You are on page 1of 19

ANTHROPOGENIC CAUSES OF GLOBAL ENVIRONMENTAL CHANGE

Wei-Chyung Wang
Atmospheric Sciences Research Center, State University of New York at Albany, USA
Keywords :Global environmental change, greenhouse effect, aerosols, ozone, global warming
Contents
1. Introduction
2. Atmospheric Constituents and Global Environment
3. Changes in Atmospheric Composition
4. Global Environmental Changes
5. Concluding Remarks
Related Chapters
Glossary
Bibliography
Biographical Sketch
Summary
Since the Industrial Revolution, significant progress in economic development and
technological achievement has been made to advance the welfare of humankind. A major
downside of the economic development, in addition to the rapid depletion and sometime
wasteful use of the limited resources, is the degradation of environment and changes in
climate.
The impact of human activity on the atmosphere is now widely recognized: the observed
increases of greenhouse gas concentrations (notably carbon dioxide) and ozone depletion as
well as the increased particle loading are the outstanding features. These compositional
changes in the atmosphere will most likely lead to changes in the environment and climate on
a global scale. This theme provides a focused presentation of the core issues concerning the
global environmental change and the current approach to these issues. Specifically, the
presentation focuses on the history of changes in atmospheric trace constituents (the
greenhouse gases and aerosols), and on the modeling studies of the associated climate
changes.
1. Introduction
With the advent of industrialization in the late eighteenth century, we have been drastically
changing the environment around us. One major by-product of this new lifestyle has been the
introduction of various trace constituents (gases and particles) into the atmosphere. Some of
these gases, such as carbon dioxide (CO2), methane (CH4), and nitrous oxide (N2O), were
naturally present in the atmosphere, although in small concentrations (parts per million by
volume (ppmv) in mixing ratio); other gases, such as chloroflurocarbons (CFCs) with even
smaller concentrations of parts per billion by volume (ppbv), are new. Increasingthe
concentrations of these gases is enhancing the "greenhouse effect," yielding a warming of the
earths climate referred to as global warming. Increases of tropospheric aerosols (microscopic
airborne particles or droplets) affect visibility and modulate clouds. In addition, increasing
CFCs, and the emissions of oxides of nitrogen (NOx) and carbon monoxide (CO) from

aviation and biomass burning can also change atmospheric O3, a greenhouse gas that absorbs
most of the harmful ultraviolet radiation; for example, the observed "stratospheric O3
depletion" can be attributed to the use of CFCs.
The main difference between 1750 and today is the degree of industrialization. It is therefore
quite clear that human activities are responsible for the changes in the concentration of these
constituents in the global environment. The environment is not static, and will respond to
changes that are made to it. There have been large, concerted efforts to develop general
circulation models (GCMs) of the earths climate system to examine the climate responses to
the observed changes in concentrations of greenhouse gases, O3, and aerosols. Though
attribution can never be certain, the GCM simulations of climate changes for the past century
are not inconsistent with the observed climate changes. These GCMs are also used to assess
future climate changes for the next century, based on the projected changes of the atmospheric
constituents. The simulations show large regional climate changes. Global warming,
stratospheric O3 depletion, and tropospheric aerosols are, therefore, recognized as potentially
destructive influences on the global environment. It is likely that these global environmental
changes will be detrimental to human society.
This theme describes the changes in composition of the atmosphere in this century, in
particular the greenhouse gases, O3, and sulfate aerosols, concentrating on aspects of globalscale change that have clearly been brought about by human activities. Two aspects need to be
addressed: those concerned with natural systems and those concerned with social systems.
These two aspects clearly interact in many ways, but in general, understanding of the former
is best derived through the physical and natural sciences. However, analysis of global change
in social systems involves using the methods of sociology, psychology, economics, and
political science, which will not be addressed here. Extensive documentation of the study of
human-induced climatic and environmental changes can be found in IPCC (2001) and WMO
(1999).
2. Atmospheric Constituents and Global Environment
The sun provides the earths only external source of heat, solar radiation being in the visible
and near-infrared spectra, while most of the harmful ultraviolet radiation is absorbed by O3 in
the stratosphere. The earth also radiates energy back into space and, because of a much colder
temperature than the Sun, the energy is in the form of infrared radiation. A balance between
solar radiation and infrared radiation yields the present-day mean temperature, about 18C,
of the earth atmospheresurface climate system.
The thermal structure of the atmosphere is influenced by the presence of trace gases and
clouds (including aerosols), both of which modulate solar radiation and thermal emission. The
principal gaseous absorbers of solar radiation are water vapor (H2O) in the troposphere and O3
in the stratosphere. H2O absorbs primarily in the near-infrared spectral region, while O3 is of
major importance in maintaining the thermal structure in the stratosphere through its
absorption of solar radiation in the ultraviolet and visible regions. It is believed that
stratospheric O3 is largely responsible for the existence of the tropopause, a nearly isothermal
region separating the radiative equilibrated stratosphere from the more dynamically controlled
troposphere. On a global and annual mean basis, about 100 Watts per square meter (Wm2) of
the 340 Wm2 incident solar radiation at the top of the atmosphere are reflected back to space

