Вы находитесь на странице: 1из 118

THE OZONE LAYER

The ozone layer is a deep layer in the stratosphere, encircling the Earth,
which has large amounts of ozone in it. The layer shields the entire Earth
from much of the harmful ultraviolet radiation that comes from the sun.
Interestingly, it is also this ultraviolet radiation that forms the ozone in the
first place. Ozone is a special form of oxygen, made up of three oxygen
atoms rather than the usual two oxygen atoms. It usually forms when some
type of radiation or electrical discharge separates the two atoms in an
oxygen molecule (O2), which can then individually recombine with other
oxygen molecules to form ozone (O3).
The ozone layer became more widely appreciated by the public when it was
realized that certain chemicals mankind manufactures, called
chloroflurocarbons, find their way up into the stratosphere where, through a
complex series of chemical reactions, they destroy some of the ozone. As a
result of this discovery, an international treaty was signed in 1973 called the
Montreal Protocol, and the manufacture of these chemicals was greatly
reduced.
The ozone layer has since begun to recover somewhat as a result of these
efforts, but there is some science which now suggests that the major
volcanic eruptions (mainly El Chichon in 1983 and and Mt. Pinatubo in 1991)
which have occurred since we started monitoring ozone with satellites in the
late 1970's, could have also contributed to the ozone depletion.
The amount of stratospheric ozone overhead on any given day and at any
given location varies quite a bit. Because of vertical circulations of air in both
the troposphere and the stratosphere, there can be greater or lesser
amounts of ozone protecting you from ultraviolet radiation. Also, living at
higher elevations exposes people to more UV radiation than at low
elevations.
While stratospheric ozone which protects us from the sun is good, there is
also ozone produced near the ground from sunlight interacting with
atmospheric pollution in cities that is bad for human health. It causes
breathing problems for some people, and usually occurs in the summertime

when the pollution over a city builds up during stagnant air conditions
associated with high pressure areas

HISTORY OF THE OZONE LAYER


As mid-May brings on the onset of winter, the Antarctic stratosphere cools
and descends closer to the surface. The Coriolis effect (caused by the earths
rotation) sets up a strong westerly circulation around the south pole, forming
an oblong vortex which varies in size from year to year.
s temperatures in the lower stratosphere cools below -80'C, Polar
Stratospheric Clouds (PSC's) start to form.
In the area over Antarctica, there are stratospheric cloud ice particles that
are not present at warmer latitudes. Reactions occur on the surface of the ice
particles that accelerate the ozone destruction caused by stratospheric
chlorine. Polar regions get a much larger variation in sunlight than anywhere
else, and during the 3 months of winter spend most of time in the dark
without solar radiation. Temperatures hover around or below -80'C for much
of the winter and the extremely low Antarctic temperatures cause cloud
formation in the relatively ''dry'' stratosphere. These Polar Stratospheric
Clouds (PSC's) are composed of ice crystals that provide the surface for a
multitude of reactions, many of which speed the degradation of ozone
molecules. This phenomenon has caused documented decreases in ozone
concentrations over Antarctica.
In fact, ozone levels drop so low in spring in the Southern Hemisphere that
scientists have observed what they call a "hole" in the ozone layer. The
ozone destruction process requires conditions cold enough for stratospheric
clouds to form. Once these stratospheric clouds form the process can take
place, even in warmer conditions.
Ground based measurements of Ozone were first started in 1956, in at Halley
Bay, Antarctica. Satellite measurements of ozone started in the early 70's,
but the first comprehensive worldwide measurements started in 1978 with
the Nimbus-7 satellite.
Chloroflourocarbons were first created in 1928 as non-toxic, non-flamable
refrigerants, and were first produced commercially in the 1930's by DuPont.
The first Chlorofluorocarbon was CFC-12, a single carbon with two chlorines
and two Fluorines attached to it.
In 1974 M.J.Molina and F.S.Rowland published a laboratory study
demonstrating the ability of CFC's to catalytically breakdown Ozone in the
presence of high frequency UV light. Further studies estimated that the

ozone layer would be depleted by CFC's by about 7% within 60yrs and based
on such studies the US banned CFC's in aerosol sprays in 1978. Slowly
various nations agreed to ban CFC's in aerosols but industry fought the
banning of valuable CFC's in other applications.

Dr. Shigeru Chubachi, Meteorological Research Institute, Japan, measures low


ozone and an ozone hole over Syowa, Antarctica (reported at Ozone
Commission meeting in Halkidiki, Greece in Sept 1984)
In 1984 British Antarctic Survey scientists, Joesph Farman , Brian Gardiner,
and Jonathan Shanklin, discovered a recurring springtime Antarctic ozone
hole .Their paper was published in Nature , May 1985, the study
summarized data that had been collected by the British Antarctic Survey
showing that ozone levels had dropped to 10% below normal January levels
for Antarctica.
Dr Richard D. McPeters, Principal Investigator, Earth Probe TOMS at
Goddard Space Flight Center explains "Our software had flags for ozone that
was lower than 180 DU, a value lower than had ever been reliably reported
prior to 1983. In 1984, before publication of the Farman paper, we noticed a
sudden increase in low value from October of 1983. We had decided that the
values were real and submitted a paper to the conference the following
summer when Joe's paper came out, showing the same thing. As the first one
in print, he gets full credit for discovery of the ozone hole. It makes a great
story to talk about how NASA "missed" the ozone hole, but it isn't quite true."
In the 1980's the first measurements of this loss were actually documented.
In 1984, when the British first reported their findings, October ozone levels
were about 35 percent lower than the average for the 1960s. When the first
measurements were taken the drop in ozone levels in the stratosphere was
so dramatic that at first the scientists thought their instruments were faulty.
The U.S. satellite Nimbus-7 quickly confirmed the results, and the term
Antarctic ozone hole entered popular language.
The ozone hole appeared first over the colder Antarctic because the ozonedestroying chemical process works best in cold conditions. The Antarctic
continent has colder conditions than the Arctic, which has no land-mass. As
the years have gone by the Ozone Hole has increased rapidly and is as large
as the Antarctica continent. The hole lasts for only two months, but its timing
could not be worse. Just as sunlight awakens activity in dormant plants and

animals, it also delivers a dose of harmful ultraviolet radiation. After eight


weeks, the hole leaves Antarctica, only to pass over more populated areas,
including The Falkland Islands, South Georgia and the tip of South America.
This biologically damaging, high-energy radiation can cause skin cancer,
injure eyes, harm the immune system, and upset the fragile balance of an
entire ecosystem. News about the ozone hole that forms over Antarctica
each October has spread around the world. The ozone hole can be as big as
1.5 times larger than the United States.

Below is a table of geographic areas to be used as references in


conceptualizing just how large the ozone hole can get.

Following are referable areas:


Australia

8,923,000 Sq Km

United
States

9,363,130 Sq Km

Europe

10,498,000 Sq Km

Antarctica

13,340,000 Sq Km

Russia

17,078,000 Sq Km

North
America

25,349,000 Sq Km

Africa

30,335,000 Sq Km

S. Pole to 70
15,300,000 Sq Km
S
Source: NOAA Climate Prediction Center

However, less-well-known is that ozone depletion has been measured


everywhere outside the tropics, and that it is, in fact, getting worse. in the

middle latitudes (most of the populated world), ozone levels have fallen
about 10% during the winter and 5% in the summer. Since 1979, they have
fallen about 5% per decade when averaged over the entire year. Depletion is
generally worse at higher latitudes, i.e. further from the Equator.
The severity of the ozone hole varies somewhat from year to year.
Ozone Hole Maximum Area Over Time
The Antarctic ozone hole reached its annual peak size on Sept. 11, 2014
according to scientists from NASA and the National Oceanic and Atmospheric
Administration (NOAA). The size of this years hole was 24.1 million square
kilometers (9.3 million square miles) an area roughly the size of North
America.
The ozone hole over Antarctica was slightly smaller in 2013 than the average
for recent decades, according to data from the Ozone Monitoring Instrument
(OMI) on NASAs Aura satellite and the Ozone Monitoring and Profiler Suite
(OMPS) on the NASA-NOAA Suomi NPP satellite. The average size of the hole
in SeptemberOctober 2013 was 21.0 million square kilometers (8.1 million
square miles). The average size since the mid 1990s is 22.5 million square
kilometers (8.7 million square miles). The single-day maximum area reached
24.0 million square kilometers (9.3 million square miles) on September 16
an area about the size of North America. credit NASA
The ozone hole reached its maximum size Sept. 22, 2012 covering 8.2
million square miles (21.2 million square kilometers), or the area of the
United States, Canada and Mexico combined. The average size of the 2012
ozone hole was 6.9 million square miles (17.9 million square kilometers).
The 2011 Ozone Hole reached it's largest size of 26 million square
kilometers on September 12,2011
2010 ozone hole smaller than usual
MACC - Monitoring Atmospheric Composition and Climate-The destruction of
ozone in the ozone layer over Antarctica this year is about 40 to 60% less
compared to the previous five years, according to MACC analyses based on
observations from the SCIAMACHY instrument on board the European
ENVISAT satellite. Less ozone destruction is consistent with the expectation
that the ozone layer will recover during the coming decades due to reduction
in the stratospheric amount of chlorine. However, such a large change
cannot be attributed to the slow decrease of stratospheric chlorine of 0.5-1%

per year. This years reduced ozone destruction turns out to be caused by
unusual meteorological conditions.
he Ozone Hole of 2010 was the second smallest in over a decade. It reached
it's largest size of 22.6 million square kilometers on September 25,2010
which is below the decade average of 25.7 million square kilometers.
Sudden Stratospheric Warming
In July and August a phenomenon known as a Sudden Stratospheric Warming
occurred in the stratosphere above Antarctica: a sudden fast warming in the
ozone layer. The stratosphere is the atmospheric layer between 15 and 35
km altitude which contains large amounts of ozone that protects us from
harmful ultraviolet radiation. The exact mechanisms of these of warmings
are still not fully understood, but they happen quite often and have their
origin in the troposphere, the lowest layer of the atmosphere where our
weather occurs.
Effect of nitrogen-oxides on chlorine release
The ozone hole appears because of the destructive power of chlorine which
is present in the ozone layer but under normal conditions chemically bonded
with nitrogen-oxides. However, via complex chemical processes the chlorine
can be released during the polar winter above Antarctica. At temperatures
below -78C the nitrogen-oxides are removed from the stratosphere. The
remaining chlorine can no longer be bonded and when the sun becomes
stronger during spring the chlorine can start to do its damaging work.
Although the temperature in the Antarctic stratosphere sinks below -78C
every July and August, it does not get much colder. A small temperature rise
of only a few degrees thus can result in considerably more nitrogen-oxides
remaining in the stratosphere, and thus less effective ozone destruction.
Recent measurements of the American MLS instrument on the AURA satellite
indeed show that there are still nitrogen-oxides present in the stratosphere
above Antarctica, unlike previous years.
Late onset of ozone desctruction in 2010
This years ozone destruction consequently only started to become visible
during September, whereas under normal conditions the first signs already
become visible about halfway through August. The observations by
SCIAMACHY show that the amount of ozone decreased steadily towards
October, but the destruction continued to lag behind that of previous five
years. At the beginning of October the amount of ozone destruction was still
40-60% less than in a typical year. In the second half of October the ozone

loss became more comparable to previous years. Ozone destruction usually


ceases in October after which the ozone hole starts to fill up.

The 2009 ozone hole is essentially over, with most of the continent
experiencing a stratospheric spring warming. The residual vortex is over the
Dronning Maud Land and here minimum ozone values are around 240 DU
and ozone depletion exceeds 30%. Ozone values outside the polar vortex
have dropped to near 400 DU, and inside the vortex ozone values are rapidly
increasing as the atmosphere warms, though the summer circulation is not
yet fully established. The temperature of the ozone layer over Antarctica is
now too warm for polar stratospheric clouds (PSCs) to form. During the early
winter, the polar vortex was often rather more elliptical than it was in 2008,
and this lead to some early depletion in circumpolar regions as stratospheric
clouds became exposed to sunlight. It reverted to a more circular circulation
as winter progressed and this led to another relatively slow start to the
growth of the ozone hole (as measured by NASA/SBUV2), with the "hole" not
beginning until mid August. The vortex became more elliptical again in late
August, with South Georgia being affected by the fringes of the ozone hole
between September 2 and 6. The hole grew to reach an area of around 24
million square kilometres by mid September, and then lasted until the end of
November. From mid August to mid November it remained around the
average area seen over the past decade. The tip of South America and
South Georgia were affected by the fringes of the ozone hole from
September 24 to September 30, from October 3 to October 7 and from
November 8 to 25. South Georgia was affected on November 6.
In 2009, the ozone hole reached its 10th largest measured size since careful
measurements began in 1979.The daily maximum ozone hole area for 2009
was 24 million km2 on 17 September.

The 2008 ozone hole season is past its peak and ozone levels over Antarctica
are slowly rising, with the ozone hole slowly shrinking. The polar vortex has
been very stable throughout the season, however the final spring warming is
expected towards the end of the month. During its initial stages, the ozone
hole was much smaller than has been usual for August, but it grew rapidly as
stratospheric clouds were exposed to sunlight. It covered over 25 million
square kilometres in mid September, about the same as last year and
remained at around 24 million square kilometres until early October. It is

now around 14 million square kilometres, which is the largest on record for
this time of year. The temperature of the ozone layer over Antarctica is
rising as the spring warming takes hold, and is now generally too warm for
polar stratospheric clouds (PSCs) to continue to exist. The temperature
within the polar vortex is generally a little below the normal. Ozone values
are above 350 DU in parts of the circum-polar regions, which is a bit lower
than at this time last year. Lowest values, near 160 DU, are offset slightly
from the pole towards the Weddell Sea. The vortex briefly showed a more
elliptical circulation pattern, and the ozone hole extended over the tip of
South America, the Falkland Islands and South Georgia between October
28th and 30th. PSCs were seen from Rothera on several occasions and also
at Halley.

On September 12, 2008, the Antarctic ozone hole reached its maximum size
for the year. Represented by blues and purples in this image from the Ozone
Monitoring Instrument (OMI) on NASA Aura satellite, the ozone hole covered
about 27 million square kilometers, making it larger than North America,
which is about 25 million square kilometers. Though larger than it was in
2007, the 2008 ozone hole was still smaller than the record set in 2006.

Situation at 2007 December 28 British Antarctic Survey Ozone Bulletin The


ozone hole of 2007 is over. The temperature of the ozone layer is now
highest over Antarctica and cools northwards. It is too warm for polar
stratospheric clouds (PSCs) to exist. Generally across Antarctica, ozone
values are around 300 DU. Values are now slowly dropping towards the
normal autumn normal minimum. In general the circum-polar stratospheric
vortex was more disturbed this year than it was last year, however there
were also periods of stability. Some areas of Antarctica saw ozone values
down to ozone hole levels (less than 220 DU) in mid June, suggesting the
possibility of early chemical depletion, combined with some dynamic
processes. In mid July the minimum temperature of the Antarctic
stratosphere at 70 & 50 hPa was close to the normal, although that at 30 hPa
was below the normal, and I suggested the likelihood of another strong
ozone hole year. Early August saw the largest ozone hole recorded for this
time of year, although at the same time very high ozone levels existed over
the northern Antarctic Peninsula. The vortex was more circular in mid
September but returned to an elliptical shape and initially warmed slowly. In

mid September the ozone hole area reached a maximum of just over 24
million square kilometres, but it then slowly shrank as the vortex warmed
further. A spring warming took place over the Pacific coast of Antarctica and
the Antarctic Peninsula in late October, however this subsided as the ozone
hole became more symmetric again. A second major warming took place
towards the end of November, but subsided in early December. Around
December 8 the ozone hole was briefly the largest on record for the date.
The tip of South America and the Falkland Islands saw ozone levels below
250 DU on August 24, with South Georgia experiencing similar levels on
September 4. The fringes of the ozone hole were over South Georgia on
September 11. The area was intermittently affected from September 22 to
October 22, with South Georgia particularly affected on September 25, when
values dropped below 175 DU. The Falklands and South Georgia were
affected for a final time between November 23 and 25.

Data from NASA's Earth-observing Aura satellite show that the ozone hole
peaked in size on Sept. 13,2007 reaching a maximum area extent of 9.7
million square miles(24.7Area (million sq. km)) just larger than the size of
North America. That's "pretty average," says Paul Newman, an atmospheric
scientist at NASA Goddard Space Fight Center, when compared to the area of
ozone holes measured over the last 15 years. Still, the extent this year was
"very big," he says, compared to 1970s when the hole did not yet exist.
2007 Climate Prediction Center summary-Extensive ozone depletion
was again observed over Antarctica during the Southern Hemisphere winter
of 2007, with widespread total ozone anomalies of 45 percent or more below
the 1979 to 1986 base period. The area covered by extremely low total
ozone values of less than 220 Dobson Units, defined as the Antarctic Ozone
Hole area, in September reached maximum size of greater than 24 million
square kilometers, the 7th largest over all 29 years of continuous satellite
monitoring of Antarctic ozone. However, after reaching its maximum size, the
Ozone Hole diminished in size such that it was just the 17th largest for the
month of October. Vertical profiles of ozone amounts, measured by balloonsondes over the South Pole, showed near-complete destruction of ozone in
the 13 to 21 km region for a relatively short period of time due to the Ozone
Holes displacement off of the pole towards the Atlantic quadrant. Minimum
total ozone values observed at the South Pole were higher than those seen
during recent years. Lower stratospheric temperatures over the Antarctic
region in the winter of 2007 were again well below -78C, and were

sufficiently low for polar stratospheric cloud formation, promoting chemical


ozone loss. The size of the area of very low temperatures in 2007 was,
however, average or below average for the past ten years. The polar vortex
persisted until mid-December when it and the remains of the Ozone Hole
slowly diminished as the polar circulation changed over into its summer
pattern.
3 October2006 World Meteorological Organization This years hole in the
Antarctic ozone layer was the most serious on record exceeding that of 2000.
Not only was it the largest in surface area (matching 2000) but also suffered
the most mass deficit, meaning that there was less ozone over the Antarctic
than ever previously measured.

European Space Agency

Measurements were taken from instruments on both NASA and European


Space Agency (ESA) satellites. These are validated by surface based
observations of the WMO Global Atmosphere Watch (GAW) ozone network.
Each agency uses different instruments hence the slightly different values.
NASA instruments showed that, on 25 September 2006, the area of the hole
reached 29.5 million km2, compared to 29.4 million km2 reached in
September 2000.

2006 Climate Prediction Center summary-Very low ozone values were


again observed over Antarctica in the winter of 2006. Ozone depletion of
more than 45 percent was observed over Antarctica, compared to total
ozone amounts observed in the early 1980's. Vertical soundings over the
South Pole during August, September and October 2006 again showed
strongest destruction of ozone at altitudes between 13 and 21 km. Lower
stratosphere temperatures in the winter of 2006 over the Antarctic region
were lower than in 2005 and near record levels. Associated with this, the
ozone hole was larger than in other years. The ozone hole in 2006
diminished in size and depth along with warming stratospheric conditions in

October, and by early December, total ozone over Antarctica had increased
to levels above ozone hole values.
The stability of the stratospheric polar vortex structure, which kept it
generally centered over the South Pole, and the very persistent, anomalously
cold temperatures in the presence of halogen levels that remain at high
levels (though no longer at their highest levels) were the prime contributors
to the record setting depletion.
2005 Climate Prediction Center summary-In the winter of 2005-2006,
positive anomalies of total ozone were prevalent in the high latitudes of the
Northern Hemisphere. The positive anomalies in total ozone were associated
with the meteorological conditions of positive anomalies of lower
stratosphere temperature. Arctic temperatures were not sufficiently low for
the formation of polar stratospheric clouds and consequent chemical ozone
depletion within the polar vortex. Chlorine and other ozone destroying
chemicals in the lower stratosphere reached peak values around 1997-98,
and have remained at high levels. Lower stratosphere ozone destruction is
strong when meteorological conditions of a strong polar vortex and cold polar
temperatures prevail. Those cold conditions were not present in the lower
stratosphere in the winter of 2005-2006.
Total ozone declined over mid-latitudes of the Northern Hemisphere at the
rate of about 2 to 4 percent per decade from 1979 to 1993. In recent years
the strong rate of decline of Northern Hemisphere total ozone has not
continued, but current stratospheric ozone amounts continue to be below the
amounts measured before the early 1980s. A full explanation of ozone and
temperature anomalies must include all aspects of ozone photochemistry
and meteorological dynamics. Continued monitoring and measurements are
essential toward this end.
2004 Climate Prediction Center summary-Very low ozone values were
again observed over Antarctica in the winter/spring of 2004. Ozone depletion
of more than 45 percent was observed over Antarctica, compared to total
ozone amounts observed in the early 1980's. Vertical soundings over the
South Pole during August, September and October 2004 again showed
strongest destruction of ozone at altitudes between 15 and 20 km. Lower
stratosphere temperatures in the winter of 2004 over the Antarctic region
were higher than in 2003. Associated with this, the ozone hole area was
smaller than in 2003. The ozone hole in 2004 diminished in size and depth
along with warming stratospheric conditions in September and October, and,

by mid-November, total ozone over Antarctica had increased to levels above


ozone hole values.
Observations of chloroflourocarbons and of stratospheric hydrogen chloride
support the view that international actions are reducing the use and release
of ozone depleting substances (WMO, 1999; Anderson et al., 2000). However,
chemicals already in the atmosphere are expected to continue to impact the
ozone amount for many decades to come. Further, changing atmospheric
conditions that modulate ozone can complicate the task of detecting the
start of ozone layer recovery. The eruption of the Pinatubo volcano provided
an example of such a complication in the 1990s. Based on an analysis of 10
years of South Pole ozone vertical profile measurements, Hofmann et al.,
(1997) estimated that recovery in the Antarctic ozone hole may be detected
as early as the coming decade. Indicators include: 1) an end to springtime
ozone depletion at 22-24 km, 2) 12-20 km mid-September column ozone loss
rate of less than 3 DU per day, and 3) a 12-20 km ozone column of more
than 70 DU on September 15. An intriguing aspect of recent observations of
the Antarctic stratosphere had been the apparent trend towards a later
breakup of the vortex in years since 1990, relative to the 1980s. The limited
duration and size of the 2004 ozone hole is attributed in part to
meteorological conditions. A full explanation of such meteorological
anomalies is not yet available. Continued monitoring and measurements,
including total ozone and its vertical profile, are essential to achieving the
understanding needed to identify ozone recovery.
2003 Climate Prediction Center summary-Very low ozone values were
again observed over Antarctica in the winter/spring of 2003. Ozone depletion
of more than 40 percent was observed over Antarctica, compared to total
ozone amounts observed in the early 1980's. Vertical soundings over the
South Pole during August, September and October 2003 again showed
strongest destruction of ozone at altitudes between 15 and 21 km. Lower
stratosphere temperatures in the winter of 2003 over the Antarctic region
were also much lower than average values. Associated with this, the ozone
hole area in August and September was among the largest of any previous
year. The ozone hole diminished in October and November along with
warming stratospheric conditions. So although the Antarctic ozone hole in
2003 reached unprecedented size, the duration of the extremely low ozone
conditions was limited by warmer meteorological conditions which developed
in October and November.

Observations of chloroflourocarbons and of stratospheric hydrogen chloride


support the view that international actions are reducing the use and release
of ozone depleting substances (WMO, 1999; Anderson et al., 2000). However,
chemicals already in the atmosphere are expected to continue to deplete
ozone for many decades to come. Further, changing atmospheric conditions
that modulate ozone can complicate the task of detecting the start of ozone
layer recovery. The eruption of the Pinatubo volcano provided an example of
such a complication in the 1990s. Based on an analysis of 10 years of South
Pole ozone vertical profile measurements, Hofmann et al., (1997) estimated
that recovery in the Antarctic ozone hole may be detected as early as the
coming decade. Indicators include: 1) an end to springtime ozone depletion
at 22-24 km, 2) 12-20 km mid-September column ozone loss rate of less than
3 DU per day, and 3) a 12-20 km ozone column of more than 70 DU on
September 15. An intriguing aspect of recent observations of the Antarctic
stratosphere had been the apparent trend towards a later breakup of the
vortex in most recent years. The large size of the August-September 2003
ozone hole but its limited duration in October-November is attributed in part
to meteorological conditions. A full explanation of such meteorological
anomalies is not yet available. Continued monitoring and measurements,
including total ozone and its vertical profile, are essential to achieving the
understanding needed to identify ozone recovery.
2002 Climate Prediction Center summary-Very low ozone values were
observed over Antarctica again in the Southern Hemisphere winter of 2002.
Ozone depletion of more than 40 percent was observed over Antarctica
compared to total ozone amounts observed in the early 1980's. Vertical
soundings over the South Pole during September and October 2002 again
showed strong destruction of ozone at altitudes between 15 and 20 km.
However, for the year 2002, the ozone hole declined rapidly in late
September, and had the shortest duration of any year since 1988. Lower
stratosphere temperatures in the winter and spring of 2002 over the
Antarctic region were much higher than average values. Associated with this,
the ozone hole area was among the smallest of recent years.
Observations of chloroflourocarbons and of stratospheric hydrogen chloride
support the view that international actions are reducing the use and release
of ozone depleting substances ; Anderson et al., 2000). However, chemicals
already in the atmosphere are expected to continue to deplete ozone for
many decades to come. Further, changing atmospheric conditions that
modulate ozone can complicate the task of detecting the start of ozone layer
recovery. The eruption of the Pinatubo volcano provided an example of such

a complication in the 1990s. Based on an analysis of 10 years of South Pole


ozone vertical profile measurements, estimated that recovery in the
Antarctic ozone hole may be detected as early as the coming decade.
Indicators include: 1) an end to springtime ozone depletion at 22-24 km, 2)
12-20 km mid-September column ozone loss rate of less than 3 DU per day,
and 3) a 12-20 km ozone column of more than 70 DU on September 15. An
intriguing aspect of recent observations of the Antarctic stratosphere had
been the apparent trend towards a later breakup of the vortex in most recent
years. However, the limited size and duration of the 2002 ozone hole is
attributed to highly unusual meteorological conditions this year. A full
explanation of such meteorological anomalies is not yet available. Continued
monitoring and measurements, including total ozone and its vertical profile,
are essential to achieving the understanding needed to identify ozone
recovery.
http://www.cpc.ncep.noaa.gov/products/stratosphere/winter_bulletins/sh_02/i
ndex.html
2001 Climate Prediction Center summary-Very low ozone values were
observed over Antarctica again in 2001. Ozone depletion of 10 percent to
more than 40 percent was observed over Antarctica compared to total ozone
amounts observed in the early 1980's. Vertical soundings over the South Pole
during late September and early October 2001 again showed complete
destruction of ozone at altitudes between 15 and 20 km. Lower stratosphere
temperatures in the winter and spring of 2001 over the Antarctic region were
below average values, and were sufficiently low for ozone production of polar
stratospheric clouds within the polar vortex. The ozone hole area and the
PSC area were again among the largest of all previous years. For the year
2001, the ozone hole and Southern Hemisphere polar vortex persisted into
December, again among the longest duration of years since 1982.
Observations of chloroflourocarbons and of stratospheric hydrogen chloride
support the view that international actions are reducing the use and release
of ozone depleting substances . However, chemicals already in the
atmosphere are expected to continue to deplete ozone for many decades to
come. Further, changing atmospheric conditions that modulate ozone can
complicate the task of detecting the start of ozone layer recovery. The
eruption of the Pinatubo volcano provided an example of such a complication
in the 1990s. Based on an analysis of 10 years of South Pole ozone vertical
profile measurements, estimated that recovery in the Antarctic ozone hole
may be detected as early as the coming decade. Indicators include: 1) an

end to springtime ozone depletion at 22-24 km, 2) 12-20 km mid-September


column ozone loss rate of less than 3 DU per day, and 3) a 12-20 km ozone
column of more than 70 DU on September 15. However, an intriguing aspect
of recent observations of the Antarctic stratosphere is the apparent trend
towards a later breakup of the vortex. A full explanation of such
meteorological anomalies is not yet available. Continued monitoring and
measurements, including total ozone and its vertical profile, are essential to
achieving the understanding needed to identify ozone recovery.
http://www.cpc.ncep.noaa.gov/products/stratosphere/winter_bulletins/sh_01/i
ndex.html
British Antarctic Survey summary 2000-The 2000 ozone hole reached its
greatest extent in early September at 28.4 million square kilometres and was
the largest ever-recorded ozone hole. It reached this maximum extent
unusually early, as the hole is normally largest in late September. Minimum
ozone values were measured in early October and were amongst the lowest
on record. The edge of the ozone hole passed over the tip of South America
and the Falkland Islands for significant periods between October 6 and
24. The hole also closed in unusually early and the event was essentially
over by early December. This early closure of the hole is probably linked
with the QBO/ENSO and in retrospect was to be expected.
British Antarctic Survey summary 1999- The ozone layer over Antarctica
evolved in a broadly similar way in 1999/2000 to each of the past few years.
Significant ozone depletion over Halley had already occurred by late August
and maximum depletion, of 60%, occurred in early October. The ozone hole
filled rapidly in early December, however a remnant area over central
Antarctica, with ozone values below 240 DU, persisted until early February.
Ozone values remained at around 255 DU at Halley for the rest of the season
(15% below the normal). At Vernadsky ozone values slowly fell from a peak in
mid December until the end of March but rose by 30 DU in April.
Stratospheric temperatures at Halley remained below the normal from late
September until early April.
British Antarctic Survey summary 1998-The 1998 Antarctic ozone hole
was broadly similar to those seen in recent years, although a little larger and
a little longer lasting. At Halley the final spring warming did not take place
until late December and a remnant of the ozone hole passed over the station
in late January.

Mean total ozone values declined from around 260 DU in early August to a
minimum of 120 DU in early October at a rate of around 3 DU per day. They
rose a little, but oscillated around 180 DU from late October until the final
warming began in early December. Peak ozone values of around 290 DU
occurred in late December, but this was nearly a month later and 25% lower
than the historical maximum. Values slowly declined from this peak through
the rest of the summer. Unusually low ozone values and 100 hPa
temperatures were recorded around January 21 when a remnant of the ozone
hole passed over the station. Autumn values were around 260 DU, some 15%
below normal.
British Antarctic Survey summary 1997-Overall the 1997/98 ozone
season was comparable to those of recent years, with minimum ozone
values of around 110 DU at Halley. The spring warming took place slightly
earlier than in the past two years, leading to a higher annual mean. The
January and February means for Vernadsky were the lowest on record. The
100 hPa temperature over Halley and Vernadsky was below the long term
mean for most of the season.Total ozone values fell from an estimated 300
DU in early July to around 200 DU in mid August. Values rose to around 240
DU towards the end of August but then steadily fell to reach 115 DU in early
October. Ozone values rose to around 160 DU late in October and remained
near this value until mid November when the spring warming started. Values
peaked at just under 300 DU in early December, compared to the long term
mean peak of 380 DU. Summer values declined to a plateau in early January
at around 260 DU. Daily variation began to increase at the beginning of
March, although mean values have changed little. Ozone values remained
significantly below the long term mean throughout the year.
British Antarctic Survey summary 1996-Routine ozone measurements
from Halley station do not start until towards the end of August because the
sun is not high enough to permit measurements. Mean total ozone values fell
from an estimated 260 DU in early July to around 140 DU in mid September
and then declined more slowly to reach 125 DU in early October. They
climbed to around 180 DU by late October, but remained at that level
throughout November. In early December, ozone levels began to rise and
reached a peak of around 300 DU late in the month. This is some 80 DU
lower than the pre-ozone hole average peak value. After that, values slowly
declined and had reached around 230 DU by the end of the season. Values
only briefly exceeded the lower bound of the variation from the long-term
mean late in the season. The lowest mean daily value seen (114 DU on
October 8) is comparable to that of the previous few years.

