This is one of the webpages of Libarid A. Maljian at the Department of Physics at CSLA at NJIT.
New Jersey Institute of Technology
College of Science and Liberal Arts
Department of Physics
The Earth in Space
Fall 2024
Fourth Examination lecture notes
Introduction to the Atmosphere
An atmosphere is a thin layer
of gas gravitationally held to a moon, a planet, or a star. The Earth’s atmosphere is roughly eighty
percent nitrogen, roughly twenty percent oxygen, and tiny amounts of other
gases. The tiny amounts of other gases
are quite important, as we will discuss shortly. Every second of every day of our lives, we
are breathing mostly nitrogen (roughly eighty percent) and a fair amount of
oxygen (roughly twenty percent). This
roughly twenty-percent abundance of oxygen is an enormous fraction; other
planets have nowhere nearly this much oxygen in their atmospheres. Other planetary atmospheres have only tiny
amounts of oxygen with a large abundance of carbon dioxide, as is the case with
the atmospheres of planets Venus and Mars for example. The Earth’s atmosphere has a large fraction
of oxygen but only a tiny amount of carbon dioxide. To understand why the Earth’s atmosphere is
so different from other planetary atmospheres, we must discuss the history of
the Earth’s atmosphere. When the Earth
formed roughly 4.6 billion years ago, its atmosphere was almost entirely
hydrogen gas and helium gas; this is called the Earth’s primary atmosphere. The primary atmosphere was almost entirely
hydrogen and helium because the Earth together with the entire Solar System was
born from a nebula, an enormous cloud of gas composed of mostly hydrogen and
helium. Indeed, most of the universe is
composed of hydrogen and helium. The
more massive (or heavier) an atom or molecule, the slower it moves; the less
massive (or lighter) an atom or molecule, the faster it moves. This is rather remarkable. Suppose all the air in a room is at the same
temperature, which means that all the air molecules in the room have the same
average energy. Nevertheless, the oxygen
molecules are moving slower on average since they are more massive (or
heavier), while the nitrogen molecules are moving faster on average since they
are less massive (or lighter), even though all of the air is at the same
temperature, which means both the nitrogen molecules and the oxygen molecules
have the same average energy! Hydrogen
is the least massive (lightest) atom in the entire universe, and helium is the
second least massive (second lightest) atom in the entire universe. Hydrogen and helium are so light that they
move so fast that they can escape from the Earth’s gravitational
attraction. Thus, the Earth lost its
primary atmosphere because its own gravity was too weak to hold onto hydrogen
and helium. This occurred with all four
of the inner planets orbiting the Sun (Mercury, Venus, Earth, and Mars). These four inner planets have weaker gravity
since they are smaller with less mass as compared with the four outer planets
orbiting the Sun (Jupiter, Saturn, Uranus, and Neptune). These four outer planets have stronger
gravity since they are larger with more mass; thus, they have retained their
primary (hydrogen and helium) atmospheres to the present day. Even the gravity of the outer planets is weak
compared with the gravity of the Sun; therefore, the Sun has certainly retained
its primary (hydrogen and helium) atmosphere to the present day. To summarize, the Sun and the four outer planets
have retained their primary (hydrogen and helium) atmospheres, but the four
inner planets have lost their primary (hydrogen and helium) atmospheres. After the Earth lost its primary atmosphere,
this left behind a secondary atmosphere composed of an abundance of water
vapor, nitrogen, carbon dioxide, methane, and other gases. These gases came from volcanic
outgassing. Since the Earth was born
almost entirely molten, volcanic eruptions everywhere on its surface ejected
not just lava but gases as well, as we discussed earlier in the course. These gases are significantly more massive
(or heavier) than hydrogen and helium.
Therefore, they did not move fast enough to escape from the Earth’s
gravitational attraction. As the Earth
cooled, the water vapor condensed into liquid water which precipitated back
down onto the planet for such a long period of time that most of the surface of
the Earth became flooded, thus forming the oceans. At this point, the Earth may be regarded as
having a normal atmosphere with an abundance of carbon dioxide similar to other
planets such as Venus and Mars. However,
roughly one billion years after the Earth formed (roughly 3.6 billion years
ago), something extraordinary occurred in the oceans that to our knowledge did
not occur anywhere else in the entire universe: life appeared. The first lifeforms were primitive
unicellular microorganisms, such as bacteria and blue-green algae. Some of the carbon dioxide in the atmosphere
dissolved into the oceans, which these primitive lifeforms converted to
oxygen. Over the next roughly one
billion years, the rock at the ocean floor was oxidized by the oxygen that
these primitive microorganisms continually synthesized. When nearly all the rock at the ocean floor
was oxidized, the oxygen that these primitive lifeforms continued to synthesize
then began to accumulate within the oceans.
Some of this accumulating oxygen then dissolved back into the
atmosphere. After an additional roughly
two billion years, these microorganisms actually succeeded in extracting almost
all of the carbon dioxide from the atmosphere, replacing it with oxygen. Thus, as of roughly 600 million years ago,
planet Earth attained its tertiary atmosphere that we enjoy to the present day:
roughly eighty percent nitrogen, roughly twenty percent oxygen, and tiny
amounts of other gases. Again, the tiny
amounts of other gases are quite important, and we will discuss these trace
gases shortly.
The pressure of the Earth’s
atmosphere is typically a maximum at mean sea level and decreases exponentially
with increasing elevation. The equation
that describes this decreasing pressure with increasing elevation is called the
law of atmospheres, but we do not need this equation to understand why this is
the case. The Earth’s gravity pulls air
downward; therefore, the air becomes thinner as we climb the atmosphere, making
the air pressure less at higher elevations.
The average air pressure at mean sea level is called one atmosphere of
pressure, equal to 101325 pascals of pressure. One pascal of
pressure is one newton of force per square-meter of area. This average air pressure of 101325 pascals is close enough to one hundred thousand pascals that meteorologists have defined another unit of
air pressure: the bar. One bar of air
pressure is exactly one hundred thousand pascals of
air pressure. Thus, the average air
pressure at mean sea level is equal to 1.01325 bars of pressure; this is also
1013.25 millibars of pressure. When meteorologists report the air pressure
on any given day, they may report that the air pressure is several millibars above average or several millibars
below average. A device that measures
air pressure is called a barometer. To
build a primitive barometer, we insert a long, narrow container inverted into
any liquid. The air pressure will push downward
onto the liquid, thus forcing the liquid upward into the long, narrow inverted
column. If the air pressure is greater
than average, it will push downward more strongly onto the liquid, thus forcing
the liquid further upward the inverted column, making the column of liquid
taller. If the air pressure is less than
average, it will push downward less strongly onto the liquid, thus forcing the
liquid not as far upward the inverted column, making the column of liquid
shorter. Thus, by measuring the height
of the column of liquid and performing a calculation, we can determine the air
pressure that has pushed downward onto the liquid. Most barometers use the element mercury as
the liquid; at average air pressure at mean sea level, liquid mercury will be
pushed 760 millimeters (or 29.9 inches) up the narrow column. Thus, the average air pressure at mean sea
level is also equal to 760 millimeters of mercury (or 29.9 inches of mercury). When meteorologists report the air pressure
on any given day, they may report that the air pressure is several millimeters
of mercury higher than average or several millimeters of mercury lower than
average. Barometers almost always use
liquid mercury because mercury is between thirteen and fourteen times more
dense than water. In other words,
mercury is between thirteen and fourteen times heavier than water; thus,
gravity pulls mercury thirteen or fourteen times more strongly than water,
making the column of mercury only 760 millimeters (or 29.9 inches) tall. If a barometer used water instead of mercury,
the column of water would be between thirteen or fourteen times taller; this
would make barometers more than ten meters (almost thirty-four feet) tall! It is not convenient to carry such a tall
device; it is much more convenient to carry a barometer that is only 760
millimeters (or 29.9 inches) tall. This
is why almost all barometers use mercury instead of water.
Meteorologists have defined
layers of the Earth’s atmosphere based on the variation of the temperature of
the Earth’s atmosphere with elevation.
The lowest layer of the atmosphere is the troposphere. With increasing elevation, the troposphere is
followed by the stratosphere, then the mesosphere, and finally the
thermosphere. Beyond the thermosphere is
the exosphere, where the air smoothly transitions from the Earth’s atmosphere
to the surrounding outer space, as we will discuss shortly. The temperature typically cools with
increasing elevation within the troposphere, the lowest layer of the
atmosphere. However, we reach a certain
elevation at which the temperature stops becoming cooler and begins instead to
become warmer with increasing elevation.
This elevation defines the end of the troposphere and the beginning of the
stratosphere. This precise elevation is
called the tropopause. We may regard the
tropopause as the boundary between the troposphere and the stratosphere, but
the tropopause is more correctly defined as the end of the troposphere. The temperature then typically becomes warmer
with increasing elevation within the stratosphere. The reason for this warming is a heat source
within the stratosphere that we will discuss shortly. However, we reach a certain elevation at
which the temperature stops becoming warmer and begins instead to become cooler
with increasing elevation. This
elevation defines the end of the stratosphere and the beginning of the
mesosphere. This precise elevation is
called the stratopause. We may regard the stratopause
as the boundary between the stratosphere and the mesosphere, but the stratopause is more correctly defined as the end of the
stratosphere. The temperature then
typically becomes cooler with increasing elevation within the mesosphere. However, we reach a certain elevation at
which the temperature stops becoming cooler and begins instead to become warmer
with increasing elevation. This
elevation defines the end of the mesosphere and the beginning of the
thermosphere. This precise elevation is
called the mesopause.
We may regard the mesopause as the boundary between
the mesosphere and the thermosphere, but the mesopause
is more correctly defined as the end of the mesosphere. The temperature then typically becomes warmer
with increasing elevation within the thermosphere. The reason for this warming is a heat source
within the thermosphere that we will discuss shortly. However, we reach a certain elevation at
which the temperature stops becoming warmer and begins instead to become cooler
with increasing elevation. This
elevation defines the end of the thermosphere and the beginning of the
exosphere. This precise elevation is
called the thermopause. We may regard the thermopause
as the boundary between the thermosphere and the exosphere, but the thermopause is more correctly defined as the end of the
thermosphere. The temperature then
typically becomes cooler with increasing elevation within the exosphere,
smoothly transitioning into the very cold temperatures of the surrounding outer
space. To summarize, the temperature
typically becomes cooler with increasing elevation within the troposphere, the
mesosphere, and the exosphere, while the temperature typically becomes warmer
with increasing elevation within the stratosphere and the thermosphere.
It may seem reasonable to ask
for the precise elevation at which the Earth’s atmosphere ends and outer space
begins, but this is in fact an ill-defined question. The concentration of gases in the Earth’s
atmosphere becomes thinner and thinner with increasing elevation until the
concentration of gases matches the concentration of gases of the surrounding
outer space. It is a common
misconception that outer space is a perfect vacuum, but this is false. There is no such thing as a perfect vacuum;
in fact, a perfect vacuum would violate the laws of physics. In actuality, the entire universe is filled
with extremely diffuse gas. Therefore,
the Earth’s atmosphere smoothly transitions into the gases of the surrounding
outer space. We may actually interpret
the Earth’s atmosphere as extending forever, filling the entire universe. The same interpretation can be applied to all
other planetary atmospheres. In other
words, there is no well-defined exopause.
The tropopause is the end of the troposphere, the stratopause
is the end of the stratosphere, the mesopause is the
end of the mesosphere, and the thermopause is the end
of the thermosphere. If there were an
end of the exosphere (which would also be the end of the entire atmosphere),
that end would be called the exopause, but there is no well-defined exopause. Nevertheless, if we insist upon a boundary
between the Earth’s atmosphere and outer space, we may arbitrarily use the
elevation of the tropopause, since the Earth’s gravity pulls most of the air in
the entire atmosphere down into the troposphere. Indeed, the vast majority of all
meteorological phenomena (commonly known as weather) occurs within the
troposphere, the lowest layer of the atmosphere. The exceptions to this are rare. For example, the jet stream is a fast-moving
current of air around the tropopause, much higher in elevation than most
meteorological phenomena (weather). So,
we may arbitrarily regard the thickness of the Earth’s atmosphere as the
elevation of the tropopause as a rough estimate. The elevation of the tropopause is roughly
ten kilometers above mean sea level.
Using this as a rough estimate for the thickness of the Earth’s
atmosphere, we conclude that the atmosphere is extremely thin as compared with
the average radius of the Earth, roughly 6400 kilometers. To summarize, the Earth’s atmosphere extends
indefinitely far according to strict interpretations, but the Earth’s
atmosphere is only a few kilometers thick for all practical purposes. The Earth’s atmosphere keeps us alive in a
variety of different ways. After we
discuss all these ways the atmosphere keeps us alive, we will be humbled. In this vast universe, we are only able to
survive within a very thin layer of air surrounding a single planet: the
atmosphere of planet Earth.
The most obvious way the
Earth’s atmosphere keeps us alive is with its abundance of oxygen. Humans and all animals must inhale oxygen to
survive. This is because humans and
animals must chemically react oxygen with glucose (a simple sugar) to extract
the energy they need for their survival.
Humans ingest various sugars as well as complex carbohydrates such as
bread, rice, cereal, and pasta. Our
bodies digest complex carbohydrates as well as sugars, breaking them down into
glucose (a simple sugar). There is a
tremendous amount of energy stored within the chemical bonds of the glucose
molecule, which our bodies access by reacting it with oxygen. The human body is composed of roughly one
hundred trillion cells. Within these
cells, the following chemical reaction occurs: glucose plus oxygen yields
energy plus carbon dioxide and water as waste products. This chemical reaction is called cellular
respiration and is more properly written C6H12O6
+ 6 O2 → energy + 6 CO2 + 6
H2O. When we inhale, the oxygen that enters our
lungs is transferred to our blood; our blood then carries the oxygen to the
trillions of cells of our bodies. The
oxygen enters our cells and chemically reacts with glucose to yield
energy. The carbon dioxide that is produced
as a waste product from the reaction is transferred back into our blood; our
blood then carries the carbon dioxide back to our lungs, and we then
exhale. Cellular respiration not only
explains why humans and animals must inhale oxygen; cellular respiration also
explains why humans and animals must exhale carbon dioxide. Plants inhale carbon dioxide to use together
with water as the raw materials to synthesize glucose. Plants use the energy of sunlight to initiate
this reaction, which is why this chemical reaction is called photosynthesis. The photosynthesis reaction is more properly
written 6 CO2 + 6 H2O + energy → C6H12O6
+ 6 O2. Notice that the photosynthesis
reaction is precisely the reverse of the cellular respiration reaction. Note also that oxygen is a waste product of
this photosynthesis reaction. Plants
inhale carbon dioxide and exhale oxygen, the precise reverse of humans and
animals. Therefore, the relationship
between animals (including humans) and plants is a symbiotic relationship. Animals (including humans) exhale carbon
dioxide, which plants then inhale. Plants
then exhale oxygen, which animals (including humans) then inhale. Animals (including humans) then exhale carbon
dioxide, which plants then inhale, and so on and so forth.
Another way the Earth’s
atmosphere keeps us alive is by maintaining a habitable temperature for
life. Based on the Earth’s distance from
the Sun, our planet should be too cold for life to exist. The temperature of our planet should be much
colder than the freezing temperature of water; not only should all the oceans
be frozen, but the continents should be frozen over as well. However, there are tiny amounts of gases
within the Earth’s atmosphere that absorb and then emit some of the heat that
our planet radiates. These gases are
called greenhouse gases. The most
important greenhouse gas is water vapor.
Carbon dioxide, methane, and other gases are secondary greenhouse
gases. These gases absorb some of the
heat that the Earth radiates, and these gases then reradiate this heat
themselves. Although these gases
reradiate some of this heat into outer space, these gases also reradiate some
of this heat back to the Earth. This
causes the temperature of planet Earth to be significantly warmer than it would
have been otherwise, based on its distance from the Sun. In fact, the Earth is sufficiently warmed
that its average surface temperature is habitable for life. This warming is called the greenhouse effect,
since it is rather like a greenhouse that is warm even in the wintertime. The Earth’s atmosphere is mostly nitrogen and
oxygen, but neither of these gases can absorb or radiate heat efficiently. In other words, neither nitrogen nor oxygen
are greenhouse gases. Primarily water
vapor and secondarily carbon dioxide, methane, and other gases are able to
absorb and radiate heat efficiently. The
tiny amounts of water vapor, carbon dioxide, methane, and other gases in the
Earth’s atmosphere warm the planet to a habitable temperature. This is a second way the Earth’s atmosphere
keeps us alive.
The Sun not only radiates
visible light; the Sun radiates all forms of electromagnetic radiation. The Electromagnetic Spectrum is a list of all
the different types of electromagnetic waves in order as determined by either
the frequency or the wavelength.
Starting with the lowest frequencies (which are also the longest
wavelengths), we have radio waves, microwaves, infrared, visible light (the
only type of electromagnetic wave our eyes can see), ultraviolet, X-rays, and
gamma rays at the highest frequencies (which are also the shortest wavelengths). All of these are electromagnetic waves. Therefore, all of them may be regarded as
different forms of light. They all
propagate at the same speed of light through the (near-perfect) vacuum of outer
space for example. We now realize that
whenever we use the word light in colloquial English, we probably mean to use
the term visible light, since this is the only type of light that our eyes can
actually see. The visible light band of
the Electromagnetic Spectrum is actually quite narrow. Nevertheless, the visible light band of the
Electromagnetic Spectrum can be subdivided.
In order, the subcategories of the visible light band of the
Electromagnetic Spectrum starting at the lowest frequency (which is also the
longest wavelength) are red, orange, yellow, green, blue, indigo, and violet at
the highest frequency (which is also the shortest wavelength). We now realize why electromagnetic waves with
slightly lower frequencies (or with slightly longer wavelengths) than visible
light are called infrared, since their frequencies (or wavelengths) are just
beyond red visible light. In other
words, infrared light is more red than red!
We also realize why electromagnetic waves with slightly higher
frequencies (or with slightly shorter wavelengths) than visible light are
called ultraviolet, since their frequencies (or wavelengths) are just beyond
violet visible light. In other words,
ultraviolet light is more purple than purple!
The Sun radiates all of these electromagnetic waves. For example, the near ultraviolet from the
Sun causes suntans, and too much near ultraviolet from the Sun causes
sunburns. The far ultraviolet has even
more energy, and the Sun radiates sufficient far ultraviolet that we should be
killed from its far ultraviolet radiation.