mainly by clouds and surface while about 70 Wm2 are absorbed by the atmosphere and the
rest absorbed by the surface.
In the infrared, H2O effectively blocks thermal emission from the surface except for the
"window" region between 7 and 12 micrometers (m) where CO2, O3, CH4, CFCs, and N2O
with strong absorption bands contribute additional atmospheric opacity. Absorption of the
outgoing thermal radiation in the atmosphere, followed by re-radiation at local temperature
can lead to an increase of the surface temperature, the so-called "greenhouse effect." Satellite
measurements indicate that the atmospheric gases can trap up to 100200 Wm2 of longwave
radiation depending on the regions and seasons, thus keeping the global mean surface air
temperature at a habitable 15C (versus a mean temperature of 18C). The presence of
clouds will further increase the magnitude of the trapping of longwave radiation, about 30
Wm2 on the global mean basis. Note, however, that, in contrast to the greenhouse gases that
warm the climate, clouds also reflect large amounts of solar radiation, about 48 Wm2, and
thus cause a net cooling of the climate system.
A perturbation to the energy available to the climate system, defined as "radiative forcing,"
due to changes in these atmospheric constituents will lead to climate changes with positive
forcing for warming (e.g. the enhanced greenhouse effect) and negative forcing for cooling
(e.g. stratospheric O3 depletion and the tropospheric aerosols). As the atmosphere interacts
with the other parts (such as the oceans) of the climatic system, changes in the atmosphere
inevitably affect the global environment.
Global warming associated with the enhanced greenhouse effect is presently considered to
pose the greatest threat to the earths climate. The current concern over global warming is
related to two observations, first, that the greenhouse gases CO2, CH4, CFCs, and N2O have
been increasing over the past few decades (for CO2 since industrialization) and this trend is
expected to continue, and second that the global mean surface air temperature has increased
by 0.3C to 0.6 C over the past hundred years. In addition, the pattern of observed surface
warming is broadly consistent with global climate model simulations. For example, Crowley
(2000), who studied the northern hemisphere temperatures and climate forcing over the last
2,000 years, concluded that the large late twentieth century warming is in close agreement
with the climate responses simulated from greenhouse gas forcing. In addition, the simulated
temporal and spatial patterns are better correlated when other causes, such as volcanic and
anthropogenic aerosols, are included.
The link between the long-term build-up of chlorine and the decline of O3 in the upper
stratosphere is firmly established. Most of the mid-latitude column O3 decreases during the
last two decades arose because of depletion in the lower stratosphere, which is also related to
the increase in O3-depleting compounds. It is well established that the chemistry of chlorine
and bromine originally from anthropogenic activity contribute to the observed large O3 losses
in the southern hemisphere polar region during spring. Continued increases in anthropogenic
emissions of O3 precursor gases (CO, NOx and volatile organic compounds) can potentially
increase tropospheric O3.
Decreases in stratospheric O3 may cool the stratospheric temperature, a characteristic
consistent with observations. Radiosonde and satellite measurements indicate that during the
period 197994, the global, annual-mean stratospheric temperature averaged over 1621 km
had a cooling trend of about 0.6C/decade, with a larger trend of 0.75C/decade at midlatitudes. The cooling trend is substantially larger, about 3C/decade, in the polar lower

stratosphere during late winter/spring in both hemispheres. For a moderate change in


atmospheric O3 amount, the changes in UV flux at the surface are confined to the spectral
region around 300340 nm. A few ground-based spectroradiometers at mid-latitudes have
detected an increase in UV-B radiance during the period 198997 with values of 1.5
percent/year at 300 nm and 0.8 percent/year at 305 nm, which are consistent with observed O3
changes. Perhaps more importantly, there are strong regional features in radiative forcing
from O3 changes. Change in radiative forcing due to O3 changes from pre-industrial times is
largest at the northern middle latitudes where the impact on O3 from emission of pollutants is
largest. There are large latitudinal gradients in the calculated change in radiative forcing due
to reduced O3 loss in the lower stratosphere during the last two decades with large reductions
at high latitudes. Regional O3 changes in areas where there has been a strong increase in the
emissions of pollutants over the last couple of decades (e.g. South-east Asia) is calculated to
have caused a substantial increase in radiative forcing.
Unlike the greenhouse gases, which are evenly distributed, anthropogenic aerosols are not
spread evenly across the globe. The release of sulfur dioxide (SO2) is the main anthropogenic
source of sulfate aerosols. Due to their short lifetimes, which average days to weeks compared
to the decades to centuries of greenhouse gases, sulfate aerosols are concentrated locally from
where they are emitted. Thus, the increases in their concentrations are not worldwide, and are
located around the centers of their production, mainly resulting from the burning of fossil
fuels and to a lesser extent from biomass burning. Tropospheric aerosols can absorb and
reflect solar radiation and in most cases, they tend to decrease radiative forcing and thus lead
to a colder climate. In addition, aerosols serve as cloud condensation nuclei (CCN), which are
involved in the complicated processes of cloud droplet formation and precipitation removal
processes, an aspect of inadequate understanding at the present stage.
It is important to recognize that climate variations can occur, in the absence of changes in
radiative forcing associated with anthropogenic emissions. This happens as a result of
interactions on different time and spatial scales among the components of the climate system;
for example, the El Nio-Southern Oscillation (ENSO) is a natural phenomenon on the
interannual timescale resulting from the couplings between the atmosphere and oceans.
3. Changes in Atmospheric Composition
The composition of the atmosphere has changed since pre-industrial times, around 1750,
mainly as a result of human activity. A large part of this change can be seen in the
concentrations of greenhouse gases, O3, and aerosols. For some of these changes,
anthropogenic emissions create small but climatologically significant perturbations
superimposed on the large natural cycles of the gases; for others, anthropogenic emissions
dominate. All of these changes can impact the global environment.
3.1. Trace Gases
The annual mean concentrations of major greenhouse gases at pre-industrial times (1750) and
19989 are shown in Table 1. The sources and sinks of these gases are briefly summarized
here while detailed information can be found in WMO (1999) and IPCC (2001). It should be
pointed out here that, because of the different absorption characteristics, different gases
absorb and trap infrared radiation from the surface at different levels of efficiency; for
example, on a per molecule basis, CFC-12 has the same effect as about 25,000 CO2

molecules. Consequently, despite the small concentration of CFC-12 in the atmosphere, its
effect on infrared radiation is quite substantial.
Table 1. Concentration of greenhouse gases and their recent trends
Greenhouse gas
Time period

Trends (pptv/year; 1990s)

19989

1750

367

280 10

2.9 PgC/year 19907

1,745

700

8.4

314

270

0.8

268

1.4

533

4.4

34

25

3-6

35

CO2 (ppmv)*
CH4 (ppbv)*
N2O (ppbv)*
CFC-11 (ppbv)*
CFC-12 (ppbv)*
Tropospheric O3 (DU)

Stratospheric H2O (ppmv)*


Notes: *ppmv/ppbv are mixing ratios with unit in parts per million/billion by
volume.

Total column amount using Dobson Unit = 103 cm.

