Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 24

The Sun Starts it All

There is one basic reason we have weather, and that is the sun. Weather systems start because the
sun’s energy heats up some parts of Earth more than others. Most of the time the sun shines most
directly on the middle of Earth, with less heating at the north and south poles. Earth is tilted on
its axis at exactly the right angle to have seasons, with different parts of Earth being heated more
or less during different times of the year. Land heats up faster than water, wetting up temperature
differences between oceans and continents. This unequal heating creates variations in
temperature and air pressure, winds, and ocean currents.

What is Meteorology?
The branch of science concerned with the processes and phenomena of the atmosphere,
especially as a means of forecasting the weather. Meteorology is the interdisciplinary scientific
study of the atmosphere. 

The word 'meteor' is a variation on the Greek "meteoron", which is a term dealing with any
objects that originate in the sky. 
Home  Science  Encyclopedias almanacs transcripts and maps  Atmosphere, composition and
structure

http://www.encyclopedia.com/science/encyclopedias-
almanacs-transcripts-and-maps/atmosphere-composition-
and-structure

Atmosphere, composition and structure


The Gale Encyclopedia of Science 
COPYRIGHT 2008 The Gale Group, Inc.

Atmosphere, composition and structure

Composition of the atmosphere

Atmospheric structure

Earth’s atmosphere is composed of about 78% nitrogen, 21% oxygen, and 0.93% argon. The
remainder, less than 0.1%, contains such trace gases as water vapor, carbon dioxide, and ozone.
All of these trace gases have important effects on Earth’s climate. The atmosphere can be
divided into vertical layers determined by the way temperature changes with altitude. The layer
closest to the surface is the troposphere, which contains over 80% of the atmospheric mass and
nearly all the water vapor. The next layer, the stratosphere, contains most of the atmosphere’s
ozone, which absorbs high-energy radiation from the sun and makes life on the surface possible.
Above the stratosphere are the mesosphere and thermosphere. These two layers include regions
of charged atoms and molecules, or ions. The upper mesosphere and lower thermosphere are
called the ionosphere, this region is important to radio communications, because radio waves can
bounce off the layer and travel great distances. It is thought that the present atmosphere
developed from gases ejected by volcanoes. Oxygen, upon which all animal life depends,

1
probably accumulated as excess emissions from plants that produce it as a waste product during
photosynthesis. Human activities may be affecting the levels of some important atmospheric
components, particularly carbon dioxide and ozone.

Composition of the atmosphere

Major gases

The most common atmospheric gas, nitrogen (chemical symbol N2) is largely inert, meaning that
it does not readily react with other substances to form new chemical compounds. The next most
common gas, oxygen (O2), is required for the respiration (breathing) of all animal life on Earth,
from humans to bacteria. In contrast to nitrogen, oxygen is extremely reactive. It participates in
oxidation, examples of which include apples turning from white to brown after being sliced, the
rusting of iron, and the very rapid oxidation reaction known as fire. Just under 1% of the
atmosphere is made up of argon (Ar), which is an inert noble gas, meaning that it does not take
part in any chemical reactions under normal circumstances. Together, these three gases account
for 99.96% of the atmosphere. The remaining 0.04% contains a wide variety of trace gases,
several of which are crucial to life on Earth.

Important trace gases

Carbon dioxide (CO2) affects Earth’s climate and plays a large support role in the biosphere, the
collection of living things that populate Earth’s surface. Only about 0.0325% of the atmosphere
is CO2. Carbon dioxide is required by plant life for photosynthesis, the process of using sunlight
to store energy as simple sugars, upon which all life on Earth depends. Carbon dioxide is also
one of a class of compounds called greenhouse gases. These gases are made up of molecules that
absorb and emit infrared radiation, which is felt as heat. The solar energy radiated from the sun is
mostly in the visible range, within a narrow band of wavelengths. This radiation is absorbed by
Earth’s surface, then re-radiated back out to space not as visible light, but as longer wavelength
infrared radiation. Greenhouse gas molecules absorb some of this radiation before it escapes to
space, and re-emit some of it back toward the surface. In this way, these gases trap some of the
escaping heat and increase the overall temperature of the atmosphere. If the atmosphere had no
greenhouse gases, it is estimated that Earth’s surface would be 90°F (32°C) cooler.

Water vapor (H2 O) is found in the atmosphere in small and highly variable amounts. While it is
nearly absent in most of the atmosphere, its concentration can range up to 4% in very warm,
humid areas close to the surface. Despite its relative scarcity, atmospheric water probably has
more of an impact on Earth than any of the major gases, aside from oxygen. Water vapor is an
element of the hydrologic cycle, the process that moves water between the oceans, the land
surface waters, the atmosphere, and the polar ice caps. Water cycling drives erosion and rock
weathering, determines Earth’s weather, and sets up climate conditions that make land areas dry
or wet, habitable or inhospitable. When cooled sufficiently, water vapor forms clouds by
condensing to liquid water droplets, or, at lower temperatures, solid ice crystals. Besides creating
rain or snow, clouds affect Earth’s climate by reflecting some of the energy coming from the sun,
making the planet somewhat cooler. Water vapor is also an important greenhouse gas. It is

2
concentrated near the surface and is much more prevalent near the tropics than in the polar
regions.