X-rays have even greater energy; thus, the X-rays from the Sun should
kill us in a fairly short amount of time.
Something must be shielding us from the Sun’s far ultraviolet and from
the Sun’s X-rays. Our atmosphere
provides these shields. The symbol for
the oxygen atom is O. Under ordinary
temperatures and pressures, the oxygen atom will never remain by itself; it
will always chemically bond with other atoms.
The oxygen atom will even chemically bond with another oxygen atom. Two oxygen atoms chemically bonded to each
other is called the oxygen molecule, which is written O2. Molecular oxygen is also known as normal
oxygen, since oxygen is always in this state under ordinary temperatures and
pressures. Roughly twenty percent of the
Earth’s atmosphere is molecular (normal) oxygen for example, and this is the
form of oxygen that plants exhale as well as the form humans and all animals
must inhale. Notice this is the form of
oxygen appearing in both the cellular respiration reaction and the
photosynthesis reaction written above.
Whenever we use the simple word oxygen, we are not being clear. Do we mean atomic oxygen O or do we mean
molecular oxygen O2? We probably mean molecular oxygen, since this
is normal oxygen. If molecular oxygen
absorbs far ultraviolet, a chemical reaction will synthesize a strange form of
oxygen: three oxygen atoms chemically bonded to each other. This strange form of oxygen is written O3 and is called ozone. The synthesis of ozone is more properly
written 3 O2 + energy → 2 O3. Ozone is
toxic, since inhaling O3 causes severe
respiratory problems. This is ironic,
since ozone also keeps us alive. The
molecular oxygen in the Earth’s atmosphere absorbs the far ultraviolet from the
Sun, synthesizing ozone. Therefore, the
far ultraviolet from the Sun never reaches the surface of the Earth, since it
is absorbed by molecular oxygen to synthesize ozone. In fact, there is a layer of ozone in the
stratosphere below which far ultraviolet does not penetrate. This layer is commonly known as the ozone
layer, but it is more correctly called the ozonosphere. The ozonosphere is the heat source within the
stratosphere that is responsible for warming temperatures with increasing
elevation within that atmospheric layer.
Much higher in the atmosphere within the thermosphere, various atoms and
molecules absorb X-rays from the Sun.
X-rays have so much energy that absorbing them strips electrons
completely free from an atom or molecule.
In other words, atoms or molecules are ionized by X-rays. Therefore, the X-rays from the Sun never
reach the surface of the Earth, since they are absorbed by atoms and molecules
to synthesize ionized atoms and molecules.
In fact, there is a layer of ionized atoms and molecules in the
thermosphere below which X-rays do not penetrate. This layer is called the ionosphere, and it
is the heat source within the thermosphere that is responsible for warming
temperatures with increasing elevation within that atmospheric layer.
Let us summarize all the ways
the Earth’s atmosphere keeps us alive.
Firstly, humans and animals would not have oxygen to react with glucose
to extract energy for their survival if unicellular microorganisms did not
remove most of the carbon dioxide from the atmosphere, replacing it with
molecular oxygen. Secondly, planet Earth
would be too cold for life to exist without the presence of greenhouse gases
that make the planet warm enough to be habitable for life. Thirdly, life on Earth would be killed from
the far ultraviolet from the Sun if it were not for the ozonosphere. Fourthly, life on Earth would be killed from
the X-rays from the Sun if it were not for the ionosphere. As we discussed earlier in the course, the
atmosphere would be substantially ionized by the Sun’s solar wind without the
Earth’s magnetic field deflecting most of these charged particles from the Sun
that continually bombard our planet.
This is a fifth way our planet keeps us alive. If only one of these were the case, we would
not be here from the lack of the other four.
If two were the case, we would not be here from the lack of the other
three. If three were the case, we would
not be here from the lack of the other two.
If four were the case, we would not be here from the lack of the
remaining one. The fact that all five of
these are the case on the same planet is truly miraculous. Again, we are humbled. In this vast universe, we are only able to
survive within a very thin layer of air surrounding a single planet: the
atmosphere of planet Earth.
The surface temperature of
the Earth causes the Earth to radiate heat from its surface. This explains why the troposphere becomes
cooler with increasing elevation; as we climb the troposphere, we are further
and further from the surface of the Earth and thus further and further from
this source of heat. Air in the lower
troposphere (near mean sea level) is often heated by the Earth’s surface. Since hot fluids rise, this hot air may rise
to the upper troposphere (near the tropopause).
This rising air may cool, as we will discuss shortly. Since cool air sinks, air in the upper
troposphere (near the tropopause) may sink to the lower troposphere (near mean
sea level), where it may be warmed again thus causing it to rise again. In summary, there is significant convection
within the troposphere caused by continually circulating air within the
troposphere. This convection
(circulation) of air within the troposphere is ultimately responsible for
meteorological phenomena (commonly known as weather), as we will discuss. This explains why the vast majority of all
meteorological phenomena (weather) occurs within the troposphere, the lowest
layer of the atmosphere. This also
explains why the lowest layer of the Earth’s atmosphere is called the
troposphere, since the Greek root tropo- means turning. The ozonosphere is in the upper stratosphere,
near the stratopause.
This explains why the temperature warms as we climb the
stratosphere. After passing the
tropopause, we approach the ozonosphere, which serves as a heat source, thus
causing warming temperatures. After
passing the stratopause, we are further and further
from the ozonosphere. This explains why
the temperature cools as we climb the mesosphere. Note that cooler air resides in the lower
stratosphere while warmer air resides in the upper stratosphere. Since cool fluids do not rise, the cool air
in the lower stratosphere does not rise to the upper stratosphere. Conversely, since warm fluids do not sink,
the warm air in the upper stratosphere does not sink to the lower
stratosphere. Therefore, there is no
convection (circulation) of air within the stratosphere. The air in the stratosphere remains layered
based on temperature. This explains why
this layer of the atmosphere is called the stratosphere, since the air is
stratified or layered. The word stratify
is derived from a Latin word meaning layer.
As we discussed earlier in the course, the word stratum (a layer of
sedimentary rock) derives from the same Latin word. The air in the mesosphere cools as we climb the
mesosphere, just as air in the troposphere cools as we climb the
troposphere. This may lead us to
conclude that there is significant convection (circulation) of air within the
mesosphere just as in the troposphere, thus causing an abundance of meteorological
phenomena (weather) within the mesosphere.
However, the Earth’s gravity pulls most of the air in the entire
atmosphere down into the troposphere.
Although there is convection (circulation) of air within the mesosphere,
the air within this layer is too thin for this convection (circulation) to
result in an abundance of meteorological phenomena (weather) within the
mesosphere. The Greek root meso- means middle.
For example, Central America is sometimes called Mesoamerica, as in
Middle America. Therefore, the word
mesosphere simply means middle sphere or middle layer. The ionosphere is in the upper thermosphere,
near the thermopause.
This explains why the temperature warms as we climb the
thermosphere. After passing the mesopause, we approach the ionosphere, which serves as a
heat source, thus causing warming temperatures.
After passing the thermopause, we are further
and further from the ionosphere. This
explains why the temperature cools as we climb the exosphere. Note that cooler air resides in the lower
thermosphere while warmer air resides in the upper thermosphere. Since cool fluids do not rise, the cool air
in the lower thermosphere does not rise to the upper thermosphere. Conversely, since warm fluids do not sink,
the warm air in the upper thermosphere does not sink to the lower
thermosphere. Therefore, there is no
convection (circulation) of air within the thermosphere. The air in the thermosphere remains layered
based on temperature. This is similar to
the air in the stratosphere, but note that the air in the thermosphere is much
thinner than the air in the stratosphere, since the Earth’s gravity pulls air
downward. As we climb the exosphere, the
air becomes cooler and cooler, smoothly transitioning into the very cold
temperatures of the surrounding outer space.
The Greek root exo- means outside or
external. For example, an exoskeleton is
a skeleton that is outside (surrounding) an organism. Therefore, the word exosphere simply means
external sphere or external layer. The
exosphere is a layer of gas that is not strictly part of the Earth’s atmosphere
but is outside (surrounding) the Earth’s atmosphere that smoothly transitions
into the gas of the surrounding outer space.
All of us have a basic
understanding of the seasons: it is warmer in summertime and colder in
wintertime. It is a common misconception
that the seasons occur because of the varying distance of planet Earth from the
Sun. Supposedly when our planet Earth is
closer to the Sun, it is warmer causing summertime, and supposedly when our
planet Earth is further from the Sun, it is colder causing wintertime. This argument seems reasonable, but it is
completely wrong. The orbit of the Earth
around the Sun is almost a perfect circle, meaning that the Earth is roughly the
same distance from the Sun throughout the entire year. Of course, the true shape of the Earth’s
orbit around the Sun is an ellipse; sometimes the Earth is closer to the Sun
than average, and other times the Earth is further from the Sun than
average. However, the eccentricity of
the Earth’s elliptical orbit is so close to zero that the orbit is almost a
perfect circle. The eccentricity of an
ellipse quantifies the elongation of the ellipse. When the eccentricity is zero, the ellipse is
a perfect circle. When the eccentricity
is close to zero, the ellipse is almost a perfect circle. The eccentricity of the Earth’s elliptical
orbit around the Sun is so close to zero that its orbit is almost a perfect
circle. When the Earth is at perihelion
(closest to the Sun), it is roughly 2.5 million kilometers (roughly 1.5 million
miles) closer to the Sun than average.
When the Earth is at aphelion (furthest from the Sun), it is roughly 2.5
million kilometers (roughly 1.5 million miles) further from the Sun than average. These closer or further distances may seem
large, but the Earth is on average roughly 150 million kilometers (roughly 93
million miles) from the Sun. Therefore,
these closer or further distances are less than two-percent variations from the
average distance between the Earth and the Sun, and this is not enough of a
difference to cause the seasons. There
is a spectacular piece of evidence that will forever bury the misconception
that the varying distance of the Earth from the Sun causes the seasons: the
Earth is closest to the Sun in wintertime and furthest from the Sun in
summertime! The Earth is at its
perihelion on roughly January 03rd every year, but early January is in
wintertime! The Earth is at its aphelion
on roughly July 03rd every year, but early July is in summertime! Therefore, it is absolutely not the Earth’s
varying distance from the Sun that causes the seasons. Caution: we do not argue that distance from
the Sun is completely irrelevant.
Obviously, if we were to move the Earth fifty million kilometers closer
to the Sun, of course the planet would become so hot that it would no longer be
habitable for life (all life, including all of us, would die). Obviously, if we were to move the Earth fifty
million kilometers further from the Sun, of course the planet would become so
cold that it would no longer be habitable for life (all life, including all of
us, would die). However, if we were to
move the Earth only a couple million kilometers closer to or further from the
Sun, this would not be enough to affect the Earth’s average temperature. The proof of this assertion is that this
already occurs; every year as the Earth orbits the Sun on its elliptical orbit,
the Earth moves roughly 2.5 million kilometers closer to the Sun at perihelion
and roughly 2.5 million kilometers further from the Sun at aphelion, and these
variations do not affect the average temperature of the planet. In fact, planet Earth is closest to the Sun
in wintertime and furthest from the Sun in summertime!
After discussing in
tremendous detail what does not cause the seasons, we must finally discuss what
does cause the seasons. The Earth’s
rotational axis is tilted from the vertical, the vertical being defined as
perpendicular to the plane of the Earth’s orbit around the Sun. The tilt of any planet’s rotational axis is
called the obliquity of the planet. The
seasons are caused by the Earth’s obliquity, the tilt of its rotational axis. The obliquity of planet Earth is roughly 23½
degrees. As the Earth orbits the Sun,
this 23½ degrees of obliquity remains fixed to an excellent approximation. Thus, as the Earth orbits the Sun, sometimes
the Earth’s northern hemisphere will be tilted toward the Sun, causing that
hemisphere to receive more direct sunlight thus causing warmer summertime. The warmer summertime is also caused by
daytime being longer than nighttime, as we will discuss shortly. At the same time the Earth’s northern
hemisphere is tilted toward the Sun, the Earth’s southern hemisphere is tilted
away from the Sun, causing that hemisphere to receive less direct sunlight thus
causing colder wintertime. The colder
wintertime is also caused by nighttime being longer than daytime, as we will
discuss shortly. Six months later when
the Earth is on the opposite side of its orbit, the Earth’s northern hemisphere
will be tilted away from the Sun, causing that hemisphere to receive less
direct sunlight and causing that hemisphere to have longer nighttime than
daytime, thus causing colder wintertime.
At the same time the Earth’s northern hemisphere is tilted away from the
Sun, the Earth’s southern hemisphere is tilted toward the Sun, causing that
hemisphere to receive more direct sunlight and causing that hemisphere to have
longer daytime than nighttime, thus causing warmer summertime. This is remarkable; the seasons are reversed
in the two hemispheres at the same time!
As another example, when it is springtime in the northern hemisphere, it
is autumntime in the southern hemisphere at the same
time. This means that the Earth is at
perihelion (closest to the Sun) during the southern hemisphere’s summertime,
and the Earth is at aphelion (furthest from the Sun) during the southern
hemisphere’s wintertime. We may be
tempted to conclude that the southern hemisphere’s summertime is especially
hot, and the southern hemisphere’s wintertime is especially cold. This is false; the opposite is true! Summers are typically hotter in the northern
hemisphere as compared with summers in the southern hemisphere, and winters are
typically colder in the northern hemisphere as compared with winters in the
southern hemisphere! In other words,
both summers and winters are more mild in the southern hemisphere as compared
with the northern hemisphere, where both summers and winters are more severe. As we discussed earlier in the course, this
is because the continents are presently somewhat crowded together in the
northern hemisphere, making the southern hemisphere mostly covered with ocean
(water). Water has a large heat
capacity, meaning that it is difficult to change the temperature of water. Therefore, the abundance of water in the
southern hemisphere stabilizes temperatures, causing smaller temperature
variations in the southern hemisphere as compared with larger temperature
variations in the northern hemisphere.
This spectacularly emphasizes that variations in the distance from the
Sun do not determine seasonal temperatures.
Again, summers are more mild (less hot) in the southern hemisphere, even
though the Earth is closest to the Sun during summertime in the southern
hemisphere, while summers are more severe (more hot) in the northern
hemisphere, even though the Earth is furthest from the Sun during summertime in
the northern hemisphere! Similarly,
winters are more mild (less cold) in the southern hemisphere, even though the
Earth is furthest from the Sun during wintertime in the southern hemisphere,
while winters are more severe (more cold) in the northern hemisphere, even
though the Earth is closest to the Sun during wintertime in the northern
hemisphere! The large heat capacity of
liquid water is also responsible for moderating the temperature difference
between daytime and nighttime on our planet Earth. The daytime side of any planet faces toward
the Sun, while the nighttime side of any planet faces away from the Sun. For most planets, nighttime is much colder
than daytime, but the nighttime side of planet Earth is only slightly cooler
than its daytime side, thanks to the stabilizing effect of all the water that
covers most of the planet.
Extraordinarily, the nighttime side of planet Earth may at times become
warmer than the daytime side depending upon weather patterns, as we will
discuss. As another example of how water
stabilizes temperatures on our planet Earth, other planets have north poles and
south poles that are much colder than their equators. Although the Earth’s poles are cold and the
Earth’s equator is hot by human standards, the difference in temperature is
nevertheless moderate as compared with other planets. Without the abundance of water that covers
our planet Earth, our poles would be too cold and our equator would be too hot
to be habitable for life.
The moment when the Earth’s
northern hemisphere is tilted the most toward the Sun is called the summer
solstice. This occurs on average June
21st every year; some years it could occur one or two days earlier, while other
years it could occur one or two days later.
Since the northern hemisphere is tilted the most toward the Sun on the
summer solstice, the Sun appears to be highest in the sky, since the northern
hemisphere receives the most direct sunrays.
This is also the longest daytime and the shortest nighttime of the year
in the northern hemisphere. The precise
duration of daytime and nighttime depends upon our latitude. At the midlatitudes,
there are roughly fifteen hours of daytime and only roughly nine hours of
nighttime on the summer solstice. Note
that the sum of fifteen hours and nine hours is twenty-four hours. Six months later when the Earth is on the
other side of its orbit around the Sun, there is a moment when the Earth’s
northern hemisphere is tilted the most away from the Sun. This moment is called the winter solstice,
occurring on average December 21st every year; some years it could occur one or
two days earlier, while other years it could occur one or two days later. Since the northern hemisphere is tilted the
most away from the Sun on the winter solstice, the Sun appears to be lowest in
the sky, since the northern hemisphere receives the least direct sunrays. This is also the longest nighttime and the
shortest daytime of the year in the northern hemisphere. The precise duration of nighttime and daytime
depends upon our latitude, but it will always be the reverse of the summer
solstice. At the midlatitudes
for example, there are roughly fifteen hours of nighttime and only roughly nine
hours of daytime on the winter solstice.
Note again that the sum of fifteen hours and nine hours is twenty-four
hours. Halfway in between the solstices
are two other moments called the equinoxes when the Earth’s axis is not tilted
toward or away from the Sun, resulting in equal amounts of daytime and
nighttime (twelve hours each). This is
why they are called equinoxes, since there are equal amounts of daytime and
nighttime. Roughly three months after
the summer solstice (roughly three months before the winter solstice) is the
autumn equinox, occurring on average September 21st every year; some years it
could occur one or two days earlier, while other years it could occur one or
two days later. Roughly three months
after the winter solstice (roughly three months before the summer solstice) is
the vernal equinox (or the spring equinox).
The vernal equinox (spring equinox) occurs on average March 21st every
year; some years it could occur one or two days earlier, while other years it
could occur one or two days later. It is
a common misconception that since every day is twenty-four hours, supposedly
every day has twelve hours of daytime and twelve hours of nighttime. This is false for almost every day the entire
year. In fact, there are only two days
of the entire year when this is the case: the equinoxes. Once we pass the vernal equinox (spring
equinox), every day for the next six months there is more daytime than
nighttime, with maximum daytime on the summer solstice. Once we pass the autumn equinox, every day
for the next six months there is more nighttime than daytime, with maximum
nighttime on the winter solstice.