Table 1. Concentration of greenhouse gases and their recent trends


3.1.1. Carbon Dioxide
For much of the nineteenth century, CO2 concentration was 280 10 ppmv until the increase
in the twentieth century reaching 367 ppmv in 19989. The present level is the highest during
the past 420,000 years and the sustained rate of increase is the largest during the past 20,000
years.
Naturally, plant and soil respiration releases CO2, balancing the amount taken in by plants
through photosynthesis. Carbon is also taken up by phytoplankton in oceans during
production and released during the re-mineralization of organic matter. The anthropogenic
sources for CO2-increases in the atmosphere come from fossil fuel burning and to a lesser
extent cement manufacture. Deforestation also contributes by removing plants that could
remove CO2 from the atmosphere. Agricultural land occupies one-fifth of the terrestrial
surface, most of which was originally forested. The conversion of this land released carbon
into the atmosphere. Table 2 shows the global CO2 budget. Fossil fuel burning accounted for
emissions of 5.5 0.3 (peta (1015) grams of carbon) PgC/year during 19809, and 6.3 0.4
PgC/year during 19907. However, due to uptakes by oceans and terrestrial ecosystems, the

rate of increase in atmospheric CO2 concentration was smaller than the emissions with values
of 3.3 0.1 PgC/year during 19809 and 2.9 0.1 PgC/year during 19907. Considering the
major factors of continued economic development and population increase, it is expected that
the CO2 emissions from fossil fuel burning are certain to dominate the atmospheric CO2
concentration during the twenty-first century.
Table 2. Global CO2 budgets (PgC/year)*
Source/sink

19809

19907

3.3 0.1

2.9 0.1

5.5 0.3

6.3 0.4

2.0 0.6

2.4 0.5

0.2 0.7

1.0 0.6

Atmospheric increase
Emissions (fossil fuel, cement)
Oceanatmosphere flux
Landatmosphere flux
Note: *PgC/year = 1015 gram of carbon/year.
Source: after IPCC (2001).
Table 2. Global CO2 budgets (PgC/year)*
3.1.2. Methane and Nitrous Oxide
While the rate of increase of CH4 concentrations in the atmosphere has slowed somewhat
since the late 1970s, the current atmospheric level of 1,745 ppbv is the highest it has been in
at least 160,000 years, and represents a significant increase over its pre-industrial
concentration of 700 ppbv. The variability of the CH4 increases during the last two decades is
highly variable, probably caused by the changes in atmospheric chemistry that controls the
CH4 lifetime. Like CO2, CH4 occurs naturally in the atmosphere with wetlands as the major
source, while the major sinks are caused by reactions with OH in the troposphere, and to a
lesser extent losses to the soils and in the stratosphere. The main anthropogenic sources
include fossil fuel burning, landfills, ruminants, rice agriculture, and biomass burning.
Nitrous oxide is an important greenhouse gas due to its long atmospheric lifetime of 120
years, and its large radiative forcing is about 200 times that of CO2 on a per molecule basis.
Its concentration has increased from 270 ppbv in pre-industrial times to 314 ppbv in 19989.
The main sources of N2O have been tropical soils, the ocean in upwelling regions, and forests
and grasslands in temperate regions, while the sink is photo-dissociation in the stratosphere.
The major candidates for changing sources of N2O are the oceans and terrestrial soils. In the
soil, production depends on the input of organic matter, fertility, moisture status, temperature,
and oxygen status. The ocean source depends on production and the transport of N2O from
depth to the surface water. As the population continues to grow, the need for agricultural
practices and food production (such as cattle and feedlots) also increase with likely increase in
the sources. Both CH4 and N2O are sensitive to climate conditions with higher concentrations
during warmer climate conditions and vice versa. As climate change can influence both

concentrations, it is difficult to calculate their response to climate variations. In the future, a


wetter climate may lead to increased N2O production, though how much is not known.
3.1.3. Other Greenhouse Gases
In addition to these naturally occurring greenhouse gases, human activity has also introduced
into the atmosphere gases normally not present in it. The concentrations of halocarbons
containing chlorine and bromine (including CFCs) have increased rapidly through the late
twentieth century. However, as the Montreal Protocol and its Amendments regulate the
production and emission of these chemicals, their emissions have slowed. The combined
abundance of these O3-depleting compounds in the lower atmosphere peaked in about 1994
and is now slowly declining. This decline is mainly due to the reduced emissions of methyl
chloroform (CH3CCl3) under the Montreal Protocol, and has reduced the total amount of
organic chlorine in the troposphere. On the other hand, the gases used to replace these
chemicals, hydrochlorofluorocarbons (HCFCs) and hydrofluorocarbons (HFCs), also have the
potential to warm the climate and their concentrations are now growing rapidly. Thus,
chemicals containing bromine continue to grow, which means that the concentration of
organic bromine may not peak for a few more years. The increase in concentrations of
perfluorocarbons (PFCs) and sulfur hexafluoride (SF6), two other potentially important
greenhouse gases, has been slow so far, but as these gases have lifetimes of over 1,000 years,
their accumulated concentrations will become significant in the future.
3.2. Atmospheric Ozone
About 90 percent of the atmospheric O3 is located in the stratosphere. However, because of
the different chemical aspects, the O3 in the stratosphere and troposphere show different
change trends.
3.2.1. Stratospheric Ozone
A decrease in stratospheric O3 concentration has been observed since the 1970s. While O3
trends in the tropics have not been significant, O3 has been severely depleted in the mid- and
high- latitudes as well as in the polar region. From 1979 to 1997, trends in both hemispheres
in mid- and high- latitudes in all seasons were negative, large, and statistically significant,
with about 6 percent/decade in the upper stratosphere over mid-latitudes and 15
percent/decade in the lower stratosphere over the South Pole. Ozone levels declined for all
months of the year, though in the Northern Hemisphere, the trends were much larger in the
winter and spring than in summer and fall. Ozone levels fell even more steeply over the polar
region. In the Arctic vortex, extremely low O3 values were detected in late winter/spring for
six out of the last nine years. The largest depletion is over the Antarctic where there is an O3
hole. In the early 1990s, October O3 concentration values there were 5070 percent lower than
they were in the 1960s. The overall trend is about 6 percent per decade. Since 1994,
however, non-polar total O3, while variable, has not shown an overall negative trend. If the
Montreal Protocol and its Amendments are complied with, the maximum O3 depletion is
estimated to lie within the current decade or next two decades, though its identification and
evidence for the O3 layers recovery will not be recognized until even later.
3.2.2. Tropospheric Ozone