British Antarctic Survey summary 1995-Preliminary mean daily ozone


values dropped from values of around 280 Dobson Units (DU) at the
beginning of August to around 175 DU at the end of September (50%
depletion). Individual daily values dropped as low as 132 DU. The first week
of October saw a major spring warming event, with a rise in mean total
ozone to over 300 DU as the circumpolar high ozone belt moved across the
station. Mean values then fell back to around 190 DU in mid October before
rising again to over 300 DU in a second warming event at the end of October.
Mean values dropped back again to 200 DU in early November before slowly
rising to around 315 DU in the final warming event of the year in mid
December. Values slowly declined from the December peak to around 260
DU at the end of April.
Generally values in the 95/96 season were a little above those reached in
1993 and 1994 during the first half of the season and below during the
second half. The final spring warming was much later than in either year.
Values were significantly below those of the long term mean throughout the
season.
British Antarctic Survey summary 1994-Preliminary mean ozone values
from Faraday show that mean ozone values at the end of the winter were
around 290 DU dropping to a minimum of around 160 DU by mid October
(55% depletion). This is the steepest decline recorded at this station, and the
minimum mean values were about 40 DU below those at the same time last
year. The mean for September is the lowest September mean recorded at
the station; 133 DU recorded at the end of the month is the lowest ever
September mean daily value. Individual values of 130 DU in early October
represent nearly 65% depletion and are the second lowest on record for the
station. A strong spring warming episode occurred between October 20th
and 28th, with a very rapid rise to 400 DU on October 22nd; this is the
earliest the warming has occurred for several years and is also the steepest
rise (some 250 DU in three days). Following the event mean values dropped
to 180 DU (50% depletion), and then slowly rose again, reaching 230 DU by
November 23rd. The final spring warming occurred between November 21st
and 23rd with mean ozone values rising rapidly to around 340 DU by the end
of November. The November monthly mean is the second lowest on record.
Values in early December were near normal at around 360 DU, but were
generally below normal throughout the late summer and autumn.
Preliminary ozone measurements from Halley show that mean ozone values
dropped from around 180 DU at the end of August to 105 DU in early

October (65% depletion), then slowly rose to reach 160 DU by November 7th.
A minor peak occurred between November 7th and 17th, with a maximum of
230 DU on the 11th, followed by a decline to 185 DU. The spring warming
commenced on November 24th with ozone levels peaking at around 350 DU
in late November and then slowly declining. The monthly means for
September and October are between those of 1992 and 1993. Mean values
throughout the summer and autumn were significantly below those of the
long term mean and throughout the season no daily values exceeded those
of the long term mean.
The data from 1956 - 1994 suggest that the mean for October 1995 will be
around 120 DU, with minimum values below 100 DU.
The temperature at 100 hPa rose slowly from the winter value of around -82
deg C in early October to reach -67 deg C by November 22nd (some 15
degrees below normal). The final warming then took place with the
temperature reaching near normal values of around -40 deg C by the end of
the month. The temperature then slowly declined, reaching -60 deg C by the
end of April.
The ozone values from Halley and Faraday are consistent with a hole of
greater maximum extent than in 1993, but slightly shallower. NASA GSFC
images from Meteor-3/TOMS and radiosonde data suggest that the edge of
the hole was over the tip of South America and the Falkland Islands between
October 11th and 21st, and was close to the region between October 29th
and November 13th. The Falkland Islands now consider the ozone hole a
major threat to health following cases of severe sunburn during this event.
British Antarctic Survey summary 1993-Preliminary ozone values from
the British Antarctic Survey Faraday station (65 south, 64 west on the coast
of the Antarctic Peninsula) show running average ozone values dropping from
around 280 DU at the begining of September to around 200 DU at the
month's end. There is considerable day to day variation. Data from our long
term record for the station indicate normal values of around 330 DU for this
period (with a range of 280 DU to 380 DU). This year's mean September
ozone value is 245 DU, giving a reduction of about 25%. The minimum
September daily mean value of 142 DU on September 28th is the lowest ever
recorded at Faraday (next lowest 145 DU in 1992).
Preliminary ozone values from the British Antarctic Survey Halley station (76
south, 26 west on the Brunt ice shelf) show mean ozone values dropping
from around 210 DU in early September to around 110 DU in early October.

00

Data from our long term record for the station indicate normal values of
around 300 DU for this period (with a range of 260 DU to 340 DU). This
year's mean September ozone value of 159 DU is the second lowest recorded
at Halley (lowest 146 DU in 1992), giving a reduction of around 35%. The
minimum September daily mean value of 108 DU on September 30 is the
lowest recorded at Halley (next lowest 111 DU in 1992). Ozone levels in early
October are the lowest ever recorded with depletion exceeding two thirds.
This year's Antarctic ozone hole (1993) is the deepest ever. Scientists of the
British Antarctic Survey (BAS) at Halley Research Station in Antarctica have
recorded their lowest ever value of ozone, showing that over two-thirds of
the protective ozone shield has been destroyed. Less than one- third remains
intact, and the depletion phase may not be over for some days yet. BAS
scientists at Faraday Research Station, normally near the edge of the ozone
hole, have also measured record low values of ozone for their station.
The Antarctic ozone hole is an annual phenomenon which has occurred in
every (southern hemisphere) spring since the late 1970s. It is caused by the
chemical action of chlorine and bromine released from man-made
chlorofluorocarbons (CFCs) and halons. Thanks to the international Montreal
Protocol, the annual production of these substances is now less than before,
but the total amount in the atmosphere continues to rise, dashing hopes of
an early recovery for the ozone layer. The CFCs are not expected to reach
their maximum concentration in the atmosphere until around the turn of the
century, so the Antarctic ozone hole may continue to worsen until then.
BAS scientists have been measuring ozone in Antarctica on a daily basis
since 1957. There was no ozone hole at all until the late 1970s. The Antarctic
ozone layer is not expected to return to normal until the 2070s.
HISTORY OF THE OZONE LAYER
Ozone layer forms
Christian Schenbein identifies ozone in the laboratory
Auguste de la Rive and Jean-Charles de Marignac suggest ozone is a form of oxygen; confirmed
Thomas Andrews in 1856
Andrei Houzeau finds ozone present in natural air
Jean-Louis Soret proves that ozone is O3

Marie Alfred Cornu measures solar spectrum and finds sharp cutoff in ultraviolet (UV) light
Walter Hartley recognizes cutoff corresponds to UV absorption by ozone
John William Strutt (Lord Rayleigh) shows absorption is not in lower atmosphere
Charles Fabry makes first spectrometric measurements of "thickness" of ozone layer
G.M.B. Dobson develops ozone spectrophotometer and begins regular measurements of ozone
abundance (Arosa, Switzerland)
Jean Cabannes and Jean Dufay show ozone is about 10 miles high
Thomas Midgley synthesizes chlorofluorocarbons (CFC's)

Umkehr method for Dobson instrument establishes that ozone maximum is below 15 miles altitu
Sydney Chapman describes theory that explains existence of an ozone "layer"

Ozonesonde (balloon) measurements establish the ozone concentration is maximum around 12


GM develops applications for CFC's
David Bates and Marcel Nicolet propose catalytic (HOx) ozone destruction

Global network of Dobson spectrophotometers established during the International Geophysical


(IGY)

CFC market expands rapidly

0's

Catalytic destruction is necessary in order to explain ozone amounts


Boeing proposes supersonic transport (SST) fleet of 800 aircraft
Paul Crutzen discovers NOx catalytic cycle

Dept of Transportation sponsors intensive program of research, The Climatic Impact Assessmen
(CIAP)
Congress axes funding for the SST
Johnston calculates that NOx from SST's could deplete ozone layer
Rick Stolarski and Ralph Cicerone suggest catalytic capability of Cl
James Lovelock detects CFC's in atmosphere
Sherwood Rowland and Mario Molina warn of ozone depletion due to CFC's

77

First international meeting (Washington DC) to address issue of ozone depletion held by the Uni
Nations Environmental Programme (UNEP)

78

US bans non-essential use of CFC's as aerosol propellant

Total Ozone Mapping Spectrometer (TOMS) is launched aboard NIMBUS-7 spacecraft giving glob
coverage of ozone layer thickness
Renewed expansion of CFC market

Shigeru Chubachi measures low ozone over Syowa, Antarctica (reported at Ozone Commission m
in Halkidiki, Greece in Sept 1984)

British Antarctic Survey scientists discover recurring springtime Antarctic ozone hole (published
May 1985)

85

Vienna Convention for the Protection of the Ozone Layer

Montreal Protocol on Substances That Deplete the Ozone Layer (Amendments - London 1990;
Copenhagen 1992)

88

DuPont agrees to CFC production phase-out

Ten years of satellite data begin to show measurable ozone depletion globally
DuPont announces phase-out of CFC production by end of 1996
Abnormally low ozone observed globally
Crutzen, Rowland, and Molina win Nobel Prize in Chemistry
springtime Arctic ozone dent appearing
CFC production ends in US and Europe
Maximum CFC concentrations in stratosphere are reached
The Ozone Layer - Global Map

very rare in our atmosphere, averaging about three molecules of ozone


10 million air molecules. In spite of this small amount, ozone plays a vital
atmosphere. In the information below, we present "the basics" about
tant component of the Earth's atmosphere.

Where is ozone found in the atmosphere?


Ozone is mainly found in two regions of the Earth's atmosphere. Most ozone
(about 90%) resides in a layer that begins between 6 and 10 miles (10 and
17 kilometers) above the Earth's surface and extends up to about 30 miles
(50 kilometers). This region of the atmosphere is called the stratosphere. The
ozone in this region is commonly known as the ozone layer. The remaining
ozone is in the lower region of the atmosphere, which is commonly called the
troposphere. The figure (above) shows an example of how ozone is
distributed in the atmosphere.
What roles does ozone play in the atmosphere and how are humans
affected?
The ozone molecules in the upper atmosphere (stratosphere) and the lower
atmosphere (troposphere) are chemically identical, because they all consist
of three oxygen atoms and have the chemical formula O3. However, they
have very different roles in the atmosphere and very different effects on
humans and other living beings. Stratospheric ozone (sometimes referred to
as "good ozone") plays a beneficial role by absorbing most of the biologically
damaging ultraviolet sunlight (called UV-B), allowing only a small amount to
reach the Earth's surface. The absorption of ultraviolet radiation by ozone
creates a source of heat, which actually forms the stratosphere itself (a
region in which the temperature rises as one goes to higher altitudes). Ozone
thus plays a key role in the temperature structure of the Earth's atmosphere.
Without the filtering action of the ozone layer, more of the Sun's UV-B
radiation would penetrate the atmosphere and would reach the Earth's
surface. Many experimental studies of plants and animals and clinical studies
of humans have shown the harmful effects of excessive exposure to UV-B
radiation.
At the Earth's surface, ozone comes into direct contact with life-forms and
displays its destructive side (hence, it is often called "bad ozone"). Because
ozone reacts strongly with other molecules, high levels of ozone are toxic to
living systems. Several studies have documented the harmful effects of
ozone on crop production, forest growth, and human health. The substantial
negative effects of surface-level tropospheric ozone from this direct toxicity
contrast with the benefits of the additional filtering of UV-B radiation that it
provides.

What are the environmental issues associated with ozone?


The dual role of ozone leads to two separate environmental issues. There is
concern about increases in ozone in the troposphere. Near-surface ozone is a
key component of photochemical "smog," a familiar problem in the
atmosphere of many cities around the world. Higher amounts of surface-level
ozone are increasingly being observed in rural areas as well.
There is also widespread scientific and public interest and concern about
losses of ozone in the stratosphere. Ground-based and satellite instruments
have measured decreases in the amount of stratospheric ozone in our
atmosphere. Over some parts of Antarctica, up to 60% of the total overhead
amount of ozone (known as the column ozone) is depleted during Antarctic
spring (September-November). This phenomenon is known as the Antarctic
ozone hole. In the Arctic polar regions, similar processes occur that have also
led to significant chemical depletion of the column ozone during late winter
and spring in 7 out of the last 11 years. The ozone loss from January through
late March has been typically 20-25%, and shorter-period losses have been
higher, depending on the meteorological conditions encountered in the Arctic
stratosphere. Smaller, but still significant, stratospheric decreases have been
seen at other, more-populated regions of the Earth. Increases in surface UV-B
radiation have been observed in association with local decreases in
stratospheric ozone, from both ground-based and satellite-borne
instruments.
What human activities affect upper-atmospheric ozone (the stratospheric
ozone layer)?
The scientific evidence, accumulated over more than two decades of study
by the international research community, has shown that human-produced
chemicals are responsible for the observed depletions of the ozone layer. The
ozone-depleting compounds contain various combinations of the chemical
elements chlorine, fluorine, bromine, carbon, and hydrogen and are often
described by the general term halocarbons. The compounds that contain
only chlorine, fluorine, and carbon are called chlorofluorocarbons, usually
abbreviated as CFCs. CFCs, carbon tetrachloride, and methyl chloroform are
important human-produced ozone-depleting gases that have been used in
many applications including refrigeration, air conditioning, foam blowing,
cleaning of electronics components, and as solvents. Another important

group of human-produced halocarbons is the halons, which contain carbon,


bromine, fluorine, and (in some cases) chlorine and have been mainly used
as fire extinguishants.
What actions have been taken to protect the ozone layer?
Through an international agreement known as the Montreal Protocol on
Substances that Deplete the Ozone Layer, governments have decided to
eventually discontinue production of CFCs, halons, carbon tetrachloride, and
methyl chloroform (except for a few special uses), and industry has
developed more "ozone-friendly" substitutes. All other things being equal,
and with adherence to the international agreements, the ozone layer is
expected to recover over the next 50 years or so.

The ozone layer is a belt of naturally occurring ozone gas that sits 9.3 to 18.6
miles (15 to 30 kilometers) above Earth and serves as a shield from the
harmful ultraviolet B radiation emitted by the sun.
Ozone is a highly reactive molecule that contains three oxygen atoms. It is
constantly being formed and broken down in the high atmosphere, 6.2 to 31
miles (10 to 50 kilometers) above Earth, in the region called the
stratosphere.
Today, there is widespread concern that the ozone layer is deteriorating due
to the release of pollution containing the chemicals chlorine and bromine.
Such deterioration allows large amounts of ultraviolet B rays to reach Earth,
which can cause skin cancer and cataracts in humans and harm animals as
well.
Extra ultraviolet B radiation reaching Earth also inhibits the reproductive
cycle of phytoplankton, single-celled organisms such as algae that make up
the bottom rung of the food chain. Biologists fear that reductions in
phytoplankton populations will in turn lower the populations of other animals.
Researchers also have documented changes in the reproductive rates of
young fish, shrimp, and crabs as well as frogs and salamanders exposed to
excess ultraviolet B.
Chlorofluorocarbons (CFCs), chemicals found mainly in spray aerosols heavily
used by industrialized nations for much of the past 50 years, are the primary

culprits in ozone layer breakdown. When CFCs reach the upper atmosphere,
they are exposed to ultraviolet rays, which causes them to break down into
substances that include chlorine. The chlorine reacts with the oxygen atoms
in ozone and rips apart the ozone molecule.
One atom of chlorine can destroy more than a hundred thousand ozone
molecules, according to the the U.S. Environmental Protection Agency.
The ozone layer above the Antarctic has been particularly impacted by
pollution since the mid-1980s. This regions low temperatures speed up the
conversion of CFCs to chlorine. In the southern spring and summer, when the
sun shines for long periods of the day, chlorine reacts with ultraviolet rays,
destroying ozone on a massive scale, up to 65 percent. This is what some
people erroneously refer to as the "ozone hole." In other regions, the ozone
layer has deteriorated by about 20 percent.
About 90 percent of CFCs currently in the atmosphere were emitted by
industrialized countries in the Northern Hemisphere, including the United
States and Europe. These countries banned CFCs by 1996, and the amount
of chlorine in the atmosphere is falling now. But scientists estimate it will
take another 50 years for chlorine levels to return to their natural levels.
one layer, also called ozonosphere, region of the upperatmosphere, between
roughly 15 and 35 km (9 and 22 miles) above Earths surface, containing
relatively high concentrations of ozonemolecules (O3). Approximately 90
percent of the atmospheres ozone occurs in the stratosphere, the region
extending from 1018 km (611 miles) to approximately 50 km (about 30
miles) above Earths surface. In thestratosphere the temperature of
the atmosphere rises with increasing height, a phenomenon created by the
absorption of solar radiation by the ozone layer. The ozone layer effectively
blocks almost all solar radiation ofwavelengths less than 290 nanometres
from reaching Earths surface, including certain types of ultraviolet (UV) and
other forms of radiation that could injure or kill most living things.
The layers of Earths atmosphere. The yellow line shows the response of air
temperature to
LOCATION IN EARTHS ATMOSPHERE
In the midlatitudes the peak concentrations of ozone occur at altitudes from
20 to 25 km (about 12 to 16 miles). Peak concentrations are found at
altitudes from 26 to 28 km (about 16 to 17 miles) in the tropics and from
about 12 to 20 km (about 7 to 12 miles) toward the poles. The lower height

of the peak-concentration region in the high latitudes largely results from


poleward and downward atmospheric transport processes that occur in the
middle and high latitudes and the reduced height of the tropopause (the
transition region between the troposphere and stratosphere).
Most of the remaining ozone occurs in the troposphere, the layer of the
atmosphere that extends from Earths surface up to the stratosphere. Nearsurface ozone often results from interactions between certain pollutants
(such as nitrogen oxides and volatile organic compounds), strong sunlight,
and hot weather. It is one of the primary ingredients in photochemical smog,
a phenomenon that plagues many urban and suburban areas around the
world, especially during the summer months.
The production of ozone in the stratosphere results primarily from the
breaking of the chemical bonds within oxygen molecules (O2) by high-energy
solar photons. This process, called photodissociation, results in the release of
single oxygen atoms, which later join with intact oxygen molecules to form
ozone. Rising atmospheric oxygen concentrations some two billion years ago
allowed ozone to build up in Earths atmosphere, a process that gradually led
to the formation of the stratosphere. Scientists believe that the formation of
the ozone layer played an important role in the development
oflife on Earth by screening out lethal levels of UVB radiation (ultraviolet
radiation with wavelengths between 315 and 280 nanometres) and thus
facilitating the migration of life-forms from the oceans to land.
The amount of ozone in the stratosphere varies naturally throughout the year
as a result of chemical processes that create and destroy ozone molecules
and as a result of winds and other transport processes that move ozone
molecules around the planet. Over the course of several decades, however,
human activities substantially altered the ozone layer. Ozone depletion, the
global decrease in stratospheric ozone observed since the 1970s, is most
pronounced in polar regions, and it is well correlated with the increase
ofchlorine and bromine in the stratosphere. Those chemicals, once freed
by UVradiation from the chlorofluorocarbons (CFCs) and other halocarbons
(carbon-halogen compounds) that contain them, destroy ozone by stripping
away single oxygen atoms from ozone molecules. Depletion is so extensive
that so-called ozone holes (regions of severely reduced ozone coverage)
form over the poles during the onset of their respective spring seasons. The
largest such holewhich has spanned more than 20.7 million square km (8
million square miles) on a consistent basis since 1992appears annually
over Antarctica between September and November.

Two bar graphs depicting the maximum ozone hole size and the minimum
ozone coverage (in Dobson
As the amount of stratospheric ozone declines, more UV radiation reaches
Earths surface, and scientists worry that such increases could have
significant effects on ecosystems and human health. The concern over
exposure to biologically harmful levels of UV radiation has been the main
driver of the creation of international treaties such as the Montreal Protocol
on Substances That Deplete the Ozone Layer and its amendments, designed
to protect Earths ozone layer. Compliance with international treaties that
phased out the production and delivery of many ozone-depleting chemicals,
combined with upper stratospheric cooling due to increased carbon dioxide,
is thought to have resulted in slightly higher stratospheric ozone levels that
were first observed in 2014. Continued reductions in chlorine loading are also
expected to result in smaller ozone holes above Antarctica after 2040.
The ozone layer or ozone shield is a region of Earth's stratosphere that
absorbs most of the Sun's ultraviolet (UV) radiation. It contains high
concentrations ozone (O3) in relation to other parts of the atmosphere,
although still small in relation to other gases in the stratosphere. The ozone
layer contains less than 10 parts per million of ozone, while the average
ozone concentration in Earth's atmosphere as a whole is about 0.3 parts per
million. The ozone layer is mainly found in the lower portion of the
stratosphere, from approximately 20 to 30 kilometres (12 to 19 mi) above
Earth, although the thickness varies seasonally and geographically. [1]
The ozone layer was discovered in 1913 by the French physicists Charles
Fabry and Henri Buisson. Measurements of the sun showed that the radiation
sent out from its surface and reaching the ground on Earth is usually
consistent with the spectrum of ablack body with a temperature in the range
of 5,5006,000 K (5,227 to 5,727 C), except that there was no radiation
below a wavelength of about 310 nm at the ultraviolet end of the spectrum.
It was deduced that the missing radiation was being absorbed by something
in the atmosphere. Eventually the spectrum of the missing radiation was
matched to only one known chemical, ozone.[2] Its properties were explored
in detail by the British meteorologist G. M. B. Dobson, who developed a
simplespectrophotometer (the Dobsonmeter) that could be used to measure
stratospheric ozone from the ground. Between 1928 and 1958, Dobson
established a worldwide network of ozone monitoring stations, which
continue to operate to this day. The "Dobson unit", a convenient measure of
the amount of ozone overhead, is named in his honor.

The ozone layer absorbs 97 to 99 percent of the Sun's medium-frequency


ultraviolet light (from about 200 nm to 315 nm wavelength), which otherwise
would potentially damage exposed life forms near the surface.[3]
The United Nations General Assembly has designated September 16 as
the International Day for the Preservation of the Ozone Layer.
Venus also has a thin ozone layer at an altitude of 100 kilometers from the
planet's surface.[4]
The photochemical mechanisms that give rise to the ozone layer were
discovered by the British physicist Sydney Chapman in 1930. Ozone in the
Earth's stratosphere is created by ultraviolet light striking
ordinary oxygen molecules containing two oxygen atoms (O2), splitting them
into individual oxygen atoms (atomic oxygen); the atomic oxygen then
combines with unbroken O2 to create ozone, O3. The ozone molecule is
unstable (although, in the stratosphere, long-lived) and when ultraviolet light
hits ozone it splits into a molecule of O2 and an individual atom of oxygen, a
continuing process called the ozone-oxygen cycle. Chemically, this can be
described as:
O2 + uv 2O
O + O2 O3
About 90 percent of the ozone in our atmosphere is contained in the
stratosphere. Ozone concentrations are greatest between about 20 and 40
kilometres (66,000 and 131,000 ft), where they range from about 2 to 8 parts
per million. If all of the ozone were compressed to the pressure of the air at
sea level, it would be only 3 millimetres (18 inch) thick.
Although the concentration of the ozone in the ozone layer is very small, it is
vitally important to life because it absorbs biologically harmful ultraviolet
(UV) radiation coming from the sun. Extremely short or vacuum UV (10
100 nm) is screened out by nitrogen. UV radiation capable of penetrating
nitrogen is divided into three categories, based on its wavelength; these are
referred to as UV-A (400315 nm), UV-B (315280 nm), and UV-C (280
100 nm).
UV-C, which is very harmful to all living things, is entirely screened out by a
combination of dioxygen (< 200 nm) and ozone (> about 200 nm) by around
35 kilometres (115,000 ft) altitude. UV-B radiation can be harmful to the skin
and is the main cause of sunburn; excessive exposure can also cause

cataracts, immune system suppression, and genetic damage, resulting in


problems such as skin cancer. The ozone layer (which absorbs from about
200 nm to 310 nm with a maximal absorption at about 250 nm)[6] is very
effective at screening out UV-B; for radiation with a wavelength of 290 nm,
the intensity at the top of the atmosphere is 350 million times stronger than
at the Earth's surface. Nevertheless, some UV-B, particularly at its longest
wavelengths, reaches the surface, and is important for the skin's production
of vitamin D.
Ozone is transparent to most UV-A, so most of this longer-wavelength UV
radiation reaches the surface, and it constitutes most of the UV reaching the
Earth. This type of UV radiation is significantly less harmful to DNA, although
it may still potentially cause physical damage, premature aging of the skin,
indirect genetic damage, and skin cancer.
The thickness of the ozone layerthat is, the total amount of ozone in a
column overheadvaries by a large factor worldwide, being in general
smaller near the equator and larger towards the poles. It also varies with
season, being in general thicker during the spring and thinner during the
autumn. The reasons for this latitude and seasonal dependence are
complicated, which involve in atmospheric circulation patterns as well as
solar intensity.
Since stratospheric ozone is produced by solar UV radiation, one might
expect to find the highest ozone levels over the tropics and the lowest over
polar regions. The same argument would lead one to expect the highest
ozone levels in the summer and the lowest in the winter. The observed
behavior is very different: most of the ozone is found in the mid-to-high
latitudes of the northern and southern hemispheres, and the highest levels
are found in the spring, not summer, and the lowest in the autumn, not
winter in the northern hemisphere. During winter, the ozone layer actually
increases in depth. This puzzle is explained by the prevailing stratospheric
wind patterns, known as the Brewer-Dobson circulation. While most of the
ozone is indeed created over the tropics, the stratospheric circulation then
transports it poleward and downward to the lower stratosphere of the high
latitudes.[8] However, owing to the ozone holephenomenon, the lowest
amounts of column ozone found anywhere in the world are over the Antarctic
in the southern spring period of September and October and to a lesser
extent over the Arctic in the northern spring period of March, April, and May.
Brewer-Dobson circulation in the ozone layer.

The ozone layer is higher in altitude in the tropics, and lower in altitude
outside the tropics, especially in the polar regions. This altitude variation of
ozone results from the slow circulation that lifts the ozone-poor air out of the
troposphere into the stratosphere. As this air slowly rises in the tropics,
ozone is produced as the sun overhead photolyzes oxygen molecules. As this
slow circulation levels off and flows towards the mid-latitudes, it carries the
ozone-rich air from the tropical middle stratosphere to the lower stratosphere
middle and high latitudes . The high ozone concentrations at high latitudes
are due to the accumulation of ozone at lower altitudes.[8]
The Brewer-Dobson circulation moves very slowly. The time needed to lift an
air parcel by 1 km in the lower tropical stratosphere is about 2 months (18 m
per day).[9] However, horizontal poleward transport in the lower stratosphere
is much faster and amounts to approximately 100 km per day in the northern
hemisphere whilst it is only half as much in the southern hemisphere
(~51 km per day).[10]Even though ozone in the lower tropical stratosphere is
produced at a very slow rate, the lifting circulation is so slow that ozone can
build up to relatively high levels by the time it reaches 26 kilometres (16 mi).
[8]

Ozone amounts over the continental United States (25N to 49N) are
highest in the northern spring (April and May). These ozone amounts fall over
the course of the summer to their lowest amounts in October, and then rise
again over the course of the winter.[11]Again, wind transport of ozone is
principally responsible for the seasonal changes of these higher latitude
ozone patterns.[8]
The total column amount of ozone generally increases as we move from the
tropics to higher latitudes in both hemispheres. However, the overall column
amounts are greater in the northern hemisphere high latitudes than in the
southern hemisphere high latitudes. In addition, while the highest amounts
of column ozone over the Arctic occur in the northern spring (MarchApril),
the opposite is true over the Antarctic, where the lowest amounts of column
ozone occur in the southern spring (SeptemberOctober).[8]
NASA projections of stratospheric ozone concentrations if
chlorofluorocarbons had not been banned.
The ozone layer can be depleted by free radical catalysts, including nitric
oxide (NO), nitrous oxide (N2O), hydroxyl (OH), atomic chlorine(Cl), and
atomic bromine (Br). While there are natural sources for all of these species,
the concentrations of chlorine and bromine increased markedly in recent

decades because of the release of large quantities of manmade organohalogen compounds, especiallychlorofluorocarbons (CFCs)
and bromofluorocarbons.[12] These highly stable compounds are capable of
surviving the rise to thestratosphere, where Cl and Br radicals are liberated
by the action of ultraviolet light. Each radical is then free to initiate and
catalyze a chain reaction capable of breaking down over 100,000 ozone
molecules. By 2009, nitrous oxide was the largest ozone-depleting substance
(ODS) emitted through human activities.[13]
Levels of atmospheric ozone measured by satellite show clear seasonal
variations and appear to verify their decline over time.
The breakdown of ozone in the stratosphere results in reduced absorption of
ultraviolet radiation. Consequently, unabsorbed and dangerous ultraviolet
radiation is able to reach the Earths surface at a higher intensity. Ozone
levels have dropped by a worldwide average of about 4 percent since the
late 1970s. For approximately 5 percent of the Earth's surface, around the
north and south poles, much larger seasonal declines have been seen, and
are described as "ozone holes".[11] The discovery of the annual depletion of
ozone above the Antarctic was first announced by Joe Farman, Brian
Gardiner and Jonathan Shanklin, in a paper which appeared in Nature on May
16, 1985.[14]
Regulation[edit]
Main article: Ozone depletion and climate change
To support successful regulation attempts, the ozone case was
communicated to lay persons "with easy-to-understand bridging metaphors
derived from the popular culture" and related to "immediate risks with
everyday relevance". The specific metaphors used in the discussion (ozone
shield, ozone hole) proved quite useful[15] and, compared to global climate
change, the ozone case was much more seen as a "hot issue" and imminent
risk.[16] Lay people were cautious about a depletion of the ozone layer and
the risks of skin cancer.
In 1978, the United States, Canada and Norway enacted bans on CFCcontaining aerosol sprays that damage the ozone layer. The European
Community rejected an analogous proposal to do the same. In the U.S.,
chlorofluorocarbons continued to be used in other applications, such as
refrigeration and industrial cleaning, until after the discovery of the Antarctic
ozone hole in 1985. After negotiation of an international treaty (theMontreal

Protocol), CFC production was capped at 1986 levels with commitments to


long-term reductions.[17] Since that time, the treaty was amended to ban CFC
production after 1995 in the developed countries, and later in developing
countries.[18] Today, all of the world's 197 countries have signed the treaty.
Beginning January 1, 1996, only recycled and stockpiled CFCs were available
for use in developed countries like the US. This production phaseout was
possible because of efforts to ensure that there would be substitute
chemicals and technologies for all ODS uses.[19]
On August 2, 2003, scientists announced that the global depletion of the
ozone layer may be slowing down because of the international regulation of
ozone-depleting substances. In a study organized by the American
Geophysical Union, three satellites and three ground stations confirmed that
the upper-atmosphere ozone-depletion rate slowed down significantly during
the previous decade. Some breakdown can be expected to continue because
of ODSs used by nations which have not banned them, and because of gases
which are already in the stratosphere. Some ODSs, including CFCs, have very
long atmospheric lifetimes, ranging from 50 to over 100 years. It has been
estimated that the ozone layer will recover to 1980 levels near the middle of
the 21st century.[11] A gradual trend toward "healing" was reported in 2016.[20]
Compounds containing CH bonds (such as hydrochlorofluorocarbons, or
HCFCs) have been designed to replace CFCs in certain applications. These
replacement compounds are more reactive and less likely to survive long
enough in the atmosphere to reach the stratosphere where they could affect
the ozone layer. While being less damaging than CFCs, HCFCs can have a
negative impact on the ozone layer, so they are also being phased out.
[21]
These in turn are being replaced by hydrofluorocarbons (HFCs) and other
compounds that do not destroy stratospheric ozone at all.
Implications for astronomy[edit]
As ozone in the atmosphere prevents most energetic ultraviolet radiation
reaching the surface of the earth, astronomical data in these wavelengths
has to be gathered from satellites orbiting above the atmosphere and ozone
layer. Most of the light from young hot stars is in the ultraviolet and so study
of these wavelengths is important for studying the origins of galaxies. The
Galaxy Evolution Explorer, GALEX, is an orbiting ultraviolet space telescope
launched on April 28, 2003, which operated until early 2012.
The ozone layer is a layer in Earth's atmosphere which contains relatively
high concentrations of ozone (O3). This layer absorbs 93-99% of the sun's

high frequency ultraviolet light, which is potentially damaging to life on earth


[1]. Over 91% of the ozone in Earth's atmosphere is present here.[1] It is
mainly located in the lower portion of the stratosphere from approximately
10 km to 50 km above Earth, though the thickness varies seasonally and
geographically[2]. The ozone layer was discovered in 1913 by the French
physicists Charles Fabry and Henri Buisson. Its properties were explored in
detail by the British meteorologist G. M. B. Dobson, who developed a simple
spectrophotometer (the Dobson meter) that could be used to measure
stratospheric ozone from the ground. Between 1928 and 1958 Dobson
established a worldwide network of ozone monitoring stations which
continues to operate today. The "Dobson unit", a convenient measure of the
total amount of ozone in a column overhead, is named in his honor.

