Sei sulla pagina 1di 34

The Carbon Cycle

by Elizabeth W. Sulzman

ob Gl

o Pr ry tion a rim pira t P es Ne R al &

c du

tio

Atmosphere

Vegetation Soils & Detritus

a Ch

ng

ng

La

nd

Us

e
Fossil Fuel Burning & Cement Production

Surface Ocean

Marine Biota Rivers r

DOC

Immediate & Deep Ocean

Surface Sediment

University Corporation for Atmospheric Research


National Center for Atmospheric Research UCAR Office of Programs

Understanding Global Change: Earth Science and Human Impacts

The Carbon Cycle


by Elizabeth W. Sulzman National Center for Atmospheric Research

Understanding Global Change: Earth Science and Human Impacts

Understanding Global Change: Earth Science and Human Impacts The Carbon Cycle by Elizabeth W. Sulzman
An instructional module produced by the Global Change Instruction Program of the University Corporation for Atmospheric Research with support from the National Science Foundation.

GCIP Staff
Tom M.L. Wigley, Scientific Director National Center for Atmospheric Research Lucy Warner, Program Manager University Corporation for Atmospheric Research Carol Rasmussen, Editor University Corporation for Atmospheric Research Linda Carbone, Secretary University Corporation for Atmospheric Research

Advisory Committee
Arthur Few Rice University John Firor National Center for Atmospheric Research William Moomaw Tufts University Ellen Mosley-Thompson The Ohio State University Jack Rhoton East Tennessee State University John Snow University of Oklahoma

2000 by the University Corporation for Atmospheric Research. All rights reserved. Any opinions, findings, conclusions, or recommentations expressed in this publication are those of the authors and donot necessarily reflect the views of the National Science Foundation. For more information on the Global Change Instruction Program, contact the UCAR Communications office, P.O. Box 3000, Boulder, CO 80307-3000. Phone: 303-497-8600; fax: 303-497-8610; lwarner@ucar.edu or carolr@ucar.edu http://home.ucar.edu/ucargen/education/gcmod/contents.html

ii

The Carbon Cycle

A note on this series


This series has been designed by college professors to fill an urgent need for interdisciplinary materials on global change. These materials are aimed at undergraduate students not majoring in science. The modular materials can be integrated into a number of existing coursesin earth science, biology, physics, astronomy, chemistry, meteorology, and the social sciences. They are written to capture the interest of the student who has little grounding in math and technical aspects of science but whose intellectual curiosity is piqued by concern for the environment. For a complete list of materials contact UCAR Communications (see previous page).

iii

Global Change Instruction Program

Contents
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .vii Chapter 1. Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1 The Importance of the Carbon Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1 Carbon Reservoirs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1 Chapter 2. Exchanges between Reservoirs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7 Exchanges between the Atmosphere and the Terrestrial Biosphere . . . . . . . . . . . . . . . . . . . . . .7 Exchanges between the Oceans and the Atmosphere . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .11 Exchanges betweeen the Terrestrial Biosphere and the Oceans . . . . . . . . . . . . . . . . . . . . . . . .14 The Anthropogenic Carbon Budget and Associated Uncertainties . . . . . . . . . . . . . . . . . . . . .14 Chapter 3. Feedbacks in the Carbon Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .16 Terrestrial Feedbacks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .16 Oceanic Feedbacks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .18 UV-B Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .19 Chapter 4. Implications for Global Climate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .20 Future Concentrations of CO2 in the Atmosphere . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .20 Options for Mitigating CO2 Release . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .21 Questions and Discussion Topics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .24 Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .25 Suggested Reading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .27 Other References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .28

Global Change Instruction Program

Preface
Carbon cycle research integrates ecology and the earth sciences. Although the importance of CO2 as a greenhouse gas was described by Svante Arrhenius in 1896, the importance of the carbon cycle was not fully realized until quite recently. Starting in 1957, direct measurements of atmospheric CO2 levels have revealed a rapid increase. In 1988, the severe consequences of this increase were acknowledged at an international level. Today, carbon cycle research is a very active field, with many issues yet to be resolved. For example, we do not yet know where all of the carbon emitted by human activities ends up. As a result of this and other unresolved issues, whole new methodologies are being developed and explored, such as the use of satellite imagery in conjunction with model calculations and precise measurements of chemical reactions involving carbon. This module is written with the nonscience major in mind, and so the basics are covered but some of the gritty details are left out. A fairly extensive list of suggested reading materials at the end of the module should point the inspired in the right direction. Because the carbon cycle is intimately linked to the cycles of other major nutrients, including nitrogen, phosphorus, and sulfur, I recommend Fred T. Mackenzies Global Change Instruction Program module, Global Biogeochemical Cycles and the Physical Climate System, as an accompaniment to this document. As CO2 plays a key role in climate, the module by Christine A. Ennis and Nancy H. Marcus, Biological Consequences of Global Climate Change, is also recommended. This module is broken into four chapters, each of which might take two or more class periods to cover. I have tried to make each chapter as independent as possible, with separate questions and discussion topics for each. The topics covered get progressively more complex from one chapter to the next. Elizabeth Sulzman Boulder, Colorado