The term summer solstice has
at least two different yet interrelated meanings. We may interpret the summer solstice as the
location on the Earth’s orbit where its northern hemisphere is tilted the most
toward the Sun. We may also interpret
the summer solstice as the moment in time when the Earth’s northern hemisphere
is tilted the most toward the Sun, occurring on average June 21st every
year. Obviously, the Earth is located at
its orbital summer solstice at the moment of the temporal summer solstice. The terms winter solstice, vernal equinox,
and autumn equinox have similar interpretations. The term winter solstice has at least two
different yet interrelated meanings. We
may interpret the winter solstice as the location on the Earth’s orbit where
its northern hemisphere is tilted the most away from the Sun. We may also interpret the winter solstice as
the moment in time when the Earth’s northern hemisphere is tilted the most away
from the Sun, occurring on average December 21st every year. Obviously, the Earth is located at its
orbital winter solstice at the moment of the temporal winter solstice. The term vernal equinox (spring equinox) has
at least two different yet interrelated meanings. We may interpret the vernal equinox (spring
equinox) as the location on the Earth’s orbit where its axis is not tilted
toward or away from the Sun as it journeys from the orbital winter solstice
toward the orbital summer solstice. We
may also interpret the vernal equinox (spring equinox) as the moment in time
occurring on average March 21st every year when the Earth’s axis is not tilted
toward or away from the Sun after the temporal winter solstice but before the
temporal summer solstice. Obviously, the
Earth is located at its orbital vernal equinox at the moment of the temporal
vernal equinox. Finally, the term autumn
equinox has at least two different yet interrelated meanings. We may interpret the autumn equinox as the
location on the Earth’s orbit where its axis is not tilted toward or away from
the Sun as it journeys from the orbital summer solstice toward the orbital
winter solstice. We may also interpret
the autumn equinox as the moment in time occurring on average September 21st
every year when the Earth’s axis is not tilted toward or away from the Sun after
the temporal summer solstice but before the temporal winter solstice. Obviously, the Earth is located at its
orbital autumn equinox at the moment of the temporal autumn equinox.
As we discussed earlier in
the course, the latitude of any location on planet Earth is defined as its
angle north or south from the equator.
The colatitude of any location on planet Earth is defined as its angle
from the north pole. Since there are
ninety degrees of latitude from the equator to the north pole, this makes the
colatitude equal to ninety degrees minus the latitude. For example, if our latitude is ten degrees
north, this means we are ten degrees of latitude north of the equator, making
us eighty degrees from the north pole; therefore, our colatitude is eighty degrees. Indeed, ninety minus ten equals eighty. As another example, if our latitude is
seventy degrees north, this means we are seventy degrees of latitude north of
the equator, making us twenty degrees from the north pole; therefore, our
colatitude is twenty degrees. Indeed,
ninety minus seventy equals twenty. The
only location on planet Earth where our latitude and our colatitude equal the
same number is at forty-five degrees north latitude, since that would place us
halfway between (equidistant from) the equator and the north pole. Indeed, ninety minus forty-five equals
forty-five. The altitude of the Sun at
noon on the summer solstice equals our colatitude plus the obliquity. The altitude of the Sun at noon on the winter
solstice equals our colatitude minus the obliquity. The altitude of the Sun at noon on either
equinox equals our colatitude.
Everything we have discussed applies not just to planet Earth but also
to any other planet orbiting the Sun.
The obliquity of any planet is the angular tilt of its rotational axis
from the vertical direction that is perpendicular to its own orbital plane
around the Sun. The poles of any planet
are where its own rotational axis intersects the planet. The equator of any planet is halfway between
the two poles of the planet. Our
latitude on that planet would be our angle north or south from that planet’s
equator. The planet’s equator would be
0° latitude on that planet. The planet’s
north pole would be 90°N latitude on that planet, and
the planet’s south pole would be 90°S latitude on
that planet. Our colatitude on that
planet would be our angle from that planet’s north pole, which would again be
ninety degrees minus our latitude. The
summer solstice of any planet is the moment when its northern hemisphere is
tilted the most towards the Sun. The
winter solstice of any planet is the moment when its northern hemisphere is
tilted the most away from the Sun. The
equinoxes of any planet is halfway between the solstices when its rotational
axis is not tilted toward or away from the Sun.
The altitude of the Sun at noon on each of these dates would be the same
equations: colatitude plus obliquity on the summer solstice, colatitude minus
obliquity on the winter solstice, and colatitude on the equinoxes. The only difference in this analysis for
other planets are the actual dates of the solstices and the equinoxes. For any planet orbiting the Sun, the time
from one solstice to the next solstice (which is also the time from one equinox
to the next equinox) is one-half of the planet’s orbital period around the
Sun. The time from one solstice to the
next equinox (which is also the time from one equinox to the next solstice) is
one-quarter of the planet’s orbital period around the Sun. As an example, consider a hypothetical planet
with an obliquity of thirty degrees, and suppose we live at fifty degrees north
latitude on this hypothetical planet.
Since our latitude is fifty degrees north, our colatitude is forty
degrees, since ninety minus fifty equals forty.
Hence, the altitude of the Sun at noon on the summer solstice would be
seventy degrees, since the colatitude plus the obliquity is forty plus thirty,
which equals seventy. The altitude of
the Sun at noon on the winter solstice would be ten degrees, since the colatitude
minus the obliquity is forty minus thirty, which equals ten. The altitude of the Sun at noon on either
equinox would be forty degrees, since that is our colatitude.
It is a common misconception
that the Sun is directly overhead at noon.
This misconception comes from the phrase high noon. Of course, the Sun is highest in the sky at
noon, giving this phrase some validity.
Nevertheless, the Sun is never ever directly overhead at most locations
on Earth. For example, suppose we live
at forty degrees north latitude. Our
colatitude would be fifty degrees, since ninety minus forty equals fifty. The highest the Sun would ever be at this
location is on the summer solstice, when its altitude at noon is 73½ degrees,
since our colatitude plus obliquity is fifty plus 23½, which is indeed 73½ degrees. Although 73½ degrees is a high altitude, it
is not directly overhead. Directly
overhead would be ninety degrees of altitude.
If the Sun is not directly overhead at noon on the summer solstice, it
would only be lower in the sky every other day of the year. This shows that the Sun is never ever
directly overhead at most locations on Earth.
Is there anywhere on planet Earth where the Sun is directly overhead at
noon on the summer solstice? Yes, at a
latitude equal to the same number of degrees north of the equator as the
obliquity of planet Earth. To prove
this, suppose we live at 23½ degrees north latitude, then our colatitude would
be 66½ degrees, since ninety minus 23½ equals 66½. Thus, the altitude of the Sun at noon would
be our colatitude 66½ degrees plus the obliquity of planet Earth 23½ degrees,
but 66½ plus 23½ equals ninety degrees of altitude, directly overhead! This location of 23½ degrees north latitude
is so important that it deserves a special name: the Tropic of Cancer, as we
discussed earlier in the course. There
is only one location on planet Earth where the Sun is directly overhead at noon
on the winter solstice: 23½ degrees south latitude. This location is so important that it
deserves a special name: the Tropic of Capricorn, as we discussed earlier in
the course. The words Cancer and
Capricorn refer to astronomical constellations of the zodiac; the reason these
lines of latitude are named for astronomical constellations of the zodiac is
beyond the scope of this course. There
is only one location on planet Earth where the Sun is directly overhead at noon
on the equinoxes: the equator at zero degrees latitude. Thousands of years ago, primitive humans did
not understand that the Earth is a planet with a tilted rotational axis
orbiting the Sun. For thousands of
years, humans believed that the motion of the Sun was responsible for the
seasons. Although today we understand
that it is actually the Earth that is orbiting the Sun, we must also confess
that when we look up into the sky, it does appear as if the Sun is moving. Therefore, we should understand the seasons
from the frame of reference of the Earth, which was the only understanding of
humans for thousands of years. From the
frame of reference of the Earth, the Sun appears to be directly on top of the
Tropic of Cancer on the summer solstice.
For the next six months, the Sun appears to move south, arriving on top
of the equator three months later on the autumn equinox and arriving on top of
the Tropic of Capricorn three months after that on the winter solstice. For the following six months, the Sun appears
to move north, arriving on top of the equator three months later on the vernal
equinox (spring equinox) and arriving on top of the Tropic of Cancer three
months after that on the summer solstice.
Again, it is not the Sun that is actually moving north and south; in
actuality, the Earth is orbiting the Sun.
Nevertheless, we live on planet Earth, and so we must understand the
seasons from the frame of reference of the Earth. To summarize, it is only possible for the Sun
to appear directly overhead at noon if we live somewhere between the Tropic of
Cancer and the Tropic of Capricorn. If
we live north of the Tropic of Cancer or south of the Tropic of Capricorn, the
Sun never ever appears to be directly overhead.
The Arctic Circle is 66½
degrees north latitude, making its colatitude 23½ degrees, since ninety minus
66½ equals 23½. The altitude of the Sun
at noon at the Arctic Circle on the winter solstice would be zero degrees,
since our colatitude minus the obliquity would be 23½ minus 23½, which is
obviously zero. An altitude of zero
degrees means the Sun is on the horizon, such as during sunrise. At even more northern latitudes, the altitude
of the Sun on the winter solstice will be a negative number, which means it is
below the horizon. In other words, we
cannot see the Sun, making it nighttime even though the clock time is
noon! Before noon or after noon, the Sun
will be even further below the horizon.
Thus, the entire day is in perpetual nighttime! The same occurs on the Antarctic Circle at
66½ degrees south latitude: the altitude of the Sun at noon on the summer
solstice is zero degrees, meaning that it is on the horizon. At even more southern latitudes, the altitude
of the Sun on the summer solstice will be a negative number, which means it is
below the horizon. Again, we cannot see
the Sun, making it nighttime even though the clock time is noon! Before noon or after noon, the Sun will be
even further below the horizon. Thus,
the entire day is in perpetual nighttime!
These extreme northern latitudes and extreme southern latitudes are the
only places on Earth where the Sun may never rise on some days of the year and
may never set on other days of the year.
At the north pole, six months of continuous nighttime occurs from the
autumn equinox all the way to the vernal equinox (spring equinox), and then six
months of continuous daytime occurs from the vernal equinox (spring equinox)
all the way to the autumn equinox. The
reverse occurs at the south pole: six months of continuous nighttime occurs
from the vernal equinox (spring equinox) all the way to the autumn equinox,
while six months of continuous daytime occurs from the autumn equinox all the
way to the vernal equinox (spring equinox).
Even when the clock time is midnight, the Sun may still be in the sky
causing daytime at these extreme latitudes.
This is the origin of the phrase midnight Sun.
There are several religious
holidays that have their origins in the solstices and the equinoxes. For example, Christmas Day is observed on
December 25th every year. Notice that
this is shortly after the winter solstice.
Before Christmas Day became the celebration of the birth of Jesus
Christ, this was a pagan holiday celebrating the winter solstice. Why would pagans celebrate the day when the
Sun appeared to be lowest in the sky at noon with the most number of nighttime
hours? Primitive humans observed the Sun
appear lower and lower in the sky after the summer solstice; many primitive
humans were probably terrified that the Sun would continue to move downward
until it disappeared below the horizon.
However, by simply paying attention every year, we observe that the Sun
stops moving downward on the winter solstice, and then begins to move
upward. This was a reason to celebrate
for many ancient pagans. This pagan
celebration became the celebration of the birth of Jesus Christ, since early Christians
saw the birth of Jesus Christ as bringing more and more light into a spiritually
dark world. In actuality, Jesus was not
born on December 25th. We will never be
certain of the exact day of the birth of Jesus.
Most of the details of the vast majority of persons throughout human
history, including Jesus, were never recorded.
We will also never be certain of the exact year of the birth of
Jesus. Although it is commonly believed
that the year of the birth of Jesus was anno Domini 1, in actuality no one
recorded the year that Jesus was born.
Again, most of the details of the vast majority of persons throughout
human history, including Jesus, were never recorded. The people of the world did not declare the
year to be anno Domini 1 when Jesus was born.
The ancient Romans designated years using a few different numbering
schemes, among which was the number of years since the founding of the city of
Rome, which was more than seven centuries before the birth of Jesus. Using this particular convention, the ancient
Romans designated years using either of the Latin phrases “ab urbe condita” or “anno urbis
conditae,” meaning “in the year since the founding of the city.” Either Latin phrase was abbreviated AUC. This numbering
scheme continued to be used by Europeans even after the fall of the Western
Roman Empire. More than five centuries
after the birth of Jesus and roughly fifty years after the Western Roman Empire
fell, the Christian monk Dionysius Exiguus tried to
determine the year that Jesus was born.
This monk declared that the year AUC 1278
should be redesignated A.D. 525, where A.D. is the
abbreviation for the new Latin phrase “anno Domini” meaning “in the year of our
Lord,” effectively meaning the number of years since the birth of Jesus. Over the next few centuries, Europeans retroactively
changed the years of historical dates from AUC to
A.D. based on this declaration by Dionysius Exiguus. Europeans even changed the years of
historical dates before the birth of Jesus from AUC
to B.C., which is simply an abbreviation for “before Christ.” Modern scholarship has revealed that the year
that Jesus was born as determined by Dionysius Exiguus
is probably a few years in error. Modern
scholarship estimates that 6 B.C. is a more accurate estimate for the year of
the birth of Jesus. Many students are
offended by this assertion. Students
claim that it would be contradictory for Jesus to have been born in the year 6
B.C., since B.C. is the abbreviation for before Christ and no one can be born
before the year of their own birth!
Again, we must remember that no one recorded the date that Jesus was
born. We must remember that the belief
that the year of the birth of Jesus was A.D. 1 is based upon an estimate by a
monk who lived roughly five centuries after Jesus was born. In fact, we should marvel that the estimate
made by Dionysius Exiguus for an unrecorded event
that occurred roughly five centuries earlier was only a few years in
error! As another example of a religious
holiday that has its origin in the solstices and the equinoxes, Easter is
always the first Sunday after the first Full Moon after the vernal equinox
(spring equinox) every year. Since Jesus
was Jewish, what is traditionally called the Last Supper was in actuality a
Passover celebration. The Jewish
calendar is a lunar calendar, and hence Jewish holy days are determined by the
cycles of the Moon. Placing Easter on
the first Sunday after the first Full Moon after the vernal equinox (spring
equinox) is an attempt to keep the date of Easter as close as possible to the
date of Passover.
A thermometer is a device
that measures temperature. The operation
of a thermometer is based on the principle of thermal expansion and thermal
contraction. Most substances expand when
they become warmer, and most substances contract when they become colder. To build a primitive thermometer, we take any
object and measure its length at one temperature, and we measure its different
length at a different temperature. We
draw marks at each of these lengths, and we draw other marks between these two
marks. To determine the temperature, we
simply read off whichever mark the end of the object meets based on its length
at that temperature. Unfortunately, most
substances expand and contract by only tiny amounts when their temperature
changes. Hence, the marks are often too
close together, making differences in length difficult to measure. However, the element mercury expands by quite
a noticeable amount as it becomes warmer and contracts by quite a noticeable
amount as it becomes colder. Therefore,
most thermometers use liquid mercury, since the marks are then well separated
and easy to read. An actinometer
is a device that measures solar radiation.
To build a primitive actinometer, we take any
object and use a thermometer to measure its initial temperature. Then, we place the object in sunlight for a
certain amount of time, perhaps one hour.
As the object absorbs solar radiation, it becomes hotter. We use a thermometer to measure its hotter
temperature afterwards, and from the difference in temperature between its
hotter final temperature and its colder initial temperature, we can calculate
the amount of solar radiation the object absorbed. Note that we must wrap the object in black
cloth to ensure that it absorbs all of the solar radiation. Otherwise, the object will only become hotter
by some of the solar radiation that it absorbed, since the object will reflect
the rest of the solar radiation. It is
convenient to use a bucket of water as the object, since we know the heat
capacity of water. An actinometer would measure the greatest amount of solar
radiation on the summer solstice, and an actinometer
would measure the least amount of solar radiation on the winter solstice. However, the summer solstice is almost never
the hottest day of the year; a thermometer measures the hottest air temperature
roughly a month later, in late July in the northern hemisphere. Similarly, the winter solstice is almost
never the coldest day of the year; a thermometer measures the coldest air
temperature roughly a month later, in late January in the northern
hemisphere. These delays occur because
the Earth is mostly covered with water, which has a large heat capacity. In other words, it is difficult to change the
temperature of water. The northern
hemisphere receives the most direct sunrays on the summer solstice around June
21st, but it still takes another month for the air to warm to maximum
temperature sometime in late July. In
fact, it takes yet another month for the ocean waters to warm to maximum
temperature, in late August in the northern hemisphere. This is why most people in the northern
hemisphere take their summer vacations in August, so that they may enjoy
swimming in the ocean when it is warmest.
The northern hemisphere receives the least direct sunrays on the winter
solstice around December 21st, but the oceans have retained so much heat from
the summertime that it still takes another month for the air to cool to minimum
temperature sometime in late January. In
fact, it takes yet another month for the ocean waters to cool to minimum temperature,
in late February in the northern hemisphere.
Local (Small-Scale) Meteorological Dynamics
Aside from the seasonal
temperature variations we have discussed, many other variables affect the air
temperature on a daily basis, on an hourly basis, and even shorter
timescales. These variations in air
temperature cause variations in air pressure.
These variations in air pressure cause the meteorological phenomena
(commonly known as weather) that we will discuss. Air pressure is the force that the air exerts
per unit area. This pressure (force per
unit area) is ultimately caused by molecular collisions. Therefore, we may interpret air pressure as a
measure of how frequently air molecules collide with each other. Warm air is at a lower pressure, while cold
air is at a higher pressure. This is
because the molecules of warm air are moving faster, enabling them to move
further from one another; hence, they collide with each other less frequently
since they are further apart from one another.
Conversely, the molecules of cold air are moving slower; they cannot
move far from one another and hence they collide with each other more
frequently. Suppose all the air in a
certain room is at the same pressure, and consider a parcel of air in the
middle of the room. Since the air
pressure on either side of the parcel of air is the same, it will suffer equal
molecular collisions from either side of itself. These equal molecular collisions will balance
each other, and the parcel of air will not move. Now suppose instead that the air on one side
of the room is at a higher air pressure for whatever reason, and suppose the
air on the other side of the room is at a lower air pressure for whatever
reason. Again, consider a parcel of air
in the middle of the room. This parcel
of air will now suffer greater molecular collisions from the higher-pressure
side of the room, and the parcel of air will suffer fewer molecular collisions
from the lower-pressure side of the room.