Without ozone, life on Earth would not have evolved in the way it has. The
first stage of single cell organism development requires an oxygen-free
environment. This type of environment existed on earth over 3000 million
years ago. As the primitive forms of plant life multiplied and evolved,
Manuscript Received, 27 November, 2010. *T.Sivasakthivel and K.K.Siva
Kumar Reddy, M.Tech - Thermal Engineering,, Department of Mechanical
Engineering,, N.I.T Silchar, Assam, India *Email: sivasakthivel.t@gmail.com
they began to release minute amounts of oxygen through the photosynthesis
reaction (which converts carbon dioxide into oxygen) [3]. The buildup of
oxygen in the atmosphere led to the formation of the ozone layer in the
upper atmosphere or stratosphere. This layer filters out incoming radiation in
the "cell-damaging" ultraviolet (UV) part of the spectrum. Thus with the
development of the ozone layer came the formation of more advanced life
forms. Ozone is a form of oxygen. The oxygen we breathe is in the form of
oxygen molecules (O2) - two atoms of oxygen bound together. Normal
oxygen which we breathe is colourless and odourless. Ozone, on the other
hand, consists of three atoms of oxygen bound together (O3).
Most of the atmosphere's ozone occurs in the region called the stratosphere.
Ozone is colourless and has a very harsh odour. Ozone is much less common
than normal oxygen. Out of 10 million air molecules, about 2 million are
normal oxygen, but only 3 are ozone. Most ozone is produced naturally in the
upper atmosphere or stratosphere. While ozone can be found through the
entire atmosphere, the greatest concentration occurs at altitudes between
19 and 30 km above the Earth's surface. This band of ozone-rich air is known
as the "ozone layer". [4] Ozone also occurs in very small amounts in the

lowest few kilometres of the atmosphere, a region known as the troposphere.


It is produced at ground level through a reaction between sunlight and
volatile organic compounds (VOCs) and nitrogen oxides (NOx), some of which
are produced by human activities such as driving cars. Ground-level ozone is
a component of urban smog and can be harmful to human health. Even
though both types of ozone contain the same molecules, their presence in
different parts of the atmosphere has very different consequences.
Stratospheric ozone blocks harmful solar radiation - all life on Earth has
adapted to this filtered solar radiation. Ground-level ozone, in contrast, is
simply a pollutant. It will absorb some incoming solar radiation, but it cannot
make up for ozone losses in the stratosphere.
In some of the popular news media, as well as in many books, the term
"ozone hole" has and often still is used far too loosely. Frequently, the term is
employed to describe any episode of ozone depletion, no matter how minor.
Unfortunately, this sloppy language trivializes the problem and blurs the
important scientific distinction between the massive ozone losses in Polar
Regions and the much smaller, but nonetheless significant, ozone losses in
other parts of the world. Technically, the term "ozone hole" should be applied
to regions where stratospheric ozone depletion is so severe that levels fall
below 200 Dobson Units (D.U.), the traditional measure of stratospheric
ozone. Normal ozone Ozone Layer Depletion and Its Effects: A Review
Sivasakthivel.T and K.K.Siva Kumar Reddy International Journal of
Environmental Science and Development, Vol.2, No.1, February 2011 ISSN:
2010-0264 31 concentration is about 300 to 350 D.U [3]. Such ozone loss
now occurs every springtime above Antarctica, and to a lesser extent the
Arctic, where special meteorological conditions and very low air
temperatures accelerate and enhance the destruction of ozone loss by manmade ozone depleting chemicals (ODCs).
The ozone layer is not really a layer at all, but has become known as such
because most ozone particles are scattered between 19 and 30 kilometers
(12 to 30 miles) up in the Earth's atmosphere, in a region called the
stratosphere. The concentration of ozone in the ozone layer is usually under
10 parts ozone per million [5]. Without the ozone layer, a lot of ultraviolet
(UV) radiation from the Sun would not be stopped reaching the Earth's
surface, causing untold damage to most living species. In the 1970s,
scientists discovered that chlorofluorocarbons (CFCs) could destroy ozone in
the stratosphere. Ozone is created in the stratosphere when UV radiation
from the Sun strikes molecules of oxygen (O2) and causes the two oxygen

atoms to split apart. If a freed atom bumps into another O2, it joins up,
forming ozone (O3). This process is known as photolysis.
Ozone is also naturally broken down in the stratosphere by sunlight and by a
chemical reaction with various compounds containing nitrogen, hydrogen
and chlorine. These chemicals all occur naturally in the atmosphere in very
small amounts. In an unpolluted atmosphere there is a balance between the
amount of ozone being produced and the amount of ozone being destroyed.
As a result, the total concentration of ozone in the stratosphere remains
relatively constant. At different temperatures and pressures (i.e. varying
altitudes within the stratosphere), there are different formation and
destruction rates. Thus, the amount of ozone within the stratosphere varies
according to altitude. Ozone concentrations are highest between 19 and 23
km [6].
Most of the ozone in the stratosphere is formed over the equator where the
level of sunshine striking the Earth is greatest. It is transported by winds
towards higher latitudes. Consequently, the amount of stratospheric ozone
above a location on the Earth varies naturally with latitude, season, and from
day-to-day. Under normal circumstances highest ozone values are found over
the Canadian Arctic and Siberia, whilst the lowest values are found around
the equator. The ozone layer over Canada is normally thicker in winter and
early spring, varying naturally by about 25% between January and July.
Weather conditions can also cause considerable daily variations.
With so much worry about the rapid ozone depletion taking place in various
parts of the earth, Indian scientists are closely monitoring the ozone layer
over India for possible depletion trends. Opinions are many and varied.
According to S K Srivastava, head of the National Ozone Centre in New Delhi,
there is no trend to show total ozone depletion over India. V.Thaphyal and S
M Kulshresta of the Indian Meteorological Department also point out that for
the period 1956 to 1986 "ozone measurements exhibit year to year
variability, but do not show any increasing or decreasing trend over India."
However, former director of the National Ozone Centre, K Chatterji, now with
Development Alternatives, warns that there is no case for complacency. He
asserts that his calculations exhibit an ozone depletion trend in the upper,
layers of the stratosphere over New Delhi and Pune from 1980 to 1983 in the
month of October when the Antarctic ozone hole is at its maximum. Since
India already receives high doses of ultraviolet (UV-B) radiation, and is at the
threshold go to speak, effects of ozone layer depletion could he far more
disastrous in India. A P Mitra, former director general of the Council of

Scientific and Industrial Research, clarifies that while there is no trend in the
total ozone value, there is some evidence of ozone depletion at higher
altitudes - at about 30 to 40 km - even over the tropics. He argues, however,
that there is insufficient data and that the depletion may be due to solar
cycles and other natural phenomena.
However, the effects of CFCs and belong cannot be ruled out. Total column
ozone data has been recorded over India for a long time. A network of
stations using Dobson spectrophotometers to mea- sure total ozone, some
six times a day, covers Srinagar, New Delhi, Varanasi, Ahmedabad, Pone and
Kodaikanal. Ozone profiles are also regularly recorded using balloons. Ozone
levels are the lowest during November and December and the highest in
summer. Across the country, variations do exist. In Kodaikanal, the total
ozone is 240 to 280 Dobson units (DU), in New Delhi 270 to 320 DU and in
Srinagar 290 to 360 DU. One Dobson unit is the equivalent of 0.01 mm of
compressed gas at a pressure of 760 rare mercury and 0C.B N Srivastava of
the National Physical Laboratory, who been working on incident UVradiation
levels, says that during summer, at noon, the UV-B radiation with a
wavelength of 290 nanometer (nm) is equivalent to levels attained in the
Antarctica during the ozone hole period.
He warns that even a slight depletion of the ozone layer over India may lead
to large percentage changes in UV-B radiation over the country. According to
eminent skin specialists in New Delhi, the incidence of skin cancer in India is
low, but they admit that the surveys conducted to identify any trends are
inadequate. Controlled studies to observe the effects of changing UV- B
radiation concentrations on crops are on, they said. However no field surveys
have been done in the country as yet.
The most common stratospheric ozone measurement unit is the Dobson Unit
(DU). The Dobson Unit is named after the atmospheric ozone pioneer G.M.B.
Dobson who carried out the earliest studies on ozone in the atmosphere from
the 1920s to the 1970s. A Dobson Unit measures the total amount of ozone
in an overhead column of the atmosphere. Dobson Units are measured by
how thick the layer of ozone would be if it were compressed into one layer at
0 degrees Celsius and with a pressure of one atmosphere above it.
Every 0.01 International Journal of Environmental Science and Development,
Vol.2, No.1, February 2011 ISSN: 2010-0264 32 millimeter thickness of the
layer is equal to one Dobson Unit [8]. The average amount of ozone in the
stratosphere across the globe is about 300 DU (or a thickness of only 3mm at

0C and 1 atmospheric pressure!). Highest levels of ozone are usually found


in the mid to high latitudes, in Canada and Siberia (360DU). When
stratospheric ozone falls below 200 DU this is considered low enough to
represent the beginnings of an ozone hole. Ozone holes of course commonly
form during springtime above Antarctica, and to a lesser extent the Arctic.
November the 2009 ozone hole is now waning, with much of the continent
experiencing a stratospheric spring warming. The residual vortex is over the
Weddell Sea and Antarctic Peninsula and here minimum values are around
160 DU and depletion exceeds 50%. Ozone values outside the polar vortex
have dropped to near 400 DU, and inside the vortex ozone values are
increasing as the atmosphere warms [7]. The temperature of the ozone layer
over Antarctica is now rising, though a small area is still cold enough for
polar stratospheric clouds (PSCs) to exist. During the early winter, the polar
vortex was often rather more elliptical than it was in 2008, and this lead to
some early depletion in circumpolar regions as stratospheric clouds became
exposed to sunlight.
It reverted to a more circular circulation as winter progressed and this led to
another relatively slow start to the growth of the ozone hole (as measured by
NASA/SBUV2), with the "hole" not beginning until mid August. The vortex
became more elliptical again in late August, with South Georgia being
affected by the fringes of the ozone hole between September 2 and 6. The
hole grew to reach an area of around 24 million square kilometers by mid
September, but had declined to 12 million square kilometres by mid
November. It is now a little larger than the average for the past decade. The
tip of South America and South Georgia were affected by the fringes of the
ozone hole from September 24 to September 30 and again from October 3 to
October 7.
The ozone depletion caused by human-produced chlorine and bromine
compounds is expected to gradually disappear by about the middle of the
21st century as these compounds are slowly removed from the stratosphere
by natural processes. This environmental achievement is due to the
landmark international agreement to control the production and use of
ozone-depleting substances. Full compliance would be required to achieve
this expected recovery. Without the Montreal Protocol and its Amendments,
continuing use of chlorofluorocarbons (CFCs) and other ozone-depleting
substances would have increased the stratospheric abundances of chlorine
and bromine tenfold by the mid-2050s compared with the 1980 amounts [9].
Such high chlorine and bromine abundances would have caused very large

ozone losses, which would have been far larger than the depletion observed
at present. In contrast, under the current international agreements that are
now reducing the human-caused emissions of ozone-depleting gases, the net
troposphere concentrations of chlorine- and bromine-containing compounds
started to decrease in 1995.
Because 3 to 6 years are required for the mixing from the troposphere to the
stratosphere, the stratospheric abundances of chlorine are starting to reach
a constant level and will slowly decline thereafter. With full compliance, the
international agreements will eventually eliminate most of the emissions of
the major ozone-depleting gases. All other things being constant, the ozone
layer would be expected to return to a normal state during the middle of the
next century. This slow recovery, as compared with the relatively rapid onset
of the ozone depletion due to CFC and bromine-containing halons emissions,
is related primarily to the time required for natural processes to eliminate the
CFCs and halons from the atmosphere. Most of the CFCs and halons have
atmospheric residence times of about 50 to several hundred years [7].
Ozone depletion occurs when the natural balance between the production
and destruction of stratospheric ozone is tipped in favour of destruction.
Although natural phenomena can cause temporary ozone loss, chlorine and
bromine released from man-made compounds such as CFCs are now
accepted as the main cause of this depletion [10]. It was first suggested by
Drs. M. Molina and S. Rowland in 1974 that a man-made group of compounds
known as the chlorofluorocarbons (CFCs) were likely to be the main source of
ozone depletion. However, this idea was not taken International Journal of
Environmental Science and Development, Vol.2, No.1, February 2011 ISSN:
2010-0264 33 seriously until the discovery of the ozone hole over Antarctica
in 1985 by the Survey. Chlorofluorocarbons are not "washed" back to Earth
by rain or destroyed in reactions with other chemicals. They simply do not
break down in the lower atmosphere and they can remain in the atmosphere
from 20 to 120 years or more.
As a consequence of their relative stability, CFCs are instead transported
into the stratosphere where they are eventually broken down by ultraviolet
(UV) rays from the Sun, releasing free chlorine. The chlorine becomes
actively involved in the process of destruction of ozone. The net result is that
two molecules of ozone are replaced by three of molecular oxygen, leaving
the chlorine free to repeat the process: Cl + O3 = ClO + O2 ClO + O = Cl +
O2 Ozone is converted to oxygen, leaving the chlorine atom free to repeat
the process up to 100,000 times, resulting in a reduced level of ozone.

Bromine compounds, or halons, can also destroy stratospheric ozone.


Compounds containing chlorine and bromine from man-made compounds are
known as industrial halocarbons. Emissions of CFCs have accounted for
roughly 80% of total stratospheric ozone depletion. Thankfully, the
developed world has phased out the use of CFCs in response to international
agreements to protect the ozone layer. However, because CFCs remain in the
atmosphere so long, the ozone layer will not fully repair itself until at least
the middle of the 21st century. Naturally occurring chlorine has the same
effect on the ozone layer, but has a shorter life span in the atmosphere.
Chlorofluorocarbons or CFCs (also known as Freon) are non-toxic, nonflammable and non-carcinogenic. They contain fluorine atoms, carbon atoms
and chlorine atoms. The 5 main CFCs include CFC-11 (trichlorofluoromethane
- CFCl3), CFC-12 (dichloro-difluoromethane - CF2Cl2), CFC-113 (trichlorotrifluoroethane - C2F3Cl3), CFC-114 (dichloro-tetrfluoroethane - C2F4Cl2),
and CFC-115 (chloropentafluoroethane - C2F5Cl).CFCs are widely used as
coolants in refrigeration and air conditioners, as solvents in cleaners,
particularly for electronic circuit boards, as a blowing agents in the
production of foam (for example fire extinguishers), and as propellants in
aerosols.
Indeed, much of the modern lifestyle of the second half of the 20th century
had been made possible by the use of CFCs.Man-made CFCs however, are
the main cause of stratospheric ozone depletion [11]. CFCs have a lifetime in
the atmosphere of about 20 to 100 years, and consequently one free chlorine
atom from a CFC molecule can do a lot of damage, destroying ozone
molecules for a long time. Although emissions of CFCs around the developed
world have largely ceased due to international control agreements, the
damage to the stratospheric ozone layer will continue well into the 21st
century.
The global market for rocket launches may require more stringent regulation
in order to prevent significant damage to Earths stratospheric ozone layer in
the decades to come, according to a new study by researchers in California
and Colorado. Future ozone losses from unregulated rocket launches will
eventually exceed ozone losses due to chlorofluorocarbons, or CFCs, which
stimulated the 1987 Montreal Protocol banning ozone-depleting chemicals,
said Martin Ross, chief study author from The Aerospace Corporation in Los
Angeles. The study, which includes the University of Colorado at Boulder and
Embry-Riddle Aeronautical University, provides a market analysis for
estimating future ozone layer depletion based on the expected growth of the

space industry and known impacts of rocket launches. As the rocket launch
market grows, so will ozone-destroying rocket emissions, said Professor
Darin Toohey of CU-Boulders atmospheric and oceanic sciences department.
If left unregulated, rocket launches by the year 2050 could result in more
ozone destruction than was ever realized by CFCs.Since some proposed
space efforts would require frequent launches of large rockets over extended
periods, the new study was designed to bring attention to the issue in hopes
of sparking additional research, said Ross. In the policy world uncertainty
often leads to unnecessary regulation, he said. We are suggesting this
could be avoided with a more robust understanding of how rockets affect the
ozone layer. Current global rocket launches deplete the ozone layer by no
more than a few hundredths of 1 percent annually, said Toohey. But as the
space industry grows and other ozone-depleting chemicals decline in the
Earths stratosphere, the issue of ozone depletion from rocket launches is
expected to move to the forefront. Highly reactive trace-gas molecules
known as radicals dominate stratospheric ozone destruction, and a single
radical in the stratosphere can destroy up to 10,000 ozone molecules before
being deactivated and removed from the stratosphere.

In 1974, after millions of tons of CFCs had been manufactured and sold;
chemists F. Sherwood Rowland and Mario Molina of the University of
California began to wonder where all these CFCs ended up. Rowland and
Molina theorized that ultraviolet (UV) rays from the Sun would break up CFCs
in the stratosphere, and that the free chlorine atoms would then enter into a
chain reaction, destroying ozone[12]. Many people, however, remained
unconvinced of the danger until the mid-1980s, when a severe springtime
depletion of ozone was first monitored by the British Antarctic Survey
International Journal of Environmental Science and Development, Vol.2, No.1,
February 2011 ISSN: 2010-0264 34 above Antarctica. The depletion above
the South Pole was so severe that the British geophysicist, Joe Farman, who
first measured it, assumed his spectrophotometer must be broken and sent
the device back to England to be repaired. Once the depletion was verified, it
came to be known throughout the world through a series of NASA satellite
photos as the Antarctic Ozone Hole. Laboratory studies backed by satellite
and ground-based measurements, show that free chlorine reacts very rapidly
with ozone.

They also show that the chlorine oxide formed in that reaction undergoes
further processes that regenerate the original chlorine, allowing the
sequence to be repeated up to 100,000 times. This process is known as a
"chain reaction". Similar reactions also take place between bromine and
ozone. Observations of the Antarctic ozone hole have given a convincing and
unmistakable demonstration of these processes. Scientists have repeatedly
observed a large number of chemical species over Antarctica since 1986.
Among the chemicals measured were ozone and chlorine monoxide, which is
the reactive chemical identified in the laboratory as one of the participants in
the ozone-destroying chain reactions. The satellite maps shown in the figure
below relate the accumulation of chlorine monoxide observed over Antarctica
and the subsequent ozone depletion that occurs rapidly in a few days over
very similar areas.
A University of Waterloo scientist says that an observed cyclic hole in the
ozone layer provides proof of a new ozone depletion theory involving cosmic
rays, a theory outlined in his new study, just published in Physical Review
Letters. Qing-Bin Lu, a professor of physics and astronomy and an ozone
depletion expert, said it was generally accepted for more than two decades
that the Earth's ozone layer is depleted by chlorine atoms produced by the
sun's ultraviolet light-induced destruction of chlorofluorocarbons (CFCs) in
the atmosphere. But mounting evidence supports a new theory that says
cosmic rays, rather than the sun's UV light, play the dominant role in
breaking down ozone-depleting molecules and then ozone. Cosmic rays are
energy particles originating in space. Ozone is a gas mostly concentrated in
the ozone layer, a region located in the stratosphere several miles above the
Earth's surface. It absorbs almost all of the sun's high-frequency ultraviolet
light, which is potentially damaging to life and causes such diseases as skin
cancer and cataracts. The Antarctic ozone hole is larger than the size of
North America. In his study, Lu analyzes reliable cosmic ray and ozone data
in the period of 1980-2007, which cover two full 11-year solar cycles.
The data unambiguously show the time correlations between cosmic ray
intensity and global ozone depletion, as well as between cosmic ray intensity
and the ozone hole over the South Pole. The Schwabe solar cycle or
Schwabe-Wolf cycle is the eleven-year cycle of solar activity of the sun. It
was named after Samuel Heinrich Schwabe (October 25, 1789 April 11,
1875) a German astronomer remembered for his work on sunspots. At
periods of highest activity, known as solar maximum or solar max, sunspots
appear. Periods of lowest activity are known as solar minimum. The last solar
maximum was in 2001. The solar cycle is not strictly 11 years; it has been as

short as 9 years and as long as 14 years in recent years. Fig.6 based on


cosmic theory ozone layer thickness variation This finding not only provides
a fingerprint for the dominant role of the cosmic-ray mechanism in causing
the ozone hole, but also contradicts the widely-accepted photochemical
theory," Lu said.
"These observations cannot be explained by that photochemical model.
Instead, they force one to conclude that the cosmic ray mechanism plays the
dominant role in causing the hole. His study quantitatively predicted that the
mean total ozone in the October hole over Antarctica would be depleted to
around 187 Dobson units (DU). The latest NASA OMI satellite data sets,
released on March 13, show that the mean total ozone in the ozone hole in
October 2008 was 197 DU, within five per cent of Lu's prediction. "The total
ozone values in the ozone hole in November and December nearly reached
the minimum values in the months on record," Lu said. "The 2008 ozone hole
shrank quite slowly and persisted until the end of December, making it one
of the longest lasting ozone holes on record." He added that in earlier studies
he and former colleagues found a strong spatial correlation between cosmic
ray intensity and ozone depletion, based on the data from several sources,
including NASA satellites.
"Lab measurements demonstrated a mechanism by which cosmic rays can
cause drastic reactions of ozone-depleting halogens inside polar clouds."
Cosmic rays are concentrated over the North and South Poles due to Earth's
magnetic field, and have the highest electron-production rate at the height of
15 to 18 km above the ground -- where the ozone layer has been most
depleted. Lu says that years ago atmospheric scientists expressed doubts
about the cosmic ray mechanism, but now observed data shows which
theory is the correct one. For instance, the most recent scientific
assessments of ozone depletion by the World Meteorological Organization
and the United Nations Environment Program using photochemical models
predicted that global ozone will recover (or increase) by one to 2.5 per cent
between 2000 and 2020 and that the Antarctic springtime ozone hole will
shrink by five to 10 per cent between 2000 and 2020. In sharp contrast, the
cosmic ray theory predicted one of the severest ozone losses over the South
Pole in 2008-2009 and another large hole around 2019-2020.
IV. EFFECT OF OZONE LAYER DEPLETION A. Effects on Human and Animal
Health Increased penetration of solar UV-B radiation is likely to have
profound impact on human health with potential risks of eye diseases, skin
cancer and infectious diseases [6]. UV radiation is known to damage the

cornea and lens of the eye. Chronic exposure to UV-B could lead to cataract
of the cortical and posterior subcapsular forms. UV-B radiation can adversely
affect the immune system causing a number of infectious diseases. In light
skinned human populations, it is likely to develop nonmelanoma skin cancer
(NMSC). Experiments on animals show that UV exposure decreases the
immune response to skin cancers, infectious agents and other antigens B.
Effects on Terrestrial Plants It is a known fact that the physiological and
developmental processes of plants are affected by UV-B radiation. Scientists
believe that an increase in UV-B levels would necessitate using more UV-B
tolerant cultivar and breeding new tolerant ones in agriculture. In forests and
grasslands increased UV-B radiation is likely to result in changes in species
composition (mutation) thus altering the bio-diversity in different ecosystems
[9].
UV-B could also affect the plant community indirectly resulting in changes in
plant form, secondary metabolism, etc. These changes can have important
implications for plant competitive balance, plant pathogens and biogeochemical cycles. C. Effects on Aquatic Ecosystems While more than 30
percent of the worlds animal protein for human consumption comes from
the sea alone, it is feared that increased levels of UV exposure can have
adverse impacts on the productivity of aquatic systems. High levels of
exposure in tropics and subtropics may affect the distribution of
phytoplanktons which form the foundation of aquatic food webs. Reportedly
a recent study has indicated 6-12 percent reduction in phytoplankton
production in the marginal ice zone due to increases in UV-B. UV-B can also
cause damage to early development stages of fish, shrimp, crab, amphibians
and other animals, the most severe effects being decreased reproductive
capacity and impaired larval development.
D. Effects on Bio-geo-chemical Cycles Increased solar UV radiation could
affect terrestrial and aquatic bio-geo-chemical cycles thus altering both
sources and sinks of greenhouse and important trace gases, e.g. carbon
dioxide (CO2), carbon monoxide (CO), carbonyl sulphide (COS), etc. These
changes would contribute to biosphere-atmosphere feedbacks responsible
for the atmosphere build-up of these gases. Other effects of increased UV-B
radiation include: changes in the production and decomposition of plant
matter; reduction of primary production changes in the uptake and release of
important atmospheric gases; reduction of bacterioplankton growth in the
upper ocean; increased degradation of aquatic dissolved organic matter
(DOM), etc. Aquatic nitrogen cycling can be affected by enhanced UV-B

through inhibition of nitrifying bacteria and photodecomposition of simple


inorganic species such as nitrate.
The marine sulphur cycle may also be affected resulting in possible changes
in the sea-to-air emissions of COS and dimethylsulfied (DMS), two gases that
are degraded to sulphate aerosols in the stratosphere and troposphere,
respectively. E. Effects on Air Quality Reduction of stratospheric ozone and
increased penetration of UV-B radiation result in higher photo dissociation
rates of key trace gases that control the chemical reactivity of the
troposphere. This can increase both production and destruction of ozone and
related oxidants such as hydrogen peroxide which are known to have
adverse effects on human health, terrestrial plants and outdoor materials.
Changes in the atmospheric concentrations of the hydroxyl radical (OH) may
change the atmospheric lifetimes of important gases such as methane and
substitutes of chlorofluoro carbons (CFCs). Increased troposphere reactivity
could also lead to increased production of particulates such as cloud
condensation nuclei from the oxidation and subsequent nucleation of sulphur
of both anthropogenic and natural origin (e.g. COS and DMS).
Effects on Materials An increased level of solar UV radiation is known to have
adverse effects on synthetic polymers, naturally occurring biopolymers and
some other materials of commercial interest. UV-B radiation accelerates the
photo degradation rates of these materials thus limiting their lifetimes.
Typical damages range from discoloration to loss of mechanical integrity.
Such a situation would eventually demand substitution of the affected
materials by more photo stable plastics and other materials in future. In
1974, two United States (US) scientists Mario Molina and F.
Sherwood Rowland at the University of California were struck by the
observation of Lovelock that the CFCs were present in the atmosphere all
over the world more or less evenly distributed by appreciable concentrations.
They suggested that these stable CFC molecules could drift slowly up to the
stratosphere where they may breakdown into chlorine atoms by energetic
UV-B and UB-C rays of the sun. The chlorine radicals thus produced can
undergo complex chemical reaction producing chlorine monoxide which can
attack an ozone molecule converting it into oxygen and in the process
regenerating the chlorine atom again.
Thus the ozone destroying effect is catalytic and a small amount of CFC
would be destroying large number of ozone molecules. Their basic theory
was then put to test by the National Aeronautic Space Authority (NASA)

scientists and found to be valid, ringing alarm bells in many countries and
laying the foundation for international action. G. Effects on Climate Change
Ozone depletion and climate change are linked in a number of ways, but
ozone depletion is not a major cause of climate change. Atmospheric ozone
has two effects on the temperature balance of the Earth. It absorbs solar
ultraviolet radiation, which heats the stratosphere. It also absorbs infrared
radiation emitted by the Earth's surface, effectively trapping heat in the
troposphere. Therefore, the climate impact of changes in ozone
concentrations varies with the altitude at which these ozone changes occur.
The major ozone losses that have been observed in the lower stratosphere
due to the human-produced chlorine- and bromine-containing gases have a
cooling effect on the Earth's surface. On the other hand, the ozone increases
that are estimated to have occurred in the troposphere because of surfacepollution gases have a warming effect on the Earth's surface, thereby
contributing to the "greenhouse" effect. In comparison to the effects of
changes in other atmospheric gases, the effects of both of these ozone
changes are difficult to calculate accurately. In the figure below, the upper
ranges of possible effects for the ozone changes are indicated by the open
bars, and the lower ranges are indicated by the solid bars.
The depletion of the ozone layer leads, on the average, to an increase in
ground-level ultraviolet radiation, because ozone is an effective absorber of
ultra-violet radiation. The Sun emits radiation over a wide range of energies,
with about 2% in the form of high-energy, ultraviolet (UV) radiation. Some of
this UV radiation (UV-B) is especially effective in causing damage to living
beings, the largest decreases in ozone during the past 15 years have been
observed over Antarctica, especially during each September and October
when the ozone hole forms. During the last several years, simultaneous
measurements of UV radiation and total ozone have been made at several
Antarctic stations. In the late spring, the biologically damaging ultraviolet
radiation in parts of the Antarctic continent can exceed that in San Diego,
California, where the Sun is much higher above the horizon. In areas where
smaller ozone depletion has been observed, UV-B increases are more difficult
to detect. In particular, detection of trends in UV-B radiation associated with
ozone decreases can be further complicated by changes in cloudiness, by
local pollution, and by difficulties in keeping the detection instrument in
precisely the same condition over many years. Prior to the late 1980s,
instruments with the necessary accuracy and stability for measurement of
small long-term trends in ground-level UV-B were not available. Therefore,
the data from urban locations with older, less-specialized instruments

provide much less reliable information, especially since simultaneous


measurements of changes in cloudiness or local pollution are not available.
When high-quality measurements have been made in other areas far from
major cities and their associated air pollution, decreases in ozone have
regularly been accompanied by increases in UV-B. This is shown in the figure
below, where clear-sky measurements performed at six different stations
demonstrate that ozone decreases lead to increased UV-B radiation at the
surface in amounts that are in good agreement with that expected from
calculations (the "model" curve).
IINTERNATIONAL ACTIONS The first international action to focus attention on
the dangers of ozone depletion in the stratosphere and its dangerous
consequences in the long run on life on earth was focused in 1977 when in a
meeting of 32 countries in Washington D.C. a World plan on action on Ozone
layer with UNEP as the coordinator was adopted. As experts began their
investigation, data piled up and in 1985 in an article published in the
prestigious science journal, Nature by Dr. Farman pointed out that although
there is overall depletion of the ozone layer all over the world, the most
severe depletion had taken place over the Antarctica. This is what is
famously called as "the Antarctica Ozone hole". His findings were confirmed
by Satellite observations and offered the first proof of severe ozone depletion
and stirred the scientific community to take urgent remedial actions in an
international convention held in Vienna on March 22, 1985. This resulted in
an international agreement in 1987 on specific measures to be taken in the
form of an international treaty known as the Montreal Protocol on Substances
That Deplete the Ozone Layer. Under this Protocol the first concrete step to
save the Ozone layer was taken by immediately agreeing to completely
phase out chlorofluorocarbons (CFC), Halons, Carbon tetrachloride (CTC) and
Methyl chloroform (MCF) as per a given schedule A.
Montreal Protocol In 1985 the Vienna Convention established mechanisms
for international co-operation in research into the ozone layer and the effects
of ozone depleting chemicals (ODCs). 1985 also marked the first discovery of
the Antarctic ozone hole. On the basis of the Vienna Convention, the
Montreal Protocol on Substances that Deplete the Ozone Layer was
negotiated and signed by 24 countries and by the European Economic
Community in September 1987. The Protocol called for the Parties to phase
down the use of CFCs, halons and other man-made ODCs. The Montreal
Protocol represented a landmark in the international environmentalist
movement. For the first time whole countries were legally bound to reducing

and eventually phasing out altogether the use of CFCs and other ODCs.
Failure to comply was accompanied by stiff penalties.
The original Protocol aimed to decrease the use of chemical compounds
destructive to ozone in the stratosphere by 50% by the year 1999. The
Protocol was supplemented by agreements made in London in 1990 and in
Copenhagen in 1992, where the same countries promised to stop using CFCs
and most of the other chemical compounds destructive to ozone by the end
of 1995. Fortunately, it has been fairly easy to develop and introduce
compounds and methods to replace CFC compounds.
In order to deal with the special difficulties experienced by developing
countries it was agreed that they would be given an extended period of
grace, so long as their use of CFCs did not grow significantly. China and India,
for example, are strongly increasing the use of air conditioning and cooling
devices. Using CFC compounds in these devices would be cheaper than using
replacement compounds harmless to ozone.
An international fund was therefore established to help these countries
introduce new and more environmentally friendly technologies and
chemicals. The depletion of the ozone layer is a worldwide problem which
does not respect the frontiers between different countries. It can only be
affected through determined international co-operation.
B. Australian Chlorofluorocarbon Management Strategy It provides a
framework for the responsible management and use of CFCs in Australia. The
strategy recognizes some continuing need for these chemicals in
pharmaceutical and laboratory uses, but commits to their gradual phasing
out. C. Environmental Protection (Ozone Protection) Policy2000 This WA
policy aims to minimize the discharge of ozone-depleting substances into the
environment, and has been extended to cover use of alternative refrigerants
(where relevant). This has been done to prevent current stocks of ozonedepleting substances from being released to the.