vii

Global Change Instruction Program

Chapter 1: Overview
The Importance of the Carbon Cycle
Carbon is an element found in all living substances as well as in many inorganic materials. Both diamond and coal are nearly pure carbon, but with different structures. Carbon is a key element for life, composing almost half of the dry mass of the earths plants (that is, the mass when all water is removed). The carbon cycle is the exchange of carbon among three reservoirs or storage places: the land, the oceans, and the atmosphere. The amount of carbon in these reservoirs is so large that it is expressed in gigatons (Gt): 109 metric tons (1 metric ton equals 1,000 kilograms or 2,200 pounds). The atmosphere is the smallest pool of actively cycling carbon, that is, carbon that stays in a reservoir less than a thousand years or so. The land and its plants and animals, which scientists call the terrestrial biosphere, is the next largest reservoir of carbon. The oceans are the earths largest active carbon reservoir by far (Figure 1). The carbon budget is the balance of carbon among the three reservoirs. The carbon cycle is vitally important to life on earth. Through photosynthesis and respiration (see p. 3), it is the way the earth produces food and other renewable resources. Through decomposition, it serves as the earths waste disposal system. In addition, the carbon cycle is important because carbon-containing gases in the atmosphere affect the earths climate (see The Greenhouse Effect, p.2). Increased carbon dioxide (CO2) in the atmosphere has been responsible for more than half of the climate warming observed in recent decades. The processes by which carbon moves through the earths reservoirs take place on very different time scales. The short-term carbon cycle includes processes that transfer carbon from one reservoir to another in a matter of years, including photosynthesis, plant and animal respiration, and the movement of CO2 across the air-sea interface. Other processes, such as the transformation of carbon into limestone and its subsequent release as the rock is weathered, occur very slowly (thousands to millions of years). Scientists call these slower transfers of carbon among reservoirs the long-term carbon cycle. The carbon cycle is inextricably linked to other chemical cycles, including those of nitrogen, phosphorus, and sulfur, as well as to the global hydrological cycle. Please see the Global Change Instruction Program module Global Biogeochemical Cycles and the Physical Climate System or other works for information on these important topics.

Carbon Reservoirs
Actively cycling carbon in its three reservoirs affects human life every day. Carbon in the atmosphere serves as food for plants (in the presence of sunlight); carbon in the soil serves as energy for the growth of miniature animals called microbes; carbon in plants feeds humans and
Figure 1. Magnitudes of the reservoirs of actively cycling CO2 in gigatons of carbon. Of the ocean pool, roughly 1,000 Gt are in contact with the atmosphere in any given decade. The terrestrial biosphere consists of soil (ca. 1,600 Gt) and vegetation (ca. 600 Gt). Data from Schimel et al. (1995).(after Christensen, 1991).
754 2,190

Atmosphere Oceans Terrestrial Biosphere

39,120

Understanding Global Change: Earth Science and Human Impacts

The Greenhouse Effect


Some heat energy escapes outgoing into space.

other animals; and carbon in the ocean makes up the homes of many marine animals. Obviously, carbon is critically important in all of the reservoirs where it is found.

The Atmosphere
Incoming solar energy warms the earth's surface. Some heat energy is absorbed by greenhouse gases in the atmosphere and is reradiated back toward earth...

...keeping the surface comfortably warm. Warmed surface radiates heat energy back out toward space.

Figure 2. A simplified diagram of the greenhouse effect. Net input of solar radiation must be balanced by net output of radiation from the earths surface. About a third of incoming solar radiation is reflected, and the remainder is mostly absorbed by the surface. Some of the outgoing heat radiation is absorbed by greenhouse gases and by clouds, thereby keeping the surface about 33C warmer than it otherwise would be.