Ozone (O3) is found almost exclusively in a layer about 9–36 mi (15–60 km) in altitude. At
lower altitudes, ozone gas is irritating to eyes and skin and chemically attacks rubber and plant
tissue. Nevertheless, it is vital to life on Earth because it absorbs most of the high-energy
radiation from the sun that is harmful to plants and animals. A portion of the energy radiated by
the sun lies in the ultraviolet (UV) region. This shorter wavelength radiation is responsible for
suntans and is sufficiently powerful to harm cells, cause skin cancer, and burn skin. The ozone
molecules, along with molecules of O2, absorb nearly all the high-energy UV rays, protecting
Earth’s surface from the most damaging radiation. The first step in this process occurs high in
the atmosphere, where O2 molecules absorb very high energy UV radiation. Upon doing so, each
absorbing molecule breaks up into two oxygen atoms. The oxygen atoms eventually collide with
another O2 molecule, forming a molecule of ozone, O3 (a third molecule is required in the
collision to carry away excess energy). Ozone in turn may absorb UV of slightly longer
wavelength, which removes one of its oxygen atoms and leaves O2. The free oxygen atom, being
very reactive, will almost immediately recombine with another O2, forming more ozone. The last
two steps of this cycle repeat themselves but do not create any new chemical compounds; they
only act to absorb ultraviolet radiation. The amount of ozone in the stratosphere is small. If it
were all transported to the surface, the ozone gas would form a layer about 0.1–0.16 in (2.5–4.0
mm) thick. This layer, as thin as it is, is sufficient to shield Earth’s occupants from harmful solar
radiation.

Aerosols

In addition to gases, the atmosphere has a wide variety of suspended particles known collectively
as aerosols. These particles may be liquid or solid and are small enough that they may require
very long times to settle out of the atmosphere by gravity. Examples of aerosols include
suspended soil or desert sand particles, smoke particles from wildfires, salt particles from
evaporated ocean water, plant pollen, volcanic dust, and particles formed from the pollution
created by coal burning power plants. Aerosols significantly affect atmospheric heat balance,
cloud growth, and optical properties.

The particles in aerosols cover a wide range of sizes. Raindrops suspended in a cloud are about
0.04–0.24 in (1–6 mm) in diameter. Fine desert sand and cloud droplets range in diameter down
to about 0.0004 in (0.01 mm). Sea salt particles and smoke particles are 1/100th of this, about
0.0001 mm, or 0.1 micrometer, in diameter (1 micrometer = one thousandth of a millimeter).
Smallest of all are the particles that form when certain gases condense—that is, when several gas
molecules come together to form a stable cluster. These are the Aitkin nuclei, whose diameters
can be measured down to a few nanometers (1 nanometer = one millionth of a millimeter).

The size of some aerosol particles allows them to efficiently scatter sunlight and create
atmospheric haze. Under some conditions, aerosols act as collecting points for water vapor
molecules, encouraging the growth of cloud droplets and speeding the formation of clouds. They
may also play a role in Earth’s climate. Aerosols are known to reflect a portion of incoming solar
radiation back to space, which lowers the temperature of Earth’s surface. Current research is

3
focused on estimating how much cooling is provided by aerosols, as well as how and when
aerosols form in the atmosphere.

Atmospheric structure

The atmosphere can be divided into layers based on the atmospheric pressure and temperature
profiles (the way these quantities change with height). Atmospheric temperature drops steadily
from its value at the surface, about 290K (63°F; 17°C), until it reaches a minimum of around
220K (–64°F;–53°C) at 6 mi (10 km) above the surface. This first layer is called the troposphere
and ranges in pressure from over 1,000 millibars at sea level to 100 millibars at the top of the
layer, the tropopause. Above the tropopause, the temperature rises with increasing altitude up to
about 27 mi (45 km). This region of increasing temperatures is the stratosphere, spanning a
pressure range from 100 millibars at its base to about 10 millibars at the stratopause, the top of
the layer. Above 30 mi (50 km), the temperature resumes its drop with altitude, reaching a very
cold minimum of 180K (–135°F;–93°C) at around 48 mi (80 km). This layer is the mesosphere,
which at its top (the mesopause) has an atmospheric pressure of only 0.01 millibars (that is, only
1/100,000th of the surface pressure). Above the mesosphere lies the thermosphere, extending
hundreds of miles upward toward the vacuum of space. It is not possible to place an exact top of
the atmosphere because air molecules become scarcer until the atmosphere blends with the
material found in space.

The troposphere

The troposphere contains over 80% of the mass of the atmosphere, along with nearly all of the
water vapor. This layer contains the air we breathe, the winds we observe, and the clouds that
bring our rain. All of what we know as weather occurs in the troposphere, the name of which
means “changing sphere.”All of the cold fronts, warm fronts, high and low pressure systems,
storm systems, and other features seen on a weather map occur in this lowest layer. Severe
thunderstorms may penetrate the tropopause.

Within the troposphere, temperature decreases with increasing height at an average rate of about
11.7°F per every 3,281 ft (6.5°C per every 1,000 meters). This quantity is known as the lapse
rate. When air begins to rise, it will expand and cool at a faster rate determined by the laws of
thermodynamics. This means that if a parcel of air begins to rise, it will soon find itself cooler
and denser than its surroundings, and will sink back downward. This is an example of a stable
atmosphere in which vertical air movement is prevented. Because air masses move within the
troposphere, a cold air mass may move into an area and have a higher lapse rate. That is, its
temperature falls off more quickly with height. Under these weather conditions, air that begins
rising and cooling will become warmer than its surroundings. It then is like a hot-air balloon: it is
less dense than the surrounding air and buoyant, so it will continue to rise and cool in a process
called convection. If this is sustained, the atmosphere is said to be unstable, and the rising parcel
of air will cool to the point where its water vapor condenses to form cloud droplets. The air
parcel is now a convective cloud. If the buoyancy is vigorous enough, a storm cloud will develop
as the cloud droplets grow to the size of raindrops and begin to fall out of the cloud as rain. Thus,
under certain conditions the temperature profile of the troposphere makes possible storm clouds
and precipitation.

4
During a strong thunderstorm, cumulonimbus clouds (the type that produce heavy rain, high
winds, and hail) may grow tall enough to reach or extend into the tropopause. Here they run into
strong stratospheric winds, which may shear off the top of the clouds and stop their growth. One
can see this effect in the anvil clouds associated with strong summer thunderstorms.

