirs have percentage
levels of He due to the radioactive decay of uranium or thorium in
sediments or crystalline base rocks. The US is the largest supplier of
He in the world with the largest known (or disclosed) reserves. Zirco-
nium finds widespread use as ceramic cladding of fuel rods in nuclear
reactors, and is therefore important in nuclear weapons production,
among many other uses.
topics in environmental physics: the solar system and earth 67
light is barely visible near the western coast of Colombia, South Amer-
ica, but much brighter four hours later when it is further out west on
the Pacific Ocean. However, the view from GOES 17, which is above
the pacific is not nearly as bright, even though it was taken at the same
time as the center image. This implies that reflected sunlight is more
intense at lower incoming and viewing angles.
The effect of light intensity, albedo, and the greenhouse effect dis-
cussed in Section ?? can be illustrated with a simple experiment as
shown in Figure C.1.
a b c d
C.2 Experiment
At point “a” (Figure C.1) the floodlight was still in its off setting and no
polycarbonate spheres are present. The thermistors indicated that the
temperature of the hemispheres was between 21 and 23 °C as shown in
the graph. The black and white hemispheres are clearly visible in am-
bient light, but infrared imagery is non-distinct because the Styrofoam
hemispheres and the background all have the same temperature. The
IR processing software “stretches” the color spectrum between blue
and white for minimal temperature fluctuations.
Next, the flood light is switched to full power, and temperatures
quickly rise with a thermal equilibrium is reached at “b”. The ther-
mistor readings show that the two thermistors on both high-albedo
(white) hemispheres measure a temperature of 27.5 °C; the low-albedo
(black) thermistors reach a higher temperature: 32.5 °C. This is con-
sistent with our planetary temperature model (Eq. 1.6) which predicts
that under the same intensity of light, low-albedo planets should reach
higher temperatures.
The infrared imagery show that the white hemispheres have the
lowest emission (blue), while both black hemispheres have a much
higher emissions (red) due to their higher temperature. Both black
topics in environmental physics: the solar system and earth 73
C.3 Discussion
C.4.5 In Summary
Most natural objects are black-body emitters and therefore produce
photons in a temperature-dependent range of wavelengths and inten-
sity. Depending on material properties and geometry, one or more
apply: photons are transmitted through the object, reflected off its sur-
faces, or absorbed.
D
Earth’s Rotation and Tides
0.004
Figure D.1: Earth’s rotation relative to
Day length anomaly (s)
−0.004
1850 1900 1950 2000
Time (year)
It turns out that the earth does indeed not spin with a constant rate.
Figure D.1 provides the differences between a civilian day (86,400 s)
and the measured rotation period of Earth. It can be seen that the
earth can rotate slower (usually, positive values) or faster than 86,400
seconds (negative values). Before the reader despairs and asks “is
78 marcel g. schaap
2πmmoon re2−m
Lorbit = (D.1)
Porbit
where mmoon is the mass of Moon, re−m is the radius of the moon’s or-
bit, Porbit is its (sidereal) rotation period. Earth’s rotational momentum
is given by:
2
4πmearth rearth
Learth = (D.2)
5Pearth
topics in environmental physics: the solar system and earth 79
Where mearth is the mass of Earth, rearth is its radius and P is its (side-
real) rotation period (86,164 s). The moon is able to take Earth’s rota-
tional momentum through the tides it generates in the oceans. Because
the moon orbits slower (27.3 days) than the earth turns, the tides move
around Earth and cause a drag on Earth’s rotation. At the same time
the bulging high-tides gravitationally attract the Moon and acceler-
ate it forwards (see Figure D.2, this is also called “torque”). Because
the sum of rotational and orbital momentum must be conserved this
effectively means that a slowing down of Earth’s rotation (longer pe-
riod Pearth ), Earth’s rotational angular momentum is transferred to the
moon’s orbital angular momentum, which is only possible by increas-
ing the radius of its orbit, re−m .
Tide without
dragging
There are two high tides on Earth. The one most easily understood
is the one that “faces” the moon because the gravitational attraction by
moon counteracts that of the earth. This will “pull” the water in the
direction of the moon. High tides are also found on the opposing side
of Earth because there the ocean water is a little further away from
the center of gravity of the combined Earth-Moon system. Gravitational
interaction of the ocean water with the sun has a similar but smaller
effect.
Laser measurements using reflectors left behind by NASA’s manned
lunar program (Figure D.3) have indeed shown that the Moon is reced-
ing from Earth by 38 mm per year (Chapront et al., 2002). The transfer
of rotational momentum means that the earth rotated faster in the geo-
Figure D.3: Laser reflector left behind
logical past. Recent work has shown that day was only 18.68 hours 1.4 on the moon by the Apollo astronauts.
billion years ago Meyers and Malinverno (2018). Other estimates point Credit: NASA.
to days of only 5 hours just after the collision with Theia (Stevenson,
1987).