Sprinkled throughout the atmosphere are pale blue molecules of a toxic gas
that are essential to most life on Earth. This gas is ozone. Ozone is toxic
because it is highly reactive. This is why it can sterilize drinking water,
eliminate odors, bleach colors, and decompose rubber. Fortunately, the
amount of ozone at ground level is usually too low for these effects to be
observed.

However, high concentrations of various air pollutants and sunlight can


increase ozone levels near the ground from a few tens of molecules per
billion molecules of air (ppb) to a few hundred ppb. These levels of ozone can
damage plants, cause eye irritation, inflame mucous membranes and impair
the performance of athletes. Ozone is essential to life because it shields the
Earth from the damaging even lethal, ultraviolet radiation emitted by the
sun. This filtering ability is particularly remarkable when you consider the
relative scarcity of ozone molecules. For every billion molecules in the
atmosphere, only around 300 are ozone. Imagine you could poke a tube
through the entire atmosphere over your head and bring all the ozone
molecules in the tube down to the surface. If they were then subjected to the
same temperature and atmospheric pressure (standard temperature and
pressure or STP) as you are, they would form a layer only about 3-millimeters
thick. Although they may be formed in many different ways, all the ozone
molecules in the stratosphere are identical. An oxygen molecule (O2 ) is
composed of two oxygen atoms (O).
Ultraviolet radiation can split an O2 molecule and leave behind two free O
atoms. Various chemical reactions leave O atoms as a byproduct. In either
case, the free O atom can merge with an O2 molecule to form triatomic
oxygen (O3 ), more commonly known as ozone. The Two Ozone Layers The
term ozone layer generally refers to a relatively high concentration of
ozone in the stratosphere, a layer of very dry air around 15 to 35 kilometers
(9 to 22 miles) above the Earths surface. However, about 10 percent of the
total ozone is found in the troposphere, the lowest portion of the
atmosphere.
The troposphere contains 90 percent of the atmosphere and nearly all of the
atmospheres water vapor. Storms form in the troposphere and usually stay
there. But the tops of especially powerful thunderstorms occasionally poke
into the lower stratosphere. The tropopause, the border between the
troposphere and the stratosphere, ranges from around 10 to 15 kilometers
above the surface. The heat of summer increases the height of the
tropopause; the height is reduced in winter.Other things being equal, when
the tropopause is low, the amount of stratospheric ozone is high and vice
versa.
The ozone between the surface and the tropopause forms only a fraction of
the ozone over most locations. Nevertheless, it absorbs solar UV more
efficiently than an equal amount of stratospheric ozone. This is because
scattering caused by dust and aerosols increases the distance that rays of

sunlight travel on their way to the surface. In spite of this benefit,


tropospheric ozone is often referred to as bad ozone because of its adverse
effects in high concentrations. If the same ozone were somehow to drift into
the stratosphere, it would be called good ozone. Forming and Destroying
Tropospheric Ozone Tropospheric ozone is produced in many ways. Some is
formed by lightning or by UV radiation from the sun. Most is formed by
chemical reactions which take place in the presence of sunlight. One such
reaction is the conversion of nitrogen dioxide (NO2 ) into nitrogen oxide (NO)
in the presence of solar UV. The O atom left over from this reaction combines
with an O2 molecule to form an ozone molecule. Various other gases in the
atmosphere can combine with NO to form more NO2 , which then can cause
a buildup of ozone. Gaseous organic chemicals, in the presence of nitrogen
oxides and sunlight, can also contribute to ozone production. The same
photochemistry that forms ozone can also destroy it. Indeed, some
photochemical processes in the atmosphere are called do-nothing
reactions since they destroy as much ozone as they create. Ultraviolet
radiation leaking through the ozone high in the stratosphere can also create
and destroy ozone in the troposphere.
There are several ways to measure tropospheric ozone. Since ozone is a
strong oxidizer, it changes the color of some chemical compounds and
solutions. For example, paper soaked in a mixture of starch and potassium
iodide will change color when exposed to ozone. The reaction of ozone with
various chemicals, gases, and even some lubricating oils causes a faint
luminescence that can be detected by a sensitive photomultiplier tube.
Detection systems such as this are known as chemiluminescence detectors.
Since ozone absorbs ultraviolet radiation so effectively, many kinds of ozone
detectors incorporate a UV lamp and a detector. Air is passed through a
chamber, and any attenuation is assumed to have been caused by ozone. A
problem with this method is that attenuation can also be caused by dust.
Therefore, its common practice to use two chambers, one of which receives
air from which any ozone has been scrubbed by a catalytic converter.
Alternatively, scrubbed and unscrubbed air can be passed in sequence
through the same chamber. Either way, the error caused by dust and other
contaminants in the ozone-free sample can then be determined by
subtraction. Sulfur dioxide and other chemicals can interfere with the
chemical and UV detection of ozone. When scientists at the Montsouris
observatory near Paris became aware of this problem in 1905, they built a
second chemical ozone detector. The air inlet for the new detector was fitted
with a 4-meter (13-feet) hose of natural rubber, which completely destroyed

any ozone passing through it. In this way any errors in the original detector
caused by gases other than ozone could be eliminated. Several times my son
Eric and I have measured the amount of tropospheric ozone between the
bottom and top of mountains in New Mexico. We do this by measuring the
total amount of ozone in the atmosphere with a UVsensitive instrument that
is pointed at the sun. One of us goes to the top of a mountain while the other
stays at the base. We then make a series of observations at prearranged
times. Later, we subtract the ozone measured at the mountaintop from the
measurements made at the base to determine the amount of ozone in
between. So far our results (a few Dobson Units per vertical kilometer; see
Figure 5) have agreed with measurements made from balloons by the
National Oceanic and Atmospheric Administration (NOAA).
Earth is rarely constant. Consider the diurnal or daily ozone cycle over
Albuquerque, New Mexico. Early on a July morning at the base of the Sandia
Mountain aerial tramway just northeast of the city, ground-level ozone
concentration might be, say, 20-30 ppb. As the sun rises high in the sky,
photochemical ozone production increases, especially when the wind is from
the southwest and the clean mountain air is spiked with nitrogen oxides and
hydrocarbons from automobile exhaust. Although little or no photochemical
smog may be visible, the ozone concentrations might reach 40-60 ppb by
late afternoon.
As the sun sinks behind the volcano cinder cones west of Albuquerque, the
ozone level also falls. Late that evening, the ozone returns to its normal
background level. Tropospheric Ozone Trends and Effects The ozone
measured at the ground near Paris, France, from 1876-86 was only around a
third to a half of what is usual in unpolluted areas today. The increase since
then is generally believed to be caused by human activities. At least twothirds of the nitrogen oxides are believed to come from the burning of fossil
fuels, wood, forests and agricultural wastes. Nitrogen oxides are also
produced naturally by lightning, forest fires, and soil. Organic chemicals,
such as methane and hydrocarbons, can be byproducts of plants, animals,
and human activity.
It's interesting to compare the amount of ozone in a vertical column of air
adjacent to a mountain with that over a city. In the summer of 1989, Eric and
I measured 5.8 DU (Dobson Units) of ozone between the base and crest of
Sandia Mountain, an altitude difference of 1,164 meters (3,819 feet). A few
weeks earlier I had measured about 5 DU of ozone between street level and
an observation deck atop the 110th floor of the World Trade Center in New

York City (420 meters or 1,377 feet). There was much more ozone in the air
above New York City than in the air near Sandia Mountain. Surprisingly,
however, much of the air at street level has less ozone than you might
expect. A government official who makes ozone measurements at fixed sites
told me that ozone levels can be much higher in the air over New York and in
the surrounding regions than at ground level in the city. Apparently the high
number of air pollutants, people, rubber tires, and the like, suppresses ozone
concentrations at street level. When California forced a significant reduction
of hydrocarbon emissions from automobiles, the ozone in downtown Los
Angeles fell. Ozone levels downwind, however, continued to rise. The
scientists who puzzled over this dilemma noticed that there is considerably
more vegetation downwind from the central city.
They concluded that ozone is reduced only when both hydro-carbons and
nitrogen oxides are reduced. The relationship of trees and ozone is
particularly interesting. Too much ozone can damage or even kill trees.
Ironically, trees emit hydrocarbons that participate in chemical reactions that
produce ozone. Several years ago William Chameides of the Georgia Institute
of Technology studied satellite images of Atlanta and found that 57 percent
of the city was wooded. He and his co-workers concluded that Atlantas trees
emitted at least as many hydrocarbons as the citys cars, trucks, buses and
factories.
Stratospheric Ozone Most references to the ozone layer mean the ozone
found in the stratosphere. There it forms a vaporous shield that protects life
on Earth from the lethal effects of the suns UV radiation. If youve flown in
the Concorde, then you have probably travelled through the bottom of the
stratospheric ozone layer. Forming and Destroying Stratospheric Ozone
Ozone in the stratosphere is formed by a natural photochemical process
when ultraviolet radiation from the sun splits molecules of oxygen into the
individual oxygen atoms from which they are formed.
The free O atoms soon react with O2 molecules to form O3 . This process
works both ways: O3 molecules that are unlucky enough to be struck by UV
radiation are split back to an O2 molecule and a free O atom. The free O
atom can merge with an O2 molecule to once again form an O3 molecule.
Ozone molecules that drift lower down in the stratosphere are protected from
the destructive effects of UV radiation by the overlying blanket of ozone
molecules. But even molecules of ozone deep in the ozone layer are not
entirely safe, for they can be destroyed by reactions involving sunlight and
oxides of nitrogen, hydrogen, chlorine, and bromine.

Although all these chemicals can arise from natural sources, they can also
arise from human activity. For example, manufactured chlorofluorocarbons
(CFCs) have been a concern since 1970 when James E. Lovelock detected
their presence in air. In 1974, Sherwood Rowland and Mario Molina proposed
that CFC molecules could eventually drift into the stratosphere, where UV
radiation would break them down into chlorine monoxide (ClO) and other
ozone-destroying compounds. In recent years, evidence has been
accumulating that CFCs may indeed be causing a gradual reduction of ozone,
particularly in regions near the poles during early spring. The question now
being researched is how much ozone they might ultimately destroy.
Fortunately solar UV is constantly creating new ozone. Therefore, CFCs
cannot destroy all the ozone.
Ozone in the stratosphere can be measured directly using instruments on
aircraft, rockets, andespeciallyballoons. Many of the same kinds of sensing
systems used for measuring ozone at the surface have been modified for
these roles. Thanks to ozones well-known ability to absorb ultraviolet
radiation, the total amount of ozone (troposphere plus stratosphere) can be
measured indirectly from the surface or from space. Several kinds of optical
instruments have been developed for measuring ozone from the surface,
including the Dobson spectrophotometer and various instruments that use
filters or diffraction gratings to measure narrow bands of ultraviolet. The
Dobson spectrophotometer plays a key role in ground-based ozone
monitoring efforts. Invented in the late 1920s by G. M. B. Dobson, this
instrument divides sunlight into a spectrum with a prism and measures the
ratio of two UV wavelengths about 20 nanometers (nm) apart. Dust and
aerosols can cause errors in ozone observations by scattering one
wavelength more than another. Dobson observations are usually made at
two pairs of wavelengths to cancel out this error. The Dobson instrument is
expensive, nearly 2 meters (6 ft) long, and heavyabout 40 kilograms (85
pounds). Since it measures the ratio of two or more ultraviolet signals, it
provides no informa tion about the amount of solar ultraviolet. The Brewer
ozonometer uses a diffraction grating to separate the suns ultraviolet
wavelengths.
(A diffraction grating consists of several parallel grooves that split light up
into several wavelengths. A compact disk produces an effect much like a
diffraction gratingexcept that the rainbow colors are produced by parallel
rows of pits instead of grooves.) This expensive instrument is smaller than
the Dobson and wellsuited for automated data taking. It also measures sulfur
dioxide, a gas that can interfere with ozone measurements. Some scientists

believe that the Brewer measures ozone with higher precision than the
Dobson. Indeed, Canada has switched from Dobsons to Brewers. A third kind
of instrument uses optical filters to measure two or more UV wavelengths.
Such instruments are cheaper, smaller, and easier to use than the Dobson
and the Brewer. They also provide information about the suns direct
ultraviolet.
For these reasons, I selected the filter approach when designing an
instrument to measure total ozone several years ago. Some sensors on
satellites can measure the amount of ozone at various altitudes by observing
the sun as it rises and sets through the atmosphere above the Earths limb
(an astronomical term referring to the edge of a planetary bodys disk). Other
satellite sensors measure the amount of the suns ultraviolet that is
scattered back into space from the atmosphere below. Since these
wavelengths are absorbed by ozone, processing the backscatter from one or
more pairs of wavelengths permits one to estimate the amount of ozone
between the satellite and the ground.
Its important to note the latter kind of instrument cannot measure all the
ozone between the top of the atmosphere and the surface. Clouds can get in
the way, and little or none of the UV backscattered from the lowest few
kilometers above the surface can penetrate back through the ozone layer. An
estimate of lower tropospheric ozone is added to the satellite ozone equation
to correct this problem. The estimate is based on measurements made from
the ground during annual calibration checks and measurements made by
balloon sensors. Stratospheric Ozone Cycles Superimposed on the daily
ozone cycle near the ground are seasonal changes in the amount of
stratospheric ozone. In the northern hemisphere, the total amount of ozone
is lowest during winter. The amount of ozone begins to rise rapidly during
spring and gradually diminishes in the summer and fall.
This gradual seasonal variation in ozone is marked by sharp spikes and dips
associated with weather systems. Passage of a cold front, for example, may
cause the amount of ozone to increase 20 percent or more for a day or two.
A warm front may cause an comparable decrease. Meterologists refer to
regions of diminished ozone as ozone minimums, and regions of high ozone
as ozone maximums.
Ill never forget the giant ozone maximum that passed over South Texas on
March 16, 1990 (See Figure 6, page 8). That morning I made a few ozone
observations around 11:00 a.m. and was surprised to find the highest

amount of ozone I had ever measured, around 360 DU. Since the ozone
amount kept climbing as noon approached, I assumed something was wrong
with the instrument. But both filters were clean, no wires were dangling in
front of them, and the battery was fresh. After 1:00 p.m. the ozone amount
climbed even faster than it had before noon. By 1:30 p.m. it was more than
440 DU and shortly before 2:00 p.m. it reached 460 DU. The ozone amount
then began a sharp slide to pre-noon levels. Several months later, data from
the TOMS (Total Ozone Mapping Spectrometer) instrument aboard the
Nimbus-7 satellite confirmed the extraordinarily high ozone levels of March
16. My son Eric wrote a Pascal program that transformed the TOMS data into
a color-coded ozone map of the United States. The map disclosed an
enormous tongue of high ozone reaching down from Canada and ending in
South Texas. A check of weather records revealed that this ozone maximum
was associated with a giant weather system that moved in from the Pacific
and crossed the United States over a 3-day period.
Recall that most ground observations of the ozone layer measure the total
amount of ozone in a column between the instrument and the top of the
atmosphere. Therefore, these measurements include the total amount of
both tropospheric and stratospheric ozone. Daily measurements of the
amount of total column ozone have been made at Arosa, Switzerland since
1926. The total amount of ozone back to 1912 has been determined by a
careful analysis of solar measurements made by the Smithsonian Institution.
Since 1957, more than 70 Dobson spectrophotometers have made regular
measurements of ozone. These measurements show that the total amount of
ozone varies in cycles that may last a decade or more. For example, during
the 1960s, scientists at NOAA found that ozone over North America
increased by about 5 percent.
Since 1970, however, ozone over the northern hemisphere has declined
around 5 percent. Since this decline is seen in both measurements from the
ground and from satellites, there is little disagreement that it is real. But
there is considerable disagreement concerning the reason for the decline,
and its significance. Some scientists believe the decline is primarily a
byproduct of natural meterological cycles, a changing climate and possibly
the solar cycle. They support their case by pointing to ozone cycles over the
past 50 or more years. Others believe the decline is in large part caused by
contamination of the atmosphere by pollutants that contribute to the
destruction of ozone particularly CFCs.

Some believe the eruption of major volcanoes such as El Chichon in 1982


and Pinatubo in 1991 exacerbate the problem. Still others believe ozone is
impacted both by natural meteorological trends and ozone-destroying
chemicals. These issues are being studied and debated by the scientists who
study ozone. Meanwhile, as has been widely reported in the press, there has
been considerable political fallout over the issue of declining ozone. Thats
one reason why more than 70 nations have signed the Montreal Protocol, an
agreement to eventually ban production of most CFCs. Because it is believed
that CFSs can remain in the atmosphere for decades, even a total ban will
not restore the ozone to its pre-1970 conditionif indeed CFCs are the
principal culprit. Instead, ozone will continue to decline at, perhaps, a few
percent or so per decade until CFCs are no longer present. Ultraviolet
Radiation If the spectral sensitivity of a honey bees eyes could somehow be
added to yours, rainbows would have an additional streak of color adjacent
to the violet band. This invisible band of black light is known as ultraviolet
radiation. Ultraviolet radiation is divided into three bands.
The wavelengths below 290 nm are referred to as UV-C. The wavelengths
between 290 and 320 nm are referred to as UV-B. And the wavelengths
between 320 and 340 nm are known as UV-A. The ozone layer blocks all UVC. The UVB that leaks through is what causes sunburn. The Ultraviolet Sky
The sky is blue because most of its molecules are just the right size to
scatter the blue wavelengths of sunlight. These molecules also scatter UV
wavelengths. This means that the entire daytime sky is a gigantic source of
UV-A and UV-B (remember that all the UV-C is absorbed by ozone). The UV
radiation from the sky can be described as direct, diffuse, or global. Direct
radiation is that which comes directly from the sun. Diffuse radiation is that
scattered from clouds and molecules of air. Global is the sum of direct and
diffuse UV radiation. Most natural materials reflect UV rather poorly. But
snow is an excellent UV reflector; and so is water. In short, exposed parts of
your body can receive a significant does of UV even when shaded from the
direct sun.
Ultraviolet Hazards The energy of electromagnetic radiation is inversely
related to its wavelength. In other words, radiation with short wavelengths
like xrayshas much higher energy than radiation with longer wavelengths,
such as visible light. For this reason, ultraviolet radiation is more energetic
than visible light. Thats why the suns UV-B radiation readily causes sunburn
while UV-A and visible sunlight do not, even though much more UV-A and
visible light reach the earth. (Of course, even visible sunlight will cause
sunburn if it is concentrated with a lens or reflector.)

The suns UV-B can also damage the chromosomes in human skin cells
which can eventually lead to skin cancer. Excessive UV-B can also cause
cataracts and alter the immune system. Plants can be damaged by excessive
UV-B. So can organisms that live at least part of their life cycle near the
surfaces of lakes, rivers and oceans. More research is needed to better
understand the nature of such damage and the amount and wavelengths of
UV that are responsible. Ultraviolet Benefits The public has been frequently
reminded about the dangers of UV-B exposure. Its important to realize,
however, that solar UV also plays an essential role in human health.
Perhaps the most important contribution of solar UV is its ability to stimulate
the production of vitamin D in the outer layers of the skin. This gives UV the
ability to prevent and to cure rickets and to maintain a healthy skeleton. Both
natural and artificial UV radiation are also used to treat psoriasis. Measuring
Ultraviolet Radiation If youve ever retrieved a rolled up newspaper which
has lain in the summer sun for a few hours, you know that newsprint darkens
when exposed to solar UV. So does freshly cut or sanded pine and other
woods. Colored construction paper and fabrics are bleached by solar UV. A
few years ago during a field trip to New Mexico, my son and I tacked a strip
of freshly sanded pine atop the crate bolted in the back of our pickup that
carried our instruments and supplies.
We covered the wood with a strip of tape, several centimeters of which we
removed each day. After 10 days on the UV-drenched highways of Texas and
New Mexico, the strip of wood was divided into 10 segments, each slightly
darker than the next. Some chemicals will also change color when exposed
to solar UV. But an electronic instrument is necessary to quantitatively
measure the intensity of UV. The two principle kinds of UV sensors used with
such instruments are phototubes and solid-state detectors. Specialized
phototubes known as photomultipliers can be made exceptionally sensitive
to UV. But they are expensive, fragile, and require a high operating voltage.
Solid-state detectors are not as sensitive.
In addition, they are are cheaper, sturdier, and much easier to use. An
optical filter can be placed between a UV detector and the sun to enable the
detector to respond only to specific regions of the UV spectrum.
Alternatively, a prism or grating can be used to select specific wavelengths.
Various methods are used to measure direct, diffuse and global UV. Direct UV
is easily measured by placing the detector inside a collimator tube that limits
its field of view. Ideally, the field of view of the collimator should not exceed
a few degrees. Global UV is measured by placing a diffuser plate (such as

ground silica or diffuse UV-transmitting plastic) over the detector and its
filter. Diffuse UV can be measured with a global detector by placing a small
disk so that its shadow completely covers the detector. This blocks direct
sunlight so that the detector signal is entirely from the diffuse sky.
Subtracting the diffuse amount from the global value gives the direct UV.

Ozone vs. Sunlight Ozone absorbs some of the suns infrared radiation. It
even weakly absorbs the wavelengths around 600 nm, which appear As
noted, its important to understand that the scattering caused by dust and
aerosols can cause ozone near the ground to absorb more UV than an equal
amount of ozone in the stratosphere. orange to the human eye. Its most
important absorption, of course, occurs at UV wavelengths below around 340
nm. The absorption increases so rapidly below 320 nm that little or no
measurable radiation below 295 nm reaches the surface at sea level. As
noted, its important to understand that the scattering caused by dust and
aerosols can cause ozone near the ground to absorb more UV than an equal
amount of ozone in the stratosphere. This helps explain why some scientific
studies have found a slight decrease in UV-B over the past decade, a time
during which the amount of stratospheric ozone has decreased and the
amount of tropospheric ozone has increased. Conclusion Evaluating claims
about the ozone layer requires basic knowledge not only about ozone and its
effects on ultraviolet, but about the research methods that have been used
to understand the ozone layers dynamics. Hopefully, this article has
provided some of that knowledgeand perhaps even sparked some interest
in doing some independent investigating. If so, be sure to check out some of
the resources mentioned on page 10 (Readings about Ozone and
Ultraviolet.) Also, be sure to read the other articles in this issue, on the
ozone hole and on the health effects of UV.
In recent years, much attention has been given to the ozone hole over
Antarctica. This phenomenon is observed each year in October during the
Antarctic spring. After several weeks, the Antarctic vortex, a whirling weather
system that encircles and isolates the South Pole during winter, breaks up
and ozone levels rapidly rise. In meterological terms, the Antarctic ozone
hole is a significant ozone minimum and not a literal hole through the
entire ozone layer. Nevertheless, for a brief time ozone levels within the hole
can plummet to 100 DU. (Normal levels are about 300 DU.) At the same
time, the ozone levels in a broad belt encircling the hole are the highest on
earth. Because various scientific studies have concluded that the ozone hole

is caused in part by chlorine believed to come from manufactured chemicals


especially CFCs some scientists, politicians, and government agencies have
sounded an alarm about the prospect of severe ozone depletion leading to
ozone holes elsewhere. In a widely publicized statement two years ago, thenSen. Albert Gore raised the possiblity that an ozone hole might appear over
Kennebunkport, Maine. Although a prominent NASA scientist discounted this
possibility, other scientists held a press conference to express alarm about
possible serious ozone depletion over the Arctic. Developments like these led
to many scary reports in the media.
Fortunately, the Antarctic hole is a phenomenon preceded by the very cold
temperatures and darkness found inside the winter Antarctic vortex, which is
much stronger than the Arctic vortex. Long before the ozone hole was
identified in 1985, G.M.B. Dobson, inventor of the Dobson
spectrophotometer, discovered something very different about Antarctic
ozone. In a paper titled Forty Years Research on Atmospheric Ozone at
Oxford: A History (Applied Optics, March 1968), Dobson described an ozone
monitoring program that began at Halley Bay, Antarctica, in 1956. When the
data began to arrive, the values in September and October 1956 were
about 150 [Dobson] units lower than expected. . . . In November the ozone
values suddenly jumped up to those expected. . . . It was not until a year
later, when the same type of annual variation was repeated, that we realized
that the early results were indeed correct and that Halley Bay showed a most
interesting difference from other parts of the world. The ozone decline
reported by Dobson was not nearly as severe as the one that characterizes
the Antarctic ozone hole today.
However, two scientists who reviewed old ozone records recently reported
that the ozone amount in the spring of 1958 fell to only 110 DU at the French
Antarctic Observatory at Dumont dUrville. In Annales Geophysicae
(November, 1990), P. Rigaud and B. Leroy observed that in 1958, the
Antarctic vortex, where the most significant ozone depletion occurs, was
centered over Dumont dUrville, on the opposite side of the South Pole from
Halley Bay. They reported that the concentration of CFCs in the atmosphere
in 1958 was much lower than it is today and concluded that natural
phenomena, such as volcanic aerosols in the stratosphere, may also lead to
ozone destruction. Writing in a recent issue of Science, however NASA
scientist Paul Newman has convincingly refuted Rigauds and Leroys ozone
measurement methods. He pointed out that Rigaud and Leroy relied on
photographic plates, an unreliable method for measuring stratospheric
ozone. Meanwhile, the chemistry and dynamics of the atmosphere inside the

Antarctic and Arctic vortices remain the subjects of extensive research using
various kinds of ground-based instruments, instrumented balloons, highflying aircraft, and satellites.
Ozone (O3) is a molecule made up of three atoms of oxygen (O), and is
mostly found in the stratosphere, where it protects us from the Suns harmful
ultraviolet (UV) radiation. Although it represents only a tiny fraction of the
atmosphere, ozone is crucial for life on Earth. Ozone in the stratospherea
layer of the atmosphere between 15 and 50 kilometers (10 and 31 miles)
above usacts as a shield to protect Earths surface from the suns harmful
ultraviolet radiation. Without ozone, the Suns intense UV radiation would
sterilize the Earths surface. With a weakening of this shield, more intense
UV-B and UV-A radiation exposure at the surface would lead to quicker
sunburns, skin cancer, and even reduced crop yields in plants. However, near
the surface where we live and breathe, ozone is a harmful pollutant that
causes damage to lung tissue and plants. This bad ozone forms when
sunlight initiates chemical reactions in the air involving pollutants,
particularly a family of gases called nitrogen oxides (released from vehicles
and industry during the combustion process) and with volatile organic
compounds (carbon-containing chemicals that evaporate easily into the air,
such as petroleum products).
There are natural processes that create and destroy ozone in the
stratosphere. These processes regulate a balance of ozone and form the
ozone layer. Ozone is created primarily by sunlight. When high-energy
ultraviolet rays (UV-C) strike an oxygen molecule (O2), they split the
molecule into two single oxygen atoms, known as atomic oxygen. A freed
oxygen atom then combines with another oxygen molecule to form a
molecule of ozone (O3). Because there is so much oxygen in our atmosphere,
this ozone-oxygen cycle is continuously absorbing high-energy ultraviolet
radiation (UV-C) and completely blocking it from reaching the Earths surface.
This process creates heat which warms the upper part of the stratosphere
Ozone is very reactive, and attacks other molecules in the air, often
regenerating oxygen in the process. Alsoand this is why ozone is important
to usozone in the stratosphere absorbs much of the suns UV-B rays,
splitting back into molecular and atomic oxygen. No matter how the oxygen
atoms are produced, they almost always quickly react with oxygen
molecules, reforming ozone. So, while ozone is continually being replenished,
it is also continually being destroyed. Sometimes an ozone molecule reacts
with an oxygen atom, creating two oxygen molecules, thus ending the cycle.
If the rate of ozone creation is equal to the rate of destruction, the total

amount will remain the same. This is like a leaky bucket: If you pour water
into the bucket at the same rate that its leaking out, the level of water in the
bucket will stay the same. Scientists have found that ozone levels change
periodically with regular natural cycles such as the changing seasons, winds,
and long time scale sun variations. Ozone also responds to some sporadic
solar events such as flares. Moreover, volcanic eruptions may inject
materials into the stratosphere that can lead to increased destruction of
ozone. In the 1970s, scientists suspected that reactions involving man-made
chlorine-containing compounds could upset this balance leading to lower
levels of ozone in the stratosphere. Think again of the leaky bucket. Putting
additional ozone-destroying compounds into the atmosphere is like
increasing the size of the holes in our bucket of ozone. The larger holes
cause ozone to leak out at a faster rate than ozone is being created.
Consequently, the level of ozone protecting us from ultraviolet radiation
decreases. The ozone destroyed by manmade emissions is comparable or
more than the amount destroyed by natural processes. Human production of
chlorine-containing chemicals, such as chlorofluorocarbons (CFCs), has
added an additional factor that destroys ozone. CFCs are molecules made up
of chlorine, fluorine and carbon. Because they are extremely stable
molecules, CFCs do not react with other chemicals in the lower atmosphere,
but exposure to ultraviolet radiation in the stratosphere breaks them apart,
releasing chlorine atoms. Free chlorine (Cl) atoms then react with ozone
molecules, taking one oxygen atom to form chlorine monoxide (ClO) and
leaving an oxygen molecule (O2).
If each chlorine atom released from a CFC molecule destroyed only one
ozone molecule, CFCs would pose very little threat to the ozone layer.
However, when a chlorine monoxide molecule encounters a free atom of
oxygen, the oxygen atom breaks up the chlorine monoxide, stealing the
oxygen atom and releasing the chlorine atom back into the stratosphere to
destroy another ozone molecule. These two reactions happen over and over
again, so that a single atom of chlorine, acting as a catalyst, destroys many
molecules (about 100,000) of ozone. Fortunately, chlorine atoms do not
remain in the stratosphere forever. Free chlorine atoms react with gases,
such as methane (CH4), and get bound up into hydrogen chloride (HCl)
molecules. These molecules eventually end up back in the troposphere
where they are washed away by rain. Therefore, if humans stop putting CFCs
and other ozone-destroying chemicals into the stratosphere, stratospheric
ozone will eventually return to its earlier, higher values.