Carbon composes much less than 1% of the atmosphere. Even this small percentage adds up to 750 Gt of carbon, which is a lot (although, as noted earlier, much less than in the other two reservoirs). Carbon in the atmosphere occurs almost entirely as CO2; small amounts of methane (CH4), carbon monoxide (CO), and chlorofluorocarbons (CFCs) are also present. These gases are all radiatively active, that is, they trap heat near the surface of the earth. So they also are considered greenhouse gases. (See The Greenhouse Effect, below.)

The Greenhouse Effect


Like the glass walls of a greenhouse, certain gases trap heat. How do they do this? Shortwave radiation from the sun passes through the earths atmosphere and is absorbed by the surface, thereby warming it. Some of this absorbed energy is reradiated back toward space in a different form: as infrared radiation or heat. The greenhouse gases in the atmosphere absorb the heat energy and keep it from escaping back to space (Figure 2). Thus the heat stays in the atmosphere and warms the earth. The greenhouse effect is a natural process that has gone on for millions of years. Without it, life on earth would be impossible, because the earths temperature would be about 33C (60F) colder than it is today. Whereas the greenhouse effect is a good thing, an enhanced greenhouse effect might not be. Scientists are concerned that the rapid increase in the amount of CO2 and many other greenhouse gases in the atmosphere has implications for global climate. Figure 3 shows the relationship between atmospheric CO2 concentration and temperature over the last 160,000 years. The strong connection suggestsalthough it alone does not provethat the level of CO2 in the atmosphere plays an important role in determining the earths temperature. However, there is not a simple linear relationship between the amount of greenhouse gases in the atmosphere and the temperature at the surface of the earth. Complex mathematical models of the earths climate system are needed to study potential future consequences of our continued dependence on fossil-fuel energy. Carbon dioxide, water vapor, methane, nitrous oxide, and ozone are the greenhouse gases that occur naturally. The concentrations of all of these gases in the atmosphere are increasing as a result of human activities. The chlorofluorocarbons (CFCs) are also greenhouse gases, but they were created by humans; they did not exist in the atmosphere before the 1950s. The CFCs were manufactured for use as aerosol propellants, refrigerants, and solvents. Although their manufacture has been banned, their replacements are also very powerful greenhouse gases.

The Carbon Cycle

Photosynthesis and respiration In the short-term carbon cycle, photosynthesis and respiration are the primary processes that involve the atmosphere. Photosynthesis is the process by which green plants make their food. In this process, plants combine CO2 and water using light energy to make carbon-containing compounds (sugars and starches, called carbohydrates). Oxygen is produced during the reaction and released to the atmosphere. On land, plants use CO2 from the atmosphere (Figure 4). In the oceans, phytoplankton use CO2 dissolved in seawater; much of this dissolved CO2 also originally came from the atmosphere. Photosynthesis also releases oxygen to the atmosphere. Thus, thanks to the activity of photosynthetic organisms, all other forms of life on the earth have oxygen to breathe and food to eatsince even carnivores feed on animals that eat plants. Respiration is the chemical process by which carbon-containing compounds are broken down within cells. It is essentially the opposite of photosynthesis: oxygen and carbohydrates react to produce CO2 and water, releasing energy during the process. Living organisms use the energy released by respiration to power everything they do. For example, reading this page requires energy, and that energy is supplied by respiration.

280
Concentration of Carbon Dioxide in the Atmosphere (ppm)

240 220 200 180 +2.5 0 -2.5 -5.0 -7.5 Temperature Change 40 0 160 120 80 Thousands of Years Before Present -10.0

Figure 3. Variations in global temperature (bottom line) and atmospheric CO2 concentration (top line) over the last 160,000 years. The strong pattern of higher temperatures when atmospheric CO2 levels are high and cooler temperatures when concentrations are lower has been used to suggest that future increases in atmospheric CO2 concentration could lead to warmer temperatures. From Barnola et al. (1987), p. 410.

Figure 4. The flow of CO2 into plants via photosynthesis and out of plants and animals via respiration. Photosynthesis removes approximately as much CO2 from the atmosphere as respiration adds to it. This cycle keeps the atmospheric level of CO2 fairly constant.