The stratosphere

The beginning of the stratosphere is defined as that point where the temperature reaches a
minimum and the lapse rate abruptly drops to zero. This temperature structure has one important
consequence: it inhibits rising air. Any air that begins to rise will become cooler and denser than
the surrounding air. The stratosphere is therefore very stable.

The stratosphere contains most of the ozone found in Earth’s atmosphere, and the presence of
ozone is the reason for the temperature profile found in the stratosphere. Ozone and oxygen gas
both absorb short wave solar radiation. In the series of reactions that follow, heat is released.
This heat warms the atmosphere in the layer at about 12–27 mi (20–45 km) and gives the
stratosphere its characteristic temperature increase with height.

The ozone layer has been the subject of concern. In 1985, scientists from the British Antarctic
Survey noticed that the amount of stratospheric ozone over the South Pole fell sharply during the
spring months, recovering somewhat as spring turned to summer. An examination of the
historical records revealed that the springtime ozone losses had begun around the late 1960s and
had grown much more severe by the late 1970s. By the mid-1980s virtually all the ozone was
disappearing from parts of the polar stratosphere during the late winter and early spring. These
ozone losses, dubbed the ozone hole, were the subject of intense research both in the field and in
the laboratory.

Although the stratosphere has very little water, clouds of ice crystals may form at times in the
lower stratosphere over the polar regions. Early Arctic explorers named these clouds nacreous or
mother-of-pearl clouds because of their iridescent appearance. More recently, very thin and
widespread clouds have been found to form in the polar stratosphere under extremely cold
conditions. These clouds, called polar stratospheric clouds, or PSCs, appear to be small crystals
of ice or frozen mixtures of ice and nitric acid. PSCs play a key role in the development of the
ozone hole.

The understanding that has emerged implicates chlorine as the chemical responsible for ozone
destruction in the ozone hole. Chlorine apparently gets into the stratosphere from
chlorofluorocarbons, or CFCs—industrial chemicals widely used as refrigerants, aerosol
propellants, and solvents. Laboratory experiments show that after destroying an ozone molecule,
chlorine is tied up in a form unable to react with any more ozone. However, it can chemically
react with other chlorine compounds on the surfaces of polar stratospheric cloud particles, which
frees the chlorine to attack more ozone. In other words, each chlorine molecule is recycled many
times so that it can destroy thousands of ozone molecules. The realization of chlorine’s role in
ozone depletion brought about an international agreement in 1987, the Montreal Protocol, which
committed the participating industrialized countries to begin phasing out CFCs.

5
The mesosphere and thermosphere

The upper mesosphere and the lower thermosphere contain charged atoms and molecules (ions)
in a region known as the ionosphere. The atmospheric constituents at this level include nitrogen
gas, atomic oxygen, nitrogen (O and N), and nitric oxide (NO). All of these are exposed to strong
solar emission of ultraviolet and x-ray radiation, which can result in ionization, knocking off an
electron to form an atom or molecule with a positive charge. The ionosphere is a region enriched
in free electrons and positive ions. This charged particle region affects the propagation of radio
waves, reflecting them as a mirror reflects light. The ionosphere makes it possible to tune in
radio stations very far from the transmitter. Even if the radio waves coming directly from the
transmitter are blocked by mountains or the curvature of Earth, one can still receive the waves
bounced off the ionosphere. After the sun sets, the numbers of electrons and ions in the lower
layers drop drastically, because the sun’s radiation is no longer available to keep them ionized.
Even at night, however, the higher layers retain some ions. The result is that the ionosphere is
higher at night, which allows radio waves to bounce for longer distances. This is the reason that
one can frequently tune in to more distant radio stations at night than during the day.

The upper thermosphere is also where the bright nighttime displays of colors and flashes known
as the aurora occur. The aurora are caused by energetic particles emitted by the sun. These
particles become trapped by Earth’s magnetic field and collide with the relatively few gas atoms
present above about 60 mi (100 km), mostly atomic oxygen (O) and nitrogen gas (N2). These
collisions cause the atoms and molecules to emit light, resulting in spectacular displays.

The past and future of the atmosphere

If any atmosphere was present after Earth was formed about 4.5 billion years ago, it was
probably much different than that of today. Most likely it resembled those of the outer planets—
Jupiter, Saturn, Uranus, and Neptune—with an abundance of hydrogen, methane, and ammonia
gases. The present atmosphere did not form until after this primary atmosphere was lost. One
theory holds that the primary atmosphere was blasted from Earth by the Sun. If the Sun is like
other stars of its type, it may have gone through a phase where it violently ejected material
outward toward the planets. All of the inner planets,

KEY TERMS

Infrared radiation —Radiation similar to visible light but of slightly longer wavelength. We


sense infrared radiation as heat.

Ionosphere —Region of the atmosphere above about 48 mi (80 km), with elevated


concentrations of charged atoms and molecules (ions).

Lapse rate —The rate at which the atmosphere cools with increasing altitude.

Mesosphere —The third layer of the atmosphere, lying about 30–48 mi (50–80 km) in height
and characterized by small lapse rate.

Ozone hole —The sharp decease in stratospheric ozone over Antarctica that occurs every spring.

6
Stratosphere —A layer of the upper atmosphere above an altitude of 5–10.6 mi (8–17 km),
which extends to about 31 mi (50 km), depending on season and latitude. Within the
stratosphere, air temperature changes little with altitude, and there are few convective air
currents.

Thermosphere —The top layer of the atmosphere, starting at about 48 mi (80 km) and
stretching up hundreds of miles into space. Due to bombardment by very energetic solar
radiation, this layer can possess very high gas temperatures.

Troposphere —The layer of air up to 15 mi (24 km) above the surface of the Earth, also known
as the lower atmosphere.