The net result is that the parcel of air will be pushed from the higher
pressure toward the lower pressure. The
parcel of air will move, since it is pushed by a pressure imbalance. This pressure imbalance is so important that
it deserves a name; it is called the pressure gradient force. If the air pressure throughout the room were
the same, there would be no pressure gradient and hence no force; the air would
not move. If there are variations in
pressure, the pressure gradient force pushes air from higher pressure toward
lower pressure. Moving air is called
wind. Hence, the pressure gradient force
causes wind to blow.
A geometrical curve
connecting locations of equal air pressure is called an isobar. The pressure gradient force is always zero
along any isobar, since every point on an isobar is at the same pressure. Since the pressure gradient force is always
zero along any isobar, the pressure gradient force must point directly
perpendicular to isobars. If the
pressure gradient force did not point directly perpendicular to isobars, then
we would be able to break the force into two components: one component directly
perpendicular to the isobars and the other component along the isobars. However, the component along the isobars must
be zero, as we just argued. Therefore,
the pressure gradient force can only have one component: the component directly
perpendicular to the isobars. The
pressure gradient force does not have two components; it only has one component
that is directly perpendicular to the isobars.
Every point on an isobar is at the same air pressure, but two different
isobars will of course be at two different pressures. Suppose as we move from one isobar to a
neighboring isobar, the pressure always drops by a definite amount, perhaps ten
millibars. If
isobars are closely spaced to each other, this means that the pressure drops by
ten millibars over a narrow distance. In other words, the pressure gradient will be
steep, thus causing strong winds. If the
isobars are widely spaced from each other, this means that the pressure drops
by ten millibars over a wide distance. In other words, the pressure gradient will be
shallow, thus causing light winds. This
is remarkable, since the pressure drop from one isobar to the neighboring
isobar is always a fixed amount: ten millibars in
these examples. Nevertheless, the ten-millibar pressure drop is steep if the isobars are closely
spaced to each other, while the ten-millibar pressure
drop is shallow if the isobars are widely spaced from each other. Again, this is remarkable: the same ten millibar pressure drop is steep causing strong winds to
blow in one case, while the same ten millibar
pressure drop is shallow causing light winds to blow in another case.
An anemometer is a device
that measures the velocity of wind, meaning that an anemometer measures both
the speed and the direction of wind. An
anemometer is essentially a wind vane together with a flag. As the wind blows, the wind vane turns with a
certain angular speed. From that angular
speed, we can calculate the speed with which the wind blows. A wind vane shaped like a rooster is called a
weathercock. The flag reveals the
direction with which the wind blows; whichever way the flag flutters is the
direction the wind is blowing. The
Beaufort scale is a wind scale, named for the Irish oceanologist/oceanographer
Francis Beaufort. The Beaufort scale
uses numbers from zero (for no winds) to twelve (for hurricane-speed
winds). A low number on the Beaufort
would be a light wind, which is called a breeze. A middle number on the Beaufort scale would
simply be called a wind. A high number
on the Beaufort scale would be a strong wind, which is called a gale.
We have already discussed
enough basic meteorology to analyze some simple weather patterns. Suppose we are at the beach in the daytime
when the Sun warms the Earth. Since
water has a large heat capacity, the ocean does not become as warm as the
continent. All of us have experienced
this while at the beach in the daytime; no matter how hot the daytime
temperature, the ocean water is not as warm as the sand. Therefore, the air above the continent is
warmer than the air above the ocean.
Thus, the air above the continent is at a lower pressure as compared
with the air above the ocean, which is at a relatively higher pressure. Since the pressure gradient force pushes air
from high pressure toward low pressure, wind will blow from the ocean toward
the continent. This is called the sea
breeze. In meteorology, we always name
wind based on the direction it is blowing from, which is the opposite of the
direction the wind is blowing toward.
For example, a wind blowing from the north (which means it is blowing
toward the south) is called a north wind.
As another example, a wind blowing from the southwest (which means it is
blowing toward the northeast) is called a southwest wind. The sea breeze is blowing from the ocean
toward the continent; hence, it is called the sea breeze. We often feel this sea breeze while at the
beach in the daytime. The sea breeze is
a steady wind blowing from the ocean toward the continent during the
daytime. In the nighttime, the opposite
occurs. Suppose we are at the beach in
the nighttime when the Earth cools.
Since water has a large heat capacity, the ocean does not become as cold
as the continent. Perhaps some of us
have experienced this while at the beach in the nighttime; no matter how cool
the nighttime temperature, the ocean water is not as cold as the sand. Therefore, the air above the continent is
colder than the air above the ocean.
Thus, the air above the continent is at a higher pressure as compared
with the air above the ocean, which is at a relatively lower pressure. Since the pressure gradient force pushes air
from high pressure toward low pressure, wind will blow from the continent
toward the ocean. This is called the
land breeze. Again, we always name wind
based on the direction it is blowing from, which is the opposite of the
direction the wind is blowing toward.
The land breeze is blowing from the continent toward the ocean; hence,
it is called the land breeze. Perhaps
some of us have felt this land breeze while at the beach in the nighttime. The land breeze is a steady wind blowing from
the continent toward the ocean during the nighttime. If we are facing the ocean, we feel the land
breeze upon our backs; if we turn our backs to the ocean, we feel the land
breeze upon our fronts. Similar to the
sea breeze and the land breeze is the valley breeze and the mountain
breeze. In the daytime, the Sun warms
the air. Hot air is less dense, and so
hot air will be buoyed upward by the surrounding air. This is why hot air rises. Therefore, daytime winds will blow from a
valley up toward a mountain. This is
called the valley breeze. Again, we
always name wind based on the direction it is blowing from, which is the
opposite of the direction the wind is blowing toward. The valley breeze is blowing from the valley
up toward the mountain; hence, it is called the valley breeze. Perhaps some of us have felt this valley
breeze while on a mountain in the daytime.
The valley breeze is a steady wind blowing from the valley up toward the
mountain during the daytime. In the
nighttime, the opposite occurs. The air
cools in the nighttime. Cold air is more
dense, and so cold air will descend into the surrounding air. This is why cold air sinks. Therefore, nighttime winds will blow from a
mountain down into a valley. This is
called the mountain breeze. Again, we
always name wind based on the direction it is blowing from, which is the
opposite of the direction the wind is blowing toward. The mountain breeze is blowing from the
mountain down into the valley; hence, it is called the mountain breeze. Perhaps some of us have felt this mountain
breeze while in a valley in the nighttime.
The mountain breeze is a steady wind blowing from the mountain down
toward the valley during the nighttime.
To summarize, during the daytime the sea breeze blows from the ocean
toward the continent, while during the nighttime the land breeze blows from the
continent toward the ocean. During the
daytime the valley breeze blows from the valley up toward the mountain, while
during the nighttime the mountain breeze blows from the mountain down toward
the valley.
Fictitious forces or pseudoforces are forces that do not actually exist; they
only seem to exist in certain frames of reference. For example, suppose we are in a stationary
car waiting at a red traffic light. When
the red traffic light turns green, we place our foot upon the car’s accelerator
pedal. As the car accelerates forward,
everyone and everything in the car feels a backward force. We actually feel ourselves pulled backward
into the backrest of our chair. Anything
hanging from the rearview mirror also swings backward. This backward force is a fictitious force or
a pseudoforce.
It does not exist; it only seems to exist within the car as the car
accelerates forward. Although everyone
and everything within the car feels this backward force, it nevertheless does
not actually exist. In actuality,
everyone and everything within the car remains stationary for a moment as the
car and its chairs accelerate forward, and hence the backrests of the chairs
accelerate forward and collide with our own backs. This is amusing: within the car we feel
pulled backward into the backrests of the chairs, but in actuality we remain
stationary while the backrests of the chairs accelerate forward into our
backs! Although we feel a backward force
within the car, we nevertheless conclude that this backward force is a
fictitious force or a pseudoforce. It does not actually exist; it only seems to
exist within the car as the car accelerates forward. As another example, suppose we are in a
moving car when we see a green traffic light turn yellow, and so we place our
foot upon the car’s brake pedal. As the
car slows down, everyone and everything in the car feels a forward force. We actually feel ourselves pulled forward off
of the backrest of our chair. Anything
hanging from the rearview mirror also swings forward. In extreme cases, we may feel pulled forward
so strongly that our heads may collide with the windshield. This forward force is a fictitious force or a
pseudoforce.
It does not exist; it only seems to exist within the car as the car
slows down. Although everyone and
everything within the car feels this forward force, it nevertheless does not
actually exist. In actuality, everyone
and everything within the car remains in motion for a moment as the car and its
chairs and its windshield slow down, and hence the backrests of the chairs move
away from our own backs while the windshield moves toward our heads. This is amusing: within the car we feel
pulled forward off of the backrests of the chairs and toward the windshield,
but in actuality the backrests of the chairs move away from our backs and the
windshield moves toward our heads!
Although we feel a forward force within the car, we nevertheless
conclude that this forward force is a fictitious force or a pseudoforce. It does not actually exist; it only seems to
exist within the car as the car slows down.
As yet another example, suppose we are in a moving car when we see that
the highway ramp ahead curves to the left, and so we turn the steering wheel to
the left so that the car will remain on the highway ramp. As the car turns left, everyone and
everything in the car feels a rightward force.
We actually feel ourselves pulled rightward away from the driver’s side
of the car and toward the passenger’s side of the car. Anything hanging from the rearview mirror also
swings rightward and continues to remain suspended rightward in apparent
defiance of the Earth’s downward gravity as the car turns left! This rightward force is a fictitious force or
a pseudoforce.
It does not exist; it only seems to exist within the car as the car
turns left. Although everyone and
everything within the car feels this rightward force, it nevertheless does not
actually exist. In actuality, everyone
and everything within the car remains in forward motion as the car turns left,
and hence the driver’s side of the car turns away from us while the passenger’s
side of the car turns toward us. This is
amusing: within the car we feel pulled rightward toward the passenger’s side of
the car, but in actuality we remain in forward motion while the passenger’s
side of the car turns leftward toward us!
Although we feel a rightward force within the car, we nevertheless
conclude that this rightward force is a fictitious force or a pseudoforce. It does
not actually exist; it only seems to exist within the car as the car turns
left. As a fourth example, projectiles
will appear to suffer from deflections within a rotating frame of reference. This deflecting force is a fictitious force
or a pseudoforce.
It does not exist; it only seems to exist within the rotating frame of
reference. In actuality, the projectiles
are not deflected; the projectiles in fact continue moving along straight
paths. The frame of reference is
rotating, and the rotation of the entire frame of reference seems to cause
projectiles to deviate from straight trajectories. This particular fictitious force or pseudoforce is called the Coriolis force, named for the
French physicist Gaspard-Gustave de Coriolis who first derived the mathematical
equations describing this particular fictitious force or pseudoforce. The Coriolis force appears to cause rightward
deflections in frames of reference rotating counterclockwise, and the Coriolis
force appears to cause leftward deflections in frames of reference rotating
clockwise. The Coriolis force appears to
cause stronger deflections if the frame of reference is rotating faster and
appears to cause weaker deflections if the frame of reference is rotating slower. The Coriolis force appears to vanish if the
frame of reference stops rotating. The
Coriolis force only appears to cause deflections; it does not cause projectiles
to speed up or slow down.
If the Earth were not
rotating, the study of the Earth’s atmosphere would be relatively simple. The pressure gradient force would simply push
wind perpendicular to isobars from high pressure toward low pressure. However, the Earth is rotating, and the
rotation of the Earth causes gross complications in the Earth’s
atmosphere. The Earth rotates from west
to east. Therefore, the northern
hemisphere rotates counterclockwise when viewed from above the north pole,
while the southern hemisphere rotates clockwise when viewed from above the
south pole. We conclude that there is a
Coriolis force on planet Earth that appears to cause rightward deflections in
the northern hemisphere and appears to cause leftward deflections in the
southern hemisphere. The Coriolis force
would appear to be stronger if the Earth were rotating faster, while the
Coriolis force would appear to be weaker if the Earth were rotating
slower. The Coriolis force would appear
to vanish if the Earth were to stop rotating altogether. The Coriolis force only appears to cause
deflections; it does not cause projectiles to speed up or slow down. Finally, the Coriolis force is weak near the
equator (the Coriolis force is in fact zero at the equator), and the Coriolis
force becomes stronger and stronger as we move away from the equator toward the
poles (the Coriolis force is in fact strongest at the poles). The pressure gradient force still pushes air
perpendicular to isobars from high pressure toward low pressure, but in
addition the Coriolis force causes deflections to the right in the northern
hemisphere and deflections to the left in the southern hemisphere. Since the Coriolis force appears to cause
these deflections, wind will not blow directly perpendicular to isobars; wind
will not blow directly from high pressure toward low pressure.
A thermal is a parcel of air
in the Earth’s atmosphere. If we have a
low-pressure thermal surrounded by high pressure, the pressure gradient force
will push wind from the surrounding high-pressure air inward toward the
low-pressure thermal. At the same time,
the Coriolis force will cause deflections to the right in the northern
hemisphere and to the left in the southern hemisphere. The net result of the pressure gradient force
together with the Coriolis force is an inward circulation of wind. Winds will blow inward while circulating
counterclockwise in the northern hemisphere, and winds will blow inward while
circulating clockwise in the southern hemisphere. If we instead have a high-pressure thermal
surrounded by low pressure, the pressure gradient force will push wind from the
high-pressure thermal outward toward the surrounding low-pressure air. At the same time, the Coriolis force will
cause deflections to the right in the northern hemisphere and to the left in
the southern hemisphere. The net result
of the pressure gradient force together with the Coriolis force is an outward
circulation of wind. Winds will blow
outward while circulating clockwise in the northern hemisphere, and winds will
blow outward while circulating counterclockwise in the southern hemisphere. The weather pattern around a low-pressure
thermal is called a cyclone. Tornadoes
and hurricanes are extreme examples of cyclones, as we will discuss. The weather pattern around a high-pressure
thermal is called an anticyclone. A
beautiful clear day is an extreme example of an anticyclone, as we will also
discuss. In summary, a cyclone is the
weather pattern around a low-pressure thermal, where winds blow inward while
circulating counterclockwise in the northern hemisphere and clockwise in the
southern hemisphere; an anticyclone is the weather pattern around a
high-pressure thermal, where winds blow outward while circulating clockwise in
the northern hemisphere and counterclockwise in the southern hemisphere.
At the center of a cyclone is
a low-pressure thermal; this low-pressure thermal has a low density and is
therefore buoyed upward by the surrounding air.
As an alternative argument, the low pressure is caused by hot
temperatures, and hot air must rise. Either
argument leads us to conclude that the low-pressure thermal at the center of a
cyclone rises. At the center of an
anticyclone is a high-pressure thermal; this high-pressure thermal has a high
density and therefore sinks downward into the surrounding air. As an alternative argument, the high pressure
is caused by cold temperatures, and cold air must sink. Either argument leads us to conclude that the
high-pressure thermal at the center of an anticyclone sinks. As the low-pressure thermal at the center of
a cyclone rises, the surrounding air pressure at higher elevations decreases in
accord with the law of atmospheres.
Therefore, the rising low-pressure thermal expands as its own pressure
pushes the surrounding lower-pressure air.
As the high-pressure thermal at the center of an anticyclone sinks, the
surrounding air pressure at lower elevations increases in accord with the law
of atmospheres. Therefore, the
high-pressure thermal contracts as the surrounding high-pressure air compresses
the sinking thermal. In summary, a
cyclone is the weather pattern around a low-pressure thermal, where winds blow
inward while circulating counterclockwise in the northern hemisphere and
clockwise in the southern hemisphere; the low-pressure thermal then rises to
higher elevations and expands.
Conversely, an anticyclone is the weather pattern around a high-pressure
thermal that sinks to lower elevations and contracts, then the winds blow
outward while circulating clockwise in the northern hemisphere and
counterclockwise in the southern hemisphere.
The most obvious way to
change the temperature of a gas is through the addition or extraction of
heat. If we add heat to a gas, we expect
it to become warmer; if we extract heat from a gas, we expect it to become
cooler. However, it is possible to
change the temperature of a gas without adding or extracting heat. If a gas expands, it must become cooler even
if no heat was extracted. This is
because the expanding gas must push the surrounding gas. This requires work, and work is a form of
energy. The gas extracts this energy
from its own internal energy content, and so the gas becomes cooler. Conversely if a gas contracts, it must become
warmer even if no heat was added. This
is because the surrounding gas performs work on the gas while compressing the
gas. This work is added to the internal
energy content of the gas, and so the gas becomes warmer. This is remarkable; we can actually change
the temperature of a gas without adding or extracting any heat. We can demonstrate this by performing all of
these experiments while the gas is wrapped within a thermal insulator, which
will not permit any heat to be added or extracted. Nevertheless, the gas becomes warmer when we
compress it, and the gas becomes cooler when we expand it. We are forced to conclude that heat and
temperature are two completely different physical concepts. Most people naïvely believe that heat and
temperature are essentially the same thing.
After all, when we add heat to an object it often becomes warmer, and
when we extract heat from an object it often becomes cooler. Nevertheless, we have just discussed
circumstances where we can actually change the temperature of a gas without the
addition or extraction of any heat. In
fact, it is possible for a gas to become warmer under certain circumstances
when we have extracted heat from the gas!
It is also possible for a gas to become cooler under certain
circumstances when we have added heat to the gas! These extraordinary examples persuade us that
heat and temperature are indeed two completely different physical
concepts. Therefore, we must use
different terms to describe one process where there is no temperature change
and another process where there is no heat exchanged, since heat and
temperature are two completely different physical concepts. A process where there is no temperature
change is called an isothermal process.
A process where there is no heat exchanged is called an adiabatic
process. These are two different
processes. Simply because a process is
adiabatic (no heat exchanged) does not mean that it is necessarily isothermal
(no temperature change). Simply because
a process is isothermal (no temperature change) does not mean that it is
necessarily adiabatic (no heat exchanged).
We just discussed two examples of adiabatic processes that are not
isothermal. A gas that expands
adiabatically (no heat exchanged) becomes cooler; since the temperature is
changing, this is not an isothermal process.
A gas that contracts adiabatically (no heat exchanged) becomes warmer;
since the temperature is changing, this is not an isothermal process
either. These are two examples of
processes that are adiabatic yet not isothermal, meaning the temperature
changes even though no heat was exchanged.
There are other examples of processes that are isothermal yet not
adiabatic, meaning heat was exchanged even though the temperature did not
change. Obviously, there are processes
that are neither adiabatic nor isothermal.
Also note that a process where there is no pressure change is called an
isobaric process.