Departure of Temperature from Current Level (C)

260

Carbon Dioxide

Understanding Global Change: Earth Science and Human Impacts

Globally, photosynthesis and respiration tend to be in balance in nature. In other words, photosynthesis takes up about as much CO2 from the atmosphere each year as is released by respiration. However, extreme weather eventssuch as prolonged flooding or drought over large areas or extended periods of unusually hot or cold temperatures can temporarily offset the delicate balance and lead to either a release of CO2 to the atmosphere or a high uptake of CO2 from the atmosphere. Increases in atmospheric carbon since 1750 As Table 1 shows, the carbon-containing compounds in the atmosphere have been increasing since the Industrial Revolution (about 1750 AD). At that time, humans began releasing more of these gases into the atmosphere by burning fossil fuels (e.g., gas, coal, oil), cutting down or burning forests to make agricultural land, and other processes. Although scientists have only been able to measure the atmospheric concentrations of these gases for about 40 years, they can estimate the concentrations in earlier periods from growth rings in trees, air trapped in polar ice cores, and other sources. Since 1957, the amount of carbon dioxide in the atmosphere has been measured directly at the South Pole; since 1958, at Mauna Loa, Hawaii; and starting later, at over 20 other locations around the world, including sites in Table 1 Atmospheric concentrations of important carbon-containing gases
Estimated preIndustrial Revolution Current concentration concentration Percentage (ca. 1750) (1999) of increase 280 ppm* 367 ppm (773 Gt) 1.8 ppm (5.1 Gt) 0.05 ppm (0.1 Gt) 31%

Antarctica, Australia, several maritime islands, and high northern latitudes. (At present, there are no continuous measurement sites in Africa or South America.) Figure 5 shows the records from these stations. Atmospheric CO2 has increased about 25% just since 1957, from 315 parts per million in 1957 to 367 ppm in 1999. The increase in atmospheric CO2 since 1957 is actually greater than the estimated increase over the preceding 200-plus years from 1750 to 1957. Atmospheric CO2 records of the past Measurements of CO2 concentrations in air extracted from polar ice cores are currently the best means to extend the CO2 record through the geologically recent past (last 200,000 years). The transformation of snow into ice traps air bubbles that are used to record the atmospheric CO2 concentration. Provided that the cores are extracted carefully from well-chosen sites and the analysis is done correctly, the ice record provides reliable information on past atmospheric CO2 concentrations. The ice-core record reveals an average value of 280 ppm in the level of atmospheric CO2 over the 1,000 years before the Industrial Revolution, with fluctuations up to 10 ppm. The largest of
Figure 5. The atmospheric CO2 increase between 1958 and 1994, recorded on Mauna Loa, Hawaii. The figure shows both the increase in atmospheric CO2 concentration since the start of measurements in 1958 and the seasonal cycle (zigzag pattern) of change in atmospheric CO2 concentration (see Why the Zigzags?, next page). Data from C.D. Keeling.
370 360 CO2 Concentration (ppm)

Gas Carbon dioxide

350

340 330 320 310 1950

Methane

0.7 ppm

157%

Carbon monoxide ~0.05 ppm

*ppm = parts per million by volume; out of 1 million molecules in the atmosphere before the Industrial Revolution, 280 molecules were carbon dioxide.

1960

1970 Year

1980

1990

2000

The Carbon Cycle

Why the Zigzags?


The zigzag pattern of increase shown in Figure 5 reflects the seasonal uptake and release of CO2 by plants. Levels of CO2 are high when plants are respiring more than they are photosynthesizing (winter) and are low when photosynthesis exceeds respiration (summer). Since Northern and Southern Hemisphere winters and summers occur at opposite times of the year, you might expect the opposing patterns from the two hemispheres to cancel each other out in the global record. However, the pattern for the midlatitude Northern Hemisphere prevails because its larger land masses support more seasonally active photosynthesis. In the seasonless tropics, photosynthesis goes on year round. The size of the zigzag, that is, the height of each peak (its amplitude), has generally increased over the past few decades. We are not sure why this is, because the trend is not well correlated with the increase in atmospheric CO2 concentration. The change could reflect either increased photosynthesis or increased respiration.

these fluctuations, which occurred between roughly 1200 and 1400 AD, was small compared to the 75 ppm increase the earth has experienced since the Industrial Revolution. Figure 6 illustrates how well the ice-core record of atmospheric CO2 concentrations matches the direct atmospheric record. It also shows both how stable the level of atmospheric CO2 was before 1850 and how rapidly that level has risen since then.

The Oceans
The oceans absorb and store about 39,000 Gt of carbonmore than 60 times as much as the atmosphere (see Figure 1). Over 95% of oceanic carbon is in the form of dissolved inorganic carbon (dissolved CO2 and bicarbonate and carbonate ions); the remainder comprises various forms of organic carbon (living organic matter and particulate and dissolved organic carbon). Scientists differentiate between the inorganic and organic forms of carbon because the two forms go

Figure 6. Historical levels of CO2 in the atmosphere. This figure combines information from Keelings direct atmospheric record (see Figure 4) and from air bubbles trapped in Antarctic ice cores. For comparison, the figure also shows the amount of CO2 released by human use of fossil fuels. Note that for a thousand years before 1850 the amount of CO2 in the atmosphere was quite constant, but that it began to increase rapidly after the Industrial Revolution, following the increased use of fossil fuels. From IPCC (1996), p. 16.