Ultraviolet radiation —Radiation similar to visible light but of shorter wavelength, and thus
higher energy.

X-ray radiation —Light radiation with wavelengths shorter than the shortest ultraviolet; very
energetic and harmful to living organisms.

Including Earth, would have lost their gaseous envelopes. A secondary atmosphere began to
form when gases were released from the crust of the early Earth by volcanic activity. These
gases included water vapor, carbon dioxide, nitrogen, and sulfur or sulfur compounds. Oxygen
was absent from this early secondary atmosphere.

The large amount of water vapor released by the volcanoes formed clouds that continually rained
on the early Earth, forming the oceans. Because carbon dioxide dissolves easily in water, the new
oceans gradually absorbed most of it. Nitrogen, being unreactive, was left behind to become the
most common gas in the atmosphere. The carbon dioxide that remained began to be used by
early plant life in the process of photosynthesis. Geologic evidence indicates this may have
begun about two to three billion years ago, probably in an ocean or aquatic environment. Around
this time, there appeared aerobic (oxygen using) bacteria and other early animal life, which
consumed the products of photosynthesis and emitted CO2. This completed the cycle for CO2 and
O2; the two gases stayed in balance as long as all plant material was consumed by an oxygen-
breathing organism. However, some plant material was inevitably lost or buried before it could
be decomposed. This effectively removed carbon dioxide from the atmosphere and left a net
increase in oxygen. Over the course of billions of years, a considerable excess built up this way,
so that oxygen now makes up over 20% of the atmosphere (and carbon dioxide makes up less
than 0.033%). All animal life thus depends on the oxygen accumulated gradually by the
biosphere over the past two billion years.

Future changes to the atmosphere are difficult to predict. There is currently growing concern that
human activity may be altering the atmosphere to the point that it may affect Earth’s climate.
This is particularly the case with carbon dioxide. When fossil fuels such as coal and oil are dug
up and burned, buried carbon dioxide is released back into the air. Because carbon dioxide is a
greenhouse gas, it acts to trap infrared (heat) energy radiated by Earth, warming up the
atmosphere.

See also Atmospheric circulation; Greenhouse effect.

7
http://www.encyclopedia.com/science/encyclopedias-almanacs-transcripts-and-
maps/atmosphere-composition-and-structure

https://www.e-education.psu.edu/meteo300/node/534
4.1 Atmospheric Composition
The major gases that comprise today's atmosphere are in the table below. The mixing ratio of a
gas X is defined as the fraction of total moles that are the moles of gas X. For instance, 78 moles
of every 100 total moles of air is nitrogen, so nitrogen's mixing ratio is 0.78. Note that in
atmospheric composition, the mixing ratio is the moles of the gas divided by the total moles of
air. In contrast, the water vapor mixing ratio is the mass of water vapor divided by the mass of
dry air.

Key features of the gases include their compressibility (i.e., ability to expand or shrink in
volume), their transparency in the visible, their momentum, and their heat capacity. Water vapor
has the additional important feature of existing in the vapor, liquid, and solid phases in the

8
atmosphere and on Earth’s surface. The most important properties of small particles include their
ability to dissolve in water in order to be Cloud Condensation Nuclei (CCN) or to maintain a
lattice structure similar to ice in order to be Ice Nuclei (IN), as well as their ability to absorb and
scatter sunlight. These properties depend completely on the particle size and composition. Most
atmospheric gases participate in the atmosphere's chemistry, which is initiated by sunlight, as
you will soon see.

 note on the units used when quantifying atmospheric composition


Three different units are typically used when specifying the amounts of gases. One is the mass
mixing ratio, which is the mass of a chemical species divided by the total mass of air. You have
already encountered this with the specific humidity of water vapor. A second is the volume
mixing ratio, which is just the number of molecules of a chemical species in a unit volume
divided by the total number of all molecules in a unit volume. For gases with relatively large
fractions like nitrogen, oxygen, and argon, we use percent to indicate this fraction. For minor
gases like carbon dioxide and ozone, we use parts per million (10-6) ppmv or parts per billion (10-
9
) ppbv by volume (meaning by number not mass). Lastly we need to use the concentration, or
number per unit volume, to calculate reaction rates and lifetimes.
To convert between volume mixing ratios and concentrations, use the following procedure. For a
species X, to convert from a mixing ratio, notated χX, to a concentration, notated [X], use the
Ideal Gas Law to find the number of total molecules in a cm3 and then multiply by χX, expressed
as a fraction. Suppose p = 960 hPa (or mb) and T = 296K, and X = 60 ppbv, then

https://www.e-education.psu.edu/meteo300/node/606
4.2 Changes in Atmospheric Composition
Since the rise of oxygen, 2 billion years ago, the nitrogen and oxygen fractions in the atmosphere
have been stable.
Water vapor is highly variable but, on average, appears to also have been fairly stable. Recent
data from satellites and sondes indicate that perceptible water (the total amount of water that is in
a column from the surface to space) has increased 1.3 ± 0.3% per decade over the oceans in the
past 25 years (Trenberth et al., Climate Dynamics, 2005).
Historical (up to 500 years before the present) changes in CO2 and CH4 show large, rapid
variations. Note that the historical range for CO2 is 200-300 parts per million (10-6), and for
methane is 350-400 parts per billion (ppbv). These changes in gas amounts have been driven by
changes in Earth's temperature, which come from changes in Earth's orbit, the axis of Earth's
rotation, and volcanoes. Until recently, changes in Earth's temperature caused changes in these
gas amounts, which then reinforced the warming. In the past century, changes in the gas amounts
have been driving the observed change in Earth's temperature.