Air is a poor conductor of heat. A spectacular illustration of the poor
conduction of heat through air is the moderate heat we feel from the intense
temperature of charcoal during a barbecue.
When a piece of charcoal begins glowing red, its temperature is a couple
thousand degrees! Yet, we can place our
hand within just a few inches of the charcoal; although we feel moderate heat,
our hand is not in danger from the extreme temperature of the charcoal. How can our hand be within a few inches of an
object at a couple thousand degrees of temperature and yet not be in any
danger? We conclude that the air between
our hand and the hot charcoal is a poor conductor of heat. Since air is such a poor conductor of heat,
we always assume thermals in the Earth’s atmosphere neither gain heat from
their surroundings nor lose heat to their surroundings. That is, we always assume thermals in the
Earth’s atmosphere do not exchange heat with their surroundings. This is called the adiabatic approximation,
since an adiabatic process involves no exchange of heat. Of course, thermals do exchange some heat
with their surroundings, but air is such a poor conductor of heat we can ignore
the small amounts of heat that are exchanged between a thermal and its surroundings. In other words, the adiabatic approximation
is an excellent approximation for analyzing most meteorological processes. As we discussed, the air at the center of a
cyclone rises and expands. As an
excellent approximation, the rising thermal expands adiabatically, by the
adiabatic approximation. If the thermal
expands adiabatically, then it must cool.
As we discussed, the air at the center of an anticyclone sinks and
contracts. As an excellent
approximation, the sinking thermal contracts adiabatically, by the adiabatic
approximation. If the thermal contracts
adiabatically, then it must warm. In
summary, a cyclone is the weather pattern around a low-pressure thermal where
winds blow inward while circulating counterclockwise in the northern hemisphere
and clockwise in the southern hemisphere; the low-pressure thermal then rises,
expands adiabatically, and cools.
Conversely, an anticyclone is the weather pattern around a high-pressure
thermal that sinks, contracts adiabatically, and warms, then the winds blow
outward while circulating clockwise in the northern hemisphere and
counterclockwise in the southern hemisphere.
The low-pressure air at the
center of a cyclone rises, expands adiabatically, and cools, but cold air is at
a high pressure. We conclude that
low-pressure air at lower elevations in the troposphere transitions to
high-pressure air at higher elevations in the troposphere. In brief, low-pressure air near mean sea
level becomes high-pressure air aloft.
The word aloft simply means up toward the sky. The loft of a house is the highest room in
the house (often the attic), the choir loft of a church is high above the pews,
and lofty dreams are high goals. If
winds blow inward toward low air pressure near mean sea level, these winds must
eventually blow outward from the high air pressure aloft. This circulation of air is a necessary
feature of convection, as we discussed earlier in the course. Cyclones may initially form from low air
pressure at lower elevations in the troposphere (near mean sea level) or from
high air pressure at higher elevations in the troposphere (aloft), but
generally cyclones initially form from both occurring in conjunction with each
other, from the appropriate vertical motion of air throughout the
troposphere. Similarly, we conclude that
high-pressure air at lower elevations in the troposphere transitions to
low-pressure air at higher elevations in the troposphere. In brief, high-pressure air near mean sea
level becomes low-pressure air aloft. If
winds blow outward from high air pressure near mean sea level, these winds must
blow inward toward the low air pressure aloft.
Again, this circulation of air is a necessary feature of convection, as
we discussed earlier in the course.
Anticyclones may initially form from high air pressure at lower
elevations in the troposphere (near mean sea level) or from low air pressure at
higher elevations in the troposphere (aloft), but generally anticyclones
initially form from both occurring in conjunction with each other, from the
appropriate vertical motion of air throughout troposphere.
Relative humidity is a
concept that everyone believes that they understand, but in actuality almost no
one correctly understands this concept of relative humidity. Most people believe that the relative
humidity of air is the amount of moisture in the air. This is such a gross simplification of the
correct definition of relative humidity that it is actually an incorrect
understanding. Firstly, air is only able
to hold a maximum amount of moisture. We
may demonstrate this with the following experiment. We place a lid upon a cup of water; a simple
piece of paper will serve as a satisfactory lid. Liquid water continuously evaporates into
water vapor, but the lid will confine the water vapor to the trapped air
between the lid and the liquid water.
When this confined air holds the maximum amount of water vapor that it
is able to hold, some of the water vapor must condense back into liquid water
so that additional liquid water may evaporate into water vapor. Some of this water will condense back into
the liquid within the cup, but some of this water will condense into drops of
liquid water on the sides of the cup and even underneath the lid. When the air holds the maximum amount of
moisture that it is able to hold, the air is said to be saturated. To saturate anything means to fill it to
capacity; the air is saturated when it holds the maximum moisture that it is
able to hold. The saturation amount of
air is itself a function of temperature.
Warm air has a greater saturation amount since warm air is able to hold
a greater quantity of moisture, while cold air has a lesser saturation amount
since cold air is not able to hold as much moisture as warm air. The strict definition of the relative humidity
of air is the amount of moisture in the air as a fraction of the saturation
amount at the given temperature. Let us
devote some time to carefully understand this definition. Firstly, the relative humidity of air is
directly related to the amount of moisture in the air. Adding moisture to air increases the relative
humidity, while subtracting moisture from air decreases the relative
humidity. However, it is possible to
change the relative humidity of air without adding or subtracting water. By simply changing the temperature of the
air, we change the saturation amount of the air and thus we change the fraction
of the moisture to the new saturation amount.
If air becomes warmer, the saturation amount is greater, making the
amount of moisture that is actually within the air a smaller fraction of that
greater saturation amount. In other
words, warming air decreases its relative humidity. If air becomes colder, the saturation amount
is lesser, making the amount of moisture that is actually within the air a
greater fraction of that lesser saturation amount. In other words, cooling air increases its
relative humidity. An analogy would be
helpful to understand these processes. Imagine
a large bucket and a small cup. A large
bucket is able to hold a large amount of water, while a small cup is only able
to hold a small amount of water. The
large bucket is analogous to hot air, since hot air has a large saturation
amount, meaning it is able to hold more moisture. The small cup is analogous to cold air, since
cold air has a small saturation amount, meaning it is not able to hold as much
moisture as hot air. Now suppose the
small cup is mostly full. If we pour
this water into a large empty bucket, the large bucket will be mostly
empty. If we take a large bucket that is
mostly empty and pour this water into a small cup, the small cup will be mostly
full. This is remarkable: it is the same
amount of water in the small cup and the large bucket. Nevertheless, this same amount of water makes
the small cup mostly full and makes the large bucket mostly empty. We always keep in mind that the small cup is
analogous to cold air and the large bucket is analogous to hot air. If we take cold air and warm it, this
decreases the relative humidity, since this is analogous to taking a small cup
that is more full and pouring its water into a large bucket that will now be
more empty. If we take warm air and cool
it, this increases the relative humidity, since this is analogous to taking a
large bucket that is more empty and pouring its water into a small cup that
will now be more full. This is
remarkable: we have not changed the amount of moisture in the thermal. We are changing its relative humidity without
adding or extracting water; we are changing its relative humidity by changing
its temperature. We now realize that
regarding relative humidity as simply the amount of moisture in the air is a
grossly incorrect understanding of relative humidity. As another remarkable example, consider two
thermals both at fifty percent relative humidity. Does this mean they hold the same amount of
moisture? Isn’t fifty percent equal to
one-half? Therefore, are not both
thermals holding moisture equal to half of their respective saturation amounts? This is certainly true, but two thermals will
almost always have two different temperatures.
The hotter thermal has a greater saturation amount, while the colder
thermal has a lesser saturation amount.
Half of a greater amount is a larger number, and half of a lesser amount
is a smaller number. Thus, the warmer
thermal actually holds more moisture and the cooler thermal actually holds less
moisture, even though both thermals have the same relative humidity! This is analogous to a large bucket and a
small cup that are both half full. The
large bucket holds more water and the small cup holds less water, even though
both are half full! We always keep in
mind that the small cup is analogous to cold air and the large bucket is
analogous to hot air. If a large bucket
and a small cup are both half full and yet the large bucket holds more water
and the small cup holds less water, we conclude that two thermals both at fifty
percent relative humidity hold different quantities of moisture. The warmer thermal holds more moisture
(analogous to the large bucket), while the cooler thermal holds less moisture
(analogous to the small cup). As an
extreme example, consider two thermals: one at ninety percent relative humidity
and the other at ten percent relative humidity.
Which thermal holds more moisture, and which thermal holds less moisture? We are tempted to conclude that surely it
must be the ninety-percent humid thermal that holds more moisture, and we are
tempted to conclude that surely it must be the ten-percent humid thermal that
holds less moisture. In actuality, we
cannot draw any conclusions about the moistures of the two thermals without
knowing their temperatures. Again, we
imagine a large bucket and a small cup.
Suppose the large bucket is only ten-percent full, while the small cup
is ninety-percent full. Nevertheless, suppose
the large bucket is so large that it still holds more water at ten-percent
capacity than the small cup at ninety-percent capacity. We always keep in mind that the small cup is
analogous to cold air and the large bucket is analogous to hot air. Suppose we have two thermals: one at ninety
percent relative humidity and the other at ten percent relative humidity. Now suppose that the ten-percent-humid
thermal is so warm that its saturation amount is so large that ten percent of
that large saturation is nevertheless more moisture than the
ninety-percent-humid cold thermal.
Therefore, it is possible for a ten-percent-humid thermal to hold more
moisture than a ninety-percent-humid thermal if the ten-percent-humid thermal
is sufficiently warm and if the ninety-percent-humid thermal is sufficiently
cold.
Since warming air decreases
the relative humidity, the least humid time of the day is typically in the late
afternoon before sunset, since the Sun has spent the entire daytime warming the
air. Since cooling air increases the
relative humidity, the most humid time of the day is typically in the very
early morning hours before sunrise, since the air has spent the entire
nighttime cooling in the darkness. Just
before sunrise, the air may have cooled sufficiently for the relative humidity
to increase to one hundred percent. That
is, the air has become saturated with water vapor. At one hundred percent relative humidity
(saturation), water vapor must condense into liquid water so that additional
liquid water may evaporate. The
temperature to which we must cool air until it becomes saturated is called the
dew point, since the water vapor that condenses into liquid water is called
dew. In the early morning, we may see
the leaves of trees and the surface of our car covered with water as if it had
rained overnight. In actuality, it
became sufficiently cold overnight that the dew point was achieved. The air became saturated, and water vapor
began condensing into liquid water. Even
in the summertime, the nighttime air may become sufficiently cold that the dew
point is achieved, thus forming dew.
Condensation is the changing
of state from water vapor to liquid water.
The condensing water must liberate heat to its surroundings to condense;
this liberated heat warms the surroundings.
Therefore, condensation is a warming process. Evaporation is the changing of state from
liquid water to water vapor. The
evaporating water must extract heat from its surroundings to evaporate; this
extracted heat cools the surroundings.
Therefore, evaporation is a cooling process. This is why we feel chilly immediately after
taking a shower. Our bodies are covered
with water that evaporates; the evaporating water extracts the heat needed for
that evaporation from our bodies, thus cooling our bodies. Drying our bodies with a towel removes water
that would have evaporated; this is why we feel less chilly whenever we dry
ourselves with a towel. This is also why
humans and some animals perspire (sweat).
The act of perspiring (sweating) covers our bodies with water that
evaporates; the evaporating water extracts the heat needed for that evaporation
from our bodies, thus cooling our bodies.
Suppose the surrounding air is very humid, perhaps close to
saturation. Some of the water vapor in
the air must condense to liquid water so that additional liquid water may
evaporate. Whereas perspiration (sweat)
on our skin may evaporate which cools our bodies since evaporation is a cooling
process, some water vapor in the surrounding air condenses to liquid water onto
our skin, adding heat back to our bodies since condensation is a warming
process. In this case, our bodies cannot
cool effectively, and we feel uncomfortable.
As a result, humid air feels warmer than its actual temperature. We can convert this discomfort into an
effective air temperature that is warmer than the actual air temperature. This effective air temperature is called the
heat stress index (or the heat index for short). For example, a meteorologist may report in
the summertime that the actual temperature today will be ninety degrees fahrenheit, but it will feel like ninety-five degrees fahrenheit. The
ninety degrees fahrenheit is the true air
temperature, while the ninety-five degrees fahrenheit
is the heat stress index (or simply the heat index). In the wintertime, wind makes the air feel
colder than its actual temperature. We
can convert this discomfort into another effective air temperature that is
colder than the actual air temperature.
This effective air temperature is called the windchill. For example, a meteorologist may report in
the wintertime that the actual temperature today will be thirty-five degrees fahrenheit, but it will feel like twenty-five degrees fahrenheit. The
thirty-five degrees fahrenheit is the true air
temperature, while the twenty-five degrees fahrenheit
is the windchill.
We may use all of these principles to construct a hygrometer, a device
that measures the relative humidity of the air.
A hygrometer is simply two thermometers.
One thermometer is wrapped in a wet cloth; this is called the wet-bulb
of the hygrometer. The other thermometer
that is not wrapped in a wet cloth is called the dry-bulb of the
hygrometer. Water will evaporate from
the wet-bulb thermometer. Since
evaporation requires heat, the evaporating water will extract heat from the
wet-bulb thermometer, giving it a colder temperature than the dry-bulb
thermometer. Again, evaporation is a
cooling process. From the difference in
temperature between the wet-bulb of the hygrometer and the dry-bulb of the
hygrometer, we can calculate the relative humidity of the air.
We now apply everything we
have discussed about relative humidity to cyclones and anticyclones. As we discussed, the low-pressure,
low-density thermal at the center of a cyclone rises, expands adiabatically,
and cools. Since it cools, its relative
humidity increases. As we also
discussed, the high-pressure, high-density thermal at the center of an
anticyclone sinks, contracts adiabatically, and warms. Since it warms, its relative humidity
decreases. In summary, the thermal at
the center of a cyclone becomes more humid as it rises, while the thermal at
the center of an anticyclone becomes less humid (or more dry) as it sinks. If the thermal at the center of a cyclone
becomes more humid as it rises, the dew point could be achieved, causing water
vapor to condense into liquid water.
However, even if the dew point is achieved, water vapor cannot condense
into liquid water in midair. The liquid
water requires a surface upon which to condense, such as the surface of the
leaves of trees or the surface of our car.
Fortunately, the atmosphere is not just air; there are tiny pieces of
dust and silt and salt in the atmosphere.
When the dew point is achieved, the water vapor can condense into liquid
water around these tiny pieces of dust and silt and salt, forming a microscopic
drop of water around each tiny piece of dust or silt or salt. For this reason, this tiny piece of dust or
silt or salt is called a condensation nucleus, since it is at the center of the
microscopic drop of water. The center of
anything is called its nucleus. For
example, the center of a biological cell is called the cellular nucleus, the
center of an atom is called the atomic nucleus, and the center of an entire
galaxy is called the galactic nucleus.
If the dew point is achieved causing water vapor to condense into
microscopic drops of water around these condensation nuclei, the thermal
becomes opaque. Ordinarily, air is
transparent, as we know from our daily experience. Almost every second of every day of our
lives, we effortlessly see through the air around us, since air is ordinarily
transparent. However, liquid water is
opaque. Actually, a small quantity of
liquid water is transparent; we can see through a glass of water for
example. However, larger and larger
quantities of liquid water become less and less transparent and more and more
opaque. It is rather difficult seeing
through a fish tank for example, and it is hopeless trying to see through the
ocean to the seafloor. Therefore, when
the dew point is achieved causing water vapor to condense into liquid water,
the thermal does indeed become opaque.
We can no longer see through the air; the thermal has turned from
invisible to visible. This is so
remarkable that the thermal deserves a special name. A thermal that has achieved the dew point
causing its water vapor to condense into liquid water and thus the thermal has
turned from transparent to opaque (from invisible to visible) is called a
cloud. We conclude that clouds form when
thermals rise, expand adiabatically, cool, and become more humid until the dew
point is achieved. This dew point is a
specific temperature. Therefore,
thermals must be lifted to a specific elevation to cool to the dew point. This elevation is called the lifting
condensation level (or the condensation level for short). We can almost always see the lifting
condensation level (or simply the condensation level) with our own eyes, since
clouds often have flat bottoms. This
flat bottom is the lifting condensation level (or simply the condensation
level). Below this elevation, the dew
point has not been achieved, and the thermals are still transparent
(invisible). Above this elevation, the
dew point has been achieved, and the thermals are opaque (visible) clouds.
We can categorize clouds into
three broad types: cumulus clouds, cirrus clouds, and stratus clouds. Cumulus clouds have the appearance of
cauliflower or puffs of cotton. Cirrus
clouds have the appearance of individual wisps or feathers. Finally, if there are so many clouds in the
sky that they all blend together to form one giant layer of cloud covering the
entire sky, this is a stratus cloud. The
word stratus is derived from a Latin word meaning layer. As we discussed, the word stratify (meaning
layered) is derived from the same Latin word.
The word stratum (a layer of sedimentary rock) also derives from the
same Latin word, as we discussed earlier in the course. Note that there are other cloud types in
addition to these three. For example,
many cirrus clouds that seem to almost blend together into one giant layer of
cloud covering the entire sky is called a cirrostratus cloud, meaning
intermediate between cirrus clouds and stratus clouds. As another example, many cumulus clouds that
seem to almost blend together into one giant layer of cloud covering the entire
sky is called a stratocumulus cloud, meaning intermediate between cumulus
clouds and stratus clouds. If it is
precipitating (raining or snowing) out of a cumulus cloud, then it is called a
cumulonimbus cloud. If it is
precipitating (raining or snowing) out of a stratus cloud, then it is called a
nimbostratus cloud. If the air becomes
sufficiently cold that it achieves the dew point without having to be pushed up
to higher elevations, a cloud will form at or near the ground. This type of cloud is called fog.
We have already discussed
enough meteorology to somewhat reliably predict weather over a timescale of a
few hours using only a barometer. The
rising or the falling of the air pressure as indicated by the barometer is
called the barometric tendency. If the
barometric tendency is falling, then low-pressure, low-density thermals must be
rising, expanding adiabatically, cooling, and becoming more humid. The relative humidity may increase
sufficiently for the dew point to be achieved, forming clouds and perhaps even
precipitation (rain or snow).