380 CO2 concentration (ppm) 6 360 4 340 2 320 0 300 280 1850 1900 1950 2000

CO2 concentration (ppm)

360 340 320


Fossil

300 280 260 800

Ice core record Direct atmospheric record

Year

1000

1200

1400

1600

1800

Fossil CO2 emissions (Gt C/yr)

380

2000

Year
5

Understanding Global Change: Earth Science and Human Impacts

through different chemical reactions and also are affected in different ways by physical ocean processes. The physical, chemical, and biological processes that affect carbon in the oceans are described in Chapter 2.

The Terrestrial Biosphere


Within this reservoir, the largest pool of carbon (roughly 65.5 million Gt) is in sedimentary rocksmostly limestone, a carbonate rock, and also organic sediments that contain fossil fuels. However, aside from human use of these fuels (which will be discussed in the next chapter), rocks do not play a role in the short-term carbon cycle. Carbon storage in plants and soils is more important. Plants and soils store and release carbon on time scales ranging from seasons to thousands of years, and they contain about 2,200 Gt carbon (C)almost three times as much total carbon as the atmosphere and more organic (but not more total) carbon than in the oceans (Figure 7). Not only is the amount of carbon in the terrestrial biosphere large, but it is readily modified by human activity. Therefore this reservoir may

be thought of as the most active part of the global carbon cycle. It is difficult, however, to pin down the role of the terrestrial biosphere in the global carbon cycle because of the complex biology underlying carbon storage, the great heterogeneity of vegetation and soils, and the somewhat unpredictable nature of human land use and land management. We will discuss some of these concerns in the next chapter.
Figure 7. Comparison of the land and oceanic carbon pools that are important in the short-term carbon cycle.

Oceanic Biomass 3 Gt C Oceanic Organic Carbon 1,020 Gt C

Land Biomass 610 Gt C

Soil Organic Carbon 1,580 Gt C

Global Change Instruction Program

Chapter 2: Exchanges between Reservoirs


In this chapter we will focus on how humans have affected the natural movement of carbon between its three reservoirs. This movement is called a flux or flow. Sources of carbon are releases from a particular reservoir to the atmosphere, and sinks are places where carbon is removed (taken up) from the atmosphere to another reservoir. Figure 8 depicts the reservoirs of the global carbon cycle and the flows of carbon between these reservoirs.

Exchanges between the Atmosphere and the Terrestrial Biosphere


Release of Fossil Fuels to the Atmosphere
When carbon that is already in the short-term carbon cycle changes form or flows to a different reservoir (for example, moving from the land to the atmosphere when organic matter is burned), the total amount of carbon in the cycle remains the same. The short-term carbon cycle is roughly

Figure 8. The global carbon cycle, showing pools of carbon (boxes) and exchanges between pools (arrows). The two processes controlling the carbon flow between the terrestrial biosphere and the atmosphere are changing land use and global net primary production (plant growth) and respiration. All flows are two-way except for the release of CO2 to the atmosphere from fossil-fuel burning and cement production and the burial of carbon in deep ocean sediments. DOC is dissolved organic carbon. From Schimel et al. (1995), p. 41.

ob Gl

u od Pr ry tion a rim pira t P es Ne R al &

cti

on

Atmosphere

Vegetation Soils & Detritus

Ch

an

g in

d an

Us

Fossil Fuel Burning & Cement Production

Surface Ocean

Marine Biota Rivers r

DOC

Immediate & Deep Ocean

Surface Sediment

Understanding Global Change: Earth Science and Human Impacts

balanced, and the amount of carbon-containing greenhouse gases is steady. Fossil fuels, on the other hand, are extracted from underground, so they are not a natural part of the short-term cycle. Thus, when fossil fuels are burned, carbon from long-term storage is added to the short-term cycle. When this happens, not only does the greenhouse effect increase because there are more carbon-containing greenhouse gases in the atmosphere, but the balance of the cycle is disrupted. Under current conditions, this means that either the terrestrial biosphere, or the oceans, or both may respond to the additional carbon in ways that we cannot yet predict. They may even store less carbon than they do today. Until about 150 years ago, fossil fuels were not an important part of the short-term cycle. Now, they must be included in any consideration of this cycle. The fossil-fuel reservoir is estimated at 5,000 Gt Calmost eight times the amount that is currently in the atmosphere. Figure 9 shows how the use of fossil fuels grew from 1950 to 1992. In 1950, approximately 1.6 Gt C from fossil fuels was burned. Current fossil-fuel emissions are about 6.7 Gt C per year. From 1950 to 1973, fossil-fuel use increased an average of 4.4% per year. Starting in the mid-1970s, the influence of world events on emission rates is dramatically evident in Figure 9. The oil embargo of 1973, during which most Middle Eastern countries stopped selling oil to the United States, led to a brief drop in emissions. A weak global economy in the early 1980s, especially in the former Soviet Union