9
The recent changes in carbon dioxide show a fairly constant increase over the past 50 years.
There is a smaller seasonal cycle imposed on this trend. This seasonal behavior occurs because
CO2 is taken up by plants in the northern hemisphere in summer, since most of the plants are in
the northern hemisphere. Note that the current increase to above 400 ppm now extends well
above any other time in the past half million years. Much of this CO2 increase can be linked to
fossil fuel combustion. We will examine the scientific consequences of these CO2 levels in the
lesson on applications of radiation.

https://www.e-education.psu.edu/meteo300/node/607
4.3 Other Trace Gases
Hundreds of different trace gases have been measured in the atmosphere and perhaps thousands
more have yet to be measured. Many of these are volatile organic compounds
(VOCs). Volatile means that the compound may exist in the liquid or solid phase but that it
easily evaporates. Organic means that the compound contains carbon but is not carbon dioxide,
carbon monoxide, or carbides and carbonates found in rocks. There are also other chemicals like
the nitrogen oxides (e.g., nitric oxide (NO), nitrogen dioxide (NO2), nitric acid (HNO3)), sulfur
compounds (e.g., sulfur dioxide (SO2), sulfuric acid (H2SO4)) and halogen compounds (e.g.,
natural methyl chloride (CH3Cl), human-made chlorofluorocarbons (CCl2F2)). If we pay
attention, we can often smell and identify many of these chemicals, even at trace levels, although
some, like methane, carbon monoxide (CO), and chlorofluorocarbons are odorless. We enjoy
smelling the VOCs emitted by trees in a forest – aah, that fresh pine smell – but we hold our nose
to escape the smells of a stagnant swamp.
In addition to these thousands of chemicals that are emitted into the atmosphere every day, there
are also some very reactive compounds that are created by atmospheric chemistry and play the
important role of cleaning the atmosphere of many gases. The most important reactive gases are

10
ozone (O3) and hydroxyl (OH). We will focus the discussion of atmospheric chemistry on these
two.
The Atmosphere’s Oxidizing Capacity
Earth’s atmosphere is an oxidizing environment. This term means what you think it would: gases
that are emitted into the atmosphere react in a way that increases their oxygen content. Gases that
contain oxygen tend to be “stickier” on surfaces and more water soluble, which means that they
stick when they hit a surface or they can be readily taken up in clouds and rain drops and be
deposited on Earth’s surface. We call gases hitting the surface and sticking “dry deposition” and
gases being taken up in precipitation and rained out “wet deposition.”
Let’s consider a natural gas that is very important in our lives – methane (a.k.a., natural gas).
More and more methane is being extracted from below Earth’s surface and used to run our
electrical power plants, heat our homes, cook our food, and, increasingly, to run our
transportation vehicles. Methane is a simple molecule – CH4 – in which each of carbon’s four
bonds is made with a hydrogen atom. Energy comes from heating methane to high enough
temperatures that cause it to react, giving off energy as more stable molecules are formed. In
complete combustion, each methane molecule is converted into CO2 and two H2O. In the
process, four oxygen atoms or two oxygen molecules are consumed.
This same process occurs in the atmosphere, but at much lower temperatures and at a much
slower rate. In both cases, the first step in the methane oxidation sequence is the reaction with
the hydroxyl radical (OH). In water, hydroxyl loses an electron and is ionized (OH-), but in the
atmosphere, hydroxyl is not ionized. We call OH a free radical because it has an odd number of
electrons (eight for oxygen and one for hydrogen). Any gas with an odd number of electrons is
reactive because the electrons want to be paired up in molecules because that makes them more
stable.
Where does OH come from?
Before we tackle this question, let’s first look at where ozone (O3) comes from. We will start
with the stratosphere (a.k.a, good ozone because it blocks solar UV that harms humans, other
animals, agriculture, and ecosystems) and then eventually we will consider tropospheric ozone
(a.k.a., bad ozone, which is the ozone that hurts our health when we breathe it and that damages
plants and their fruit).

Lesson 5: Cloud Physics


 Overview
 5.1 Do you recognize these clouds, drops, and snowflakes?
 5.2 What are the requirements for forming a cloud drop?
 5.3 How can supersaturation be achieved?
 5.4 Curvature Effect: Kelvin Effect
 5.5 Solute Effect: Raoult’s Law
 5.6 Koehler Theory
 5.7 Vapor Deposition
 5.8 Did you know most precipitation comes from collision-coalescence?
 5.9 An Unusual Way to Make Precipitation in Mixed-Phase Clouds
 5.10 Looking at the Whole Cloud
 Summary and Final Tasks

11
Overview
Clouds and precipitation are integral to weather and can be difficult to forecast accurately.
Clouds come in different sizes and shapes that depend on atmospheric motions, their
composition, which can be liquid water, ice, or both, and the temperature. While clouds and
precipitation are being formed and dissipated over half the globe at any time, their behavior is
driven by processes that are occurring on the microscale, where water molecules and small
particles collide. We call these microscale processes “cloud microphysics” and microphysics is
the focus of this lesson. Three ingredients are required for the formation of clouds: moisture,
aerosol, and cooling. If any one of these is missing, a cloud will not form. Over eighty years ago,
a simple hypothesis was developed to explain the formation of clouds. This hypothesis has been
thoroughly tested and validated and is now called Koehler Theory. We will learn the elements of
Koehler Theory and how to use them to determine when clouds will form and when they will
not, becoming only haze. Clouds do not automatically precipitate. In fact, most clouds do not.
We will learn about the magic required for precipitation to form. Thus, cloud formation through
precipitation is a series of microsteps, each of which is necessary, but not sufficient, to achieve
precipitation.
5.1 Do you recognize these clouds, drops, and snowflakes?
Clouds have fascinated people for millennia, but it wasn’t until 1802 that Luke Howard first
classified clouds with the terms that are used today. His classification scheme was formalized
later in the 19th century and has 10 basic cloud types with many minor variations (see figure
below).