Scientists have been measuring ozone since the 1920s using ground-based
instruments that look skyward. Data from these instruments, although useful
in learning about ozone, only tell us about the ozone above their sites, and
do not provide a picture of global ozone concentrations. To get a global view
of ozone concentrations and its distribution, scientists use data from
satellites. The principle of measuring ozone is simple. We know from
measurements how much incoming UV-B sunlight arrives at the top of the
Earths atmosphere every second. We also know how much light (including
UV-B) is scattered by air molecules in the atmosphere--this is called
Rayleigh scattering, and this is the same phenomenon that makes the sky
appear blue.
So we can calculate the amount of UV-B sunlight that would make it to the
Earths surface if there were no ozone to absorb it. When we measure the
amount of UV-B sunlight from the ground, we find it is much less than what
we calculated. The difference is the amount of UV-B absorbed by ozone, and
from that we can calculate the amount of ozone. Measuring ozone from
space is similar, but you have to know the amount of the solar UV-B light that
is backscattered, or bouncing, off molecules in the atmosphere (Rayleigh
scattering, again) in the direction of the satellite. This measurement
technique is depicted in the figure on left. We can calculate how much UV-B
light the space-based instrument would observe if there were no ozone.
However, the amount of UV-B measured is much less because UV-B is
passing through the atmosphere a second time. Again, from the amount of
UV-B thats missing, we can calculate the amount of ozone. The longest
satellite record of ozone data has been from instruments using this
backscatter method. The first measurements were taken by the BUV
instrument on Nimbus-4 satellite in 1970 followed by the Total Ozone
Mapping Spectrometer (TOMS) instruments on the Nimbus-7, Earth Probe,
and Meteor-3 satellites, several SBUVs on NOAA satellites, the Ozone
Monitoring Instrument (OMI) onboard the EOS Aura satellite, and the Ozone
Mapping and Profiler Suite (OMPS) on the Suomi NPP satellite. The Europeans
also flew BUV-type instruments on their environmental satellites.
Another type of instrument uses occultation to measure ozone, such as the
Stratospheric Aerosol and Gas Experiment (SAGE) instruments including the
upcoming SAGE-III-ISS onboard the International Space Station, the
Atmospheric Chemistry Experiment (ACE), and Halogen Occultation
Experiment (HALOE) on the UARS satellite. Other techniques include
measuring the radiation emitted from the atmosphere in the infrared and
microwave wavelengths where ozone also absorbs and allows instruments to

detect the radiation change. In addition to the OMI instrument, the Aura
satellite carried three other ozone-measuring instruments, the High
Resolution Dynamics Limb Sounder (HIRDLS), the Tropospheric Emission
Spectrometer (TES) using infrared emission, and the Microwave Limb
Sounder (MLS). (Note: All these instruments measure several other
atmospheric constituents in addition to ozone.)
The Ozone Hole is not really a hole but a thinning of the ozone layer over
the south polar region. Every year, since at least 1978, there is a sudden,
rapid decrease in the stratospheric ozone levels at the end of the Antarctic
winter. During the long winter months of darkness over the Antarctic,
atmospheric temperatures drop, creating unique conditions for chemical
reactions that are not found anywhere else in the atmosphere. The wind in
the stratosphere over the polar region intensifies and forms a polar vortex,
which circulates around the pole. The transition from inside to outside the
polar vortex creates a wind barrier that isolates the air inside the vortex and
also results in very cold temperatures.
At temperatures below -78C, thin clouds made of mixtures of ice, nitric
acid, and sulfuric acid form in the stratosphere. Chemical reactions on the
surfaces of these ice crystals convert chlorine-containing compounds like
HCl, which is harmless to ozone, into more reactive forms. When the sun
rises over the Antarctic in the Spring (September), light rapidly releases free
chlorine atoms into the stratosphere. A new ozone destroying cycle begins.
The chlorine atoms react with ozone, creating ClO. The ClO molecules
combine with each other, forming a compound called a dimer. Sunlight
releases chlorine atoms from the dimer, and the cycle begins again. The
polar vortex keeps the ozone-depleted air inside from mixing with the
undepleted air outside the vortex.
The ozone destruction continues within the polar vortex until the ozone
levels approaches zero at the altitudes where reactions on the thin clouds
have released chlorine atoms. Once ozone has reached such a low level, the
chlorine atoms react with methane, filling the vortex with HCl. The low ozone
persists until the vortex weakens and breaks apart. Then the ozone levels in
the polar stratosphere begin to return to pre-September levels due to the
increase in solar UV and the mixing of polar and nonpolar air.
This activity will introduce students to the use of color maps to visualize data
about stratospheric ozone. Scientists use colors and other representations for
data to help interpret and visualize information. Data are mapped to colors

and other representations to help the mind interpret this information.


Sometimes this means creating an image that looks much like an aerial
photo of the planets surface, but other data are best mapped to a color
scale. In this lesson, students will discover that selecting a good color scale is
both essential to understanding data and to accurately communicating
science. Objectives After completing this activity, students should be able to
describe why color maps are used to visualize data interpret data using a
color mapped image compare and evaluate different color scales Standards
(Grades 9-12) NGSS: Practice 4 Analyzing and Interpreting Data AAAS:
12E/H2 Check graphs to see that they do not misrepresent results by using
inappropriate scales. AAAS: 11C/H4 Graphs and equations are useful ways
for depicting and analyzing patterns of change. NSES: Unifying Concepts and
Processes Standard: Evidence, models, and explanation. NSES: Content
Standard E: Understandings about science and technology Engage Ask
questions about the front of the poster. When was the ozone hole the
smallest? (1979) When did the ozone hole grow the fastest? (19811985,
pattern of growth, no shrinking) What year had the largest ozone hole?
Ozone is one of the most important trace species in the atmosphere.
Therefore, the history of research on ozone has also received a good deal of
attention. Here a short overview of ozone research (with a focus on the
stratosphere) is given, starting from the first atmospheric measurements and
ending with current developments. It is valuable to study the history of ozone
research, because much can be learned for current research from an
understanding of how previous discoveries were made. Moreover, since the
1970s, the history of ozone research has also encompassed also the history
of the human impact on the ozone layer and thus the history of policy
measures taken to protect the ozone layer, notably the Montreal Protocol and
its amendments and adjustments. The history of this development is
particularly important because it may serve as a prototype for the
development of policy measures for the protection of the Earths climate.
In 1839, in a lecture to the Naturforschende Gesellschaft Basel, Christian
Friedrich Schonbein announced the discovery of a new substance that he
would later call ozone (SCHONBEIN , 1839). In the years after its
discovery, the new substance aroused the interest of the scientific
community; e.g., Justus von Liebig said . . . betrachte ich die Entdeckung
des ozonisirten Sauerstoffs fur eine der merkw urdigsten, die je gemacht
wur- den1 (KAHLBAUM and THON, 1900, p. 18). Interestingly, however, it
took many years before the name ozone, which we use quite naturally
today, was generally accepted for the new substance (see HERRMANN

(1982) and RUBIN (2001) for a more extensive report on Schonbeins work).
Ozone is one of the most important trace species in the atmosphere.
The presence of an ozone layer in the lower stratosphere is a key factor
determining the vertical temperature profile of the Earths atmosphere.
Further, UV-B radiation from the sun is absorbed in the ozone layer; if this
radiation was not absorbed, it would reach the Earths surface to an extent
that would be harmful to a variety of life forms (including humans). Excess
ozone at the Earths surface (photochemical smog) occurs because of air
pollution and can be harmful to plants, animals, and humans. On the other
hand, a certain ozone level in the lower atmosphere is beneficial, because
without ozone the chemical cycles that remove pollutants from the
atmosphere cannot work. Owing to the importance of ozone for the Earths
atmosphere, the history of atmospheric ozone research has received much
attention. Over the years, many papers and books have appeared describing
various aspects of ozone research and in various levels of detail (DOBSON,
1968; KHRGIAN, 1975; TUCK, 1978; DOTTO and SCHIFF, 1978; HERRMANN,
1982; CRUTZEN, 1988; FARMAN, 1989; CRUTZEN, 1996; SOLOMON, 1999;
RUBIN, 2001; STOLARSKI, 2001; SOLOMON, 2004; FARRELL, 2005)
. The history of ozone research in the last several decades is furthermore a
good example of the knowledge cycle in geosciences today, where field
measurements and laboratory studies are combined to advance the
understanding of processes leading to an improvement of numerical models,
and the results and forecasts of the numerical models are then confronted
again with measurements (BRASSEUR, 2008). A feature of research on
atmospheric issues that becomes obvious when studying the history of
ozone research is that the scarcity of observational data has al- ways been a
fundamental problem hindering progress. And in spite of all the technological
advances achieved so far, it still is. It is difficult and expensive to obtain
measurements several kilometres above the ground and the gases to be
measured are only present in the atmosphere in trace amounts (for example
the mixing ratios of some chlorofluorocarbons (CFCs) in the stratosphere are
orders of magnitude below one particle in 1012).
Measurements in the atmosphere are never repeatable so that long-term,
consistent time series are very important. A prominent example is the time
series of total column ozone measurements in Halley Bay, Antarctica, that
was started in 1956 and which formed the basis for the discovery of the
ozone hole by FARMAN et al. (1985). Moreover, the interpretation of
atmospheric measurements is always complicated, because advection,

mixing, chemistry, and radiation are all important for the composition of the
atmosphere. Arguably, the most eminent example of the impact of human
activities on nature today is the ozone hole in the Antarctic (see Section 5.1
below). The discovery of the ozone hole by FARMAN et al. (1985) was a key
factor leading to the Montreal Protocol on Substances that Deplete the
Ozone Layer, which was signed in 1987. Consequently, the history of
research on stratospheric ozone depletion has recently also aroused interest
as an example of how the dynamics of scientific learning and scientific
assessment processes can be studied (BRASSEUR, 2008; CRUTZEN and
OPPENHEIMER, 2008), which is an important issue in the current debate
about climate change.
The outline of the paper is as follows. The early observations of atmospheric
ozone are introduced in Section 2, the discussion starts with the discovery of
ozone and the first tropospheric measurements, covers the discovery of the
role of ozone as an absorber of solar radiation, the first measurements of
total column ozone, and the findings from the early 1930s that the height of
the ozone layer is 25 km rather than 4050 km as had been hitherto
thought. Section 3 presents the development of the understanding of the
chemistry of stratospheric ozone, from the early work by CHAPMAN (1930) to
the discovery of the importance of catalytic cycles in destroying ozone. The
development of concerns about anthropogenic effects on the ozone layer is
described in Section
4. The effects discussed are nuclear weapons tests in the stratosphere, the
emissions of nitrogen oxides from supersonic aircraft, the use of artificial
fertilisers in agriculture, followed by a review of the warning by MOLINA and
ROWLAND (1974) that the accumulation of anthropogenic CFCs in the
atmosphere might ultimately lead to a depletion of the ozone layer. The
discovery of the ozone hole (FARMAN et al., 1985; STOLARSKI et al., 1986;
WMO, 2007) is described in Section 5, the related issue of polar ozone loss in
the Arctic in late winter and spring is introduced, Figure 1: Erich Regener in
1954. He is shown holding an automatic air sampler for balloon-borne
operation. (Photo courtesy of the Max Planck Society). and developments
towards an increasing understanding of the phenomenon are described.
Finally, in the conclusions (Section 6), the timeline of discoveries in ozone
research is compared to milestones in the history of physics.
Schonbein himself believed that ozone played an im- portant role in the
Earth system and suggested in 1853 that long-term ozone measurements in
the atmosphere should be performed on an international scale: Geneigt zu

glauben, das atm. Ozon spiele im Haushalte der Erde eine wichtige Rolle,
halte ich es fur w unschenswerth, dass moglichst zahlreiche, sowohl
grosse Zeitr aume als bedeutende Landerstrecken umfassende,
untereinan- der vergleichbare Beobachtungen uber die Ver anderung-
en des Ozongehaltes der Atmosphare angestellt wer- den. . . 2 (in a letter
to Justus von Liebig, KAHLBAUM and THON, 1900, p. 10). Indeed numerous
measurements were made during the second half of the 19th century using a
method that Schonbein had developed (for a detailed discussion of the
early measurements of atmospheric ozone see e.g., CRUTZEN, 1988;
LONDON and LIU, 1992; SONNEMANN, 1992). Unfortunately, measurements
using Schonbeins method are only semi- quantitative because of poor
standardisation and the in- fluence of humidity and wind speed on the values
obtained (LINVILL et al., 1980; VOLZ and KLEY, 1988).
The only historic measurements of tropospheric ozone that proved reliable
enough to allow a reanalysis and a meaningful comparison with modern
measurements were taken at the Observatoire de Montsouris, located on the
outskirts of Paris (VOLZ and KLEY, 1988). At Montsouris a quantitative wetchemical method was established in 1876 and used continuously for 34
years. This early interest in ozone was motivated to a large extent by the
fact that ozone, at that time, was considered to be an indicator of clean,
healthy air and that ozone-poor air would lead to illness. Measurement series
of ozone were performed in spa towns with the intention of demonstrating
how beneficial to human health a stay would be (LENDER, 1872).
This view persisted for many decades and was held by leading scientists in
the field; as late as 1946 Erich Regener, one of the early pioneers in ozone
research (Figure 1), stated Anwesenheit von Ozon ist ein Indikator fur gute
Luft. 3 (REGENER, 1946). Today, of course, it is known that high
concentrations of tropospheric ozone are detrimental to human health and,
similarly, harmful to other living systems. Since the early 1950s, it has been
understood that high ozone concentrations in the troposphere (referred to as
photochemical smog) are caused by pollution; ozone is chemically formed
as a result of photochemical reactions involving NOx and non-methane
hydrocarbons from automobile exhausts and similar combustion processes
(HAAGEN-SMIT, 1952).
Still in the 19th century, another important aspect of ozone was discovered,
namely its importance as an absorber of light. In 1879 the French physicist
Alfred Cornu recognised that solar radiation with wavelengths below 300 nm
does not penetrate to the Earths surface (CORNU, 1879a) 4 . Only three

years later, HARTLEY (1881b) suggested that ozone is responsible for this
observation; the ozone absorption bands in this UV region are therefore
today referred to as Hartley bands.
Already HARTLEY (1881a) believed that ozone is a normal constituent of the
higher atmosphere (see KHRGIAN (1975) for more details on Hartleys
work). The discovery of the spectroscopic properties of ozone opened up the
possibility of making measurements of the total thickness of the atmospheric
ozone (i.e., the total atmospheric column of ozone) including the ozone in the
stratosphere. The first detailed measurements of the total atmospheric
column of ozone were made by FABRY and BUISSON (1921), who firmly
established that ozone was responsible for the observed strong cut-off in
solar UV spectra at about 300 nm towards shorter wavelengths. FABRY and
BUISSON (1921) speculated (correctly as we know today) that ozone was
formed by solar UV radiation. They also suggested that the ozone layer was
situated at an altitude of about 50 km.
And first measurements of the height of the ozone layer indeed con- firmed
an altitude of the ozone layer of 4853 km (CABANNES and DUFAY, 1925;
LAMBERT et al., 1926). In 1924, G. M. B. Dobson designed a spectrograph to
measure the total ozone column that was more suited for routine out-of-door
use and that was cheaper to build than the FABRY and BUISSON
spectrograph. The aim was to develop an instrument that allowed regular
measurements to be made over extended time periods. This development
was very successful; today, the standard unit for the total atmospheric
column of ozone bears Dobsons name: the Dobson unit (DU).
One DU is de- fined as 2.687 1016 ozone molecules per square centimetre.
The Dobson unit describes the thickness of a layer of pure ozone, if the total
amount of ozone in the atmosphere were brought to standard conditions (15
degrees Celsius and 1013 hPa). For example, an atmospheric ozone column
of 300 DU (a typical atmospheric value) brought down to the surface of the
Earth would occupy a 3 mm thick layer of pure ozone. The first observations
at Oxford 1924-1925 showed a marked annual variation of ozone. Likewise,
the data exhibited a strong day-to-day variability that was closely connected
to meteorological conditions (DOBSON and HARRISON, 1926; DOBSON,
1968).
Between July 1926 and November 1927 seven instruments were built at
Oxford and distributed throughout Europe covering the latitude range 68N
to 47N and one instrument was operated at 22S in Chile (DOBSON et al.,

1929; DOBSON, 1930). After the 1926-1927 observation period, the


instruments were sent to Spitzbergen (via an Italian airship expedition),
California, Egypt, India, and New Zealand (see DOBSON, 1968; FARMAN,
1989; BRONNIMANN et al., 2003, for further details on the history of early
total ozone measurements). In the former USSR, the first measurements of
total ozone were taken in 1933; later, in 1959, the M-83 ozonometer was
developed, which became the backbone of the USSR ozone station network
(GUSHCHIN, 1995).
Even today, ground-based total ozone measurements are essential as a
reference measurement for satellite data and thus for monitoring the ozone
content of the atmosphere. Further, the longest records of continu- ous
reliable measurements are available from stations equipped with Dobson
spectrophotometers (e.g., JONES and SHANKLIN, 1995; STAEHELIN et al.,
1998), whose design has remained essentially unchanged since 1930
(DOBSON, 1968). Such long data records are invaluable for time series
analysis and for trend detection (e.g., FIOLETOV et al., 2002; JANOSI and
MULLER , 2005; WMO, 2007). In the late twenties, the ozone layer was still
assumed to be located in the upper stratosphere. GOTZ and DOBSON
(1928), based on their measurements, reported that the . . . average height
[of the ozone layer] seems to be between 3040 km above sea-level.
BJERKNES believed in 1929 that . . . le fait qui parait aussi etre bien etabli,
que la couche dozone se trouve a une altitude de ` 4050 kilometres ` 5
(BJERKNES, 1929). When GOTZ and DOBSON (1928) and BJERKNES (1929)
refer to the height of the ozone layer they mean the altitude of the maximum
ozone concentration in number density (measured in particles per unit
volume or in DU per km).
A different, greater altitude results from considering the maximum of the
molar mixing ratio (measured in micromole ozone per mole of air, commonly
referred to as ppm). In ground-based measurements in Spitzbergen in 1929,
GOTZ (1931) discovered the so called Umkehreffekt, which allows a lowresolution vertical ozone pro- file to be deduced (e.g., PAETZOLD and
REGENER, 1957; DUTSCH and STAEHELIN, 1992). He reported an average
height of the ozone layer for the period 10 July to 28 August 1929 in
Spitzbergen of 27.6 8 km. At the time, GOTZ put forward these new ideas
rather carefully . . . es fragt sich nun ob die geophysikalisch interessante
tiefere Lage der Ozonschicht in der hohen Breite Spitzbergens, vor allem
aber auch ihre Schwankung bis herunter zur minimalen Hohe des 26. August
[11 km] als genugend gesichert angesehen werden d urfen 6 .

However, two years later the Spitzbergen measurements were confirmed by


mid-latitude measurements and GOTZ et al. (1933) stated with much more
confidence that the average height [of the ozone in the atmosphere] at
Arosa now appears to be about 20 km, which is much below the former
estimates. The first attempt to directly measure the vertical pro- file of the
atmospheric ozone concentration (and other atmospheric parameters) was
made by James Glaisher and Henry Tracey Coxwell on a manned balloon.
On 5 September 1862 they made a balloon ascent from Wolverhampton,
England, and measured ozone using Schonbeins method. Although the
balloon reached an altitude of 11 km, the last ozone measurement was
only performed at an altitude of 4.8 km because the balloonists lost
consciousness at greater altitudes. This balloon flight was described by
LENDER (1873) and later in the New York Times, 27 June (1909); see also
HOINKA (1997) for a report on the flight. Between 1862 and 1866 Glaisher
made numerous further balloon ascents in order to measure the temperature
and humidity of the atmosphere at greater heights. He was also a founder
member of the Meteorological Society in 1850, which later (in 1883) became
the Royal Meteorological Society. The first successful balloon-borne
measurements of the vertical distribution of ozone were conducted soon
after the first Umkehr measurements.
In a balloon flight on 31 July 1934 from Stuttgart, REGENER and REGENER
(1934) recorded the solar UV spectrum in the stratosphere and deduced an
ozone profile from these measurements. These data independently
confirmed the first Umkehr observations. The payload for this balloon
experiment weighted only 2.7 kg and was launched to an altitude of about
31 km using two standard meteorological balloons. Table 1 lists the original
values (ozone column above the balloon) reported by REGENER and
REGENER (1934) and Figure 2 shows these data as they were presented in
the original paper. In Figure 3, the vertical profile of the ozone mixing ratio
deduced from these early measurements is compared with a corresponding
ozone climatology deduced from measurements of the Halogen Occultation
Experiment (HALOE) on the Upper Atmosphere Research Satellite (UARS) in
the years 1991-2002 (RUSSELL et al., 1993; GROOSS and RUSSELL, 2005)
demonstrating a rather good agreement of the early measurements with a
current climatology. In the early fifties, PAETZOLD (1950) developed a
method for determining the vertical profile of ozone during lunar eclipses up
to altitudes that cannot be reached by balloons (up to about 45 km). In this
method, the ozone profile is determined from the ozone absorption in the

Chappius band measured at the edge of the Earths shadow on the moon.
The effectiveness of this method stems from the large effective ozone c
olumn (up to 15000 DU) due to the tangential optical path through the
Earths atmosphere. This effect is visible during lunar eclipses as a small
area at the edge of the umbra showing a greenish colour rather than the
reddish colour 0 2 4 6 8 10 12 14 Ozone (DU/km) 0 10 20 30 40 50 Altitude
(km) Figure 4: The first vertical ozone profile measured from a rocket
experiment on 10 October 1946 above White Sands Proving Ground, New
Mexico (32.33N,106.44W). Data are taken from Table 1 in JOHNSON et al.
(1951). The quantity measured is the ozone column overhead of a rocketborne spectrograph so that the deduced ozone concentrations have a rather
large uncertainty. in the core of the umbra. This greenish colour is caused by
the extinction of the solar light reaching the moon in the yellow-red spectral
region by the Chappius band of ozone (PAETZOLD, 1951). Erich Regener was
forced to retire from his chair at the University of Stuttgart in 1937 by the
Nazi regime. Nonetheless, he continued his scientific work; he first founded a
private research institute on 1 January 1938, his Forschungsstelle fur Physik
der Stratosph are 7 , which was integrated later that year into the
Kaiser Wilhelm Society for the Advancement of Science (which was
refounded in 1948 as the Max Planck Society).
He developed the first scientific payload for a rocket that allowed it to reach
high altitudes (about 150 km); the rocket was known as the A4. He designed
an airand water-tight container containing an assembly of instruments, in
particular temperature and pressure gauges and a spectrometer to register
solar UV radiation (ORDWAY III et al., 2007). This payload, referred to as the
Regener-Tonne (Regener barrel) was never flown, but the first scientific
payload (a Geiger counter, which however failed to work) was launched on
an A4 rocket from New Mexico as early as 16 April 1946 (GREEN, 1954).
Many more flights followed, including flights of A4 rockets carrying UV
spectrometers for ozone measurements (JOHNSON et al., 1951; ORDWAY III
et al., 2007). The first in situ measurement of an ozone profile above balloon
altitudes (Figure 4) was made on 10 October 1946 by a UV spectrograph
mounted in a tail fin of an A4 rocket (JOHNSON et al., 1951). Simpler and less
expensive rockets were soon developed for upper atmosphere research (VAN
ALLEN et al., 1948). For many years, rocket measurements were the key
experimental tool for exploring ozone chemistry in the upper stratosphere
and in the mesosphere.

In the early sixties, the first satellite measurements of the vertical


distribution of ozone in the stratosphere and mesosphere were made using
the same technique as for the lunar eclipses but replacing the moon as a
reflector by an artifical satellite (VENKATESWARAN et al., 1961). Nonetheless,
rocket measurements remained an important scientific tool for decades to
come (e.g., MILLER and RYDER, 1973; VAUGHAN, 1982, 1984; GERNANDT et
al., 1989). An important feature of the vertical ozone profile, the secondary
ozone maximum at about 90 km, was discovered both in ozone
measurements from rocket payloads and in ozone concentrations in the
upper atmosphere deduced from stellar occultation measurements made
from a satellite (MILLER and RYDER, 1973; HAYS and ROBLE, 1973).
When DOBSON and HARRISON first published their measurements of total
ozone in 1926, the formation mechanism of stratospheric ozone was still
unclear. They discussed whether the ozone is formed in the extreme upper
atmosphere by ultra-violet radiation from the sun, or by electrical discharges
in auror. . . . Four years later, CHAPMAN (1930) proposed the first
photochemical theory of the formation of ozone; four reactions that involved
only allotropes of oxygen and that are today known as the Chapman
reactions.
He suggested that stratospheric ozone is produced by the photolysis of
molecular oxygen (O2) at ultraviolet wavelengths below 242 nm, R1: O2 +h
2O( 3P), where h denotes an ultraviolet photon. The atomic oxygen
(O( 3P), hereafter denoted by O) produced in reaction R1 reacts rapidly with
molecular oxygen to form ozone (O3) R2: O+O2 +M O3 +M, where M
denotes a collision partner (N2 or O2) that is not affected by the reaction.
Ozone is photolysed rapidly R3: O3 +h O+O2.
Through the reaction R4: O+O3 2O2. an O atom and an O3 molecule are
lost. Because O and O3 are in rapid photochemical equilibrium, the loss of
one oxygen atom effectively implies the loss of an ozone molecule, so that
R4 destroys two molecules of odd oxygen. The Chapman reactions R1-R4
were accepted for decades as sufficient to describe the observed vertical
distribution of ozone.
However, new measurements of the rate constants for the Chapman
reactions (BENSON and AXWORTHY, 1957; JONES and DAVIDSON, 1962)
demonstrated that a quantitative understanding of ozone chemistry was
lacking. It became clear that the destruction of ozone by reaction R4 alone
cannot explain the observed ozone abundances in the stratosphere (e.g.,

HUNT, 1966b; SCHIFF, 1969); as HUNT (1966a) put it . . . the photochemical


reaction scheme normally considered is no longer adequate, and there must
be reactions occurring in the atmosphere which destroy ozone but which
have been neglected in the past. BATES and NICOLET (1950) had originally
suggested that reactions involving OH and HO2 radicals would lead to a
significant catalytic ozone loss.
Based on this work, HOx-catalysed reactions were suggested as an ozone
destruction mechanism by HAMPSON (1964, 1966) and, using laboratory
measurements by NORRISH and WAYNE (1965), implemented in a model
calculation by HUNT (1966b). CRUTZEN (1969) found that the mechanism
involving OH and HO2 could not explain ozone observations in the middle
stratosphere between 3035 km and, a year later, proposed that NO- and
NO2- catalysed reactions control the ozone concentrations in the middle
stratosphere (CRUTZEN, 1970). At that time, the very first measurements of
nitrogen compounds in the stratosphere had just become available
(MURCRAY et al., 1968; RHINE et al., 1969). The possibility that nitrous oxide
(N2O) is decomposed in the stratosphere via the reaction R5: N2O+O( 1D)
2NO. and thus constitutes a possible source for stratospheric NO and
NO2 had already been suggested by HAMPSON (1964)8 . At that time he
concluded that it is unlikely that the nitrogen oxide reactions can affect total
ozone but it is possible that observation of them, knowing the rates of the
reactions involved, can assist our understanding of the distribution and
behaviour of ozone (HAMPSON, 1964, p. 267). Again a few years later, the
possibility of chlorine catalysed ozone loss was put forward by STOLARSKI
and CICERONE (1974). Their original motivation for investigating the impact
of chlorine on ozone was studying the possible effects on the stratosphere of
HCl originating from space shuttle exhaust, but soon the focus shifted to
chlorine injected into the stratosphere by volcanoes (STOLARSKI, 2001). At
that time, laboratory studies of chlorine reactions had been begun employing
the latest technique.
STOLARSKI (2001) recalls that as early as 1972 he heard laboratory chemist
Don Stedman say in a conversation: Chlorine destroys ozone. Everybody
knows that!. The different catalytic ozone loss cycles can be summarised in
the form R6: XO+O X+O2 R7: X+O3 XO+O2 C1: Net: O+O3 2O2
where the net reaction is identical to reaction R4. For the cycle catalysed by
hydrogen radicals, X = H, OH, for the nitrogen radical cycle, X = NO, and for
the chlorine radical cycle, X = Cl. Today the relative importance of the
various ozone loss cycles is well understood. Because of the large increase of
atomic oxygen (O) with altitude, the rates of all cycles increase strongly

between 25 and 40 km and the same is true of the rate of ozone production
through reaction R1 (e.g., CRUTZEN et al., 1995; GROOSS et al., 1999).
Further, the relative importance of the cycles for ozone loss varies
considerably with altitude. Between 2540 km the NOx cycle is the dominant
ozone loss process, whereas above 45 km HOx catalysed ozone loss
dominates (see Figure 5). The HOx cycle is also the strongest loss cycle
below about 25 km. The loss through the ClOx cycle (which also depends on
the stratospheric chlorine loading) peaks at 40 km. Furthermore,
interestingly, the importance of heterogeneous reactions (albeit no
heterogeneous reactions involving chlorine species) in the stratosphere was
already suggested in the mid-seventies (CADLE et al., 1975). Based on
laboratory experiments, heterogeneous reactions were, however, for a long
time thought to be unimportant.
This situation changed only in 1986 when SOLOMON et al. suggested that
heterogeneous chemistry could greatly enhance the ability of chlorine to
destroy ozone in polar regions (see Section 5 below). The paper by SOLOMON
et al. prompted numerous laboratory studies and today, a large number of
heterogeneous reactions are known that are important for stratospheric
Figure 5: The vertical distribution of the relative importance of the individual
contributions to ozone loss by the HOx, ClOx, and NOx cycles as well as the
Chapman loss cycle (R4). The calculations are based on HALOE (V19)
satellite measurements and are for overhead sun (23S, January) and for
total inorganic chlorine (Cly) in the stratosphere corresponding to 1994
conditions.