The budget of tropospheric O3 includes three components, the influx of stratospheric air, in
situ tropospheric photochemical production and loss, and surface deposition. The
stratospheretroposphere exchange is estimated to be 475 tera (1012) grams (Tg)(O3) although
the value could be as high as 1,400 Tg(O3). Photochemical production is closely related to
anthropogenic activity such as biomass burning and aircraft emissions, and the O3 loss
involves mainly the reaction with OH. The O3 mixing ratio in the troposphere has values
ranging from less than 10 ppbv over remote tropical oceans, 100 ppbv in the upper
troposphere, and to more than 100 ppbv in downstream regions of the polluted metropolitan
areas. The current tropospheric O3 amount is estimated to be 370 Tg(O3), which yields a
globally averaged column amount of 34 DU or a mean mixing ratio of 50 ppbv.
Studies suggest that tropospheric O3 has increased from a global mean value of 25 DU during
the pre-industrial period to the present-day value of 34 DU. However, this change is locally a
variable with a doubling of concentration (+25 ppbv) in the Northern Hemisphere and a
decrease at the South Pole since 1985. In fact, O3 has decreased at about a rate of 0.7 percent
per year since 1975 over the South Pole. The largest increases are reported over Europe and
Japan, while decreases are identified over Canada with no significant trends in the tropics. It
is expected that continued increases in anthropogenic (surface and aircraft) emissions of O3
precursor gases (CO, NOx, and hydrocarbons) can potentially increase tropospheric O3,
especially in the upper troposphere and lower stratosphere.
3.3. Tropospheric Aerosols
The major anthropogenic aerosols with global implications are the sulfate aerosols, which are
produced by chemical reactions in the atmosphere involving sulfur dioxide (SO2). Since SO2
emissions are mostly related to fossil fuel combustion and biomass burning, the source
distribution and magnitude for this trace gas is reasonably known. However, the chemical
pathway of conversion of precursors to sulfate is important because it changes the radiative
effects. The SO2 is converted to sulfate either in the gas phase or in cloud droplets that later
evaporate. Both processes produce sulfate mostly in submicron aerosols that are efficient
scatters for solar radiation, but the size distribution of sulfates is different for the gas phase
and aqueous production. This perhaps presents the most uncertainty in estimating the sulfate
loadings in the atmosphere.
In addition, unlike the greenhouse gases, which are evenly distributed, sulfate aerosols are not
spread evenly across the globe. Due to their short lifetimes, which average days to weeks
compared to the decades to centuries of greenhouse gases, sulfate aerosols are concentrated
locally where they are emitted. Thus, the increases in their concentrations are not worldwide,
and are located around the centers of their production. Given the fact that fossil fuel
consumption will likely be increasing in the future, the sulfate aerosols loadings will most
likely be increasing as well. There are many on-going efforts to compile the data for different
regions that will eventually be used in GCMs for assessing their effects on radiative forcing
and climate response.
4. Global Environmental Changes
In recent years, the GCM has been used extensively to study global climate changes due to
human activities. The GCMs are based on the numerical solution of the fundamental
equations governing the dynamical, physical, and chemical processes of the earth climate

system. The models have been demonstrated to simulate reasonably well the observed largescale climate features of the mean climate as well as the variability. It is believed that GCMs
are the best scientific tools available to understand the climate system and to assess future
global climate change.
In the models, the climatic effects due to increasing atmospheric greenhouse gases CO2, CH4,
CFCs, and N2O, changes of atmospheric O3, and increasing tropospheric sulfate aerosols are
initiated by a perturbation to the radiation energy balance of the earths atmospheresurface
climate system, the radiative forcing, and subsequently affected by climate feedbacks such as
due to changes in moisture and clouds in response to the radiation energy perturbation.
Because of the strong dynamical coupling between the troposphere and surface, the radiative
forcing of the climate system is generally defined as the net solar and longwave radiation flux
at the tropopause, around 12 km on the global, annual mean basis. Given a temperature
distribution, the radiative forcing depends on the concentration change and absorption
characteristics of the constituents, both of which are reasonably known. Therefore,
uncertainties associated with model calculated radiative forcing is smaller than the climate
responses simulated from GCMs that include complex interactions among dynamics, physics,
and chemistry of the climate system.
Below, the radiative forcing calculated for increasing the atmospheric constituents between
1750 and present is summarized to compare the relative importance of the individual
constituents. The GCM simulated climate responses of several important climate parameters
are then described to provide the potential changes in climate.
4.1. Radiative Forcing
Changes in atmospheric gases and aerosols can have direct and indirect effects on radiative
forcing. For the direct effects, increases in the atmospheric concentrations of CO2, CH4, CFCs
and N2O increase the longwave radiative forcing (a warming effect) while increases in
atmospheric sulfate aerosols are expected to decrease the solar radiative forcing (a cooling
effect). However, the effect on radiative forcing of changes in atmospheric O3 depends on the
compensating changes in both the solar and longwave radiative forcing. Atmospheric O3 and
sulfates are chemically-active species, and changes in their concentrations can induce indirect
effects on the radiative forcing. Atmospheric O3 is formed and destroyed by chemical
processes in the atmosphere due to interaction involving a large number of source gases (e.g.
H2O, NOx, CO, NMHC, N2O, CH4, and the CFCs), with several of these being important
greenhouse gases. Increases in atmospheric sulfate aerosols are expected to alter cloud
properties with subsequent effect on the radiative forcing. These direct and indirect effects are
summarized in Table 3 to provide an indication of the relative importance.
Table 3. Global and annual mean radiative forcing (Wm2) from 1750 to present
Atmospheric constituents
17502000

200050

CO2
1.46
CH4
0.48

0.27

N2O
0.15

0.17

0.07

0.04

0.17

0.6

0.15

0.35

0.42

0.40

0.29

0 to 2.0

0.20

0.30

CFC-11
CFC-12
Stratospheric O3*
Tropospheric O3
Direct (sulfates) aerosol
Indirect aerosol (first effect)
Land use (albedo only)
Solar
Notes: *for the period 197997.

For the period 200030.

The values are for SRES-A2 scenario.


Source: IPCC (2001).
Table 3. Global and annual mean radiative forcing (Wm2) from 1750 to present
4.2. Well-Mixed Greenhouse Gases
The radiative forcing due to well-mixed greenhouse gases during the periods 17502000 and
200050 are summarized in Table 3. Note that the values calculated for the future are based
on the emission scenario A2 of the Special Report on Emission Scenarios (SRES). The
forcing due to CO2 dominates in both periods while CH4 and N2O contribute substantially to
the total forcing. On the other hand, the decreases in the concentration of CFC-11 and CFC-12
caused by the emission controls lead to negative radiative forcing.
4.2.1. Atmospheric Ozone
Ozone change over the last couple of decades has been recently estimated to have an impact
on globally and annually averaged radiative forcing. These estimates give values of: 1) 0.2
0.15 Wm2 for changes in stratospheric O3 since the late 1970s; and 2) +0.35 0.15 Wm2 for
tropospheric O3 increases since pre-industrial times. From a broader perspective, these
observed O3 changes could have important climate implications. For example, on a global
mean basis, stratospheric O3 losses during 1980 and 1990 offset about 30 percent of the
radiative forcing attributed to an increase of well-mixed greenhouse gases during that period.
On the other hand, tropospheric O3 increases since pre-industrial times could augment
radiative forcing from all other greenhouse gases by about 1020 percent over the same
period.