Cloud physics goes beyond the classification of clouds to determine the actual physical and
chemical mechanisms that create clouds and cause their evolution over time. There are two
aspects of cloud physics. One is the physics on the cloud scale, which is tens to hundreds of

12
meters in size. This physics is driven in part by behavior in the cloud’s environment, such as the
wind shear or the location of a front, and determines the evolution of the cloud and the cloud’s
size and shape. All of this action, however, is not possible without the physics that is occurring
on the microscale, which is less than a few centimeters in size.

A cloud is defined as a (visible) suspension of small particles in the atmosphere. For a water
cloud, there are a number of types of particles that we are interested in.

Note the wide range in size, volume, and number of particles in the figure above. The smallest,
the cloud condensation nuclei (CCN), can have rather little water vapor and are made up of
substances to which water can attach (called hydrophilic, water loving). The other particles grow
by adding water molecules but still contain the original CCN upon which they formed.
We can specify the amount of water that is in liquid form by using the liquid water content
(LWC). The liquid water content can be defined as:

Typical LWC are 0.1- 0.9 g m-3, but a few g m-3 are possible for wetter conditions.

A cloud drop is typically 10 µm in diameter, while a raindrop, which comes from a collection of
cloud drops, is typically 1 mm (1000 µm) in diameter.

Some clouds exist in regions where the temperature is below 0oC and thus can be made of ice.
Ice is simply water in an organized crystalline form. While on the molecular scale the
arrangement of water molecules in the ice matrix is the same, the visible shape can vary
dramatically. In fact, certain shapes are favored in different temperature regimes (see figure

13
below). Excess vapor density is simply the amount of water vapor (in terms of milligram per
cubic meter of air) in excess of ice saturation, not liquid water saturation, which is shown as the
dashed line. At temperatures above -3oC, simple hexagonal plates occur. Between -3oC and -8oC,
columns form. Between -8oC and -22oC, plates are formed, with the more complex plates
occurring with a greater excess of water vapor. At temperatures below -22oC, columns once
again dominate and go from simpler to more complex as the excess water vapor increases.

5.2 What are the requirements for forming a cloud drop?


There are three requirements for forming a cloud drop:
1. Moisture
2. Aerosol
3. Cooling
If any one of these three is missing, a cloud cannot form. We have talked about moisture and
aerosol and now need to consider ways that the air can be cooled. The air needs to be cooled so
that the water vapor pressure initially equals and then exceeds the water saturation vapor
pressure.
Saturation occurs when e=es, w=ws, and condensation = evaporation. At saturation, RH/100 =
e/ee  ~ w/ws = 1, or in terms of percent, 100%. When we find the lifting condensation level (LCL)

14
on a skew-T, we are finding the pressure level at which T (as determined from the dry adiabat)
= Td (as determined from the constant water vapor mixing ratio), or when w = ws.
Let’s define two new variables that are useful in discussing the cloud drop formation.

 
 
s = 0 at saturation; s < 0 for a subsaturated environment; s > 0 for a supersaturated environment.
Note that s and S are both unitless.
This equation applies only for a flat surface of pure water. When we get into situations where we
have curvature or a solute, we need to think about the supersaturation relative to the equilibrium
value of e, eeq, which can be different from es. So, depending on the circumstances, eeq can be es,
ei, esc, esol, or some combination. We will see that a small supersaturation is actually needed to
form clouds.

https://www.e-education.psu.edu/meteo300/node/624
Radiative Cooling
All matter radiates infrared energy, as we will see in the next lesson. When an air mass radiates
this energy, it cools down, but the amount of water vapor does not change.
We can understand this process by using the water phase diagram (see figure below). Initially,
the air mass is at the position of the blue dot. As the air parcel cools and the temperature drops,
the air parcel temperature moves to the left on the diagram but the water vapor pressure does not
change. However, because the temperature drops, es drops. When es becomes slightly less than e,
a cloud forms.
Summary:
 e is constant as T decreases.
 Since es depends only on T, es also decreases until es < e.
 When es becomes slightly less than e, a cloud forms.

15
An example of radiative cooling in action is radiation fog, which occurs overnight when Earth's
surface and the air near it cool until a fog forms (see figure below).

16
Mixing
Assume two air parcels with different temperatures and water vapor partial pressures are at the
same total pressure. If these two parcels mix, then the temperature and the water vapor partial
pressure is going to be a weighted average of the T and e of the two parcels. The weighting is
determined by the fraction of moles that each parcel contributes to the mixed parcel.
Mathematically, for parcel 1 with e1, T1, and N1(number of moles) and parcel 2 with e2, T2,
and N2, the e and T of the mixed parcel are given by the equations:
e=N1N1+N2e1+N2N1+N2e2  T=N1N1+N2T1+N2N1+N2T2or 
approximatelye=M1M1+M2e1+M2M1+M2e2  T=M1M1+M2T1+M2M1+M2T2This
equation is not rendering properly due to an incompatible browser. See Technical Requirements
in the Orientation for a list of compatible browsers.
[5.4]
where M1 and M2 are the masses of the air parcels. On the phase diagram, these give straight
lines for different proportions of the mixed parcel being from parcel 1 (0% to 100%) and parcel 2
(100% to 0%), as in the figure below.
Note that both of these two air parcels are unsaturated. So how does a cloud form? Think about a
single warm, moist air parcel mixing into the environment of colder, drier air. As the warm
mixes into more and more of the drier air, it gets increasingly diluted but the mixed parcel
continues to grow. As the amount of environmental air in the mixture increases, the
average e and  T of the mixed air parcel decreases to be closer to the environmental values and
the mixed parcel’s e and T follow a mixing line. Starting in the upper right near the warmer
parcel, as the mixed parcel continues to grow, eventually the e and T will hit the Clausius-
Clapeyron curve. As it continues to push into the liquid portion of the phase diagram and become
supersaturated, a mixing cloud will form. The cloud will stay as long as the mixed e and T put
the parcel to the left of the Clausius-Clapeyron curve. However, once the mixed parcel comes to
the right of the curve, the cloud will evaporate.