Conversely, if the barometric tendency is rising, then high-pressure,
high-density thermals must be sinking, contracting adiabatically, warming, and
becoming less humid. The relative
humidity may decrease sufficiently for liquid water to evaporate back into
water vapor. In other words, thermals
will turn from opaque (visible) clouds to transparent (invisible) air; we will
have a clear day. In summary, a falling
barometric tendency is an indication of what is commonly considered to be bad
weather, while a rising barometric tendency is an indication of what is
commonly considered to be good weather.
The rate at which a rising
thermal cools before it becomes a cloud is called the dry adiabatic rate of
cooling (or the dry adiabatic rate for short).
The rate at which a rising thermal cools after it becomes a cloud is
called the wet adiabatic rate of cooling (or the wet adiabatic rate for
short). A thermal becomes a cloud when
water vapor condenses into liquid water.
This liberates heat, thus making the thermal warmer. Again, condensation is a warming
process. Therefore, the wet adiabatic
rate of cooling is always more shallow than the dry adiabatic rate of
cooling. That is, the dry adiabatic rate
of cooling is always more steep than the wet adiabatic rate of cooling. The rate at which the surrounding atmosphere
cools with rising elevation is called the environmental lapse rate of cooling
(or the environmental lapse rate for short).
Suppose the environmental lapse rate is more shallow than the wet adiabatic
rate which itself must be more shallow than the dry adiabatic rate. In other words, both adiabatic rates are more
steep than the environmental lapse rate.
In this case, either before or after a thermal becomes a cloud, its rate
of cooling is very steep. The rate of
cooling of the thermal may be sufficiently steep that the thermal becomes so
cold and so dense that it is forced to sink to lower elevations. This is called absolute stability, and what
is commonly considered to be bad weather such as clouds or precipitation (rain
or snow) will be less likely.
Conversely, suppose the environmental lapse rate is steeper than the dry
adiabatic rate which itself must be steeper than the wet adiabatic rate. In other words, both adiabatic rates are more
shallow than the environmental lapse rate.
In this case, either before or after a thermal becomes a cloud, its rate
of cooling is shallow. Thermals will
probably not cool sufficiently to become dense enough to sink to lower
elevations. In other words, thermals are
more likely to rise to higher elevations.
This is called absolute instability, and what is commonly considered to
be bad weather such as clouds and perhaps even precipitation (rain or snow)
will be more likely. It is also possible
for the environmental lapse rate to be steeper than the wet adiabatic rate but
more shallow than the dry adiabatic rate.
In this case, a thermal before it becomes a cloud may cool to attain
sufficiently high density to sink, resulting in what is commonly considered to
be good weather. However, if the thermal
reaches the lifting condensation level and becomes a cloud, its rate of cooling
slows. Hence, the thermal will continue
to rise, resulting in what is commonly considered to be bad weather. This is called conditional instability, and
either good weather or bad weather may result under these circumstances.
Our discussion leads us to
conclude that weather is strongly determined by lifting, the rising of
thermals. There are three mechanisms
that could cause lifting: orographic lifting, convergence lifting, and frontal
wedging. Orographic lifting is caused by
mountains pushing air aloft. This term
is derived from the Greek root oro- for mountain, as
we discussed earlier in the course. When
winds encounter a mountain, much of the air blows up over the mountain. As the air rises, it expands adiabatically,
cools, and becomes more humid. If the
dew point is achieved, clouds form, and precipitation may occur. Therefore, we expect a humid climate on the
windward side of a mountain range. The
windward side of anything is the side that faces the wind. The opposite of the windward side of anything
is its leeward side, which faces away from the wind. The adjective leeward is derived from the
noun lee, which means shelter. For
example, the lee of a building or the lee of a rock faces away from the wind
and hence provides shelter from the wind.
On the leeward side of a mountain range, air sinks, contracts
adiabatically, warms, and becomes less humid.
Therefore, we expect an arid (dry) climate on the leeward side of a
mountain range. Actually, we expect an
arid (dry) climate for an additional reason: any moisture that was in the air
probably precipitated out of the air on the windward side of the mountain
range. With moisture subtracted and in
addition warming temperatures from sinking air, we expect an extremely arid
(dry) climate on the leeward side of mountain ranges. These are called rainshadow
deserts. For example, the contiguous
United States is at the midlatitudes, and the
prevailing winds at the midlatitudes blow from the
west, as we will discuss shortly.
Therefore, the west side of the Rocky Mountains is its windward side,
while the east side of the Rocky Mountains is its leeward side. The leeward side (the east side) of the Rocky
Mountains is the Great Plains of the United States, which is a rainshadow desert.
Although there is agriculture in the Great Plains, the soil is not as
productive as the farmland of the midwestern United
States, which is further east of the Great Plains. Convergence lifting is caused by crowded
winds pushing air aloft. Consider an
island or a peninsula surrounded on many sides by water. Every day, sea breezes will blow from the
surrounding waters toward the island or peninsula, as we discussed. These breezes become crowded and thus push
each other upward. As the air rises, it
expands adiabatically, cools, and becomes more humid. If the dew point is achieved, clouds form,
and precipitation may occur. Therefore,
we expect islands and peninsulas to have humid climates. Actually, we expect the climate to be
extremely humid, since the winds originally came from the surrounding waters,
where evaporation added significant moisture to the sea breezes. With moisture added and in addition cooling
temperatures from rising air, we expect extremely humid climates on islands and
peninsulas. For example, Florida is a
peninsula in the southeastern United States.
Every day, a sea breeze blows from the Gulf of Mexico from the west
towards Florida. Every day, a sea breeze
blows from the Atlantic Ocean from the east towards Florida. Every day, a sea breeze blows from the
Caribbean Sea from the south towards Florida.
These sea breezes were already humid, since they came from bodies of
water where evaporation added moisture to the winds. In addition, these sea breezes become crowded
over Florida and thus push each other upward.
The air rises, expands adiabatically, cools, and becomes even more
humid. The dew point is achieved, clouds
form, and rain occurs. This explains why
Florida has an extremely humid climate.
In fact, the entire peninsula is infested with amphibians and reptiles
as a result of this extreme humidity.
Frontal wedging is caused by one air mass pushing another air mass
aloft. This is the most important type
of lifting that determines weather patterns, as we will discuss shortly.
The lightest type of liquid
precipitation is called mist. Heavier
than mist is drizzle, and the heaviest liquid precipitation is called
rain. The lightest freezing precipitation
is called snow. Heavier than snow is
freezing drizzle. Heavier than freezing
drizzle is called sleet. Even heavier
than sleet is called graupel, and hail is the
heaviest freezing precipitation. Hail is
quite dangerous; many people have been killed from falling hail. Snow is very light because it is composed of
individual snowflakes, which are themselves composed of mostly air. Since clouds form aloft (at higher
elevations) in the troposphere where the air temperature is colder,
precipitation almost always begins in the frozen state, such as snow or
sleet. On its way down to lower
elevations, the precipitation warms and melts into liquid precipitation such as
rain. This is usually the case even in
the summertime; warm rain in the summertime most likely began as snow or sleet
from clouds at higher elevations in the troposphere that melted into rain on
its way down toward lower elevations in the troposphere. We can personally experience this extreme
temperature difference between the lower troposphere and the upper troposphere
with a sufficiently tall mountain. As we
climb the mountain, the air temperature becomes colder and colder until we
reach the summit of the mountain, where it may be so cold that it is snowing. This is usually the case even in the summertime. The air temperature at the bottom of the
mountain may be quite hot in the summertime.
Nevertheless, the air temperature becomes colder and colder as we climb
the mountain until (if the mountain is sufficiently tall) the air temperature
is so cold that it is snowing at the summit of the mountain, even in summertime
when the base of the mountain is still quite hot!
Global (Large-Scale) Meteorological Dynamics
The Coriolis force caused by
the Earth’s rotation causes the global circulation of air in the atmosphere to
be complex. In order to emphasize the
complications caused by the rotation of the Earth, let us first suppose that
the Earth were not rotating. In this
case, the global circulation of air in the atmosphere would be simple. Since the equator is hot throughout the
entire year, the air at the equator is at low pressure. Since the poles are cold throughout the
entire year, the air at the poles is at high pressure. The pressure gradient force would then push
air from high pressure at the poles toward low pressure at the equator. The result is that winds would blow from the
north in the northern hemisphere and from the south in the southern
hemisphere. Since we always name wind
based on the direction it is blowing from, the winds in the northern hemisphere
would be a north wind, while winds in the southern hemisphere would be a south
wind. These are known as the prevailing
winds. Caution: wind does not always blow
in the directions of these prevailing winds; variations in pressure may cause
winds to blow in various different directions.
The prevailing winds are the directions in which the wind generally or
usually blows, not the directions in which the wind always blows. If the Earth were not rotating, winds in the
northern hemisphere would generally or usually be a north wind from the north
pole toward the equator, and winds in the southern hemisphere would generally
or usually be a south wind from the south pole toward the equator. At the equator, the low-pressure, low-density
air would rise, becoming high pressure aloft.
The high pressure at the poles is low pressure aloft. Again, the pressure gradient force pushes air
from high pressure toward low pressure.
Hence, the pressure gradient force would push the risen air at the
equator toward the poles, where the air would sink until the pressure gradient
force pushes the air back toward the equator.
This overall motion is called a circulation cell. Notice there would be only one circulation
cell in each hemisphere if the Earth were not rotating. This discussion completely summarizes the
global circulation of air in the atmosphere if the Earth were not rotating.
Of course, the Earth is
rotating, causing a Coriolis force and hence tremendous complications to the
simplistic model we have just presented.
The low-pressure, low-density air still rises at the equator, and the
pressure gradient force still pushes this risen air toward the poles. However, by the time the air reaches roughly
thirty degrees latitude in each hemisphere, the air has cooled sufficiently to
sink. The pressure gradient force then
pushes this air back toward the equator, completing the tropical circulation
cells. However, the Coriolis force
causes rightward deflections in the northern hemisphere and leftward deflections
in the southern hemisphere. The net
result of the pressure gradient force together with the Coriolis force is that
the prevailing winds (near mean sea level) from roughly 30°N
latitude to 0° latitude (the equator) blow from the northeast; these are called
the northeast trade winds, since we always name wind based on the direction it
is blowing from. The prevailing winds
(near mean sea level) from roughly 30°S latitude to
0° latitude (the equator) blow from the southeast; these are called the southeast
trade winds, since we always name wind based on the direction it is blowing
from. The term trade wind is used since
these winds facilitated trade between the Old World and the New World by
pushing sailing ships across the Atlantic Ocean from Europe and Africa toward
North America and South America. The
high-pressure, high-density air still sinks at the poles, and the pressure
gradient force still pushes this air toward the equator. However, by the time the air reaches roughly
sixty degrees latitude in each hemisphere, the air has warmed sufficiently to
rise. The pressure gradient force still
pushes this risen air toward the poles where it sinks, completing the polar
circulation cells. However, the Coriolis
force causes rightward deflections in the northern hemisphere and leftward
deflections in the southern hemisphere.
The net result of the pressure gradient force together with the Coriolis
force is that the prevailing winds (near mean sea level) from 90°N latitude (the north pole) to roughly 60°N latitude blow from the northeast; these are called the
polar northeasterlies, since we always name wind
based on the direction it is blowing from.
The prevailing winds (near mean sea level) from 90°S
latitude (the south pole) to roughly 60°S latitude
blow from the southeast; these are called the polar southeasterlies,
since we always name wind based on the direction it is blowing from. Notice that air sinks at roughly thirty
degrees latitude in each hemisphere, while air rises at roughly sixty degrees
latitude in each hemisphere. Sinking air
is high-density, high-pressure air, while rising air is low-density,
low-pressure air. Therefore, we have
high pressure at roughly thirty degrees latitude in each hemisphere, and we
have low pressure at roughly sixty degrees latitude in each hemisphere. The pressure gradient force pushes air from
high pressure toward low pressure.
Hence, wind will blow from roughly thirty degrees latitude to roughly
sixty degrees latitude in each hemisphere, where the air rises and is pushed
back to thirty degrees latitude where it sinks, completing the midlatitude circulation cells. However, the Coriolis force causes rightward
deflections in the northern hemisphere and leftward deflections in the southern
hemisphere. The net result of the
pressure gradient force together with the Coriolis force is that the prevailing
winds (near mean sea level) from roughly 30°N
latitude to roughly 60°N latitude blow from the
southwest; these are called the southwesterlies,
since we always name wind based on the direction it is blowing from. The prevailing winds (near mean sea level)
from roughly 30°S latitude to roughly 60°S latitude blow from the northwest; these are called the
northwesterlies, since we always name wind based on
the direction it is blowing from. To
summarize, there are three prevailing winds in each hemisphere, and there are
three circulation cells in each hemisphere.
In the northern hemisphere, the prevailing winds (near mean sea level)
are the northeast trade winds near the equator, the southwesterlies
at the midlatitudes, and the polar northeasterlies near the north pole. In the southern hemisphere, the prevailing
winds (near mean sea level) are the southeast trade winds near the equator, the
northwesterlies at the midlatitudes,
and the polar southeasterlies near the south
pole. The circulation cells are called
the two Hadley cells (one in each hemisphere) near equator, named for the
British meteorologist George Hadley, the two Ferrel
cells (one in each hemisphere) at the midlatitudes,
named for the American meteorologist William Ferrel,
and the two polar cells (one in each hemisphere) near the poles. If the Earth rotated faster, the Coriolis
force would be stronger, thus causing more prevailing winds and more
circulation cells in each hemisphere. If
the Earth rotated slower, the Coriolis force would be weaker, thus causing
fewer prevailing winds and fewer circulation cells in each hemisphere. If the Earth stopped rotating, the Coriolis
force would vanish, and there would be only one prevailing wind and only one
circulation cell in each hemisphere, as we discussed with our simplistic
non-rotating model. A spectacular
example of the effects of a strong Coriolis force on the global circulation of
air is the planet Jupiter, which rotates more than twice as fast as the
Earth. In fact, Jupiter is the fastest
rotating planet in the Solar System.
Therefore, Jupiter has the strongest Coriolis force out of all the
planets in the Solar System. This very
strong Coriolis force has divided Jupiter’s atmosphere into many prevailing
winds and many circulation cells. We can
actually see these winds in photographs of Jupiter. We can even see these winds if we look at
Jupiter with our own eyes through a sufficiently powerful telescope.
At the equator, there is
little to no wind, since the air is rising; this is called the equatorial low,
since low-pressure, low-density air rises.
This rising air expands adiabatically, cools, and becomes more humid. The dew point may be achieved, forming clouds
and rain. Indeed, there is a perpetual
band of clouds at the equator, and the perpetual rain from these clouds causes
tropical rainforests at and near the equator, such as the Amazon rainforest in
northern South America, the Congo rainforest in central Africa, and the
Indonesian rainforests. Sailing ships
that found themselves at the equatorial low would become stuck, since there are
no winds to push ships. For this reason,
the equatorial low is also called the doldrums.
Sailors would pray that their ship happens to drift slightly to the
north or slightly to the south to catch one of the trade winds that would push
them again. At least the sailors could
drink the perpetual rainwater while stuck at the equatorial low (the
doldrums). As we discussed earlier in
the course, the lack of wind at the equatorial low permits the oceanic
equatorial countercurrents to flow virtually unhindered against the direction
of other oceanic surface currents near the equator. At roughly thirty degrees latitude in both
hemispheres, there is also little to no wind, but for the opposite reason. The air is sinking; these are called the
subtropical highs, since high-pressure, high-density air sinks. This sinking air contracts adiabatically,
warms, and becomes less humid (more dry).
Hence, we do not have clouds or rain.
Indeed, there is a perpetual band of clear skies free of clouds at
roughly thirty degrees latitude in both hemispheres. The perpetual lack of rain causes hot deserts
at and near roughly thirty degrees latitude in both hemispheres, such as the
Basin and Range in southwestern United States and northwestern Mexico
(including the Mojave Desert, the Sonoran Desert, and the Chihuahuan
Desert), the Sahara in northern Africa, the Arabian Desert in the Arabian peninsula,
the Gobi in China and Mongolia, the Patagonian Desert in Argentina, the
Kalahari in southern Africa, and the Great Australian Desert in Australia
(including the Great Victoria Desert, the Great Sandy Desert, the Tanami
Desert, the Simpson Desert, and the Gibson Desert). Sailing ships that found themselves at the
subtropical high in either hemisphere would become stuck, since there are no
winds to push ships. There would also be
no rain for the sailors to drink.
Therefore, not only would sailors pray that their ship happens to drift
slightly to the north or slightly to the south to catch prevailing winds that
would push them again, but the sailors would also kill their horses to stretch
out their limited supply of drinking water.
For this reason, the subtropical highs are also called the horse
latitudes. At roughly sixty degrees
latitude in both hemispheres, there is little to no wind, since the air is
rising; these are called the subpolar lows, since low-pressure, low-density air
rises. This rising air expands
adiabatically, cools, and becomes more humid.
The dew point may be achieved, forming clouds and rain. Indeed, there is a perpetual band of clouds
at roughly sixty degrees latitude in both hemispheres, and the perpetual rain
from these clouds causes boreal forests (cold forests or taigas) at and near
roughly 60°N latitude, including the Canadian boreal
forests, the Scandinavian boreal forests, and the Russian boreal forests. Theoretically, there would be boreal forests
(cold forests or taigas) at and near roughly 60°S
latitude if there were land at these latitudes.
At the poles, there is also little to no wind since the air is sinking;
these are called the polar highs, since high-pressure, high-density air sinks. This sinking air contracts adiabatically,
warms, and becomes less humid (more dry).
Hence, we do not have clouds or rain.
Indeed, there is a perpetual area free of clouds at and near the poles
in both hemispheres. To summarize, we
have the equatorial low (the doldrums) at the equator, we have the subtropical
highs (the horse latitudes) at roughly thirty degrees latitude in both
hemispheres, we have the subpolar lows at roughly sixty degrees latitude in
both hemispheres, and we have the polar highs at ninety degrees latitude in both
hemispheres. At the lows, we have rising
air, causing humid climates from perpetual clouds and rain. At the highs, we have sinking air, causing
arid (dry) climates from the perpetual absence of clouds and rain. Actually, we expect arid climates at the
highs for an additional reason: any moisture that was in the air precipitated
out of the rising air at the equatorial low and at both subpolar lows before
being pushed toward the highs where the air sinks. With moisture subtracted and in addition warming
temperatures from sinking air, we expect extremely arid (dry) climates at both
subtropical highs and at both polar highs.
An air mass is an enormous
mass of air that has roughly the same temperature and pressure throughout its
volume at a given elevation. We can
classify air masses based on their temperature.