Perhaps more importantly, there are strong regional features in radiative forcing from O3
changes. Changes in radiative forcing due to O3 changes since pre-industrial times is largest at
the northern middle latitudes where the impact on O3 from emission of pollutants is largest.
There are large latitudinal gradients in the calculated change in radiative forcing due to
reduced O3 loss in the lower stratosphere during the last two decades with large reductions at
high latitudes. Regional tropospheric O3 changes in areas where there has been a strong
increase in the emissions of pollutants over the last couple of decades (e.g. South-east Asia) is
calculated to have caused an increase in radiative forcing in part of the region of up to 0.5
Wm2 (IPCC, 2001).
4.2.2. Atmospheric Aerosols
It has been demonstrated that the direct effect of increasing sulfate aerosols in the troposphere
is likely to have a significant effect on solar radiative forcing. Recent studies suggest that
sulfate aerosols contribute a global and annual mean "direct" radiative forcing of 0.2 to 0.8
Wm2. However, uncertainties of these estimates are large because of the uncertainty in the
distribution and magnitude of sulfate aerosol emission, in the radiative properties in terms of
sulfate aerosols size distribution, and the scavenging processes that determine their lifetime in
the atmosphere. For example, estimates in sulfate aerosol emission might be as large as 20
30 percent, and a lack of understanding of the appropriate sulfate aerosol size distribution may
lead to a factor of 25 percent in the direct forcing. In addition, the uncertainty in sulfate
aerosol size distribution is related to that of other aerosols, serving as CCN, which are all
involved in the complex processes of cloud droplet formation and precipitation removal.
Moreover, the treatment of the growth of the extinction coefficient with relative humidity may
lead to an uncertainty of 50 percent. This treatment depends on factors such as the
contribution of organics to this growth, the contribution of nitrate in aerosols, and the
frequency with which the aerosol encounters regions of high relative humidity. Overall, the
uncertainty associated with the calculation of optical depth is judged to be 100 percent.
Aerosols alter cloud properties, and in recent studies the effects have been categorized into
two aspects: cloud formation processes (the "first" effect) and precipitation efficiency (the
"second" effect). While study of the second indirect effect is limited, the first indirect effect
has been examined using either "empirical" or "parameterized" approaches. The empirical
approach assumes a certain relationship between the number of cloud droplets and mass or
number of sulfate aerosols. Clearly, the approach has limitations, such as the effects of CCN
versus sulfate aerosol size distribution and aerosols precipitation scavenging processes on the
empirical relation. For the parameterized approach, a microphysics cloud model is generally
used to develop the relationship between CCN size distribution and the number of cloud
droplets. However, there are still many other issues. For example, the percentage of cloud
droplet concentration depends partly on a prescribed pre-existing aerosol particle size
distribution, and furthermore the relationship varies with different assumptions on processes
that form sulfate aerosols. Studies using this approach indicate that the "first" indirect effect (a
reduction of the cloud droplet size at constant liquid water content) is calculated to have a
range from 0.2 to 1.8 Wm2, while the total (first and second effects) indirect radiative
forcing has even larger uncertainty, from 0.3 to 3.0 Wm2. Therefore, the increased sulfate
aerosols could yield a surface cooling similar in magnitude to warming from greenhouse
gases over large regions.
Current efforts in simulating the effects of anthropogenic sulfate aerosols mainly focus on the
large and global scales. However, it is known that sulfate aerosols, due to their short lifetime,

concentrate in industrial regions, causing the most cooling in the northern mid-latitudes land
areas. Both observation and GCM studies highlight the importance of their regional effects.
For example, satellite data has shown that urban and industrial air pollution can suppress
precipitation from clouds that have temperatures at their tops of about 10C over large areas.
In addition, the process that dominates the relationship between cloud droplet concentration
and soluble aerosols, acting as CCN, depend on the updraft velocity. Nevertheless, most of the
GCMs cannot simulate detailed meso-scale structures with its coarse horizontal resolution.
Thus it is necessary to study the effects of sulfate aerosols using regional climate models
(RCMs) with higher horizontal resolution and more detailed representation of the cloud and
precipitation processes.
4.3. Global Warming
Two types of global GCMs have been used to study "global warming," the climatic effects of
increasing atmospheric greenhouse gases; the first type is an Atmospheric GCM (AGCM)
coupled with a mixed-layer ocean (without dynamics), while the second type is based on an
AOGCM that couples an AGCM with an Oceanic GCM (OGCM). So far, the available GCM
climate simulations of the greenhouse effect have been obtained using either the first type to
study the equilibrium climate responses between one and two times the pre-industrial CO2
concentration, or the second type to study the transient climate response based on a 1
percent/year increase in CO2; for the transient experiments, about half of the 1 percent/year
increase in CO2 was used as a surrogate to mimic the radiative effects of CH4, N2O, and CFCs.
In AGCM with a mixed-layer ocean, the horizontal and/or vertical transports of heat and salt
are prescribed from climatology and therefore they were not allowed to change in response to
a climate change. On the other hand, the AOGCM allows the simulation of the effects of
increased CO2 on ocean circulation and the storage of heat in the deep ocean that have been
found to be important in climate simulations. There have been many GCM simulations of the
greenhouse effect corresponding to a CO2 doubling, which occurs at around 70-years using
the 1 percent CO2 growth rate. Highlights of climate changes are summarized here.
The model simulations of changes in temperature and precipitation due to a doubling of CO2
are summarized in Table 4. For AOGCM, the time of a CO2 doubling corresponds to year-70
in the transient simulation with 1 percent CO2 growth rate. Note that the surface warming
simulated in AGCM for CO2 doubling is 3.4C, which is larger than 1.68C for AOGCM
mainly because of the smaller ocean heat capacity. Note also that the range of surface
warming among the different models is large, indicating the large uncertainties associated
with these simulations.
Table 4. Changes in global and annual mean climate parameters
due to a doubling of CO2
AGCM