17
Water phase diagram for mixing, with two air parcels at (T1,e1) and (T2,e2). When the two
air parcels mix, the temperature and vapor pressure of the mixed parcel lies along the
mixing line between the two parcels. If a small, warm, moist parcel is mixed into a colder,
drier environment, then as the warm moist parcel mixes with more environmental air, the
size of the mixed air parcel grows and the temperature and vapor pressure follow the
mixing line toward the environmental temperature and vapor pressure (Eq. 5.4).

Uplift
The uplift of air can lead to cloud formation, as we know from the skew-T. Uplift is generally the
same as adiabatic ascent. This adiabatic ascent can be driven by convection, by a less dense air
mass overriding a more dense one, or by air flowing up and over a mountain. The following
happens:
 The water vapor mixing ratio remains the same, but e drops as p drops, thus reducing the
possibility that RH = e/es will reach 100%.
 The temperature drops in accordance with Poisson’s relations so that es also drops.
The question is “Does e or es drop faster so that eventually e equals es?" It turns out that es drops
faster. As a result, in uplifted air, e and es converge at the lifting condensation level (LCL) and a
cloud forms just at that level (see figure below).

Water phase diagram for uplift, with an air parcel starting at the beginning of the arrow.
As the parcel ascends both e and T (and thus es) decrease, but es decreases faster than e so
that eventually e > es. Once the cloud forms as the line passes into the liquid part of the
phase diagram, the water line tries to achieve saturation.

18
The arrow on the figure above shows the changes in e and T (and thus es) as an air parcel rises.
Once es <= e, then s > 0 and the air parcel is supersaturated. This supersaturated situation is not
stable; the water vapor in excess of es forms liquid. As the uplift continues, more water vapor is
converted into liquid water and the vapor pressure remains close to es. All convective clouds, that
is clouds with vertical extent, form this way. An example of adiabatic uplift is a cumulus cloud,
as seen in the figure below.

A cumulus cloud over the ocean.


Credit: JanneG via pixabay.com
Why is supersaturation required for a cloud drop to form?
I thought that cloud drops formed when w = ws. Why is supersaturation required for a cloud drop
to form?

https://www.e-education.psu.edu/meteo300/node/676

5.4 Curvature Effect: Kelvin Effect


Let’s look at curvature effect first (see figure below). Consider the forces that are holding a water
drop together for a flat and a curved surface. The forces on the hydrogen bonding in the liquid
give a net inward attractive force to the molecules on the boundary between the liquid and the
vapor. The net inward force, divided by the distance along the surface, is called surface tension,
σ. Its units are N/m or J/m2.

19
Sketches of the curvature effect. Left is a flat surface of pure water; right is a curved
surface of pure water.
Credit: W. Brune
If the surface is curved, then the amount of bonding that can go on between any one water
molecule on the surface and its neighbors is reduced. As a result, there is a greater probability
that any one water molecule can escape from the liquid and enter the vapor phase. Thus, the
evaporation rate increases. The greater the curvature, the greater the chance that the surface
water molecules can escape. Thus, it takes less energy to remove a molecule from a curved
surface than it does from a flat surface.

Cloud drops start as very small spherical drops, but the vapor pressure required for them to form
is much greater than es until they get closer to 10-2 μm in size. The Kelvin effect is important
only for tiny drops; it is important because all drops start out as tiny drops and must go through
that stage. As drops gets bigger, their radius increases and esc approaches es.
So, is it possible to form a cloud drop out of pure water? This process is called homogeneous
nucleation.The only way for this to happen is for two molecules to stick together, then add
another, then another, etc. But the radius of the nucleating drop is so small that the vapor
pressure must be very large. It turns out that drops probably can nucleate at a reasonable rate
when the relative humidity is about 440%. Have you ever heard of such a high relative humidity?

5.5 Solute Effect: Raoult’s Law


On the other hand, the atmosphere is not very clean either. There are all kinds of dirt and other
particles in the atmosphere. Some of these are hydrophilic (i.e., they like water) and water
soluble (i.e., they dissolve in water). So let’s see what the effect of soluble CCN might be on the
water evaporation rate for a flat water surface. We’ll then put the curvature and the solute effects
together.
First, here are some important definitions:
Solvent: The chemical that another chemical is being dissolved into. For us, the solvent is H2O.
Solute: The chemical that is being dissolved in the solvent.

20
Sketch of a flat liquid surface with a solvent (water, red dots) and a solute (black dots).
Credit: W. Brune
The simplest view of this effect is that solute molecules are evenly distributed in the water
(solvent). Therefore, some solute molecules occupy surface sites that would otherwise be
occupied by water molecules. Thus, the solute prevents water molecules from evaporating from
those sites. Adding more solute means that more surface sites would be occupied by solute
molecules and water vapor would have even less opportunity to break hydrogen bonds and
escape the liquid. The real view is more complicated by the electrostatic interactions between
water and solute molecules that cause an attraction between water and solute molecules, but the
basic result is the same as the simple view.
Because the evaporation rate is lowered, that means that there will be net condensation until the
water vapor flux to the surface matches the water vapor flux leaving the surface. When
equilibrium is established between the lower evaporation and condensation, the condensation
will be less, which means that the water saturation vapor pressure will be lower. The equilibrium
vapor pressure is less than es, which, remember, is the saturation vapor pressure over a flat
surface of pure water. As the amount of solute is increased, the equilibrium vapor pressure of the
solution will be even less.                