(USSR) and Eastern Bloc, caused a reduction in emission levels. As the economy strengthened, emission rates rose again. The downward trend of the early 1990s may again have something to do with weak economic conditions in several parts of the industrialized world. Table 2 shows how much CO2 was released in one year by five countries, industrialized and nonindustrialized, through the burning of fossil fuels. Note that although Chinas total emissions were the greatest, the United States had a much higher emission level per capita. Releases in industrialized countries were as high as 5 tons C for each
Figure 9. Global CO2 emissions from fossil-fuel burning, cement production, and gas flaring from 1950 through 1992. Although there was a downward trend in the early 1990s, emissions for 1996 (not depicted) were again near levels for the late 1980s. Data from G. Marland. 6.5 Emissions (Gt C/yr) 5.5 4.5 3.5 2.5 1.5 1950

1960

1970 Year

1980

1990

Table 2 Fossil-fuelderived CO2 emissions per capita for five countries, 1987
Country Brazil China India United States USSR
1million metric tons.

Coal1 10 480 110 430 370

Oil1 38 840 35 540 340

Natural Gas1 2 7 3 240 300

Total Emissions 50 1390 148 1210 1010

Population 143 million 1.09 billion 800 million 243 million 273 million

Emissions per Capita 0.34 1.28 0.19 4.98 3.70

Emissions data from World Resource 19901991, Oxford University Press, 1990; population data from U.S. Census Bureau International Data Base, http://www.census.gov/ipc/www/idbnew.html.

The Carbon Cycle

person during the mid-1980s; in developing countries, per capita emissions at that time were 0.2 to 0.6 tons C.

Evidence That the Increased Emissions Come from Fossil Fuels


How do we know that the increased CO2 in the atmosphere comes from fossil-fuel burning and other human actions rather than from natural processes? An important indicator is provided by the isotopes (different chemical forms) of carbon (see Isotopes, below). Compared to atmospheric carbon, carbon originating in the terrestrial biosphereincluding fossil-fuel carbon, which started life as plant materialis also low in the isotope 13C relative to 12C. Measurements from ice cores show a general decrease over the last century in atmospheric levels of 13C relative to 12C, indicating that the increased carbon comes from either fossil-fuel burning, biomass burning, or both. Fossil-fuel carbon does not contain 14C, whereas both the ocean waters and the plants that were alive during the nuclear tests of the 1950s and 60s do. Thus the addition of fossil-fuel carbon to the active carbon cycle reduces the proportion of 14C in the atmosphere and also in plants, which use atmospheric carbon in photosynthesis. In fact, the 14C concentration measured in tree rings decreased

by about 2% from 1800 to 1950. This isotopic decrease, known as the Suess effect, is one of the most clear demonstrations that the increase in atmospheric CO2 is due to fossil inputs.

Other Emissions from the Terrestrial Biosphere


Plants, soils, and animals all release CO2 to the atmosphere. As we have seen, respiration by terrestrial plants is one of the primary means by which carbon is transferred naturally from the land to the atmosphere. The breakdown of dead plants into organic matter in soil and their further decomposition also release large amounts of CO2 to the atmosphere. Decomposition alone releases, on average, about 60 Gt C per year. Changes in land use and land management, especially deforestation, cause significant losses of terrestrial carbon. When land is converted from its natural state to cropland, approximately 90% of the carbon stored in the natural vegetation is lost, and 25% of the carbon in the top meter of soil is lost as well. Global land-use changes from 1850 to 1999 are estimated to have caused about 124 Gt C to be emitted to the atmosphere. Globally, carbon emissions from land-use changes dominated those from fossil-fuel burning until 1959. Until the 1940s, these emissions came principally from the cutting of temperate forests in the Northern Hemispheres middle and high