AOGCM

(13 models)

(18 models)

3.4 (2.04.8)

1.68 (1.102.15)

6.5 (215)

2.4 (0.94.6)

Parameters
Surface temperature ( C)
Precipitation (%)

Note: The numbers are global and annual mean values averaged over all the models while the
range of the model values are included in the parentheses. The values for AOGCM
simulations are the 20-years mean centered on year-70 in the transient experiments using
1percent CO2 growth rate.
Source: IPCC (2001).
Table 4. Changes in global and annual mean climate parameters due to a doubling of CO2
4.3.1. Temperature Change
As indicated in Table 4, the resulting global and annual mean temperature change from
AOGCM simulation is about +0.24C/decade, though this does not apply equally everywhere
around the world. Because of smaller land heat capacity, warming over land will be greater
than over the oceans. In high latitudes, snow melting due to the initial radiative heating further
enhances the surface warming, thus producing a positive feedback. Minimum warming
around the northern North Atlantic Ocean and Antarctica are identified because of the large
inertia of the ocean associated with deep convection and large-scale vertical motions. There
will be a decrease in the strength of the meridional circulation in the northern North Atlantic
oceanic circulation. In the high northern latitudes and the Arctic, warming will be at a
maximum in late autumn and winter, as sea ice forms later in this warmer climate. In summer,
the sea ice will be removed, preventing a high degree of warming. In both the low latitudes
and in the southern circumpolar ocean, there is little seasonal variation of warming.
Note that while the magnitude of the warming does not seem to be a high number, one must
remember that over the past 10,000 years, the largest change the earth experienced was less
than 1C over a single century. Note also that temperature will slowly rise for centuries after
greenhouse gas stabilization, as there is a lag between changes in greenhouse gas
concentrations and the actual global mean temperature.
4.3.2. Precipitation Changes
The AOGCMs simulate an increase in the global and annual mean precipitation of about 2.4
percent with a strong geographical dependence. The warming of the atmosphere leads to a
higher water vapor content, enhanced poleward water vapor transport into the northern high
latitudes, and enhanced water vapor convergence and precipitation. Consequently,
precipitation increases in the high latitudes in winter and in most cases, the mid-latitudes. The
precipitation rate measures the intensity of the hydrological cycle, and influences the ocean
thermohaline circulation through the fresh water flux; warmer temperatures are usually
associated with larger precipitation rates. The other characteristics are the increases in the
extreme precipitation events in areas with increasing precipitation, in, for example, middle
and high latitudes.
It is projected that warming will cause a drier land surface (with regard to soil moisture) in
summer in the northern mid-latitudes (southern Europe, especially) due to enhanced
evaporation, though this will be strongly affected by aerosols, and may actually result in
increased soil moisture in high latitudes in winter. Summer precipitation and soil moisture
decrease over South-east Asia and increase over Europe and Asia, a reversal of the changes
due to the greenhouse gases alone. While some of the changes in hydrology predicted to occur

with increases in greenhouse gases are less extreme in the short term, this may not prevail as
sulfate concentrations are maintained only as long as sulfur emissions continue.
4.3.3. Sea Level Rise
Sea levels most likely will rise with the increased temperature. Though the oceans have a
large thermal inertia, they will still respond to an increase in temperature, causing them to
expand. The melt rates of glaciers and ice caps, and to a lesser extent ice sheets, increase,
adding more water to the oceans.
The lag between changes in greenhouse gas concentrations and changes in sea level is even
greater. As the oceans have a high thermal inertia, they will initially resist changes in
temperature. Once the oceans begin to warm, though, they will not stop immediately in
response to a stabilization in greenhouse gas concentration, or even in response to a
stabilization of the global mean land surface temperature. In fact, changes in greenhouse gas
emissions will affect the sea level for the next fifty years, at which point it will rise at the rate
of 36 cm per decade. By the year 2100, it is projected that the sea level will have risen 49
cm, with a range from 1694 cm. Most of this will be due to the expansion of the oceans,
though increased melting of glaciers, ice caps, and ice sheets will contribute as well. Thus, the
rise in sea level is the major problem to deal with in the case of increased greenhouse gas
concentrations.
4.4. Modification of Global Warming by Aerosols
Aerosols lessen warming where they are produced, especially over southern Asia because of
the sulfate aerosols caused by SO2 emissions associated with fossil fuel consumption. Sulfate
aerosols also decrease diurnal temperature range by increasing cloud coverage, obstruct
daytime sunshine, and prevent the escape of longwave radiation at night. Nights will then
warm more than days, and minimum temperatures will increase faster than maximum
temperatures. However, sulfate aerosols can also cool climate leading to a reduction in
evaporative cooling of the surface and an increase in global mean diurnal range. This will
occur only in regions of maximum aerosol loading.
Model simulations suggest that including the direct effect of sulfate aerosols reduces global
mean mid-twenty-first century warming by about 20 percent, although the spatial distribution
is highly inhomogeneous. For example, in northern mid-latitudes, the warming is substantially
reduced and the landsea temperature contrast is decreased with weakened Asian summer
monsoon.
4.5. Climate Feedbacks
Although the GCMs are useful tools, there are many other considerations that models have yet
to take into account. Among these are the feedbacks involved with climate changes. Just as
changes in climate influence the environment, changes in the environment will be reflected in
the climate. This feedback can be positive, heightening the temperature increase in global
warming, or they can be negative, minimizing the change.
The greatest degree of uncertainty about feedback lies with clouds. They absorb and reflect
solar radiation, or absorb and emit longwave radiation, depending on height, thickness, and
radiative properties. The reflected solar radiation dominates for low clouds with negative