An air mass that forms near the equator will be warm; these are called
tropical air masses, which we label with the uppercase (capital) letter T for
tropical. An air mass that forms near
the poles will be cold; these are called polar air masses, which we label with
the uppercase (capital) letter P for polar.
We can also classify air masses based on their moisture. An air mass that forms over the ocean or any
body of water will be humid, since evaporating water will add moisture to the
air mass; these are called maritime air masses, which we label with the
lowercase letter m for maritime. An air
mass that forms over a continent will be dry, since there is little water on
the continent to evaporate to add moisture to the air mass; these are called
continental air masses, which we label with the lowercase letter c for
continental. To summarize, there are
four different types of air masses. An
air mass that forms over a body of water near the equator will be humid and
warm; these are called maritime tropical air masses, which we label with the
symbol mT. An
air mass that forms over a body of water near the poles will be humid and cold;
these are called maritime polar air masses, which we label with the symbol mP. An air mass that
forms over a continent near the equator will be dry and warm; these are called
continental tropical air masses, which we label with the symbol cT. Finally, an air
mass that forms over a continent near the poles will be dry and cold; these are
called continental polar air masses, which we label with the symbol cP. We must
emphasize that once an air mass is born of a certain type, it does not remain
that type permanently. In other words,
the particular type of an air mass can change.
For example, an air mass that forms near the equator will be warm. This would be a tropical air mass, but if
this air mass happens to move toward one of the poles, it may become colder and
colder until we must reclassify it as a polar air mass. The reverse can occur. An air mass that forms near one of the poles
will be cold. This would be a polar air
mass, but if this air mass happens to move toward the equator, it may become
warmer and warmer until we must reclassify it as a tropical air mass. As another example, an air mass that forms
over a continent will be dry. This would
be a continental air mass, but if this air mass happens to move over the ocean
or any body of water, it may become more and more humid as evaporating water
adds more and more moisture to the air mass.
Eventually, we must reclassify it as a maritime air mass. The reverse can occur. An air mass that forms over the ocean or any
body of water will be humid. This would
be a maritime air mass, but if this air mass happens to move over a continent,
it may lose more and more moisture through precipitation that will not be
replenished, since there is little water on the continent to evaporate. The air mass becomes less and less humid
until we must reclassify it as a continental air mass. A spectacular example of the changing of an
air mass is lake-effect snow. The five
Great Lakes are between the United States and Canada, two countries in the North
American continent. Cities on the
windward side of the Great Lakes may experience very little snow, since air
masses that form over either Canada or the United States would be continental
(dry) air masses. However, a city on the
leeward side (the opposite side of the windward side) of the Great Lakes may
experience enormous amounts of snow.
This is because a continental air mass that moves over the Great Lakes
will become more and more humid as water evaporates from the Great Lakes. By the time the air mass has crossed the
Great Lakes, the air mass has become so humid that it is now a maritime air
mass. The humid maritime air mass then
precipitates snow onto these cities on the opposite side of the Great Lakes
from cities that experienced no snow from the same air mass when it was formerly
a continental (dry) air mass before crossing the Great Lakes. As a result, two cities that are not
particularly distant from each other may nevertheless experience vastly
different amounts of precipitation, since these two cities are on two opposite
sides of a large body of water.
The Bjørgvin Theory of Meteorology
The fundamental theory of
meteorology was formulated by the Norwegian meteorological physicist Vilhelm Bjerknes and other
meteorologists in Bjørgvin, Norway. Consequently, we will refer to the
fundamental theory of meteorology as the Bjørgvin
Theory of Meteorology. According to the Bjørgvin Theory of Meteorology, the Earth’s troposphere
(the lowest layer of the atmosphere) is divided into many pieces called air
masses. These air masses are pushed by
the prevailing winds, and much meteorological activity (commonly known as
weather) occurs at the boundary between two air masses, which is called a
front. This Bjørgvin
Theory of Meteorology, the fundamental theory of meteorology, is remarkably
similar to the Theory of Plate Tectonics, the fundamental theory of
geology. As we discussed earlier in the
course, the Theory of Plate Tectonics states that the Earth’s lithosphere (the
uppermost layer of the geosphere) is divided into many pieces called tectonic
plates. These tectonic plates are pushed
by convection cells in the asthenosphere (underneath the lithosphere), and much
geological activity occurs at the boundary between two tectonic plates. These two fundamental theories have further similarities. Just as there are different types of tectonic
plate boundaries that cause different types of geological activities as we
discussed earlier in the course, there are different types of fronts (air mass
boundaries) that cause different types of meteorological activities (commonly
known as weather). A cold air mass
pushing on a warm air mass is called a cold front. The symbol for a cold front on a weather map
is triangles along the front pointing in the direction in which the cold front
is moving. A warm air mass pushing a
cold air mass is called a warm front.
The symbol for a warm front on a weather map is semicircles along the
front again pointing in the direction in which the warm front is moving. Cold fronts move faster than warm fronts, as
we will discuss shortly. Therefore, a
faster-moving cold front can catch up to and merge with a slower-moving warm
front. This is called an occluded
front. We will discuss the meaning of
the term occluded shortly. The symbol
for an occluded front on a weather map is both triangles and semicircles along
the front again pointing in the direction in which the occluded front is
moving. A front that does not move for
several days or perhaps even a couple weeks is called a stationary front. The symbol for a stationary front on a
weather map is both triangles and semicircles along the front, but the
triangles and the semicircles point in two opposite directions.
The meteorological term front
is borrowed from military terminology. Vilhelm Bjerknes and other
meteorologists formulated the Bjørgvin Theory of
Meteorology during and shortly after the Great War (commonly known as the First
World War or World War I) roughly one hundred years ago. The Great War (the First World War or World
War I) was the most global and most horrific war in human history up to that
time, compelling many people throughout the world to often draw military
analogies. A military front is the
boundary between two opposing armies. If
one army advances over (or pushes) the other army, the military front will move
with the advancing army. If two armies
are equally matched, the military front will not move. This is called a stationary military front,
the textbook example being the western front of the Great War (the First World
War or World War I). The western front
remained stationary for most of the years of the Great War since the combined
British and French armies on the western side of the western front equally
matched the German army on the eastern side of the western front. The western front did not move until the United
States joined the British and the French toward the end of the Great War. The combined British, French, and American
armies now had sufficient momentum to advance upon the German army, finally
pushing the western front eastward. As Vilhelm Bjerknes and other
meteorologists formulated the Bjørgvin Theory of
Meteorology during and shortly after the Great War, they imagined air masses
pushing each other as if they were opposing armies. It is for this reason that meteorologists to
the present day refer to the boundary between two air masses as a front. The
term for a meteorological front that does not move, a stationary meteorological
front, was literally copied from the term for a military front that does not
move, which is again a stationary military front.
Since warm air rises and cold
air sinks, the actual front between two air masses is not a perfectly vertical
wall. The actual front between the two
air masses is an inclined wall, since the rising warm air will be above the
sinking cold air. In other words, the
sinking cold air will be below the rising warm air. Therefore, a cold front is inclined backward
as the cold air mass pushes the warm air mass, while a warm front is inclined
forward as the warm air mass pushes the cold air mass. Moreover, since cold air is more dense than
warm air, the cold air mass can strongly push the warm air mass, making the
cold front more vertical than a warm front.
That is, a warm front is more shallow than a cold front. Since warm fronts are more shallow, it takes
a longer duration of time for a warm front to move over any particular
location. Since cold fronts are more
vertical, it takes a shorter duration of time for a cold front to move over any
particular location. Along both cold
fronts and warm fronts, rising hot air will expand adiabatically and cool thus
becoming more humid; the dew point could be achieved, causing clouds and
possibly precipitation along the front.
Since a warm front is more shallow, all of the precipitation will be
spread over a larger area; consequently, the precipitation along a warm front
is often mild. The usual weather
associated with a warm front is gentle precipitation over a long duration of
time (often many hours) followed by warmer temperatures as compared with the
temperatures before the warm front arrived.
Since a cold front is more vertical, all of the precipitation will be
concentrated over a smaller area; consequently, the precipitation along a cold
front is often severe. The usual weather
associated with a cold front is intense precipitation over a brief duration of
time (often only a few minutes), followed by colder temperatures as compared
with the temperatures before the cold front arrived.
As a concrete application of
the Bjørgvin Theory of Meteorology, consider weather
patterns in the contiguous United States, which is at the midlatitudes
of the northern hemisphere. The
prevailing winds of the midlatitudes of the northern
hemisphere are the southwesterlies. Therefore, weather patterns (both good
weather and bad weather) are pushed from the west toward the east by these southwesterlies.
This explains why weather patterns move across the United States from
the west toward the east. Philadelphia
is west of New York City, and Chicago is further west from Philadelphia. A weather pattern in Chicago will move from
Chicago toward Philadelphia, and the weather pattern will continue to move from
Philadelphia toward New York City. Now
consider a low-pressure system being pushed from the west toward east by the southwesterlies.
Winds will blow inward toward this moving low-pressure system. Winds from the south will carry warmer air,
since they are from equatorial latitudes.
Winds from the north will carry cooler air, since they are from polar
latitudes. Since the United States is in
the northern hemisphere, the Coriolis force deflects all of these winds to the
right ultimately circulating them counterclockwise around this moving
low-pressure system. Hence, the warmer
winds from the south will be deflected to the east and will collide with cooler
air. Warm air pushing cold air is a warm
front; hence, there will be a warm front to the east of the moving low-pressure
system. Also, the cooler winds from the
north will be deflected to the west and will collide with warmer air. Cool air pushing warm air is a cold front;
hence, there will be a cold front to the west of the moving low-pressure
system. As this low-pressure system is
pushed by the southwesterlies, any given geographical
region of the United States will first be attacked by the warm front, often
bringing many hours of gentle precipitation followed by warmer
temperatures. Then, the same
geographical region will be attacked by the cold front, often bringing only a
few minutes of intense precipitation followed by colder temperatures. Since cold fronts move faster than warm
fronts, the cold front to the west of the low-pressure system may catch up to
and merge with the warm front to the east of the low-pressure system, forming an
occluded front. The cold air to the west
of the former cold front merges with the cold air to the east of the former
warm front, thus squeezing the warm air between them and wedging it upward,
since warm air rises. Hence, the
formation of the occluded front begins the dispersion of the entire
low-pressure system. This is the reason
these fronts are called occluded fronts.
In colloquial English, the verb to occlude means to stop or to obstruct
or to close.
As rising air and sinking air
rub against each other, electrons are transferred from one thermal to
another. This may create an electric
field between the clouds and the ground.
Usually, air is a poor conductor of electricity; air is usually an
electrical insulator. However, all
electrical insulators will conduct electricity if subjected to electric fields
of sufficiently enormous strength. The
threshold electric field at which an electrical insulator becomes an electrical
conductor is called the dielectric breakdown of the material. The dielectric breakdown of air is roughly
three million volts per meter. If the
electric field in air exceeds roughly three million volts per meter, the air
actually becomes an electric conductor.
In this case, electrons can flow between the clouds and the ground. This flow of electrons is called lightning. There is an enormous quantity of energy
associated with lightning. Some of this
energy is transferred to the air itself, causing a loud, explosive sound called
thunder. In brief, lightning causes
thunder. The light from the lightning
propagates at the speed of light, which is almost one million times faster than
the speed of sound. In other words,
sound propagates almost one million times slower than the speed of light. The speed of light is so fast that we never
notice its propagation in our daily experiences; light seems to propagate
instantaneously fast. However, sound
propagates sufficiently slow that we notice its propagation in some of our
daily experiences. For example, some of
us notice while sitting near the outfield of a baseball stadium that there is a
delay between seeing and hearing a baseball bat crack a baseball. Some of us notice while sitting near the
infield of a baseball stadium that there is a delay between seeing and hearing
a baseball land in the baseball mitt of an outfielder. Some of us notice that there is a delay
between seeing and hearing a hockey stick strike a hockey puck. In all such examples, we see the event first,
then we hear the event second. Again,
light seems to propagate instantaneously fast, while sound propagates slow
enough that the sound arrives noticeably after the light. The speed of sound through air is roughly one
mile per five seconds, which we may restate as roughly five seconds per mile. We can use this relatively slow propagation
of sound to estimate how far away a storm is occurring from our location. We simply count the number of seconds
starting from when we see lightning until we hear thunder. For every five seconds we count, the storm is
roughly one mile distant. For example,
if we see lightning and count fifteen seconds until we hear thunder, the storm
is roughly three miles distant, since every five seconds we counted corresponds
to roughly one mile of distance. If we
count many seconds after seeing lightning but never hear thunder, this means
that the storm is very far away. Thunder
propagates outwards in all directions, spreading its total energy thinner and
thinner. By the time the thunder arrives
at our location, the sound energy was too diluted for our ears to hear. At the opposite extreme, suppose we see
lightning and immediately thereafter we hear thunder; in other words, suppose
we did not have the opportunity to count to even one second before hearing
thunder. This means that the storm is
very close; we are probably located within the storm itself.
A tornado is a continental
storm with fast, circulating winds around an extremely low-pressure
thermal. Most tornadoes are a few dozen
meters across. A large tornado could be
a couple hundred meters across. Enormous
tornados that are one kilometer across are very rare. Most of the tornadoes in the world occur in
the midwestern United States. This is because cP
air masses (continental polar air masses) form over Canada, since Canada is in
the North American continent and is near the north pole, while cT air masses (continental tropical air masses) form over
Mexico, since Mexico is also in the North American continent but near the
equator. Moreover, there are two
mountain ranges along both coasts of North America: the Rocky Mountains along
the Pacific coast (the west coast) and the Appalachian Mountains along the
Atlantic coast (the east coast). These
two mountain ranges tend to confine air masses between them. Hence, cP air
masses that form over Canada and cT air masses that
form over Mexico tend to collide over the country that is between Canada and
Mexico; that country is the United States.
For all these reasons, most of the tornadoes in the entire world occur
in the midwestern United States.
The Fujita scale (or F-scale)
is a tornado wind-speed scale, named for the Japanese-American meteorologist
Tetsuya Theodore Fujita who formulated this scale. The weakest tornadoes are designated F0. More powerful
than F0 would be called F1
followed by F2, F3, and F4. The most powerful
tornados are designated F5. Even an F0 tornado
is powerful enough to destroy entire towns; many people have been killed by F0 tornados, the weakest scale of tornado. We must always seek shelter during a tornado
warning, regardless of the Fujita-scale designation of the tornado.
The largest storms in the
entire world form from low-pressure mT air masses
(maritime tropical air masses). These
storms are called hurricanes if they form in the Atlantic Ocean, and they are
called typhoons if they form in the Pacific Ocean. Other than their oceanic location, there is
no difference between a hurricane and a typhoon. The development of a hurricane/typhoon is as
follows. A slightly low-pressure mT air mass is called a tropical disturbance. If a tropical disturbance happens to form at
or near the equator, the Coriolis force will be too weak to cause any
circulation of winds, and the tropical disturbance will quietly disperse. However, if a tropical disturbance happens to
form significantly north or south of the equator, the Coriolis force may be
strong enough to circulate the winds.
When the winds are sufficiently strong, the tropical disturbance becomes
a tropical depression. On rare
occasions, the winds are so strong that the tropical depression may become a
tropical storm. At this point, the
tropical storm is given a human name, as we will discuss. On very rare occasions, the winds become so
extraordinarily strong that the tropical storm becomes a
hurricane/typhoon. In this case, the
hurricane/typhoon retains its tropical-storm human name, as we will
discuss. To summarize, first we have a
tropical disturbance, then we have a tropical depression, then we have a
tropical storm, then we have a hurricane/typhoon.
Since a hurricane/typhoon is
a low-pressure system, the winds in a hurricane/typhoon circulate
counterclockwise in the northern hemisphere but circulate clockwise in the
southern hemisphere. The winds circulate
around the eye of the hurricane/typhoon, where there is calm weather and clear
skies. When a northern-hemisphere
hurricane/typhoon attacks a continent, the winds to the right of the eye push
the ocean waters onto the continent.
This is called the storm surge.
The winds to the left of the eye push the ocean waters away from the
continent; thus, there is no storm surge to the left of the eye. Therefore, most of the destruction to the
left of the eye is from the winds themselves.
To summarize, most of the devastation from a northern-hemisphere
hurricane/typhoon is from the storm surge to the right of the eye, but most of
the devastation from a northern-hemisphere hurricane/typhoon is from the winds
to the left of the eye. These directions
are reversed in the southern hemisphere, but note that hurricanes/typhoons are
rare in the southern hemisphere. This is
because winds would circulate clockwise around hurricanes/typhoons in the
southern hemisphere, causing the winds on the southern edge of the storm to
blow from the east. However, the
prevailing winds at the midlatitudes of the southern hemisphere
blow from the west, thus weakening the storm winds and preventing tropical
storms from strengthening to hurricanes/typhoons. Of course, we could present the same argument
for the northern hemisphere. Winds
circulate counterclockwise around hurricanes/typhoons in the northern
hemisphere, causing the winds on the northern edge of the storm to again blow
from the east. Again, the prevailing
winds at the midlatitudes of the northern hemisphere
blow from the west, which should again weaken the storm winds and prevent
tropical storms from strengthening to hurricanes/typhoons. Indeed, this is an important reason why the
strengthening of tropical storms to hurricanes/typhoons is rare even in the
northern hemisphere. However, the
Antarctic Circumpolar Current at the midlatitudes of
the southern hemisphere is the strongest oceanic surface current in the entire
world, as we discussed earlier in the course.
Although the Antarctic Circumpolar Current is caused by the prevailing
winds at the midlatitudes of the southern hemisphere,
this oceanic surface current is so strong that it actually pushes back on the
atmosphere, making the midlatitude prevailing winds
in the southern hemisphere stronger than the midlatitude
prevailing winds in the northern hemisphere.
The stronger prevailing winds at the midlatitudes
of the southern hemisphere make the formation of hurricanes/typhoons in the
southern hemisphere much less likely than the formation of hurricanes/typhoons
in the northern hemisphere. In either
hemisphere, most of the devastation inland from a hurricane/typhoon is from
flooding from rain. As we will discuss
later in the course, flooding is the most common and the most destructive of
all natural disasters.