Isotopes
Isotopes are two or more forms of an element that have the same number of protons in the nucleus of the atom (called the atomic number) but a different number of neutrons. Isotopes have similar chemical properties, but since they differ in atomic weight (the number of protons plus the number of neutrons), they act differently in chemical reactions. The most abundant and stable form of carbon is 12C, or carbon-12 (98.98% of all natural C). The number 12 refers to the atomic weight of this carbon atom. Carbon-12 atoms have six protons and six neutrons. There are two common isotopes of carbon, 13C (carbon-13) and 14C (carbon-14). Carbon-13 has six protons and seven neutrons. Carbon-14 has six protons and eight neutrons and is radioactive; it is often called radiocarbon. The half-life of 14C, or amount of time it takes for half of the 14C radioactivity to decay, is roughly 5,700 years. Thus 14C survives in materials such as tree rings and corals that are tens of thousands of years old. It is useful for dating things (e.g., buried artifacts), and it is also used extensively as a tracer of ocean circulation. Carbon-13 is not radioactive, but because it is heavier than carbon-12, it is also a useful tracer, especially of physiological processes, such as photosynthesis by plants. Photosynthesis favors the lighter 12C over the heavier 13C. Because plants use more 12C than 13C, the carbon in the atmosphere contains more 13C than it would if there were no plants. Oxygen isotopes (18O, 16O) are also indicators of marine and terrestrial activity. As one unit of oxygen is produced for every unit of CO2 consumed during photosynthesis, measurement of oxygen isotopes reveals information about the cycling of carbon.

Understanding Global Change: Earth Science and Human Impacts

latitudes. Since then, the conversion of forests to agricultural land in the Northern Hemisphere has decreased, and forests are regrowing in previously harvested areas and on abandoned agricultural lands. In the tropics, however, emissions from land-use changes have been increasing since the 1950s (see sidebar, below).

Terrestrial Uptake from the Atmosphere


Changes in photosynthesis and respiration rates The biosphere takes up CO2 from the atmosphere naturally through photosynthesis at a rate of about 61 Gt per year. It is not certain how this rate would change with the increases in atmospheric CO2 that are expected to occur in the next century. Laboratory and some field studies of agricultural and wild plant species have shown typically 2040% higher photosynthesis and faster growth if the amount of CO2 in the atmosphere is doubled relative to todays concentrations and if the plants can get enough nutrients, light, and water (Figure 11). Although this effect (called the CO2 fertilization effect) is relatively strong, carbon storage in the earths ecosystems may be less responsive to increased levels of atmospheric CO2 than is plant growth. Various field studies have shown that after a while, increases in plant growth slow down, with the final rate being either the same as that of today or slightly higher. Moreover, whether increased levels of photosynthesis can be sustained in the real world, where factors other than CO2 (such as nutrients

and water) limit plant growth, remains a question. Few long-term field experiments have been conducted to address this issue. The results of those that have been carried out for more than three years seem to suggest that with unlimited nutrients and water (as in the case of marsh plants and irrigated, fertilized crops), high growth levels are sustained, whereas when water, nutrients, or both are limited, the initial high growth rates decline. We can use tree rings to study how plant growth has already changed over the long term in response to elevated CO2 and changing climate.
Figure 10. Reduction in the amount of carbon stored in forests of South and Southeast Asia as a result of land-use changes from 1850 to 1990. The solid curve shows carbon loss as a result of deforestation alone, and the dashed curve shows losses as a result of both deforestation and logging. From Houghton (1994), p. 309.
Carbon in Forests of Tropical Asia (Gt) 70 60 50 40 30 20 10 0 1850

1870 1890

1910 1930 1950 1970 1990 Year

How Deforestation Affects the Carbon Budget


Forests cover about 4.1 billion hectares (10.1 billion acres) of the earths surface, with 43% of those forests located in the tropics. During the 1980s, about 15 million hectares of forest were lost to deforestation per year in tropical areas. Although tropical deforestation has slowed since the 1980s, it is still growing at approximately 4% per year. When forests are cut to clear land for agricultural use, much carbon is released to the atmosphere when the wood is burned. Soil carbon storage decreases as a result of erosion. Furthermore, the removal of trees means that less carbon can be stored in wood in the terrestrial biosphere. Figure 10 shows that roughly 15 to 35 Gt C can no longer be stored in forests of South and Southeast Asia as a result of selective timber harvesting and complete deforestation. Fields store much less carbon than forests because they have less organic matter (no leaf litter); more aeration (as a result of plowing), which increases decomposition rates; and in some cases, wind or water erosion. One of the primary drivers of deforestation is the need for more agricultural land. Unfortunately, because of unsustainable agricultural practices, only half of the land converted from tropical forest to cropland contributes to an increase in agricultural productivity. The other half replaces land that has been degraded and is no longer productive. Thus, deforestation of the tropics yields a net loss of carbon and a loss of land suited to sustain most human or animal needs.