THE IMPORTANCE OF OZONE LAYER


Ozone is found in the stratosphere, a layer of air that encircles the globe
approximately 6 to 30 miles above its surface. When ultraviolet radiation
strikes this layer of air, it interacts with the ozone and is chemically
decomposed. Even with the ozone filter, some ultraviolet radiation reaches
the Earth. This radiation is responsible for skin cancer and stunting of plant
growth.
In the 1980s, inquiry into the cause of the depletion of the ozone layer led to
investigation of chlorofluorocarbons, which are chemicals used in
refrigerants, insulating foams and solvents. Once chlorofluorocarbons are
released into the atmosphere, winds buffet them upward where they interact
with ozone, destroying the ozone molecules. In addition to
chlorofluorocarbons, pesticides containing methyl bromide, halons used in

fire extinguishers and methyl chloroform found in industry solvents are


known to destroy ozone.
Earth is an extraordinary planet.Complex interactions between the land,
oceans, and atmosphere created conditions that are favorable for life.One
species, man, has managed to alter the environment on a global scale.In
order to fully comprehend the impact of our actions, we must view the planet
as a whole and understand the relationship between its basic components;
land, water, and air.
This web site discusses the chemical composition and evolution of Earth's
atmosphere, focusing on the protective layer of ozone in the stratosphere.
The destructive properties of tropospheric ozone are also
presented.Diagrams and animation sequences are used to visually depict the
delicate structure of the ozone molecule and the chemical reactions involved
in its formation and destruction.Ozone destroying pollutants were first
identified in 1973.Since that time there has been a considerable amount of
controversy surrounding the subject of ozone depletion.More than 20 years
of ozone-related scientific studies, international meetings, and global
industrial agreements are summarized in the last section of this site.
One billion years ago, early aquatic organisms called blue-green algae began
using energy from the Sun to split molecules of H2O and CO2 and recombine
them into organic compounds and molecular oxygen (O2).This solar
energyconversion process is known as photosynthesis.Some of the
photosynthetically created oxygen combined with organic carbon to recreate
CO2 molecules.The remaining oxygen accumulated in the atmosphere,
touching off a massive ecological disaster with respect to early existing
anaerobic organisms.As oxygen in the atmosphere increased, CO2 decreased.
High in the atmosphere, some oxygen (O2) molecules absorbed energy from
the Sun's ultraviolet (UV) rays and split to form single oxygen atoms.These
atoms combined (27k jpeg) with remaining oxygen (O2) to form ozone (O3)
molecules, which are very effective at absorbing UV rays.The thin layer of
ozone that surrounds Earth acts as a shield, protecting the planet from
irradiation by UV light.
The amount of ozone required to shield Earth from biologically lethal UV
radiation, wavelengths from 200 to 300 nanometers (nm), is believed to
have been in existence 600 million years ago.At this time, the oxygen level
was approximately 10% of its present atmospheric concentration.Prior to this
period, life was restricted to the ocean.The presence of ozone enabled

organisms to develop and live on the land.Ozone played a significant role in


the evolution of life on Earth, and allows life as we presently know it to exist.
Depending on where ozone resides, it can protect or harm life on Earth.When
it is close to the planet's surface, in the air we breathe, ozone is a harmful
pollutant that causes damage to lung tissue and plants, and is considered to
be "bad ozone."It is a powerful photochemical oxidant that damages rubber,
plastic, and all plant and animal life.It also reacts with hydrocarbons from
automobile exhaust and evaporated gasoline to form secondary organic
pollutants such as aldehydes and ketones.The peroxyacyl nitrates are
especially damaging photochemical oxidants that are very irritating to the
eyes and throat.
Ozone pollution originating in urban areas can extend into surrounding rural
and forested areas that are hundreds of kilometers downwind.Episodes of
elevated ozone concentrations are associated with warm, slow moving high
pressure systems and contain between 30 and 50 parts per billion by
volume. Concentrations 3 to 8 times greater than natural background levels
have been observed.During the summer heat wave of 1988, record ozone
concentrations were recorded in the United States.Even Acadia National Park
in Maine, and the Shenandoah mountains of Virginia, were affected by
dangerous levels of ozone pollution.These rural areas are far removed from
industrial regions and polluted cities. The ozone pollution recorded in Acadia
most likely originated in New York City.That in Virginia may have migrated
from refineries on the Gulf Coast.
Photochemical oxidants are the most significant cause of agricultural loss in
the United States.Their damaging effects on vegetation and crops have been
confirmed in the eastern United States, adjacent areas in Canada, and much
of Europe.Ozone alone, or in combination with sulfur dioxide (SO2) and
nitrogen dioxide (NO2), accounts for 90% of the annual crop losses in the U.S.
that are caused by air pollution.
Ozone Production and Destruction
Stratospheric ozone is created and destroyed primarily by ultraviolet (UV)
radiation. The air in the stratosphere is bombarded continuously with UV
radiation from the Sun.When high energy UV rays strike molecules of
ordinary oxygen (O2), they split the molecule into two single oxygen
atoms.The free oxygen atoms can then combine with oxygen molecules (O2)
to form ozone (O3) molecules.

O2 + UV light 2 O
O + O2 + M O3 + M (where M indicates conservation of energy and
momentum)
The same characteristic of ozone that makes it so valuable its ability to
absorb a range of UV radiation also causes its destruction. When an ozone
molecule is exposed to UV energy it may revert back into O2and O. During
dissociation, the atomic and molecular oxygens gain kinetic energy, which
produces heat and causes an increase in atmospheric temperature.
Ozone production is driven by UV radiation of wavelengths less than 240 nm.
Ozone dissociation typically produces atomic oxygen (O) that is stable when
exposed to longer UV wavelengths, up to 320 nm, and visible light
wavelengths of 400-700 nm. Longer wavelength photons can penetrate
deeper into the atmosphere, creating regions of ozone production and
destruction. When an ozone molecule absorbs even low energy UV, it splits
into an ordinary oxygen molecule and a free oxygen atom.
O3 + UV, visible light O + O2
The free oxygen atom may then combine with an oxygen molecule, creating
another ozone molecule, or it may take an oxygen atom from an existing
ozone molecule to create two ordinary oxygen molecules.
O + O2 O3 or O3 + O O2 + O2
Processes of ozone production and destruction, initiated by ultraviolet
radiation, are often referred to as Chapman Reactions.
Most O3 destruction takes place through catalytic processes rather than
Chapman Reactions.Ozone is a highly unstable molecule that readily donates
its extra oxygen molecule to free radical specie,s such as nitrogen, hydrogen,
bromine, and chlorine.These compounds naturally occur in the stratosphere,
released from sources such as soil, water vapor, and the oceans.
O3 + X XO + O2 ( where X may be O, NO, OH, Br or Cl)
Anthropogenic Destruction
Manufactured compounds are also capable of altering atmospheric ozone
levels. Chlorine, released from CFCs, (15k JPG) and bromine (Br), released
from halons, are two of the most important chemicals associated with ozone
depletion. Halons are primarily used in fire extinguishers. CFCs are used
extensively in aerosols, air conditioners, refrigerators, and cleaning solvents.

Two major types of CFCs are trichlorofluorocarbon (CFC13), or CFC-11,and


dichlorodifluoromethane (CF2Cl2), or CFC-12. Trichlorofluorocarbon is used in
aerosols, while dichlorodifluoromethane is typically used as a coolant.
CFCs were originally created to provide a substitute for toxic refrigerant
gases and to reduce the occupational hazard of compressor explosions. Near
Earth's surface, chloroflourocarbons are relatively harmless and do not react
with any material, including human skin. For 50 years they appeared to be
the perfect example of a benign technical solution to environmental and
engineering problems, with no negative side effects. While CFCs remain in
the troposphere, they are virtually indestructible. They are not water soluble
and cannot even be washed out of the atmosphere by rain. We now
understand that the very quality that made them seem so safe their
stability is what makes them so dangerous to the Earth system. CFCs
remain in the troposphere for more than 40 years before their slow migration
to the stratosphere is complete. Even if we were to end their production and
use at this very moment, they will continue to contribute to ozone
destruction far into the future.
In the stratosphere, high energy UV radiation causes the CFC molecules to
break down through photodissociation. Atomic chlorine, a true catalyst for
ozone destruction, is released in the process. Chlorine initiates and takes
part in a series of ozone-destroying chemical reactions and emerges from the
process unchanged. The free chlorine atom initially reacts with an unstable
oxygen containing compound (such as ozone) to form chlorine monoxide
(ClO):
Cl + O3 ClO + O2
The ClO molecule then reacts with atomic oxygen to produce molecular
oxygen (O2) and more atomic chlorine. The regenerated Cl atom is then free
to initiate a new cycle:
ClO + O Cl + O2
This destructive chain of reactions will continue over and over again, limited
only by the amount of chlorine available to fuel the process.
Chlorine occurs naturally in the oceans. However, the majority of chlorine in
the atmosphere has originated with man-made chemicals. Without the
dissociation of manufactured chlorofluorocarbons, there would be almost no
chlorine in the stratosphere. CFC-12 concentrations were less than 100 parts
per trillion by volume when they were first measured in the 1960s. Between

1975 and 1987, concentrations more than doubled from less than 200 parts
per trillion by volume to more than 400 parts per trillion by volume. The
amount of chlorine in the stratosphere also increased by a factor of 2 to 3.
Scientists believe that continued buildup of CFCs could lead to severe ozone
loss (61k JPG) worldwide. Thus, ongoing studies are essential to provide a
necessary understanding of the causes of ozone depletion.The history of
CFCs demonstrates that human activities can have unexpected long-term
effects on the environment.
ver Earth's lifetime, natural processes have regulated the balance of ozone in
the stratosphere. Scientists are finding that ozone levels change periodically
as part of regular natural cycles such as seasons, periods of solar activity,
andchanges in wind direction.Concentrations are also affected by isolated
events that inject materials into the stratosphere, such as volcanic eruptions.
Polar regions reflect the greatest changes in ozone concentrations, especially
the South Pole.The topography of Antarctica is such that a stagnant whirpool
of extremely cold stratospheric air forms over the region during the long
polar night.The air stays within this polar vortex all winter, becoming cold
enough to allow the formation of polar stratospheric clouds.
Polar stratospheric clouds speed up the natural process of ozone destruction
by providing ice crystal surfaces on which the destructive reactions take
place.After the long polar winter, ozone within this extremely cold vortex is
very vulnerable to the arrivalof sunlight. As spring arrives, major ozone
losses occur. In the southern hemisphere, the area of most severe ozone
depletion is localized above Antarctica and is generally referred to as the
ozone hole.Thehole appears in the southern spring, following the continent's
coldest season and polar night.
Ozone depletion over the Arctic is not as well defined as in Antarctica.The
rugged topography results in an air circulation pattern that is quite different
from that of the South Pole, but expeditions have shown that the
atmospheric chemistry of the two polar regions is very similar.In the Northern
Hemisphere, the polar vortex is not as strong.It can break up and reform
several times during the course of winter.One air mass after another enters
the polar darkness and soon emerges back into low sunshine. Thus, a bit of
ozone is lost from each parcel of air, rather than a large amount from one
parcel as in the southern hemisphere.
The end result is that we are losing ozone in both hemispheres. (26k jpeg)
Ozone levels in the atmosphere have been monitored from the ground since

the 1950s and by satellite since the 1970s.Regional total ozone levels
measured from satellites over Antarctica have decreased 30-50% since their
monitoring began.
Since ozone is created and destroyed by solar UV radiation, there is some
correlation of ozone concentration with 11-year sunspot cycles.Sunspots emit
high levels of electromagnetic radiation. The increased UV radiation
contributes to ozone production.Sunspot variations only account for 2 to 4 %
of the total variation in ozone concentrations.Natural cycles in ozone
variation are also associated with the quasi-biennial oscillation in which
tropical winds switch from easterly to westerly every 26 months.This cyclic
change in wind direction accounts for approximately 3 % of the natural
variation in ozone concentration.
In 1973, two scientists from the University of California at Irvine, Mario
Molina and F. Sherwood Rowland, first discovered that man-made substances
called chlorofluorocarbons (CFCs) could play a major role in the destruction
of stratospheric ozone.Their findings were published in the journal Nature in
June 1974.Since that time there has been much controversy surrounding the
subject of ozone depletion.Researchers have struggled to understand the
nature and severity of the problem through numerous scientific
studies.Nations from all over the world have come together and agreed to
establish international industrial regulations in hope of protecting the ozone
layer.The following timeline is a combination of material quoted from Sharon
Roan's book entitled "Ozone Crisis; The 15 Year Evolution of a Sudden Global
Emergency", and up-to-date information from current research scientists. It
demonstrates the course of events that has taken place since the first ozone
destroying pollutants were identified more than 20 years ago.
Because of uncertainty about how global environmental systems work, and
because the people affected will probably live in circumstances very much
different from those of today and may have different values, it is difficult to
predict how present-day actions will affect future generations.To project or
forecast the human consequences of global change at some point in the
relatively distant future, one would need to know at least the following:
the
the
the
the

future state of the natural environment


future of social and economic organization
values held by the members of future social groups
proximate effects of global change on those values

the responses that humans will have made in anticipation of global change
or in response to ongoing global change
International Agreements
Even the value systems, and technological advancements, of present
daynations are extremely different.Nonetheless, efforts to predict and
protect are underway.Despite their differences, the international community
made significant progress in addressing ozone depletion as a serious global
environmental problem.Through the 1985 Vienna Convention for the
Protection of the Ozone Layer, the 1987 Montreal Protocol on substances that
deplete the ozone layer, and the 1990 London Amendments to the protocol,
members from nations around the world have committed to phasing out the
production and consumption of CFCs, and a number of related chemicals, by
the year 2000.
Ozone depletion control started in the early 1970s, when the United States,
along with a handful of other Western countries, expressed concern over
emissions from supersonic transport (SST) aircraft and aerosol spray
cans.Environmental groups organized opposition to the development of the
SST and to the extensive use of aerosols.Public response led to a sharp drop
in the sales of aerosol products.The U.S. Congress, prodded by government
studies supporting the CFC ozone depletion theory and its links to skin
cancer, approved the Toxic Substances Control Act of 1976, which gave the
Environmental Protection Agency (EPA) authority to regulate CFCs.In 1978,
the United States became the first country to ban the nonessential use of
CFCs in aerosols.However, the EPA ruled that other uses of CFCs, such as
refrigeration, were essential and lacked available substitutes.
Ozone depletion emerged as a major international issue in the 1980s. This
occurred primarily as a result of initiatives by the United Nations
Environmental Programme and actions of the international scientific and
environmental communities.A United Nations Environment Program to
protect the ozone layer was signed in Vienna in 1985, and a protocol
outlining proposed protective actions followed.The Vienna convention of
1985 embodied an international environmental consensus that ozone
depletion was a serious environmental problem.
However, there was no consensus on the specific steps that each nation
should take.The Montreal Protocol, signed in September 1987,stated that
there would be a 50 % cut back in CFC production by 2000.The United States
ratified the Montreal Protocol in 1988.The 1990 London Amendments to the

protocol state that production of CFCs, CCl4, and halons will be completely
halted by the year 2000.The phaseout schedule for other compounds was
accelerated by 4 years by the 1992 Copenhagan agreement.
All human activity potentially contributes directly or indirectly to global
change.Earth's atmosphere consists of a delicate balance of gases essential
to life.Throughout the history of the planet, the atmospheric gases have
been influenced by Earth processes and by the living organisms from both
the oceans and land, and natural changes have occurred in the type of gases
and their concentrations. Anthropogenic activities are now believed to be
causing rapid changes in atmospheric composition on an accelerated time
scale. Due to extended human life expectancies and greater population
densities, the influence of humans will continue to grow.
Scientists are now confident that stratospheric ozone is being depleted
worldwide.However, how much of the loss is the result of human activity, and
how much is the result of fluctuations in natural cycles, still need to be
determined.To understand global atmospheric changes, we need to
understand the composition and chemistry of Earth's atmosphere and how
they are affected by human activity.To create accurate models, scientists
must account for all of the factors affecting ozone creation and destruction,
and conduct simultaneous, global studies over the course of many years.
Stratospheric ozone has been depleted over the last 25 years following
anthropogenic emissions of a number of chlorine- and bromine-containing
compounds (ozone-depleting substances, ODSs), which are now regulated
under the Montreal Protocol. The Protocol has been effective in controlling
the net growth of these compounds in the atmosphere. As chlorine and
bromine slowly decrease in the future, ozone levels are expected to increase
in the coming decades, although the evolution will also depend on the
changing climate system.
Emissions of chlorofluorocarbons (CFCs) and hydrochloro- fluorocarbons
(HCFCs) have depleted stratospheric ozone. Globally, ozone has decreased
by roughly 3% since 1980. The largest decreases since 1980 have been
observed over the Antarctic in spring (the ozone hole), where the monthly
column-ozone amounts in September and October have been about 4050%
below pre-ozone-hole values. Arctic ozone shows high year-to-year variability
during winterspring due to variability in dynamical transport and chemical
loss (which act in concert). Chemical losses of up to 30% are observed during
cold winters, whereas they are small for dynamically active warm years.

Due to the control of ODSs under the Montreal Protocol and its Amendments
and Adjustments, the abundance of anthropogenic chlorine in the
troposphere has peaked and is now declining. The tropospheric abundance of
anthropogenic bromine began decreasing in the late 1990s. Stratospheric
concentrations of ODSs lag those in the troposphere by several years.
Stratospheric chlorine levels have approximately stabilized and may have
already started to decline.
Changing ODS concentrations will be a dominant factor controlling changes
in stratospheric ozone for the next few decades. Stratospheric ozone
depletion is likely to be near its maximum, but ozone abundance is subject to
considerable interannual variability. Assuming full compliance with measures
adopted under the Montreal Protocol, ozone should slowly recover. Currently,
there is no unequivocal evidence from measurements in the atmosphere that
the onset of ozone recovery has begun. Models suggest that in Antarctica
the peak in springtime ozone depletion may already have occurred or should
occur within the next few years. Because of greater variability, predicting the
timing of the peak in springtime Arctic ozone depletion is more uncertain but
models suggest it should occur within the first two decades of the 21st
century. Based on these model calculations, we do not expect an Arctic
ozone hole similar to that observed over the Antarctic.
The ozone layer will not necessarily return to its pre-depleted state, even
when the abundance of stratospheric chlorine and bromine returns to
previous levels. Recovery of the ozone layer is a complex issue: it depends
not just on the extent to which the ODSs are replaced by non-ODSs, but also
on emissions of gases (including ODS substitutes) that affect the climate
system directly. There are many complex two-way interactions between
stratospheric ozone and climate. Changes in stratospheric temperature and
transport affect the concentration and distribution of stratospheric ozone;
changes in tropospheric climate affect stratospheric circulation; changes in
stratospheric ozone influence the radiative forcing of the atmosphere, and
hence surface climate, as well as the chemistry of the troposphere. While our
understanding is still far from complete, new evidence about some of these
interactions has emerged in recent years.

Stratospheric ozone depletion has led to a cooling of the stratosphere. A


significant annual-mean cooling of the lower stratosphere over the past two
decades (of approximately 0.6 K per decade) has been found over the mid-

latitudes of both hemispheres. Modelling studies indicate that changes in


stratospheric ozone, well-mixed greenhouse gases and stratospheric water
vapour could all have contributed to these observed temperature changes. In
the upper stratosphere the annual mean temperature trends are larger, with
an approximately uniform cooling over 19791998 of roughly 2 K per decade
at an altitude of around 50 km. Model studies indicate that stratospheric
ozone changes and carbon dioxide changes are each responsible for about
50% of the upper stratospheric temperature trend.
The southern polar vortex, which creates the dynamical setting for the
Antarctic ozone hole, tends to persist longer now than in the decades before
the appearance of the ozone hole. During the period 19902000, the vortex
break-up in the late spring to early summer has been delayed by two to
three weeks relative to the 19701980 period. Future temperature changes
in the stratosphere could either enhance or reduce stratospheric ozone
depletion, depending on the region. Increases in the concentrations of wellmixed greenhouse gases, which are expected to cool the stratosphere, could
reduce the rate of gas-phase ozone destruction in much of the stratosphere
and hence reduce stratospheric ozone depletion. However ozone depletion in
the polar lower stratosphere depends on the low temperatures there. Until
stratospheric ODS abundances return to pre-ozone-hole levels, temperature
reductions in polar latitudes could enhance polar ozone depletion; this effect
is expected to be most important in the Arctic late winter to spring where a
small decrease in temperature could have a large effect on ozone.
The dynamical feedbacks from greenhouse-gas increases could either
enhance or reduce ozone abundance in these regions; currently, not even
the sign of the feedback is known. Furthermore, dynamical variability can
affect ozone in these regions on decadal time scales. Statistical and
modelling studies suggest that changes in stratospheric circulation regimes
(e.g., between strong and weak polar vortices) can have an impact on
surface climate patterns. Because future changes in ozone will have an
impact on the circulation of the stratosphere, these changes could also
influence the troposphere.
In particular there are indications that the Antarctic ozone hole has led to a
change in surface temperature and circulation including the cooling that
has been observed over the Antarctic continent, except over the Antarctic
Peninsula where a significant warming has been observed. Halocarbons are
particularly strong greenhouse gases (gases that absorb and emit thermal
infrared radiation); a halocarbon molecule can be many thousands of times

more efficient at absorbing radiant energy emitted from the Earth than a
molecule of carbon dioxide. Ozone is a greenhouse gas that also strongly
absorbs ultraviolet and visible radiation. Changes in these gases have
affected the Earths radiative balance in the past and will continue to affect it
in the future.
Over the period 17502000 halocarbon gases have contributed a positive
direct radiative forcing of about 0.33 W m2, which represents about 13% of
the total well-mixed greenhouse gas radiative forcing over this period. Over
the period 19702000 halocarbon gases represent about 23% of the total
well-mixed greenhouse gas radiative forcing. Over the same period, only
about 5% of the total halocarbon radiative forcing is from the ODS
substitutes; ODSs themselves account for 95%. Stratospheric ozone
depletion has led to a negative radiative forcing of about 0.15 W m2 with a
range of 0.1 W m2 from different model estimates. Therefore the positive
radiative forcing due to the combined effect of all ODSs (0.32 W m2) over
the period 17502000 is larger than the negative forcing due to stratospheric
ozone depletion. Assuming that the observed changes in ozone are caused
entirely by ODSs, the ozone radiative forcing can be considered an indirect
forcing from the ODSs. However, there is evidence that some fraction of the
observed global ozone changes cannot be attributed to ODSs, in which case
the indirect forcing by ODSs would be weaker.
The relative contributions of various halocarbons to positive direct and
negative indirect forcing differ from gas to gas. Gases containing bromine
(such as halons) are particularly effective ozone depleters, so their relative
contribution to negative forcing is greater than that of other ozone-depleting
gases, such as the CFCs. The same magnitudes of positive direct radiative
forcing from halocarbons and of negative indirect forcing from ozone
depletion are highly likely to have a different magnitude of impact on global
mean surface temperature. The climate responses will also have different
spatial patterns. As a consequence, the two forcings do not simply offset one
another.
Both the direct and indirect radiative forcings of ODSs will be reduced as a
result of the Montreal Protocol. Nevertheless, these forcings are expected to
remain signifi- cant during the next several decades. The precise change of
each of the forcings will be affected by several factors, including (1) the way
climate change may affect stratospheric ozone chemistry and hence the
ozone indirect radiative forcing; (2) future emissions of ODSs and their
substitutes (see Chapter 2); and (3) continuing emissions from ODS banks

(ODSs that have already been manufactured but have not yet been released
into the atmosphere).
Ozone absorbs solar ultraviolet (UV) radiation (thereby warming the
stratosphere) and is a greenhouse gas (thereby warming the troposphere). It
thus plays an important role in the climate system. Furthermore, absorption
of potentially damaging UV radiation by the stratospheric ozone layer
protects life at the Earths surface. Stratospheric ozone amounts have been
depleted in recent decades, following emission into the atmosphere of
various ozone-depleting substances (ODSs), most of which are also
greenhouse gases.
The CFCs (chlorofluorocarbons) and halons, the major anthropogenic
depleters of stratospheric ozone, are now controlled under the Montreal
Protocol and its Amendments and Adjustments. They are being replaced by a
variety of substances with lower ozone depletion potentials (ODPs) but that
are generally still greenhouse gases, often with large global warming
potentials (GWPs). The scientific connection between the different
environmental problems of climate change and ozone depletion, and
between the provisions of the Montreal Protocol and the Kyoto Protocol (and
the United Nations Framework Convention on Climate Change, UNFCCC), is
the subject of this chapter. The scientific issues are complex and the
connections between the two problems more intricate than suggested above.
For example, not only is ozone a greenhouse gas, so that any change in
ozone abundance could have an impact on climate, but changes in climate
could affect ozone in a number of different ways. So, the ozone distribution
in the future will depend, among other things, on the emission and impact of
other greenhouse gases and not just on those that deplete ozone. For this
reason, ozone recovery following the reduction in atmospheric abundances
of the ODSs will not be a simple return to the preozone-hole atmosphere
(WMO, 1999, Chapter 12). Climate change and ozone depletion have both
been the subject of recent assessments (IPCC, 2001; WMO, 2003) describing
the advances in understanding made in recent years.
It is now well established, for example, that observed polar and mid-latitude
ozone depletion is a consequence of the increase in stratospheric chlorine
and bromine concentrations. The chemical processes involving these
compounds and leading to polar ozone depletion are well understood.
Similarly, the important role of ozone and the ODSs in the climate system
has been documented. Stratospheric ozone depletion during recent decades

has represented a negative radiative forcing of the climate system; in


contrast, the increase in ODSs has been a positive radiative forcing. These
earlier reports form an important basis for this special report, which aims to
address specifically the coupling between the chemistry and climate
problems. In particular, this report will explore the scientific issues arising
from the phase-out and replacement of ODSs as they affect stratospheric
composition and climate change.
The earlier reports did not have this particular emphasis (although the issues
were addressed by a European Commission report (EC, 2003)). 1.1.2 Ozone
in the atmosphere and its role in climate Figure 1.1 illustrates a number of
important concepts concerning the ozone layer (Figure 1.1a) and its role in
the stratosphere (the region between approximately 15 and 50 km altitude in
which temperature rises with altitude, Figure 1.1b). Ozone, like water vapour
and carbon dioxide, is an important and naturally occurring greenhouse gas;
that is, it absorbs and emits radiation in the thermal infrared (Figure 1.1c),
trapping heat to warm the Earths surface (Figure 1.1d).
In contrast to the so-called wellmixed greenhouse gases (WMGHGs),
stratospheric ozone has two distinguishing properties. First, its relatively
short chemical lifetime means that it is not uniformly mixed throughout the
atmosphere and therefore its distribution is controlled by both dynamical and
chemical processes (Section 1.3). In fact, unlike the WMGHGs, ozone is
produced entirely within the atmosphere rather than being emitted into it.
Second (Figure 1.1c), it is a very strong absorber of short wavelength UV
radiation (it is also a weak absorber of visible radiation). The ozone layers
absorption of this UV radiation leads to the characteristic increase of
temperature with altitude in the stratosphere and, in consequence, to a
strong resistance to vertical motion.
As well as ozones role in climate it also has more direct links to humans: its
absorption of UV radiation protects much of Earths biota from this
potentially damaging short wavelength radiation. In contrast to the benefits
of stratospheric ozone, high surface ozone values are detrimental to human
health. The distribution of ozone in the atmosphere is maintained by a
balance between photochemical production and loss, and by transport
between regions of net production and net loss. A number of different
chemical regimes can be identified for ozone. In the upper stratosphere, the
ozone distribution arises from a balance between production following
photolysis of molecular oxygen and destruction via a number of catalytic

cycles involving hydrogen, nitrogen and halogen radical species (see Box
1.1). T
he halogens arise mainly from anthropogenic ODSs (the CFCs, HCFCs and
halons). In the upper stratosphere, the rates of ozone destruction depend on
temperature (a reduction in temperature slows the destruction of ozone, see
Section 1.3.2) and on the concentrations of the radical species. In the lower
stratosphere, reactions on aerosols become important. The distribution of the
radicals (and the partitioning of the nitrogen, hydrogen and halogen species
between radicals and reservoirs that do not destroy ozone) can be affected
by heterogeneous and multiphase chemistry acting on condensed matter
(liquid and solid particles). At the low temperature of the wintertime polar
lower stratosphere, this is the chemistry that leads to the ozone hole (see
Box 1.1).
The large-scale circulation of the stratosphere, known as the Brewer-Dobson
circulation (see Box 1.2), systematically transports ozone poleward and
downward (figure in Box 1.2). Because ozone photochemical reactions
proceed quickly in the sunlit upper stratosphere, this transport has little
effect on the ozone distribution there as ozone removal by transport is
quickly replenished by photochemical production.
The stratospheric Brewer-Dobson circulation consists of rising motion in the
tropics and sinking motion in the extratropics, together with an associated
poleward mass flux. Its effect on chemical species is complemented by
mixing (mainly quasi-horizontal) that acts to transport air parcels both
poleward and equatorward. Both processes are primarily driven by
mechanical forcing (wave drag) arising from the dissipation of planetaryscale waves in the stratosphere. These planetary waves are generated in the
troposphere by topographic and thermal forcing, and by synoptic
meteorological activity, and propagate vertically into the stratosphere.
Because of filtering by the large-scale stratospheric winds, vertical
propagation of planetary waves into the stratosphere occurs primarily during
winter, and this seasonality in wave forcing accounts for the winter
maximum in the Brewer-Dobson circulation.
In the case of ozone, the Brewer-Dobson circulation (together with the
associated horizontal mixing) transports ozone poleward and downward and
leads to a springtime maximum in extratropical ozone abundance (see
figure). Because air enters the stratosphere primarily in the tropics, the
physical and chemical characteristics of air near the tropical tropopause

behave as boundary conditions for the global stratosphere. For example,


dehydration of air near the cold tropical tropopause accounts for the extreme
dryness of the global stratosphere (Brewer, 1949). Overall, the region of the
tropical atmosphere between about 12 km and the altitude of the tropopause
(about 17 km) has characteristics intermediate to those of the troposphere
and stratosphere, and is referred to as the tropical tropopause layer (TTL).
This transport leads to significant variations of ozone in the extra-tropical
lower stratosphere, where the photochemical relaxation time is very long
(several months or longer) and ozone can accumulate on seasonal time
scales. Due to the seasonality of the BrewerDobson circulation (maximum
during winter and spring), ozone builds up in the extra-tropical lower
stratosphere during winter and spring through transport, and then decays
photochemically during the summer when transport is weaker. The
columnozone distribution (measured in Dobson units, DU) is dominated by its
distribution in the lower stratosphere and reflects this seasonality (Figure
1.2). Furthermore, planetary waves are stronger (and more variable) in the
Northern Hemisphere (NH) than in the Southern Hemisphere (SH), because of
the asymmetric distribution of the surface features (topography and land-sea
thermal contrasts) that, in combination with surface winds, force the waves.
Accordingly, the stratospheric BrewerDobson circulation is stronger during
the NH winter, and the resulting extra-tropical build-up of ozone during the
winter and spring is greater in the NH than in the SH (Andrews et al., 1987;
Figure 1.2). Variations in the Brewer-Dobson circulation also influence polar
temperatures in the lower stratosphere (via the vertical motions); stronger
wave forcing coincides with enhanced circulation and higher polar
temperatures (and more ozone transport).
Since temperature affects ozone chemistry, dynamical and chemical effects
on column ozone thus tend to act in concert and are coupled. Human
activities have led to changes in the atmospheric concentrations of several
greenhouse gases, including tropospheric and stratospheric ozone and ODSs
and their substitutes. Changes to the concentrations of these gases alter the
radiative balance of the Earths atmosphere by changing the balance
between incoming solar radiation and outgoing infrared radiation. Such an
alteration in the Earths radiative balance is called a radiative forcing. This
report, past IPCC reports and climate change protocols have universally
adopted the concept of radiative forcing as a tool to gauge and contrast
surface climate change caused by different mechanisms (see Box 1.3).
Positive radiative forcings are expected to warm the Earths surface and
negative radiative forcings are expected to cool it. The radiative forcings

from the principal greenhouse gases are summarized in Figure 1.3 and Table
1.1 (in Section 1.5.2). Other significant contributors to radiative forcing that
are not shown in Figure 1.3 include tropospheric aerosol changes and
changes in the Suns output. Changes in carbon dioxide (CO2 ) provide the
largest radiative forcing term and are expected to be the largest overall
contributor to climate change. In contrast with the positive radiative forcings
due to increases in other greenhouse gases, the radiative forcing due to
stratospheric ozone depletion is negative. Halocarbons are particularly
effective greenhouse gases in part because they absorb the Earths outgoing
infrared radiation in a spectral range where energy is not removed by carbon
dioxide or water vapour (sometimes referred to as the atmospheric window).
Halocarbon molecules can be many thousands of times more efficient at
absorbing the radiant energy emitted from the Earth than a molecule of
carbon dioxide, which explains why relatively small amounts of these gases
can contribute significantly to radiative forcing of the climate system.
Because halocarbons have low concentrations and absorb in the atmospheric
window, the magnitude of the direct radiative forcing from a halocarbon is
given by the product of its tropospheric mixing ratio and its radiative
efficiency. In contrast, for the more abundant greenhouse gases (carbon
dioxide, methane and nitrous oxide) there is a nonlinear relationship between
the mixing ratio and the radiative forcing. Since 1970 the growth in
halocarbon concentrations and the changes in ozone concentrations
(depletion in the stratosphere and increases in the troposphere) have been
very significant contributors to the total radiative forcing of the Earths
atmosphere.
Because halocarbons have likely caused most of the stratospheric ozone loss
(see Section 1.4.1), there is the possibility of a partial offset between the
positive forcing of halocarbon that are ODSs and the negative forcing from
stratospheric ozone loss. This is discussed further in Section 1.5. The climate
impacts of ozone changes are not confined to the surface: stratospheric
ozone changes are probably responsible for a significant fraction of the
observed cooling in the lower stratosphere over the last two decades
(Section 1.2.4) and may alter atmospheric dynamics and chemistry (Section
1.3.6).
Further, it was predicted that depletion of stratospheric ozone would lead to
a global increase in erythemal UV radiation (the radiation that causes
sunburn) at the surface of about 3%, with much larger increases at high
latitudes; these predicted high lat itude increases in surface UV dose have