feedback; with thin tropical cirrus clouds, the greenhouse effect dominates. Cloud
microphysical processes also affect climate. In warmer atmospheres, water clouds replace ice
clouds and cloud amount, especially at low- and mid-levels. Sulfate aerosols increase the
number of CCN, thus increasing the number of cloud drops and reducing mean particle size.
This results in an increased cloud optical thickness that increases cloud albedo, with a net
cooling effect. Shallow clouds drizzle, which depletes the cloud liquid-water and reduces
cloud reflectivity; if sulfur dioxide emissions cause cloud droplets to become smaller and
more numerous, this will impede drizzle production, leading to longer cloud lifetimes. While
it is impossible to judge even the sign of the sum of all cloud process feedback as they affect
greenhouse warming, it is unlikely that this sum is very negative or would result in more than
a doubling of the response occurring in their absence.
There are other important feedbacks when climate changes occur, including water vapor,
landocean interactions, and sea ice. The increase in temperature results in an increased
atmospheric water vapor amount, thus further enhancing the greenhouse effect. As water
vapor is mainly controlled by temperature in the lower troposphere and by transport process
in the upper troposphere, this creates positive feedback. The ocean influences the mean
climate, the annual variations of climate, and climate variations on time-scales as long as
millennia. It is through ocean surface variations that atmosphere and land are affected, and
these surface variations depend on the thermal and saline coupling between the deeper ocean
and the surface. The oceans are generally a source of local negative feedback. Sea ice reflects
solar radiation, but as the temperature warms and this ice melts, there is less reflection,
leading to an enhancement of the warming.
Current GCMs simulate the phase and amplitude of the seasonal cycles of large-scale
distribution of pressure, temperature, and circulation, although differences exist. The largest
discrepancies in the seasonal cycle of sea level pressure and surface air temperature are in the
higher latitudes, while the largest discrepancies in the seasonal cycle of precipitation are in the
tropics. Their local effects are also difficult to predict. Nevertheless, these models do work
well on a large-scale basis.
However, in recent years, concerns have been focused on the assessment of regional climate
changes on social and economic activities, and the associated policy implications. This will
require developing the capability for regional climate simulations. However, because of the
coarse grid size on the order of 5 x 104 km2 and larger in current GCMs, the information
cannot be used directly for regional impact study that requires a much higher resolution on the
order of 103 km2. Consequently, the development of regional modeling capability has received
the highest priority in climate modeling.
5. Concluding Remarks
It is clear that human activity has perturbed the global environmental system in an
unprecedented manner, in particular the changes in the concentrations of greenhouse gases
and particles. It is also expected that the trends of these activities will continue, leading to a
further increase in atmospheric constituents.
It is also clear that GCMs will be the primary tools for projecting future climate changes. In
addition to the continued effort in development and validation of GCMs, there are further
improvements in the radiative forcing representation in GCM simulation of future climate

changes. First, instead of using 1 percent CO2 increase to represent the total greenhouse gases
future growth, explicit treatment of the radiative forcing of CH4, N2O, CFCs, and other minor
gases and aerosols (in particular the sulfates) are considered. Second, changes of stratospheric
and tropospheric O3 are included in the climate change simulations. In this regard, the
development of the coupled climatechemistry models is emphasized. Thirdly, to more
realistically assess future climate changes, emission scenarios based on a range of economic
and technological assumptions (the Special Report on Emission Scenarios (SRES); IPCC,
1995) have been formulated for the period 2000100 so that future concentration levels can
be better estimated. Based on these scenarios, there has been a co-ordinated effort to use
several GCMs to simulate future climate changes so that the ensemble provide a better
assessment and quantification of potential climate changes for the twenty-first century.
The question to be addressed, therefore, is not whether, but how, the climate system will
respond. A follow-up and more important issue will be how the social and economic system
will respond to such adverse climate and environmental changes.
Related Chapters
Click Here To View The Related Chapters
Glossary
Aerosols :A suspension of fine solid or liquid particles in gas.
AGCM
:Atmospheric General Circulation Model.
C2F6
:Hexafluoroethane.
CCl2F2
:Dichlorotrifluoromethane.
CCl3F
:Trichlorofluoromethane.
CCl4
:Carbon tetrachloride.
CCN
:cloud condensation nuclei.
CF4
:Tetrafluoromethane.
CFC
:Chlorofluorocarbon.
CH3CCl3 :Methyl chloroform.
CH4
:Methane.
CHClF2
:Chlorodifluoromethane.
CO
:Carbon monoxide.
CO2
:Carbon dioxide.
ENSO
:El Nio-Southern Oscillation.
Greenhouse:Warming of the surface and lower atmosphere of a planet (as the earth or Venus)
effect
that is caused by conversion of solar radiation into heat in a process involving
selective transmission of short wave solar radiation by the atmosphere, its
absorption by the planets surface, and re-radiation as infrared that is absorbed
and partly reradiated back to the surface by atmospheric gases.
Global
:The progressive gradual rise of the earths surface temperature, thought to be
warming caused by the greenhouse effect and responsible for changes in global climate
patterns. An increase in the near surface temperature of the earth. Global
warming has occurred in the distant past as the result of natural influences, but
the term is most often used to refer to the warming predicted to occur as a result
of increased emissions of greenhouse gases.

H2O
HCFCs
HFCs
IPCC
IS92
N2O
NOx
O3
OGCM
PFCs
Radiative
forcing

RCMS
SA90
SAGE
SF6
SRES
TOMS
WMO

:Water vapor.
:Hydrochlorofluorocarbons.
:Hydrofluorocarbons.
:Intergovernmental Panel on Climate Change.
:IPCC 1992 Emission Scenarios.
:Nitrous oxide.
:Oxides of nitrogen (NO, NO2, NO3).
:Ozone.
:Oceanic General Circulation Model.
:Perfluorocarbons.
:A change in the balance between incoming solar radiation and outgoing infrared
radiation. Without any radiative forcing, solar radiation coming to the earth
would continue to be approximately equal to the infrared radiation emitted from
the earth. The addition of greenhouse gases traps an increased fraction of the
infrared radiation, radiating it back toward the surface and creating a warming
influence (i.e. positive radiative forcing because incoming solar radiation will
exceed outgoing infrared radiation).
:Regional climate models.
:IPCC 1990 Business-as-Usual Emission Scenario.
:Stratospheric Aerosol and Gas Experiment.
:Sulfur hexafluoride.
:Special Report on Emission Scenarios.
:Total Ozone Mapping Spectrometer.
:World Meteorological Organization.