https://www.e-education.psu.edu/meteo300/node/679
5.7 Vapor Deposition
The growth of the cloud drop depends initially on vapor deposition, where water vapor diffuses
to the cloud drop, sticks, and thus makes it grow. senv must be greater than sk for this to happen,
but as the drop continues to grow, sk approaches 0 (i.e., eeq approaches es), so smaller amounts of
supersaturation still allow the cloud drop to grow. Deriving the actual equation for growth is
complex, but the physical concepts are straightforward.
 The growth rate (dmd/dt) is proportional to senv – sk. Physically, this statement means
that the greater the difference between the supersaturation in the environment and
supersaturation at the particle’s surface, the faster water vapor will diffuse and stick on the
surface. For instance, if senvequaled sk, then the evaporation and condensation of water on the
particle’s surface would be equal and there would be no mass growth.
 As water vapor diffuses to the drop and forms water, energy is released (i.e., latent heat
of condensation) and this raises the temperature of the cloud drop surface so that Tsfc > Tenv.
But an outward energy flow occurs and is proportional to Tsfc -Tenv. Physically, this statement

21
means that the particle and the air molecules around it are warmed by latent heat release.
These warmer molecules lose some of this energy by colliding with the cooler molecules
further away from the particle, and warm them by increasing their kinetic energy (see figure
below).

Schematic of the two physical processes in the growth of a cloud drop by vapor deposition.
One is vapor deposition and the other is the transfer of condensational heating to the
atmosphere;

http://ffden-2.phys.uaf.edu/211_fall2010.web.dir/Levi_Cowan/development.html
Cloud Formation and Development

What exactly are clouds? A common answer from many lay people is that clouds made up of
concentrations of water vapor floating in the air. Although this answer is understandable from an
intuitive standpoint, it is wrong. Clouds are actually made up of liquid water, in the form of tiny
droplets that are so light that they are able to remain airborne without being overcome by gravity.
This makes sense because water vapor is always present in the air, but it is invisible. Clouds, on
the other hand, are very visible, as is the liquid water of which they are comprised. 

Clouds form from water vapor molecules in the atmosphere that condense into liquid droplets.
These droplets are extremely tiny, with a radius on the order of about 10 micrometers. Water is
constantly evaporating and condensing into and out of the air when the relative humidity (the
amount of moisture in the atmosphere relative to its capacity to hold water vapor) is less than
100%. However, to get a big enough concentration of water droplets in the air that eventually
forms a cloud that is visible to the naked eye requires a relative humidity of near 100%. This
generally occurs when the air is relatively cool, because the amount of water vapor that the
atmosphere can "hold" is proportional to the air temperature. The warmer the air is, the more
water the atmosphere can hold in vapor form, because there is more energy available. When the
relative humidity is 100%, the air is said to be "saturated" with as much moisture as it can hold,
and therefore water vapor can only condense back into liquid. This greatly increases the number
of water droplets floating in the air, and this is what creates clouds. However, just having the
humidity at 100% is not enough. Tiny airborne particles called "condensation nuclei" are
necessary for water vapor to condense onto. Although the air may look clean on an ordinary day,

22
as many as 150,000 particles can exist in a volume of air approximately the size of your index
finger. These particles are extremely small and light, with many having a mass less than one-
trillionth of a gram. Without them, relative humidities of several hundred percent would be
required for water vapor to condense.

Most of the time, the relative humidity is less than 100% at the earth's surface, and thus most
clouds float above ground-level. When the relative humidity of the air at the surface of the earth
is 100%, a cloud forms in contact with the ground that we call "fog."  Often the atmosphere is
fairly dry near the surface but becomes saturated at a certain height. This level is called the
"condensation level," and represents the height at which water vapor will be forced to condense ,
and thus indicates where the base of any clouds will be. 

Getting water vapor in the atmosphere to condense into clouds is often accomplished by lifting
air upwards. The two main processes that cause air to rise are "convection" and forced ascent,
usually in the form of weather fronts. Convection is a vertical circulation of air that occurs when
air gets heated from below by the ground or ocean water that has been heated by the sun. Parcels
of air that are warmer than the air around them will tend to expand and become less dense, and
therefore lighter, than the surrounding air and begin to rise. As they rise, the energy lost by the
air molecules through the effort of expanding their parcels causes the air to cool. Eventually, as
the air continues to rise (as long as it is warmer than the surrounding environment), it will cool to
the point at which the relative humidity becomes 100%. This is called the "dewpoint
temperature," and occurs at the condensation level. This is where the water vapor contained
within the air starts to condense into clouds. The phase change of water from vapor to liquid
during lifting also releases a huge amount of latent heat into the atmosphere, adding to the
convective instability and causing more air to rise, which in turn condenses more water and
releases more heat. This feedback loop is part of what helps clouds grow, especially convective
ones. The amount of energy released inside a cloud in this manner is massive. A typical
thunderstorm complex can release more energy during its lifespan than a small nuclear

23
explosion. The other way by which air rises is when it is forced upward, usually by weather
fronts. An example is a "warm front," when a warm air mass is advancing towards a colder one.
At the boundary between the two, warm, less dense air is forced to lift upwards and glide over
the top of the colder air mass. The forced lifting cools the air to its dewpoint and causes clouds to
form. Both of these situations in which air cools and becomes saturated when it rises are
examples of what is called "adiabatic cooling."

The different processes that form clouds also give them different appearances. Clouds formed
from convection have a great deal of vertical development due to vigorously rising air and have a
puffy, billowing look to them. These are called cumulus clouds, and are the most well-known
type of cloud. More gradual lifting caused by warm fronts, on the other hand, causes clouds that
are layered and flat in nature, with little vertical depth. This family of cloud types is called
"stratus," which literally means "layered." At great heights above the earth's surface where
temperatures are well below freezing, clouds can be composed entirely of ice crystals. These
clouds typically have a wispy, feathery look, and are called "cirrus." A common term used to
describe classic wisps of cirrus against a blue sky is "horse tails." Many different cloud types
exist as a combination of these three main classifications at the low, mid, and high levels of the
troposphere.

24

You might also like