indeed been observed (WMO, 2003, Chapter 5). 1.1.3 Chapter outline Our
aim in this chapter is to review the scientific understanding of the
interactions between ozone and climate. We interpret the term climate
broadly, to include stratospheric temperature and circulation, and refer to
tropospheric effects as tropospheric climate.
This is a broad topic; our review will take a more restricted view
concentrating on the interactions as they relate to stratospheric ozone and
the role of the ODSs and their substitutes. For this reason the role of
tropospheric ozone in the climate system is mentioned only briefly. Similarly,
some broader issues, including possible changes in stratospheric water
vapour (where the role of the ODSs and their substitutes should be minor)
and climate-dependent changes in biogenic emissions (whose first order
effect is likely to be in the troposphere), will not be discussed in any detail.
The relationship between ozone and the solar cycle is well established, and
the solar cycle has been used as an explanatory variable in ozone trend
analysis.
The connection between ozone and the solar cycle was assessed most
recently in Chapter 12 of WMO (1999) and will not be discussed further here.
Section 1.2 presents an update on stratospheric observations of ozone,
ozone-related species and temperature. Section 1.3 considers, at a process
level, the various feedbacks connecting stratospheric ozone and the climate
system. The understanding of these processes leads, in Section 1.4, to a
discussion of the attribution of past changes in ozone, the prediction of
future changes, and the connection of both with climate. Finally, Section 1.5
reviews the trade-offs between ozone depletion and radiative forcing,
focusing on ODSs and their substitutes. 1.2 Observed changes in the
stratosphere 1.2.1 Observed changes in stratospheric ozone Global and
hemispheric-scale variations in stratospheric ozone can be quantified from
extensive observational records covering the past 20 to 30 years.
There are numerous ways to measure ozone in the atmosphere, but they fall
broadly into two categories: measurements of column ozone (the vertically
integrated amount of ozone above the surface), and measurements of the
vertical profile of ozone. Approximately 90% of the vertically integrated
ozone column resides in the stratosphere. There are more independent
measurements, longer time-series, and better global coverage for column
ozone. Regular measurements of column ozone are available from a network
of surface stations, mostly in the mid-latitude NH, with reasonable coverage

extending back to the 1960s. Near-global, continuous column ozone data are
available from satellite measurements beginning in 1979.
The different observational data sets can be used to estimate past ozone
changes, and the differences between data sets provide a lower bound of
overall uncertainty. The differences indicate good overall agreement between
different data sources for changes in column ozone, and thus we have
reasonable confidence in describing the spatial and temporal characteristics
of past changes (WMO, 2003, Chapter 4). Five data sets of zonal and monthly
mean column ozone values developed by different scientific teams were
used to quantify past ozone changes in Chapter 4 of WMO (2003); they
include ground-based measurements covering 19642001, and several
different satellite data sets extending in time over 19792001. Figure 1.4a
shows globally averaged column ozone changes derived from each of these
data sets, updated through 2004. The analyses first remove the seasonal
cycle from each data set (mean and linear trend), and the deviations are
area weighted and expressed as anomalies with respect to the period 1964
1980. The global ozone amount shows decreasing values between the late
1970s and the early 1990s, a relative minimum during 19921994, and
slightly increasing values during the late 1990s. Global ozone for the period
19972004 was approximately 3% below the 19641980 average values.
Since systematic global observations began in the mid-1960s, the lowest
annually averaged global ozone occurred during 19921993 (5% below the
19641980 average). These changes are evident in each of the available
global data sets. The global ozone changes shown in Figure 1.4a occur
mainly in the extratropics of both hemispheres (poleward of 25 35). No
significant long-term changes in column ozone have been observed in the
tropics (25N to 25S).
Column ozone changes averaged over mid-latitudes (from 35 to 60
latitude) are significantly larger in the SH than in the NH; averaged for the
period 19972001, SH values are 6% below pre-1980 values, whereas NH
values are 3% lower. Also, there is significant seasonality to the NH midlatitude losses (4% losses in the winter-spring period and 2% in summer),
whereas long-term SH losses are about 6% year round (WMO, 2003, Chapter
4). The most dramatic changes in ozone have occurred during the spring
season over Antarctica, with the development during the 1980s of a
phenomenon known as the ozone hole (Figure 1.5).
The ozone hole now recurs every spring, with some interannual variability
and occasional extreme behaviour (see Box 1.6 in Section 1.4.1). In most

years the ozone concentration is reduced to nearly zero over a layer several
kilometres deep within the lower stratosphere in the Antarctic polar vortex.
Since the early 1990s, the average October column ozone poleward of (a 40
to 50% decrease), with up to a 70% local decrease for periods of a week or
so (WMO, 2003, Chapter 3; Figure 1.6). Compared with the Antarctic, Arctic
ozone abundance in the winter and spring is highly variable (Figure 1.6),
because of interannual variability in chemical loss and in dynamical
transport. Dynamical variability within the winter stratosphere leads to
changes in ozone transport to high latitudes, and these transport changes
are correlated with polar temperature variability with less ozone transport
being associated with lower temperatures. Low temperatures favour
halogen-induced chemical ozone loss.
Thus, in recent decades, halogen-induced polar ozone chemistry has acted in
concert with the dynamically induced ozone variability, and has led to Arctic
column ozone losses of up to 30% in particularly cold winters (WMO, 2003,
Chapter 3). In particularly dynamically active, warm winters, however, the
estimated chemical ozone loss has been very small. Changes in the vertical
profile of ozone are derived primarily from satellites, ground-based
measurements and balloon-borne ozonesondes. Long records from groundbased and balloon data are available mainly for stations over NH midlatitudes. Ozone profile changes over NH mid-latitudes exhibit two maxima,
with decreasing trends in the upper stratosphere (about 7% per decade at
3545 km) and in the lower stratosphere (about 5% per decade at 1525 km)
during the period 19792000. Ozone profile trends show a minimum (about
2% per decade decrease) near 30 km. The vertically integrated profile trends
are in agreement with the measured changes in column ozone (WMO, 2003,
Chapter 4).
As a result of reduced emissions because of the Montreal Protocol and its
Amendments and Adjustments, mixing ratios for most ODSs have stopped
increasing near the Earths surface (Montzka et al., 1999; Prinn et al., 2000).
The response to the Protocol, however, is reflected in quite different
observed behaviour for different substances. By 2003, the mixing ratios for
CFC-12 were close to their peak, CFC-11 had clearly decreased, while methyl
chloroform (CH3 CCl3 ) had dropped by 80% from its maximum (Figure 1.7;
see also Chapter 1 of WMO, 2003). Halons and HCFCs are among the few
ODSs whose mixing ratios were still increasing in 2000 (Montzka et al., 2003;
ODoherty et al., 2004). Halons contain bromine, which is on average 40 to
50 times more efficient on a per-atom basis at destroying stratospheric
ozone than chlorine. However, growth rates for most halons have steadily

decreased during recent years (Fraser et al., 1999; Montzka et al., 2003).
Furthermore, increases in tropospheric bromine from halons have been offset
by the decline observed for methyl bromide (CH3 Br) since 1998.
(Yokouchi et al., 2002; Montzka et al., 2003; Simmonds et al., 2004).
Atmospheric amounts of HCFCs continue to increase because of their use as
CFC substitutes (Figure 1.7b). Chlorine from HCFCs has increased at a fairly
constant rate of 10 ppt Cl yr1 since 1996, although HCFCs accounted for
only 5% of all chlorine from long-lived gases in the atmosphere by 2000
(WMO, 2003, Chapter 1). The ODPs of the most abundant HCFCs are only
about 510% of those of the CFCs (Table 1.2 in Section 1.5.3). Figure 1.8
contrasts the observed atmospheric abundances of CFC-11 with those of
HCFC-22 and HFC-134a. The behaviour of CFC-11 is representative of the
behaviour of CFCs in general, and the behaviour of HCFC-22 and HFC-134a is
representative of HCFCs (as well as HFCs) in general (see Figure 2.3 in
Chapter 2). The most rapid growth rates of CFC-11 occurred in the 1970s and
1980s.
The largest emissions were in the NH, and concentrations in the SH lagged
behind those in the NH, consistent with an inter-hemispheric mixing time
scale of about 1 to 2 years. In recent years, following the implementation of
the Montreal Protocol, the observed growth rate has declined, concentrations
appear to be at their peak and, as emissions have declined, the interhemispheric gradient has almost disappeared. In contrast, HCFC-22 and HFC134a concentrations are still growing rapidly and there is a marked interhemispheric gradient. Ground-based observations suggest that by 2003 the
cumulative totals of both chlorine- and bromine-containing gases regulated
by the Montreal Protocol were decreasing in the lower atmosphere (Montzka
et al., 2003).
Although tropospheric chlorine levels peaked in the early 1990s and have
since declined, atmospheric bromine began decreasing in 1998 (Montzka et
al., 2003). The net effect of changes in the abundance of both chlorine and
bromine on the total ozone-depleting halogens in the stratosphere is
estimated roughly by calculating the equivalent effective stratospheric
chlorine (EESC; Daniel et al., 1995) (Figure 1.7d). The calculation of EESC
(see Box 1.8 in Section 1.5.3) includes consideration of the total amount of
chlorine and bromine accounted for by long-lived halocarbons, how rapidly
these halocarbons degrade and release their halogen in the stratosphere and
a nominal lag time of three years to allow for transport from the troposphere
into the stratosphere.

The tropospheric observational data suggest that EESC peaked in the mid1990s and has been decreasing at a mean rate of 22 ppt yr1 (0.7% yr1)
over the past eight years (Chapter 1 of WMO, 2003). Direct stratospheric
measurements show that stratospheric chlorine reached a broad plateau
after 1996, characterized by variability (Rinsland et al., 2003).
1.2.3 Observed changes in stratospheric aerosols, water vapour, methane
and nitrous oxide In addition to ODSs, stratospheric ozone is influenced by
the abundance of stratospheric aerosols, water vapour, methane (CH4 ) and
nitrous oxide (N2 O). Observed variations in these constituents are
summarized in this section. During the past three decades, aerosol loading in
the stratosphere has primarily reflected the effects of a few volcanic
eruptions that inject aerosol and its gaseous precursors (primarily sulphur
dioxide, SO2 ) into the stratosphere. The most noteworthy of these eruptions
are El Chichn (1982) and Pinatubo (1991). The 1991 Pinatubo eruption likely
had the largest impact of any event in the 20th century (McCormick et al.,
1995), producing about 30 Tg of aerosol (compared with approximately 12 Tg
from El Chichn) that persisted into at least the late 1990s. Current aerosol
loading, which is at the lowest observed levels, is less than 0.5 Tg, so the
Pinatubo event represents nearly a factor of 100 enhancement relative to
non-volcanic levels.
The source of the non-volcanic stratospheric aerosol is primarily carbonyl
sulfide (OCS), and there is general agreement between the aerosols
estimated by modelling the transformation of observed OCS to sulphate
aerosols and observed aerosols. However, there is a significant dearth of SO2
measurements, and the role of tropospheric SO2 in the stratospheric aerosol
budget while significant remains a matter of some uncertainty. Because of
the high variability of stratospheric aerosol loading it is difficult to detect
trends in the non-volcanic aerosol component. Trends derived from the late
1970s to the current period are likely to encompass a value of zero. The
recent Stratospheric Processes And their Role in Climate (SPARC) Assessment
of Upper Tropospheric and Stratospheric Water Vapour (SPARC, 2000)
provided an extensive review of data sources and quality for stratospheric
water vapour, together with detailed analyses of observed seasonal and
interannual variability.
The longest continuous reliable data set is at a single location (Boulder,
Colorado, USA), is based on balloon-borne frost-point hygrometer
measurements (approximately one per month), and dates back to 1980.
Over the period 19802000, a statistically significant positive trend of

approximately 1% yr1 is observed at all levels between about 15 and 26 km


in altitude (SPARC, 2000; Oltmans et al., 2000). However, although a linear
trend can be fitted to these data, there is a high degree of variability in the
infrequent sampling, and the increases are neither continuous nor steady
(Figure 1.9). Long-term increases in stratospheric water vapour are also
inferred from a number of other ground-based, balloon, aircraft and satellite
data sets spanning approximately 19802000 (Rosenlof et al., 2001),
although the time records are short and the sampling uncertainty is high in
many cases.
Global stratospheric water vapour measurements have been made by the
Halogen Occultation Experiment (HALOE) satellite instrument for more than a
decade (late 1991 to 2004). Interannual changes in water vapour derived
from HALOE data show excellent agreement with Polar Ozone and Aerosol
Measurement (POAM) satellite data, and also exhibit strong coherence with
tropical tropopause temperature changes (Randel et al., 2004). An updated
comparison of the HALOE measurements with the Boulder balloon data for
the period 19922004 is shown in Figure 1.9. The Boulder and HALOE data
show reasonable agreement for the early part of the record (1992 1996),
but there is an offset after 1997, with the Boulder data showing higher
values than HALOE measurements. As a result, linear trends derived from
these two data sets for the (short) period 19922004 show very different
results (increases for the Boulder data, but not for HALOE).
The reason for the differences between the balloon and satellite data (for the
same time period and location) is unclear at present, but the discrepancy
calls into question interpretation of water vapour trends derived from short
or infrequently sampled data records. It will be important to reconcile these
differences because these data sets are the two longest and most continuous
data records available for stratospheric water vapour. It is a challenge to
explain the magnitude of the water vapour increases seen in the Boulder
frost-point data. Somewhat less than half the observed increase of about
10% per decade (through 2001) can be explained as a result of increasing
tropospheric methane (transported to the stratosphere and oxidized to form
water vapour).
The remaining increase could be reconciled with a warming of the tropical
tropopause of approximately 1 K per decade, assuming that air entered the
stratosphere with water vapour in equilibrium with ice. However,
observations suggest that the tropical tropopause has cooled slightly for this
time period, by approximately 0.5 K per decade, and risen slightly in altitude

by about 20 m per decade (e.g., Seidel et al., 2001; Zhou et al., 2001).
Although regional-scale processes may also influence stratospheric water
vapour (such as summer monsoon circulations, e.g., Potter and Holton,
1995), there is no evidence for increases in tropopause temperature in these
regions either (Seidel et al., 2001). From this perspective, the extent of the
decadal water vapour increases inferred from the Boulder measurements is
inconsistent with the observed tropical tropopause cooling, and this
inconsistency limits confidence in predicting the future evolution of
stratospheric water vapour.
The atmospheric abundance of methane has increased by a factor of about
2.5 since the pre-industrial era (IPCC, 2001, Chapter 4). Measurements of
methane from a global monitoring network showed increasing values
through the 1990s, but approximately constant values during 19992002
(Dlugokencky et al., 2003). Changes in stratospheric methane have been
monitored on a global scale using HALOE satellite measurements since 1991.
The HALOE data show increases in lower stratospheric methane during 1992
1997 that are in reasonable agreement with tropospheric increases during
this time (Randel et al., 1999). In the upper stratosphere the HALOE data
show an overall decrease in methane since 1991, which is likely attributable
to a combination of chemical and dynamical influences (Nedoluha et al.,
1998; Considine et al., 2001; Rckmann et al., 2004). Measurements of
tropospheric N2 O show a consistent increase of about 3% per decade (IPCC,
2001, Chapter 4).
Because tropospheric air is transported into the stratosphere, these positive
N2 O trends produce increases in stratospheric reactive nitrogen (NOy ),
which plays a key role in ozone photochemistry. Measurements of
stratospheric column NO2 (a main component of NOy ) from the SH (1980
2000) and the NH (19852000) mid-latitudes (about 45) show long-term
increases of approximately 6% per decade (Liley et al., 2000; WMO, 2003,
Chapter 4), and these are reconciled with the N2 O changes by considering
effects of changing levels of stratospheric ozone, water vapour and halogens
(Fish et al., 2000; McLinden et al., 2001). 1.2.4 Observed temperature
changes in the stratosphere There is strong evidence of a large and
significant cooling in most of the stratosphere since 1980. Recent updates to
the observed changes in stratospheric temperature and to the understanding
of those changes have been presented in Chapters 3 and 4 of WMO (2003).
Current long-term monitoring of stratospheric temperature relies on satellite
instruments and radiosonde analyses.

The Microwave Sounding Unit (MSU) and Stratospheric Sounding Unit (SSU)
instruments record temperatures in several 1015 km thick layers between
17 and 50 km in altitude. Radiosonde trend analyses are available up to
altitudes of roughly 25 km. Determining accurate trends with these data sets
is difficult. In particular the radiosonde coverage is not global and suffers
from data quality concerns, whereas the satellite trend data is a result of
merging data sets from several different instruments. Figure 1.4b shows
global temperature time-series in the lower and middle stratosphere derived
from satellites and radiosondes. It reveals a strong imprint of 1 to 2 years of
warming following the volcanic eruptions of Agung (1963), El Chichn (1982)
and Mt. Pinatubo (1991).
When these years are excluded from long-term trend analyses, a significant
global cooling is seen in both the radiosonde and the satellite records over
the last few decades. This cooling is significant (at the 97% level or greater)
at all levels of the stratosphere except the 30 km (10 hPa) level in the SSU
record (Figure 1.10). The largest global cooling is found in the upper
stratosphere, where it is fairly uniform in time at a rate of about 2 K per
decade (for 19791997). In the lower stratosphere (below 25 km) this longterm global cooling manifests itself as more of a step-like change following
the volcanic warming events (Figure 1.4b). The cooling also varies with
latitude.
The lower stratosphere extratropics (from 20 to 60 latitude) show a cooling
of 0.40.8 K per decade in both hemispheres, which remains roughly
constant throughout the year. At high latitudes most of the cooling, up to 2 K
per decade for both hemispheres, occurs during spring. Much recent
progress has been made in modelling these temperature trends (see Shine et
al., 2003; WMO, 2003, Chapters 3 and 4). Models range from onedimensional (1-D) fixed dynamical heating rate (FDH) calculations to threedimensional (3-D) coupled chemistry-climate models; many of their findings
are presented later in this chapter. FDH is a simple way of determining the
radiative response to an imposed change whilst fixing the background
dynamical heating to its climatological value; this makes the calculation of
temperature change simpler, as only a radiation model needs to be used.
Changes in dynamical circulations simulated by models can lead to effects
over latitudinal bands, which vary between models.
However, dynamics cannot easily produce a global mean temperature
change. Because of this, global mean temperature is radiatively controlled
and provides an important focus for attribution (Figure 1.10). For the global

mean in the upper stratosphere the models suggest roughly equal


contributions to the cooling from ozone decreases and carbon dioxide
increases. In the global mean mid-stratosphere there appears to be some
discrepancy between the SSU and modelled trends near 30 km: models
predict a definite radiative cooling from carbon dioxide at these altitudes,
which is not evident in the SSU record. In the lower stratosphere the cooling
from carbon dioxide is much smaller than higher in the stratosphere.
Although ozone depletion probably accounts for up to half of the observed
cooling trend, it appears that a significant cooling from another mechanism
may be needed to account for the rest of the observed cooling. One possible
cause of this extra cooling could be stratospheric water vapour increases.
However, stratospheric water vapour changes are currently too uncertain
(see Section 1.2.3) to pinpoint their precise role. Tropospheric ozone
increases have probably contributed slightly to lower stratospheric cooling by
reducing upwelling thermal radiation; one estimate suggests a cooling of
0.05 K per decade at 50 hPa and a total cooling of up to 0.5 K over the last
century (Sexton et al., 2003).
The springtime cooling in the Antarctic lower stratosphere is almost certainly
nearly all caused by stratospheric ozone depletion. However, the similar
magnitude of cooling in the Arctic spring does not seem to be solely caused
by ozone changes, which are much smaller than in the Antarctic; interannual
variability may contribute substantially to the cooling observed in the Arctic
wintertime (see Section 1.4.1). One mechanism for altering temperatures in
the upper troposphere and lower stratosphere comes from the direct
radiative effects of halocarbons (WMO, 1986; Ramaswamy et al., 1996;
Hansen et al., 1997). In contrast with the role of carbon dioxide, halocarbons
can actually warm the upper tropospheric and lower stratospheric region
(see Box 1.4).
In summary, stratospheric temperature changes over the past few decades
are significant (they are, in fact, substantially larger than those seen in the
surface temperature record over the same period), and there are clear
quantifiable features of the contributions from ozone, carbon dioxide and
volcanism in the past stratospheric temperature record. More definite
attribution of the causes of these trends is limited by the short timeseries.
Future increases of carbon dioxide can be expected to substantially cool the
upper stratosphere. However, this cooling could be partially offset by any
future ozone increase (Section 1.4.2). In the lower stratosphere, both ozone

recovery and halocarbon increases would warm this region compared with
the present.
Any changes in stratospheric water vapour or changes in tropospheric
conditions, such as high-cloud properties and tropospheric ozone, would also
affect future temperatures in the lower stratosphere. Furthermore, circulation
changes can affect temperatures over sub-global scales, especially at
midand polar latitudes. Several studies have modelled parts of these
expected temperature changes, and are discussed in the rest of this chapter
(see especially Section 1.4.2). 1.3 Stratospheric ozone and climate feedback
processes The distribution of ozone depends on a balance between chemical
processes, which can be affected by changes in the concentration of the
ODSs, and transport processes, which can be affected by climate change.
Climate change, in its broadest sense, can also affect ozone chemistry
directly by modifying the rates of temperature-dependent reactions. These
interactions are discussed in Sections 1.3.1 to 1.3.4. In these sections we
also discuss the possible impact on ozone of changing stratospheric water
vapour abundances, which may be regulated by climate. Changes in
stratospheric ozone can also affect climate. The impact of ozone changes on
stratospheric temperatures has already been discussed in Section 1.2.4. As
well as the direct impact of ozone as an absorber of UV radiation and as a
greenhouse gas, there are other effects that could be important. For
example, changes in ozone could affect the lifetime of reactive.
Halocarbons and carbon dioxide are greenhouse gases: they trap longwave
radiation to warm the Earths surface. However, compared with carbon
dioxide halocarbons have quite different spectral absorption characteristics
and they interact very differently with the Earths radiation field. The very
strong 15 m band dominates the role of carbon dioxide, whereas the
halocarbons tend to absorb weakly in the 813 m atmospheric window, a
region of the spectrum where other gases have only a small effect on
outgoing longwave radiation. These absorption properties, combined with a
typical vertical temperature profile, means that halocarbons usually warm
the atmosphere locally, whereas carbon dioxide generally cools it (the
atmosphere only warms as a response to the induced surface warming).
Further, this effect is largest at the tropical tropopause (about 17 km
altitude), where temperatures are most different from those of the
underlying surface (e.g., Dickinson, 1978; Wang et al., 1991). The tropical
tropopause can be defined as the height at which the coldest temperatures

are found in a vertical temperature profile (see Figure 1.1b). The effects of
halocarbon, carbon dioxide and ozone changes are contrasted in the figure in
this box. Panel (a) shows that halocarbons may have warmed the tropical
upper troposphere and lower stratosphere by as much as 0.3 K, which is
locally larger than the cooling effects of carbon dioxide.
For the calculation of a globally averaged temperature change this warming
can be thought of as partially cancelling out the cooling effect of ozone
depletion in the extra-tropical lower stratosphere. However, the patterns are
quite distinct (compare panels (a) and (b)), and as a result the equator-topole
temperature gradient in the lower stratosphere, where temperature
increases towards higher latitudes, would be reduced. Although halocarbons
are potentially important, because of their coupling to water vapour, it is
unlikely that halocarbons dominate the response of the tropical tropopause
to changing greenhouse gases, as there is some observational evidence for a
general cooling of the tropical tropopause over the last few decades (Seidel
et al., 2001; Zhou et al., 2001).
greenhouse gases, by changing the penetration of UV radiation. Changes in
the structure of the stratosphere, caused by ozone changes, could alter the
interaction between the troposphere and stratosphere and lead to further
changes in stratospheric ozone. These latter interactions are discussed in
Sections 1.3.5 and 1.3.6. Improved knowledge of these various feedback
processes is essential for informing the numerical models used to predict
future chemistry-climate interactions; these models are discussed in Section
1.4. Note that other factors, which are not discussed in detail here, can also
affect the interaction between ozone and climate. Perhaps the most obvious
is the impact of major volcanic eruptions. These can lead to an increase in
volcanic aerosol in the stratosphere, which can influence both climate and
the chemical processes controlling the ozone layer. 1.3.1 Impact of ODSs on
stratospheric ozone.
The abundance of ozone in the stratosphere at a particular location is
governed by three processes: photochemical production, destruction by
catalytic cycles, and transport processes. Photochemical production in the
stratosphere occurs mostly through the photolysis of O2 , with loss via
catalytic cycles involving hydrogen (HOx ), nitrogen (NOx ), and halogen
(ClOx , BrOx ) radicals (see Box 1.1). The relative importance of the various
loss cycles in the stratosphere varies substantially with altitude (Figure 1.11).
Above about 45 km, loss through HOx dominates, while below this altitude
NOx -catalyzed ozone loss is most important. The importance of the ClOx

-catalyzed ozone loss cycle, which varies with chlorine loading, peaks at
about 40 km. Below about 25 km HOx -driven ozone loss cycles dominate
again. 1.3.1.1 Upper stratosphere In the upper stratosphere the ozone
budget is largely understood, although uncertainties remain regarding the
rate constants of key radical reactions.
In particular, strong evidence has been accumulated that the observed
ozone depletion in the upper stratosphere is caused by increased levels of
stratospheric chlorine (WMO, 1999, Chapter 6; WMO, 2003, Chapter 4), as
originally proposed by Molina and Rowland (1974) and Crutzen (1974).
Because of the direct correspondence between the stratospheric chlorine
abundance and ozone depletion in the upper stratosphere, it has been
suggested (e.g., WMO, 1999, Chapter 12) that the response of stratospheric
ozone to the declining stratospheric chlorine levels might be first detectable
in the upper stratosphere (see also Box 1.7 in Section 1.4.2). 1.3.1.2 Polar
regions In recent decades stratospheric ozone losses have been most
pronounced in polar regions during winter and spring (WMO 2003, Chapter 3;
see also Figures 1.5 and 1.6). These losses are determined largely by three
chemical factors: (1) the conversion of chlorine reservoirs into active, ozonedestroying forms through heterogeneous reactions on the surfaces of polar
stratospheric cloud particles; (2) the availability of sunlight that drives the
catalytic photochemical cycles that destroy ozone; and (3) the timing of the
deactivation of chlorine (i.e., the timing of the conversion of active chlorine
back to the reservoir species). Temperature controls the formation and
destruction of polar stratospheric clouds (PSCs) and thus the timing of
activation (1) and deactivation (3) of chlorine. Furthermore, temperature
controls the efficiency of the catalytic cycles (especially the ClO dimer cycle;
see Box 1.1) that destroy ozone in the presence of sunlight (2).
However, polar ozone loss would not occur without a prominent dynamical
feature of the stratosphere in winter and spring: the polar vortex. In both
hemispheres, the polar vortex separates polar air from mid-latitude air to a
large extent, and within the vortex the low-temperature conditions that
develop are the key factor for polar ozone loss. These two factors are
dynamically related: a strong vortex is generally also a colder vortex. The
two crucial questions for future polar ozone are whether, in an increased
greenhouse-gas climate, the region of low stratospheric temperatures will
increase in area and whether it will persist for longer in any given year (see
Section 1.3.2).

When anthropogenically emitted ODSs are eventually removed from the


stratosphere, the stratospheric halogen burden will be much lower than it is
today and will be controlled by the naturally occurring source gases methyl
chloride (CH3 Cl) and methyl bromide (CH3 Br). Under such conditions,
dramatic losses of polar ozone in winter and spring as we see today are not
expected to occur. However, the rate of removal of anthropogenic halogens
from the atmosphere depends on the atmospheric lifetimes of CFCs (typically
50100 years) and is considerably slower than the rate at which halogens
have been increasing in the decades prior to approximately the year 2000,
when the stratospheric halogen loading peaked (Figure 1.7). A complete
removal of anthropogenic halogens from the atmosphere will take more than
a century.
During this period of enhanced levels of halogens caused by past
anthropogenic emissions, the polar stratosphere will remain vulnerable to
climate perturbations, such as increasing water vapour or a cooling of the
stratosphere, that lead to enhanced ozone destruction. 1.3.1.3 Lowerstratospheric mid-latitudes It is well established that ozone in the lower
stratosphere at midlatitudes has been decreasing for a few decades; both
measurements of ozone locally in the lower stratosphere and measurements
of column ozone (a quantity that is dominated by the amount of ozone in the
lower stratosphere) show a clear decline (WMO, 2003, Chapter 4).
However, because mid-latitude ozone loss is much less severe than polar
ozone loss, it cannot be identified in measurements in any one year but is
rather detected as a downward trend in statistical analyses of longer timeseries. It is clear that chemical loss driven by halogens is very important, but
other possible effects that may contribute to the observed mid-latitude
trends have also been identified. A definitive quantitative attribution of the
trends to particular mechanisms has not yet been achieved. Chapter 4 of
WMO (2003) reviewed this issue most recently; we will not repeat that
detailed analysis here but instead provide a brief summary.
The chemical processes that may affect trends of mid-latitude ozone are
essentially related to the ODS trends that are known to be responsible for the
observed ozone loss in the upper stratosphere and in the polar regions.
(Transport effects are discussed in Section 1.3.4.) Halogen chemistry may
lead to ozone depletion in the mid-latitude lower stratosphere through a
number of possible mechanisms, including: 1. Export of air that has
encountered ozone destruction during the winter from the polar vortex (e.g.,
Prather et al., 1990). 2. Export of air with enhanced levels of active chlorine

from the polar vortex (e.g., Prather and Jaffe, 1990; Norton and Chipperfield,
1995).
3. In situ activation of chlorine either on cold liquid sulfate aerosol particles
(e.g., Keim et al., 1996) or on ice particles (e.g., Borrmann et al., 1997;
Bregman et al., 2002). Further, the reaction of N2 O5 with water on liquid
aerosol particles at higher temperatures indirectly enhances the
concentrations of ClO at mid-latitudes (McElroy et al., 1992). 4. Ozone
depletion due to elevated levels of BrO in the lower stratosphere, possibly
caused by transport of very shortlived halogen-containing compounds or BrO
across the tropopause (WMO, 2003, Chapter 2). The first mechanism results
from transport combined with polar ozone loss, whereas the remaining three
mechanisms all involve in situ chemical destruction of ozone at midlatitudes. All four mechanisms listed above are ultimately driven by the
increase of halogens in the atmosphere over the past decades.
Thus, while Millard et al. (2002) emphasized the strong interannual variability
in the relative contributions of the different mechanisms to the seasonal midlatitude ozone loss during the winter-spring period, they nonetheless showed
that ozone loss driven by catalytic cycles involving halogens was always an
important contributor (4070%) to the simulated mid-latitude ozone loss in
the five winters in the 1990s that they studied. Similarly, Chipperfield (2003)
found that the observed midlatitude column ozone decrease from 1980 to
the early 1990s could be reproduced in long-term simulations in a 3-D
chemistry-transport model (CTM).
The modelled ozone is affected by dynamical interannual variability, but the
overall decreases are dominated by halogen trends; and about 3050% of
the modelled halogen-induced change is a result of high latitude processing
on PSCs. Under climate change the strength of the polar vortex may change
(see Section 1.3.2), but the sign of this change is uncertain. A change in the
strength and temperature of the vortex will affect chlorine activation and
ozone loss there and, through mechanisms (1) and (2), ozone loss in midlatitudes. The possibility that chlorine might be activated on cirrus clouds or
on cold liquid aerosol particles in the lowermost stratosphere (mechanism
(3)) was revisited recently by Bregman et al. (2002). Based on their model
results it seems unlikely that this process is the main mechanism for the
observed long-term decline of ozone in the mid-latitude lower stratosphere.
Very short-lived organic chlorine-, bromine-, and iodine-containing
compounds possess a potential to deplete stratospheric ozone.