Bibliography
Crowley, T. J. 2000. Causes of Climate Change over the Past 1000 Years. Science, No. 289, pp. 2707. [Climatic
changes during the last 1,000 years and the relevant causes for such changes are discussed.]
Easterling, D. R.; Meehl, G. A.; Parmesan, C.; Changnon, S. A.; Karl, T.; Mearns, L. O. 2000. Climate Extremes:
Observations, Modeling, and Impacts. Science, No. 289, pp. 206874. [Observed climatic extremes using
measurements are discussed and their spatial characteristics are used to check their consistency with climate
model simulations.]
Epstein, P. R. 2000. Is Global Warming Harmful to Health? Scientific American, No. 283, pp. 507. [The
implications of climate warming and its effect on health is discussed.]
Grassl, H. 2000. Status and Improvements of Coupled General Circulation Models. Science, No. 288, pp. 1991
7. [Progress of climate models is documented and areas for further improvement are indicated.]
Hall, A.; Manabe, S. 2000. Effect of Water Vapor Feedback on Internal and Anthropogenic Variations of the
Global Hydrologic Cycle. Journal of Geophysical Research, No. 105, pp. 693544. [Water vapor is an important
greenhouse gas and this paper presents its variability within the hydrologic cycle and its potential feedback
during anthropogenic activity.]
IPCC. 1995. Climate Change 1994: Radiative Forcing of Climate Change and an Evaluation of the IPCC IS92
Emission Scenarios. J. T. Houghton; L. G. Meira Filho; J. Bruce; H. Lee; B. A. Callander; E. Haites; N. Harris;
K. Maskell (eds.). Cambridge, Cambridge University Press. 339pp, [Radiative forcing due to various changes of
atmospheric constituents is a good indicator for comparing the subsequent global climate change. The study
examines the magnitudes of these radiative forcings and the potential future changes using IPCC scenarios.]
IPCC. 1996. Climate Change 1995: The Science of Climate Change. J. T. Houghton; L. G. Meira Filho; B. A.
Callander; N. Harris; A. Kattenberg; K. Maskell (eds.). Cambridge, Cambridge University Press. 572pp, [The

scientific basis for the climatic changes, in particular global warming, is summarized and potential future
changes of climate due to scenarios are suggested.]
IPCC. 2001. Climate Change 2001: The Scientific Basis. J. T. Houghton; Y. Ding; D. J. Griggs; M. Moguer; P. J.
Van Der Linden; L. Dai; K. Maskell; C. A. Johnson (eds.), 881pp, Cambridge, Cambridge University Press. [An
update of the IPCC, 1996.]
Karl, T. R.; Knight, R. W.; Baker, B. 2000. The Record Breaking Global Temperatures of 1997 and 1998:
Evidence for an Increase in the Rate of Global Warming? Geophysical Research. Letter, No. 27, pp. 71922.
[The measurements suggest an accelerated rate of temperatures during 1997 and 1998.]
Rosenfeld, D. 2000. Suppression of Rain and Snow by Urban and Industrial Air Pollution. Science, No. 287, pp.
17936. [The connection between air pollution and small-scale meteorology is discussed with the potential of
decreasing rain and snow amounts.]
Santer, B. D.; Wigley, T. M. L.; Gaffen, D. J.; Bengtsson, L.; Doutriaux, C.; Boyle, J. S.;. Esch, M.; Hnilo, J. J.;
Jones, P. D.; Meehl, G. A.; Roeckner, E.; Taylor, K. E.; Wehner, M. F. 2000. Interpreting Differential
Temperature Trends at the Surface and in the Lower Troposphere. Science, No. 287, pp. 122732. [The
differences in the surface temperature (from ground measurements) and tropospheric mean temperature (from
satellite measurements) are discussed and trends reconciled.]
Solomon, S. 1999. Stratospheric Ozone Depletion: A Review of Concepts and History. Review of Geophysics,
No. 37, pp. 275316. [Evolution of the findings concerning stratospheric ozone depletion is presented.]
Stott, P. A.; Tett, S. F. B.; Jones, G. S.; Allen, M. R.; Mitchell, J. F. B.; Jenkins, G. J. 2000. External Control of
20th-Century Temperature by Natural and Anthropogenic Forcings. Science, No. 290, pp. 2133 7. [The
attributes of twentieth-century temperature changes are discussed in the context of natural causes and human
induced activity.]
WMO. 1999. Albritton, D. L.; Aucamp, P. J.; Megie, G.; Watson, R. T. (eds.). Scientific Assessment of Ozone
Depletion: 1998. Global Ozone Research and Monitoring Project, Report No. 44, Geneva, World Meteorological
Organization. [The scientific basis of atmospheric ozone changes is summarized.]

Biographical Sketch
Dr Wei-Chyung Wang is a Professor of Applied Sciences at the State University of New York at Albany
(SUNYA), USA. Since September 1989, he has been the Head of the Climate System Sciences Section of the
Atmospheric Sciences Research Center at SUNYA. Prior to that time he was Vice President for Research at
Atmospheric and Environmental Research, Inc. in Cambridge, Mass. Prof. Wang received his B.Sc. from
National Cheng Kung University in Tainan, Taiwan in 1965; his M.Sc. from the State University of New York at
Buffalo in 1970; and his D.Eng.Sci. from Columbia University in New York in 1973. All three degrees were in
mechanical engineering.
Professor Wang has a broad background in atmospheric radiative transfer, climate modeling, and climate data
analysis. His research focuses on global and regional climate change due to increases of the atmospheric
constituents of greenhouse gases, CO2, O3, CH4, N2O, and CFCs and aerosols associated with human activities.
Professor Wang was the first to publish, in a 1976 issue of Science, an article identifying CH4 and N2O as
important greenhouse gases and in 1980, in Nature and Journal of the Atmospheric Sciences, articles indicating
the important climatic effect of ozone changes in the upper troposphere and lower stratosphere. Professor Wang
has been developing global and regional climate models for understanding the physical and chemical processes
concerning the greenhouse effect and stratospheric ozone depletion and for assessing future regional climate
changes. He is also engaged in research evaluating the effect and impact of climatic changes on social and
economic activities, and their policy implications. He has over 100 publications in more than twenty refereed
journals, including Science and Nature. In addition to conducting climate research, Professor Wang teaches
graduate courses and is active in graduate education related to global change. He has been a mentor of graduate
students in four SUNYA academic departments, the Atmospheric Sciences, Biology, Physics, and Political
Science.

To cite this chapter

Wei-Chyung Wang, (2004), ANTHROPOGENIC CAUSES OF GLOBAL ENVIRONMENTAL CHANGE,


in Anthropogenic Causes of Global Environmental Change, [Ed. Wei-Chyung Wang], in Encyclopedia of Life
Support Systems (EOLSS), Developed under the Auspices of the UNESCO, Eolss Publishers, Oxford ,UK,
[http://www.eolss.net] [Retrieved July 5, 2007]

You might also like