The human name of a tropical
storm in the Atlantic Ocean is chosen from six lists each containing twenty-one
alphabetized human names. For example,
the first tropical storm in the Atlantic Ocean in the year 2023 was named
tropical storm Arlene, the second was named tropical storm Bret, the third was
named tropical storm Cindy, and so on and so forth. Notice that the names are in alphabetical
order. Although the English alphabet has
twenty-six letters, these six lists each have only twenty-one names because
names beginning with the five letters Q, U, X, Y, and Z are not used, since
human names beginning with any of these five letters are rare. If a tropical storm is promoted to a
hurricane, then it retains its human name.
For example, tropical storm Franklin was promoted to hurricane Franklin
in the year 2023. These six lists of
human names are recycled every six years.
However, if a hurricane is particularly destructive, then its name is
permanently retired and is forever associated with the hurricane for that
particular year. A new human name beginning
with the same letter of the English alphabet must then replace that name for
future years. For example, the fourth
tropical storm in the year 2013 should have been named Dean, but hurricane Dean
was so destructive in the year 2007 that the name Dean was permanently retired
and replaced with the name Dorian. If
there happens to be more than twenty-one tropical storms in the Atlantic Ocean
in any given year, then the letters of the Greek alphabet are used after
reaching the end of the list of twenty-one names. For example, the twenty-second tropical storm
in the Atlantic Ocean in the year 2005 was named tropical storm Alpha, the
twenty-third was named tropical storm Beta (later promoted to hurricane Beta),
the twenty-fourth was named tropical storm Gamma, the twenty-fifth was named
tropical storm Delta, the twenty-sixth was named tropical storm Epsilon (later
promoted to hurricane Epsilon), and so on and so forth. There are other lists of names for tropical
storms in the Pacific Ocean. Again,
there are twenty-one human names in each list of names for the Atlantic Ocean,
and there are twenty-four letters in the Greek alphabet. Twenty-one plus twenty-four equals
forty-five. What do we do if there are
more than forty-five tropical storms in the Atlantic Ocean in a single
year? In this case, we run to the
nearest church, since it is probably the end of the world!
The Saffir-Simpson
scale is a hurricane/typhoon scale, named for American engineer Herbert Saffir and American meteorologist Robert Simpson who
together formulated this scale. The
weakest hurricane/typhoon is called Category 1, stronger is called Category 2,
even stronger is called Category 3, stronger is called Category 4, and the
strongest hurricane/typhoon is called Category 5. We must keep in mind that even a Category 1
hurricane/typhoon is stronger and more destructive than a tropical storm. For example, hurricane Sandy was a Category 1
hurricane when it attacked and devastated New Jersey in the year 2012. Even tropical storms, which are themselves
weaker than Category 1 hurricanes/typhoons, can destroy entire towns; many
people have been killed by tropical storms, themselves weaker than the weakest
hurricanes/typhoons. We must always seek
shelter during a tropical storm warning, and we must certainly always seek
shelter during a hurricane/typhoon warning, regardless of the Saffir-Simpson-scale designation of the hurricane/typhoon.
Climatology
The study of short-term
trends and variations in the atmosphere is called meteorology, and someone who
studies short-term trends and variations in the atmosphere is called a
meteorologist. By short-term, we may
mean a few minutes, a few hours, a few days, or a few weeks. The study of long-term trends and variations
in the atmosphere is called climatology, and someone who studies long-term
trends and variations in the atmosphere is called a climatologist. By long-term, we may mean months, years,
decades, centuries (hundreds of years), millennia (thousands of years), or even
millions of years. The study of the
atmosphere (short-term and/or long-term) is called atmospheric sciences, and
someone who studies the atmosphere (short-term and/or long-term) is called an
atmospheric scientist.
Statistics is used to study
trends and variations of any kind. Any
collection of numbers is called data, and the purpose of statistics is to
calculate two quantities about data: the central tendency of the data and the
dispersion of the data. The central
tendency of the data is a number that is a typical representative of most of
the data. The most common way of
measuring central tendency is the average, which we will call the mean. The mean of the data is the sum of the
numbers divided by the number of numbers.
The most common way of measuring dispersion is the standard deviation,
but in this course we will measure dispersion with the range. The range of the data is the difference
between the largest number and the smallest number. In atmospheric sciences, the daily
temperature mean is the average of the hottest temperature and the coldest
temperature in any given day. For
example, if the hottest temperature today is eighty degrees fahrenheit
and if the coldest temperature today is seventy degrees fahrenheit,
then the daily temperature mean for today is seventy-five degrees fahrenheit, since eighty plus seventy is one hundred and
fifty, and dividing this by two yields seventy-five. The daily temperature range is the difference
between the hottest temperature and the coldest temperature in any given
day. In the previous example, the daily
temperature range for today would be ten fahrenheit
degrees, since eighty minus seventy equals ten.
The monthly temperature mean is the average of all the daily means for
that month. For example, if a particular
month happens to have thirty days, the monthly temperature mean for that month
would be the sum of all the daily means for that month divided by thirty. The monthly temperature range is the
difference between the hottest daily mean and the coldest daily mean during that
month. The annual temperature mean is
the average of all the monthly means for that year. In other words, the annual temperature mean
is the sum of all the monthly means for that year divided by twelve, since
there are twelve months in one year. The
annual temperature range is the difference between the hottest monthly mean
(almost always July or August in the northern hemisphere) and the coldest
monthly mean (almost always January or February in the northern hemisphere) of
that year.
Generally, temperature means
are hotter at the equatorial latitudes, while temperature means are colder at
the polar latitudes. At the midlatitudes, temperature means are hotter during
summertime and colder during wintertime.
However, we must not only specify trends and variations in the
temperature, but trends and variations in the precipitation must also be
specified in any climatological analysis.
Generally, precipitation means are high at and near the equator due to
the equatorial low (the doldrums).
Precipitation means are low at and near roughly thirty degrees latitude
in both hemispheres due to the subtropical highs (the horse latitudes). Precipitation means are high at and near
roughly sixty degrees latitude in both hemispheres due to the subpolar
lows. Finally, precipitation means are
low at and near the poles due to the polar highs.
Temperature ranges are
smaller at coasts and shores due to the marine effect: the oceans stabilize
temperatures due to the relatively large heat capacity of water. Temperature ranges are larger inland due to
the continental effect: continents do not stabilize temperatures due to the
relatively small heat capacity of land.
In other words, inland winters tend to be colder and inland summers tend
to be hotter as compared with coasts and shores where both winters and summers
tend to be relatively mild. Extending
this logic across the entire planet, temperature ranges are generally smaller
in the water hemisphere (the southern hemisphere), since the abundance of
southern-hemisphere oceans stabilize temperatures in that hemisphere. Conversely, temperature ranges are generally
larger in the land hemisphere (the northern hemisphere), since the abundance of
northern-hemisphere continents do not stabilize temperatures in that hemisphere. In other words, winters in the northern
hemisphere tend to be colder and summers in the northern hemisphere tend to be
hotter as compared with the southern hemisphere, where both winters and summers
tend to be relatively mild.
For most of the history of
planet Earth, the hot temperatures at the equatorial latitudes and the cold
temperatures at the polar latitudes have been moderated by the oceans due to
the relatively large heat capacity of water.
However, the moving tectonic plates of the lithosphere slowly change the
configuration of the continents and the oceans over enormous timescales
(millions of years). If there happens to
be relatively isolated continents and/or microcontinents at the poles, the
relatively small heat capacity of these landmasses will permit the temperatures
at the poles to become extremely cold.
The result is an ice age, an extremely long period of time (millions of
years) when the temperature at the poles is so cold that enormous icecaps cover
these landmasses. There have been
several ice ages throughout the entire history of the Earth, each lasting many
millions of years. As we discussed
earlier in the course, the Current Ice Age began roughly thirty million years
ago when South America ripped off of Antarctica, completely isolating
Antarctica at the South Pole and establishing the Antarctic Circumpolar Current
surrounding Antarctica. Hence,
Antarctica became extremely cold, and our entire planet Earth plunged into the
Current Ice Age. The Current Ice Age
began roughly thirty million years ago and continues to the present day. The Current Ice Age will last many more
millions of years as long as Antarctica remains isolated at the South Pole
surrounded by the Antarctic Circumpolar Current that further isolates
Antarctica. There are only two scenarios
that can end the Current Ice Age. In one
scenario, Antarctica may move off of the South Pole, which would make it less
cold. This would also interrupt the
Antarctic Circumpolar Current, which would contribute to warming temperatures. In the other scenario, another continent may
move to the South Pole and collide with Antarctica. This would end the isolation of Antarctica,
making it less cold. This would also
interrupt the Antarctic Circumpolar Current, again contributing to warmer
temperatures. Whether Antarctica moves
away from the South Pole or another continent moves toward the South Pole, it
takes millions of years for tectonic plates to move significantly, as we
discussed earlier in the course.
Therefore, the Current Ice Age will continue for millions of more years
to come. During the Current Ice Age, the
southern icecap covers the continent Antarctica, and the northern icecap covers
the microcontinent Greenland.
Within the Current Ice Age,
there have been many periods of time when the Earth has become even
colder. These are glacial periods of the
Current Ice Age. Between two glacial
periods is an interglacial period when the Earth is not as cold. The Earth becomes so cold during a glacial
period that the icecaps expand beyond the poles and advance onto other
continents. A major glacial period lasts
roughly one hundred thousand years, while a minor glacial period lasts between
roughly twenty-five thousand years and roughly fifty thousand years. We are currently within an interglacial
period of the Current Ice Age. This
interglacial period began roughly twelve thousand years ago at the end of a
major glacial period that lasted roughly one hundred thousand years. Plate tectonics cannot be responsible for the
alternation between glacial periods and interglacial periods within the Current
Ice Age, since tectonic plates do not move appreciably over timescales of
thousands of years. The alternation
between glacial periods and interglacial periods within the Current Ice Age is
caused by the Milanković cycles, named for the
Serbian climatologist and astronomer Mulutin Milanković who formulated this theory. The Earth’s orbit around the Sun is presently
almost a perfect circle. In other words,
the eccentricity of the Earth’s orbit around the Sun is close to zero, but this
has not always been the case nor will it always be the case. Gravitational perturbations (tugs) from the
other planets, primarily Jupiter, change the eccentricity of the Earth’s
orbit. When the Earth’s orbit is
perturbed to become more elliptical, the Earth will be significantly further
from the Sun, causing the Earth to become significantly colder thus causing a
major glacial period of the Current Ice Age.
Calculations show that the eccentricity of the Earth’s orbit changes
once every one hundred thousand years (roughly), which is roughly the duration
of time of a major glacial period within the Current Ice Age. Minor glacial periods are caused by the
precession and the nutation of the Earth’s rotational axis. Precession is the turning of an axis around
another axis. Nutation is the nodding of
an axis resulting in a change in obliquity.
The Earth’s rotational axis precesses and nutates, and this changes the amount of sunlight the Earth
receives from the Sun, which in turn causes minor glacial periods within the
Current Ice Age. Calculations show that
the Earth’s rotational axis precesses once every
twenty-six thousand years (roughly), and the Earth’s rotational axis nutates once every forty-one thousand years (roughly). These are roughly equal to the duration of
time of minor glacial periods within the Current Ice Age.
Long-term variations in
global temperature (over millions of years) are caused by slowly moving
tectonic plates, resulting in ice ages when continents and/or microcontinents
happen to be relatively isolated at or near the poles, as with the Current Ice
Age. Intermediate-term variations in
global temperature (over thousands of years) cause glacial periods and
interglacial periods within the Current Ice Age. These intermediate-term variations in global
temperature (over thousands of years) are caused by the Milanković
cycles (the variations of the eccentricity of the Earth’s orbit around the Sun,
the precession of the Earth’s rotational axis, and the nutation of the Earth’s
rotational axis). However, there are
also short-term variations in global temperature, from centuries to decades,
and even as short as a few years.
Extremely short-term variations in global temperature (over a few years)
are caused by violent igneous eruptions.
As we discussed earlier in the course, a single igneous eruption can
cause global cooling for a few years, since igneous eruptions eject ash into
the atmosphere, reducing the amount of incoming sunlight to the Earth. Short-term variations in global temperature
(from decades to centuries) are caused by a combination of the Pacific Decadal
Oscillation (PDO), the Atlantic Multidecadal
Oscillation (AMO), and variations in solar activity. As we discussed earlier in the course, the
Pacific Decadal Oscillation (PDO) is the alternation
between decades of El Niño domination in the Pacific Ocean and decades of La
Niña domination in the Pacific Ocean.
The Atlantic Multidecadal Oscillation (AMO) is a similar oscillation in the Atlantic Ocean, as we
also discussed earlier in the course.
Although direct observations of the PDO and
the AMO only stretch back several decades,
measurements of the radioactive isotope carbon-fourteen within trees have revealed that the PDO and the AMO each undergo
roughly sixty-year cycles, where one complete PDO for
example consists of roughly three decades of El Niño domination followed by
roughly three decades of La Niña domination.
Quantitative observations of solar activity stretch back a few
centuries, giving us a stronger understanding of how variations in solar
activity affect global temperatures. At
times, the Sun is more active, radiating more energy that warms planet Earth;
at other times, the Sun is more quiet, radiating less energy that cools planet
Earth. These variations in solar
activity manifest themselves through sunspots, regions on the surface of the
Sun where magnetic field strengths are particularly strong. These sunspots have been directly observed
for roughly four hundred years, since the invention of the telescope. Astronomers have observed that the number of
sunspots goes through a roughly eleven-year cycle. In one complete cycle, the number of sunspots
increases then decreases over a time period of roughly eleven years. For example, the Sun experienced a period of
increasing activity during the last few years of the twentieth century and the
first few years of the twenty-first century (the current century). This period of increasing solar activity was
followed by a period of decreasing solar activity. During that particular period of decreasing
solar activity, some of the coldest monthly temperature means over the last one
hundred years occurred. Furthermore,
measurements of the radioactive isotope carbon-fourteen within trees have revealed that this roughly
eleven-year solar cycle itself goes through a roughly two-hundred-year
cycle. This is the de Vries cycle, named
for the Dutch physicist Hessel de Vries, one of the pioneers of radiocarbon
dating. According to the de Vries cycle, the Sun gradually increases in activity to
what is called a solar maximum then gradually decreases in activity to what is
called a solar minimum. Caution: the
eleven-year solar cycles continue to occur throughout each two-century de Vries cycle. Since
one complete de Vries cycle lasts for roughly two
centuries, each solar maximum and each solar minimum lasts for roughly one
hundred years. Over the past twelve
thousand years (since the beginning of the current interglacial period of the
Current Ice Age), there have been roughly sixty complete de Vries
cycles, with each de Vries cycle having one solar
maximum and one solar minimum. The
Modern Maximum occurred throughout most of the twentieth century, and the
Modern Minimum began toward the beginning of the twenty-first century (the
current century). These de Vries cycles have caused variations in global temperatures
over the past several thousand years. In
particular, a solar maximum contributed to warm temperatures lasting from the
ancient Late Roman Republic Period to the ancient Early Roman Empire Period, a
solar minimum contributed to cold temperatures during the Early Middle Ages, a
solar maximum contributed to warm temperatures during the High Middle Ages, and
a solar minimum contributed to cold temperatures during the Little Ice Age,
lasting from the Late Middle Ages to the Early Modern Ages. Most recently, the Modern Maximum that
occurred throughout most of the twentieth century contributed to warming
temperatures during that century, and the Modern Minimum that began toward the
beginning of the twenty-first century (the current century) has already caused
cooling temperatures that will continue for the rest of the current century. By combining the roughly sixty-year PDO, the roughly sixty-year AMO,
the roughly eleven-year solar cycle, and the roughly two-century de Vries cycle, we obtain a nearly perfect model of global
temperature variations over the past couple thousand years. As we just discussed, this model predicts
that there will be gradual global cooling during the current century. Caution: an unexpected violent igneous
eruption may cause additional global cooling in addition to these cyclic variations.
To summarize the climate of planet
Earth, the average temperature of the atmosphere is determined primarily by the
Earth’s distance from the Sun together with the concentration of greenhouse
gases in the atmosphere, primarily water vapor, which warms the Earth
sufficiently so that it is habitable for life.
Temperature means are hot at the equatorial latitudes, temperature means
are cold at the polar latitudes, and temperature means vary at the midlatitudes based on the seasons (warmer in the summertime
and cooler in the wintertime). The
abundance of liquid water that covers the Earth (the oceans) stabilizes global
temperatures due to the large heat capacity of water. Temperature ranges are smaller in the
southern hemisphere (the water hemisphere), while temperature ranges are larger
in the northern hemisphere (the land hemisphere). Temperature ranges are smaller at coasts and
shores due to the marine effect, while temperature ranges are larger inland due
to the continental effect. Precipitation
means are high at and near the equator due to the equatorial low (the
doldrums), precipitation means are low at and near roughly thirty degrees
latitude in both hemispheres due to the subtropical highs (the horse
latitudes), precipitation means are high at and near roughly sixty degrees latitude
in both hemispheres due to the subpolar lows, and precipitation means are low
at and near the poles due to the polar highs.
Long-term variations in global temperature (over millions of years) are
caused by slowly moving tectonic plates, resulting in ice ages when continents
and/or microcontinents happen to be relatively isolated at or near the poles,
as with the Current Ice Age that began roughly thirty million years ago and
continues to the present day. Intermediate-term
variations in global temperature (over thousands of years) cause glacial
periods and interglacial periods within the Current Ice Age. These intermediate-term variations in global
temperature (over thousands of years) are caused by the Milanković
cycles (the variations of the eccentricity of the Earth’s orbit around the Sun,
the precession of the Earth’s rotational axis, and the nutation of the Earth’s
rotational axis). We are currently in an
interglacial period of the Current Ice Age, and this interglacial period began
roughly twelve thousand years ago.
Short-term variations in global temperature (over a few decades or a few
centuries) are caused by the roughly eleven-year solar cycle, the roughly
two-century de Vries cycle, the roughly sixty-year PDO, and the roughly sixty-year AMO.
We just began a century of gradual
global cooling resulting from these short-term solar cycles and oceanic
cycles. Finally, extremely short-term
variations in global temperature (over a few years) may result from powerful
igneous eruptions.
copyeditor: Michael Brzostek (Spring2023)
Libarid A. Maljian homepage at the Department of Physics at CSLA at NJIT
Libarid A. Maljian profile at the Department of Physics at CSLA at NJIT
Department of Physics at CSLA at NJIT
College of Science and Liberal Arts at NJIT
New Jersey Institute of Technology
This webpage was most recently modified on Monday, the twenty-fifth day of November, anno Domini MMXXIV, at 03:45 ante meridiem EST.