However a quantitative assessment of their impact on stratospheric ozone is


made difficult by their short lifetime, so there is need to consider the
transport pathways from the troposphere to the stratosphere of these
compounds in detail (WMO, 2003, Chapter 2). An upper limit for total
stratospheric iodine, Iy , of 0.10 0.02 ppt was recently reported (for below
20 km) by Bsch et al. (2003). The impact of this magnitude of iodine loading
on stratospheric ozone is negligible. The disturbance of the mid-latitude
ozone budget caused by anthropogenic emissions of ODSs will ultimately
cease when the stratospheric halogen burden has reached low enough levels
(see Section 1.2.2 and Box 1.7 in Section 1.4.2). However, like polar ozone,
mid-latitude ozone will for many decades remain vulnerable to an
enhancement of halogen-catalyzed ozone loss caused by climate change and
by natural phenomena such as volcanic eruptions. 1.3.2 Impact of
temperature changes on ozone chemistry
The increasing abundance of WMGHGs in the atmosphere is expected to
lead to an increase in temperature in the troposphere. Furthermore,
increasing concentrations of most of these gases, notably CO2 , N2 O and
CH4 , are expected to lead to a temperature decrease in the stratosphere. By
far the strongest contribution to stratospheric cooling from the WMGHGs
comes from CO2 (Section 1.2.4). As noted in Section 1.2.4, temperatures in
the stratosphere are also believed to have decreased in part because of the
observed reductions in ozone concentrations; any such cooling would have a
feedback on the ozone changes.
Decreasing stratospheric temperatures lead to a reduction of the ozone loss
rate in the upper stratosphere, thereby indirectly leading to more ozone in
this region. This reduction of ozone loss is caused by the very strong positive
temperature dependence of the ozone loss rate, mainly owing to the
Chapman reactions and the NOx cycle (Figure 1.12). This inverse relationship
between ozone and temperature changes in the upper stratosphere has been
known for many years (Barnett et al., 1975). More recently, differences in
temperature between the two hemispheres have been identified (Li et al.,
2002) as a cause of inter-hemispheric differences in both the seasonal cycle
of upper-stratospheric ozone abundances and in the upper-stratospheric
ozone trend deduced from satellite measurements.
In the mid-latitude lower stratosphere, one mechanism that leads to ozone
loss and is directly sensitive to temperature changes involves heterogeneous
reactions on the surfaces of cloud and cold aerosol particles (see mechanism
(3) in Section 1.3.1.3). The rates of many of these reactions increase strongly

with decreasing temperature. Similarly, the reaction rates increase with


increasing water-vapour concentrations, so that future increases in water,
should they occur, would also have an impact (see Sections 1.2.3 and 1.3.3).
Note, however, that in the lower-stratospheric mid-latitudes the most
important heterogeneous reactions are hydrolysis of N2 O5 and bromine
nitrate, which are relatively insensitive to temperature and water vapour
concentrations.
Therefore, the expected impact of climate change on the chemical
mechanism (3) is expected to be relatively small. In addition to this impact
on heterogeneous chemistry, Zeng and Pyle (2003) have argued that a
reduction in temperature in the lower stratosphere can slow the rate of HOx
-driven ozone destruction. Polar ozone loss occurs when temperatures in a
large enough region sink below the threshold temperature for the existence
of PSCs (approximately 195 K), because chlorine is activated by
heterogeneous reactions on the surfaces of PSCs. Therefore, a cooling of the
stratosphere enhances Arctic ozone loss if the volume of polar air with
temperatures below the PSC threshold value increases. (The Antarctic in
winter and spring is already consistently below this threshold.)
Moreover, a stronger PSC activity is expected to lead to a greater
denitrification of the Arctic vortex, hence a slower deactivation of chlorine
and, consequently, a greater chemical ozone loss (Waibel et al., 1999;
Tabazadeh et al., 2000). Rex et al. (2004) have deduced an empirical relation
between observed temperatures and observed winter-spring chemical loss of
Arctic ozone (Figure 1.13). Based on this relation, and for current levels of
chlorine, about 15 DU additional loss of total ozone is expected for each 1 K
cooling of the Arctic lower stratosphere. Although the lower stratosphere is
expected to generally cool with increasing greenhouse-gas concentrations
(Section 1.2.4), the temperature changes in the lower stratosphere dur ing
the polar wintertime will be sensitive to any change in circulation associated
with dynamical feedbacks, principally from changes in planetary wave drag
(see Section 1.3.4.1).
In principle, these feedbacks could be of either sign and thus could lead to
enhanced cooling or even to warming. An early model study found enhanced
cooling, to the extent that polar ozone loss would be expected to increase
during the next 10 to 15 years even while halogen levels decreased (Shindell
et al., 1998). More recent studies with higher-resolution models have found a
much less dramatic dynamical feedback, with some models showing an
increase and some a decrease in planetary wave drag, and with all models

predicting a relatively small change in Arctic ozone over the next few
decades (Austin et al., 2003). However, it should be noted that model results
for the Arctic are difficult to assess because the processes leading to polar
ozone depletion show so much natural variability that the atmosphere may
evolve anywhere within (or even outside) the envelope provided by an
ensemble of model simulations (see Section 3.5 of WMO, 2003).
In any event, it is the Arctic that is most sensitive to the effects caused by
climate change that are discussed above. Present-day temperatures in the
Arctic lie close to the threshold value for the onset of heterogeneous chlorine
activation and thus close to the threshold value for the onset of rapid ozone
loss chemistry; in the Antarctic temperatures are much lower and thus ozone
loss is not as sensitive to temperature changes. 1.3.3 Impact of methane and
water vapour changes on ozone chemistry With the exception of the high
latitudes in winter, ozone in the upper stratosphere is in a photochemical
steady-state; photochemical reactions are sufficiently fast that ozone
concentrations are determined by a local balance between photochemical
production and loss. Nonetheless, transport has a significant indirect
influence on upper stratospheric ozone insofar as it determines the
concentrations of trace compounds such as CH4 , H2 O, CFCs and N2 O, all of
which act as precursors of the radicals that determine the ozone chemistry.
Furthermore, CH4 is of particular importance because it is the primary
mechanism for the conversion of reactive Cl to the unreactive HCl reservoir
via the reaction CH4 + Cl HCl + CH3 , and thus affects the efficiency of the
chlorine-driven ozone loss. Changes in upper stratospheric CH4 (Section
1.2.3) have important implications for upper stratospheric ozone. Siskind et
al. (1998) found that for the period 19921995, the increase in active
chlorine (Cl and ClO) resulting from the CH4 decrease was the largest
contributor to the ozone changes occurring over that time.
Indeed, a measured increase in upper-stratospheric ClO between 1991 and
1997, which is significantly greater than that expected from the increase in
the chlorine source gases alone, may be explained by the observed
concurrent decrease of CH4 (Froidevaux et al., 2000). Further, Li et al. (2002)
find that inter-hemispheric differences in CH4 are partly responsible
(together with inter-hemispheric differences in temperature) for observed
differences in upper stratospheric ozone trends between the two
hemispheres (from the SAGE I and SAGE II satellite experiments for 1979
1997). Stratospheric water vapour is the primary source of the HOx radicals
that drive the dominant ozone loss cycles in the upper stratosphere (Figure

1.11). An increase in stratospheric water vapour is therefore expected to lead


to a greater chemical loss of ozone in the upper stratosphere (e.g., Siskind et
al., 1998).
This effect has been investigated quantitatively in model simulations. In a
study of conditions for the year 2010, Jucks and Salawitch (2000) find that
above 45 km an increase of 1% in the water vapour mixing ratio would
completely negate the increase of ozone driven by the 15% decrease of
inorganic chlorine that is expected by the year 2010. At 40 km the increase
of ozone would still be reduced by about 50%. Shindell (2001) conducted a
general circulation model (GCM) study for the period 19791996. The
simulations show a significant chemical effect of water vapour increases on
ozone concentrations, with a reduction of more than 1% between 45 and 55
km, and a maximum impact of about 4% at 50 km. Further, Li et al. (2002)
found that the annually averaged (downward) ozone trend at 45S and 1.8
hPa (45 km) increases by 1% per decade for a water vapour increase of 1%
yr1. These model results point to an anticorrelation between ozone and
water vapour in the upper stratosphere and lower mesosphere.
However, such an anticorrelation is not seen in observations and ozone in
this region varies much less than predicted by models (Siskind et al., 2002).
In the mid-latitude lower stratosphere, any increase in water vapour would
also be expected to lead to ozone loss because of an intensified HOx ozone
loss cycle (see Box 1.1). The model results of Dvortsov and Solomon (2001)
predict that an increase in stratospheric water vapour of 1% yr1 translates
to an ad- 93 95 96 00 97 03 94 92 98 99 100 80 60 40 20 0 0 10 20 30 40
Ozone Column Loss (DU) VPSC (106 km3 ) Figure 1.13. The overall chemical
loss in average polar column ozone during a given Arctic winter versus the
winter average of the stratospheric volume where conditions were cold
enough (based on ECMWF temperature data) for the existence of PSCs
(VPSC). Also shown is the linear fit to the data. The calculations could not be
performed for 2001 or 2002.
Adapted from Rex et al. (2004). IPCC Boek (dik).indb 104 15-08-2005
10:52:51 Chapter 1: Ozone and Climate: A Review of Interconnections 105
ditional depletion of mid-latitude column ozone by 0.3% per decade. In the
polar lower stratosphere, the rates of the heterogeneous reactions that are
responsible for the activation of chlorine (which eventually leads to chemical
ozone destruction) increase with increasing water vapour concentrations. An
increase in stratospheric water vapour essentially means that the
temperature threshold at which PSCs form, and thus heterogeneous

reactions rates begin to become significant for chlorine activation, is shifted


to higher values.
If stratospheric water vapour were to increase it could lead to a substantially
enhanced Arctic ozone loss in the future (e.g., Kirk-Davidoff et al., 1999). In
considering the above discussion concerning stratospheric water vapour, it
needs to be borne in mind that there is considerable uncertainty about the
sign of future water vapour changes in light of the puzzling past record
(Section 1.2.3). 1.3.4 The role of transport for ozone changes Transport is a
key factor influencing the seasonal and interannual variability of
stratospheric ozone. Seasonal variations in transport force the large winterspring build-up of extra-tropical total ozone in both hemispheres, and interhemispheric differences in transport (larger in the NH) cause corresponding
differences in extra-tropical total ozone (see Box 1.2).
In both hemispheres, mid-latitude ozone decreases during summer and
returns to approximate photochemical balance by autumn, and there is a
strong persistence of the dynamically forced anomalies throughout summer
(Fioletov and Shepherd, 2003; Weber et al., 2003). The interannual variability
in the winter-spring build-up is greater in the NH, reflecting the greater
dynamical variability of the NH stratosphere. 1.3.4.1 Stratospheric planetarywave-induced transport and mixing Large-scale transport of ozone is a result
of advection by the Brewer-Dobson circulation and of eddy transport effects;
both of these mechanisms are first-order terms in the zonal mean ozone
transport equation (Andrews et al., 1987).
The strength of the Brewer-Dobson circulation is directly tied to dissipating
planetary waves forced from the troposphere, and eddy transports of ozone
are also linked to planetary wave activity (although this latter linkage is
difficult to quantify), so that net ozone transport is tied to the variability of
forced planetary waves. The amount of dissipating wave activity (also called
planetary wave drag, PWD) within the stratosphere is related to the vertical
component of wave activity entering the lower stratosphere (the so-called
Eliassen-Palm (EP) flux).
This quantity can be derived from conventional meteorological analyses and
is a convenient proxy used to quantify PWD. A significant correlation has
been found between interannual changes in PWD and total ozone build-up
during winter and spring (Fusco and Salby, 1999; Randel et al., 2002). The
fact that the effect of PWD on ozone is seasonal and has essentially no
interannual memory suggests that long-term changes in PWD can be

expected to lead to long- -term changes in winter-spring ozone, all else being
equal. However, although the basic physics of the ozone-PWD connection is
well understood, its quantification via correlations is at best crude, and this
limits our ability to attribute changes in ozone to changes in PWD.
In the NH, there have been interannual variations in various meteorological
parameters during the period 19802000 that together paint a fairly
consistent, albeit incomplete, picture. During the mid-1990s the NH exhibited
a number of years when the Arctic wintertime vortex was colder and stronger
(Graf et al., 1995; Pawson and Naujokat, 1999) and more persistent (Waugh
et al., 1999). Any dynamically induced component of these changes requires
a weakened Brewer-Dobson circulation, which in turn requires a decrease in
PWD. Such a decrease in PWD during this period has been documented
(Newman and Nash, 2000), although the results were sensitive to which
months and time periods were considered.
Randel et al. (2002) show that for the period 19792000, PWD in the NH
increased during early winter (November to December) and decreased
during mid-winter (January to February). This seasonal variation is consistent
with the Arctic early winter warming and late winter cooling seen over the
same time period at 100 hPa (Langematz et al., 2003; see also Figure 1.16 in
Section 1.4.1.2). The weakened Brewer-Dobson circulation during midwinter
implies a decrease in the winter build-up of mid-latitude ozone, and Randel
et al. (2002) estimated that the decreased wave driving may account for
about 2030% of the observed changes in ozone in the January to March
period. Changes in SH dynamics are not as clear as in the NH, primarily
because meteorological reanalysis data sets are less well constrained by
observations.
Planetary waves, in addition to affecting the temperature and chemistry of
the polar stratosphere through the processes described earlier, can also
displace the centre of the polar vortex off the pole. This has important
implications for ozone and NOx chemistry because air parcel trajectories
within the vortex are then no longer confined to the polar night but
experience short periods of sunlight (Solomon et al., 1993). Polar ozone loss
processes are usually limited by the poleward retreat of the terminator (the
boundary that delineates polar night).
Planetary wave distortion of the vortex can expose deeper vortex air to
sunlight (Lee et al., 2000) and cause ozone depletion chemistry to start
earlier than it would have otherwise. In this way, waveinduced displacements

of the vortex can drive a mid-winter start to Antarctic ozone depletion


(Roscoe et al., 1997; Bodeker et al., 2001). Because wave-induced forcing in
the stratosphere is believed to come primarily from planetary-scale Rossby
waves that are generated in the troposphere during wintertime, future
changes in the generation of tropospheric waves may influence polar ozone
abundance. However there is as yet no consensus from models on the sign of
this change (Austin et al., 2003). 1.3.4.2 Tropopause variations and ozone
mini-holes Tropospheric circulation and tropopause height variations also
affect the mid-latitude distribution of column ozone.
The relationship between local mid-latitude tropopause height and column
ozone on day-to-day time scales is well documented (e.g., Dobson, 1963;
Bojkov et al., 1993). Day-to-day changes in tropopause height are associated
with the passage of synoptic-scale disturbances in the upper troposphere
and lower stratosphere, which affect ozone in the lower stratosphere through
transport (Salby and Callaghan, 1993) and can result in large local changes
in column ozone, particularly in the vicinity of storm tracks (James, 1998).
In extreme situations, they can lead to socalled ozone mini-holes, which
occur over both hemispheres and have the lowest column ozone levels
observed outside the polar vortices, sometimes well under 220 DU (Allaart et
al., 2000; Hood et al., 2001; Teitelbaum et al., 2001; Canziani et al., 2002).
Ozone mini-holes do not primarily entail a destruction of ozone, but rather its
re-distribution. Changes in the spatial and temporal occurrence of synopticscale processes and Rossby wave breaking (e.g., storm track displacements)
that are induced by climate change or natural variability, could lead to
changes in the distribution or frequency of the occurrence of mini-hole and
low-ozone events, and thus to regional changes in the mean column ozone
(Hood et al., 1999; Reid et al., 2000; Orsolini and Limpasuvan, 2001).
Observations have shown that over the NH the altitude of the extra-tropical
tropopause has generally increased over recent decades.
Radiosonde measurements over both Europe and Canada show an increase
in altitude of about 300600 m over the past 30 years, with the precise
amount depending on location (Forster and Tourpali, 2001; Steinbrecht et al.,
2001). Consistent regional increases are also seen in meteorological
reanalyses over both hemispheres (Hoinka, 1999; Thompson et al., 2000),
although reanalysis trend studies should be viewed with care (Bengtsson et
al., 2004). Spatial patterns show increases over the NH and SH beginning at
mid-latitudes and reaching a maximum towards the poles. The magnitude of
the changes can also depend on the longitude. However, the effects of

tropospheric circulation changes on ozone and the relation of both to


tropopause height changes remain poorly understood at present.
In particular, the relationship between tropopause height and ozone
mentioned earlier applies to single stations; there is no reason to expect it to
apply in the zonal mean, or on longer time scales (e.g., seasonal or
interannual) over which ozone transport is irreversible. Because of our poor
understanding of what controls the zonal-mean midlatitude tropopause
height, or whether it is possible to consider such processes in a zonal-mean
approach, it is not clear that the ozone-tropopause height correlations
(derived from daily or monthly statistics) can be extended to decadal time
scales in order to estimate changes in ozone.
Some recent model studies suggest that lower stratospheric cooling caused
by ozone depletion, acting together with tropospheric warming due to
WMGHGs, can be a contributor to tropopause height changes (Santer et al.,
2003a,b), in which case the tropopause height changes cannot be entirely
considered as a cause of the ozone changes. One source of uncertainty in
these model calculations is that most climate models cannot resolve the
observed tropopause height changes, so these must be inferred by
interpolation; another is that the observed ozone changes are not well
quantified close to the tropopause (WMO, 2003, Chapter 4). Furthermore
there remain considerable differences between the various reanalysis
products available, as well as between the model results. 1.3.4.3
Stratosphere-troposphere exchange Stratosphere-troposphere exchange
(STE) processes also affect the ozone distribution and have the capability to
affect tropospheric chemistry. STE is a two-way process that encompasses
transport from the troposphere into the stratosphere (TST) and from the
stratosphere into the troposphere (STT) through a variety of processes (see
Holton et al., 1995, and references therein).
Some aspects of STE are implicit in the preceding discussions. The net mass
flux from the troposphere into the stratosphere and back again is driven by
PWD through the Brewer-Dobson circulation (Section 1.1.2; Holton et al.,
1995). Although the global approach can explain the net global flux, local
and regional processes need to be identified and assessed to fully
understand the distribution of trace species being exchanged. Such
understanding is relevant both for chemical budgets and for climate change
and variability studies. Although there is a net TST in the tropics and a net
STT in the extratropics, a number of mechanisms that lead to STT in the
tropics and TST in the extratropics have been identified (e.g., Appenzeller

and Davies, 1992; Poulida et al., 1996; Hintsa et al., 1998; Ray et al., 1999;
Lelieveld and Dentener, 2000; Stohl et al., 2003).
In the extratropics TST can significantly affect the composition of the
lowermost stratosphere, even if the exchange cannot reach higher into the
stratosphere, by introducing, during deep TST events, near-surface
pollutants, and ozone-poor and humid air. Similarly, deep STT events can
transport ozone-rich air into the mid- and lower troposphere. It should be
noted, however, that so-called shallow events, with exchanges near the
tropopause, remain the main feature in the extratropics. Regions of
occurrence of such exchange processes, at least in the NH (Sprenger and
Wernli, 2003), appear to be linked to the storm tracks and their variability,
that is, in the same region where mini-holes are most frequent.
The past and future variability of STE remains to be assessed. It is clear that
changes in planetary wave activity will modify the Brewer-Dobson circulation
and hence affect global transport processes (Rind et al., 1990; Butchart and
Scaife, 2001). As for regional mechanisms, seasonal and interannual
variability, driven, for example, by the North Atlantic Oscillation (NAO) and
the El Nio-Southern Oscillation (ENSO), have already been observed (James
et al., 2003; Sprenger and Wernli, 2003). Changes in the occurrence of
regional and local weather phenomena could also modify STE. Fifteen-year
studies with reanalysis products have not yielded any distinct trends
(Sprenger and Wernli, 2003). Given the discrepancies between different
approaches to evaluate STE and inhomogeneities in meteorological analyses
and reanalyses, consistent trend studies are not available as of yet.
Air enters the stratosphere primarily in the tropics, and hence the physical
and chemical characteristics of air near the tropical tropopause behave as
boundary conditions for the global stratosphere. The tropical tropopause is
relatively high, near 17 km. The tropospheric lapse rate (up to 1214 km) is
determined by radiative-convective equilibrium, whereas the thermal
structure above 14 km is primarily in radiative balance, which is
characteristic of the stratosphere (Thuburn and Craig, 2002). Overall, the
region of the tropical atmosphere between about 12 km and the tropopause
has characteristics intermediate to those of the troposphere and
stratosphere, and is referred to as the tropical tropopause layer (TTL)
(Highwood and Hoskins, 1998).
Thin (sometimes subvisible) cirrus clouds are observed over large areas of
the TTL (Wang et al., 1996; Winker and Trepte, 1998), although their

formation mechanism(s) and effects on large-scale circulation are poorly


known. In the tropics, the background clear-sky radiative balance shifts from
cooling in the troposphere to heating in the stratosphere, with the transition
occurring at around 15 km. The region of heating above about 15 km is
linked to mean upward motion into the lower stratosphere, and this region
marks the base of the stratospheric Brewer-Dobson circulation.
The TTL is coupled to stratospheric ozone through its control of stratospheric
water vapour and by the transport of tropospheric source gases and ODSs
into the lower stratosphere, so changes in the TTL could affect the
stratospheric ozone layer. If the residence time of air in the TTL before it is
transported into the stratosphere is short, then it may be possible that shortlived halogen compounds (natural or anthropogenic replacements for the
CFCs) or their degradation products could enter the lower stratosphere and
contribute to ozone loss there (WMO, 2003, Chapter 2). Air entering the
stratosphere is dehydrated as it passes through the cold tropical tropopause,
and this drying accounts for the extreme aridity of the global stratosphere
(Brewer, 1949).
Furthermore, the seasonal cycle in tropopause temperature imparts a strong
seasonal variation in stratospheric water vapour, which then propagates with
the mean stratospheric transport circulation (Mote et al., 1996). Year-to-year
variations in tropical tropopause temperatures are also highly correlated with
global stratospheric water vapour anomalies (Randel et al., 2004), although
there remain some issues concerning the consistency of decadal-scale
changes in water vapour (Section 1.2.3). However, while there is strong
empirical coupling between tropopause temperatures and stratospheric
water vapour, details of the dehydration mechanism(s) within the TTL are
still a topic of scientific debate (Holton and Gettelman, 2001; Sherwood and
Dessler, 2001).
One critical, unanswered question is whether, and how, the TTL will change
in response to climate change and what will be the resulting impact on
stratospheric ozone. 1.3.5 Stratosphere-troposphere dynamical coupling
Analysis of observational data shows that atmospheric circulation tends to
maintain spatially coherent, large-scale patterns for extended periods of
time, and then to shift to similar patterns of opposite phase. These patterns
represent preferred modes of variability of the coupled atmosphere-oceanland-sea-ice system. They fluctuate on intra-seasonal, seasonal, interannual
and decadal time scales, and are influenced by externally and

anthropogenically caused climate variability. Some of the patterns exhibit


seesaw-like behaviour and are usually called oscillations.
A few circulation modes, listed below, have been implicated in stratospheretroposphere dynamical coupling, and thus may provide a coupling between
stratospheric ozone depletion and tropospheric climate. Northern Annular
Mode (NAM) (Thompson and Wallace, 1998, 2000; Baldwin and Dunkerton,
1999): The NAM, also referred to as the Arctic Oscillation (AO), is a
hemispherewide annular atmospheric circulation pattern in which
atmospheric pressure over the northern polar region varies out of phase with
pressure over northern mid-latitudes (around 45N), on time scales ranging
from weeks to decades. Southern Annular Mode (SAM) (Gong and Wang,
1999; Thompson and Wallace, 2000): The SAM, also referred to as the
Antarctic Oscillation, is the SH analogue of the NAM. The SAM exhibits a
large-scale alternation of pressure and temperature between the midlatitudes and the polar region.
North Atlantic Oscillation (NAO) (Walker and Bliss, 1932; Hurrell, 1995): The
NAO was originally identified as a seesaw of sea-level pressure between the
Icelandic Low and the Azores High, but an associated circulation (and
pressure) pattern is also exhibited well above in the troposphere. The NAO is
the dominant regional pattern of wintertime atmospheric circulation
variability over the extra-tropical North Atlantic, and has exhibited variability
and trends over long time periods (Appenzeller et al., 2000). The relationship
between the NAO and the NAM remains a matter of debate (Wallace, 2000;
Ambaum et al., 2001; Rogers and McHugh, 2002).
The NAM extends through the depth of the troposphere. During the cold
season (January to March), when the stratosphere has large-amplitude
disturbances, the NAM also has a strong signature in the stratosphere, where
it is associated with variations in the strength of the westerly vortex that
encircles the Arctic polar stratosphere; this signature suggests a coupling
between the stratosphere and the troposphere (Perlwitz and Graf, 1995;
Thompson and Wallace, 1998; Baldwin and Dunkerton, 1999). During winters
when the stratospheric vortex is stronger than normal, the NAM (and NAO)
tends to be in a positive phase. Circulation modes can affect the ozone
distribution directly (in the troposphere and the lowermost stratosphere; see
Lamarque and Hess, 2004), and indirectly by influencing propagation of
planetary waves from the troposphere into the middle atmosphere (Ambaum
and Hoskins, 2002). Therefore, changes in circulation modes can produce
changes in ozone distribution. It has been shown that because the NAO has a

large vertical extent during the winter and because it modulates the
tropopause height, it can explain much of the spatial pattern in column
ozone trends in the North Atlantic over the past 30 years (Appenzeller et al.,
2000). Stratospheric changes may feed back onto changes in circulation
modes.
Observations suggest that at least in some cases the large amplitude NAM
anomalies tend to propagate from the stratosphere to the troposphere on
time scales of weeks to a few months (Baldwin and Dunkerton, 1999;
Christiansen, 2001). Furthermore, because of the strong coupling between
the stratospheric vortex and the NAM, the recent trend in the NAM and NAO
has been associated with processes that are known to affect the strength of
the stratospheric polar vortex, such as tropical volcanic eruptions (Kodera,
1994; Kelly et al., 1996; Rozanov et al., 2002; Stenchikov et al., 2002), ozone
depletion (Graf et al., 1998; Shindell et al., 2001a), and anthropogenic
changes in greenhouse gas concentrations (Shindell et al., 2001a; Gillett et
al., 2002a). However, some modelling studies have shown that a simulated
trend in the tropospheric NAM and SAM does not necessarily depend on
stratospheric involvement (Fyfe et al., 1999; Gillett et al., 2002b). Global
climate modelling simulations that include interactive stratospheric
chemistry suggest that one mechanism by which solar variability may affect
tropospheric climate is through solar-forced changes in upper-stratospheric
ozone that induce changes in the leading mode of variability of the coupled
troposphere-stratosphere circulation (Shindell et al., 2001b).
A number of modelling studies have examined the effect of increased
concentrations of greenhouse gases on the annular modes (e.g., Perlwitz et
al., 2000; Shindell et al., 2001a; Gillett et al., 2002b; Rauthe et al., 2004).
Most coupled atmosphereocean climate models agree in finding a positive
NAM trend under increasing greenhouse-gas concentrations, a trend which is
qualitatively consistent with the observed positive NAM trend. A more
positive NAM is consistent with a stronger stratospheric vortex. On the other
hand, an intensified polar vortex is related to changes in planetary- and
synoptic-scale wave characteristics, and may produce tropospheric
circulation anomalies similar to the positive phase of the NAM. These
potential feedbacks obscure cause and effect.
For example, whereas Rind et al. (2002) and Sigmond et al. (2004) found
that in the middle stratosphere the perturbation to the NAM from a 2 CO2
climate depends on modelled changes in sea-surface temperatures (SSTs),
Sigmond et al. (2004) suggested that perturbations in the zonal wind near

the surface (which will affect the response to SSTs) are mainly caused by
doubling of stratospheric CO2 . The stratosphere-troposphere dynamical
coupling through the circulation modes discussed in this section suggests
that stratospheric dynamics should be accounted for in the problem of
detection and prediction of future tropospheric climate change.
Because increased concentrations of greenhouse gases may cause changes
in stratospheric dynamics (and in ozone, which in turn cause changes in
stratospheric dynamics), greenhouse-gas changes may induce changes in
surface climate through stratosphere-troposphere dynamical coupling, in
addition to radiative forcing. 1.3.6 Possible dynamical feedbacks of ozone
changes The effect of ozone loss on polar stratospheric temperatures results
in concomitant changes in stratospheric zonal winds and polar vortex
structure, and also possible changes in planetarywave behaviour.
The strongest effect is seen in the Antarctic, where a large cooling of the
lower stratosphere, which is associated with the ozone hole (Randel and Wu,
1999), has resulted in an intensified and more persistent springtime polar
vortex. Waugh et al. (1999) show that since the ozone hole developed, the
break-up date of the Antarctic vortex occurs two to three weeks later
(moving from mid-November to early December) than before. Recent
research has shown that trends in surface temperatures over Antarctica
(cooling over the interior and part of the warming over the peninsula) may
be in part traceable, at least over the period 19802000, to trends in the
lower stratospheric polar vortex, which are largely caused by the ozone hole
(Thompson and Solomon, 2002).
It has been suggested that during the early summer the strengthening of
the westerly flow extends all the way to the surface (Thompson and
Solomon, 2002). The role of this mechanism which has yet to be elucidated
has been highlighted by the modelling study of Gillett and Thompson
(2003), who prescribed ozone depletion in an atmospheric climate model.
They found that the seasonality, structure and amplitude of the modelled
changes in 500 hPa geopotential height and near-surface air temperature in
the Antarctic had similar spatial patterns to observations, which were a
cooling over the Antarctic interior and a warming over the peninsula and
South America (Figure 1.14).
This result suggests that anthropogenic emissions of ozone-depleting gases
have had a distinct impact on climate not only in the stratosphere but also at
the Earths surface. These surface changes appear to act in the same

direction as changes resulting from increases in greenhouse gases (Shindell


and Schmidt, 2004). An increase in the strength of the Antarctic polar vortex
by stratospheric ozone depletion does not affect only surface winds and
temperatures. Using a 15,000-year integration of a coupled oceanatmosphere model, Hall and Visbeck (2002) showed that fluctuations of the
mid-latitude westerly winds generate ocean circulation and sea-ice variations
on interannual to centennial time scales.
Cause and effect are more difficult to separate in the Arctic stratosphere
than in the Antarctic stratosphere because of higher natural meteorological
variability, relatively smaller ozone losses and a less clear separation
between transport and chemical effects on ozone. Furthermore,
temperatures need to be suf- ficiently low in any given year to initiate polar
ozone chemistry. Nonetheless, in several years during the mid-1990s the
Arctic experienced low ozone, low temperatures in late spring and enhanced
vortex persistence (Randel and Wu, 1999; Waugh et al.,.

Вам также может понравиться