10

The Carbon Cycle

Elevated nutrients Photosynthesis

Ambient nutrients

CO2 Concentration
Figure 11. Photosynthetic response of plants to elevated CO2 under conditions of normal (ambient) nutrients and elevated nutrients. In both cases photosynthesis increases with elevated CO2 up to a point and then levels off, but the increase is much greater in plants given additional nutrients. From Bazazz (1990), p. 170.

In temperate ecosystems (e.g., the deciduous forests of the northeastern United States), each ring represents one years growth; the wider the ring, the more growth during that year. If increased atmospheric CO2 did cause trees to store more carbon, ring widths should have increased steadily over time since the Industrial Revolution (except in times of climatic extremes, e.g., severe drought). Tree ring studies of contemporary and preindustrial forest growth rates do not suggest ubiquitous or homogeneous increases in woody carbon storage with increasing CO2. Instead, they reveal a complex picture: enhanced growth in subalpine conifers, no growth enhancement in other tree species, and both increases and decreases in growth rates between species and sites. Other factors affecting terrestrial carbon storage One difficulty in understanding how humans have affected biosphere-atmosphere carbon exchanges is that we only have measurements of such processes as carbon accumulation in forests from modern times, when atmospheric concentrations of CO2 and nitrogen are already increasing. These modern measurements alone cannot

distinguish how other processes have already contributed to the carbon accumulation. With so many uncertainties, scientists do not even agree on whether forests are sources or sinks of carbon. We do know that human activities have led to increased storage of carbon in some systems. For example, whereas older forests may be nearly in balance with respect to carbon (the amount of carbon emitted through respiration and taken up through photosynthesis is almost the same), young trees that are growing rapidly take up more carbon. Thus, ecosystems that are currently recovering from a disturbance (e.g., a fire or harvest) take up more carbon than they did before they were disturbed. But dont forget that the disturbance itself released carbon, so the net effect of the disturbance plus recovery may be small. Another process that could be causing higher uptake of CO2 by plants is the deposition of nitrogen. Both pollution, including nitrogen oxides from automobile and airplane exhaust, and human use of fertilizers have increased the amount of nitrogen reaching both undisturbed land and land that has been adapted for human use. Because many ecosystems are nitrogen limited (i.e., their growth is stalled by a lack of nitrogen), this extra nitrogen allows increased plant growth, and thus greater uptake of CO2. Figure 12 shows the increase in carbon uptake that is expected to result from increased nitrogen use and release. Most of the increase is expected to occur in regions where large additions of nitrogen have taken placeexcept in agricultural areas like the central Great Plains of the United States, where little carbon can be stored because crops are harvested annually and little to no carbon is returned to the soil. Carbon storage as a result of nitrogen deposition is also low where deposition is low (e.g., South America, where little fertilizer is used) and where plants are water limited rather than nitrogen limited (e.g., North Africa).

Exchanges between the Oceans and the Atmosphere


The role of the oceans in the global carbon cycle is twofold. First, they absorb excess atmospheric CO2, and second, changes in the physical,

11

Understanding Global Change: Earth Science and Human Impacts

g C/m2 /yr 100+


+

75

50

25

0
Figure 12. Map of the global distribution of terrestrial carbon storage resulting from global deposition of nitrogen. This analysis considers fossil-fuel emissions plus 90% of nonfossil-fuel sources of nitric and nitrous oxide as well as 95% of the nitrogen deposited as a result of agricultural activities and assumes that 80% of the available nitrogen was taken up by plants. From Holland et al. (1997), p. 15, 861.

chemical, and biological state of the oceans may The unshaded portion of the loop represents the affect future atmospheric CO2. The ocean absorbs warm, shallow current, and the shaded portion excess CO2 in a three-step process: the CO2 gas represents the cold, salty, deep current. Water folpasses through the air-sea interface, it reacts lows this loop mainly because the ocean surface is chemically with dissolved inorganic carbon in the cooler at high latitudes than at low latitudes. This oceans surface waters, and then it is Figure 13. The conveyor belt movement of water around the globe. Water in the transported into the deeper waters.

Oceanic Circulation
Oceanic circulation is important in the carbon cycle because it affects carbon storage in the ocean. The ocean waters circulate mainly because of three factors: winds, which create surface currents; differences in saltiness throughout the oceans (caused by differences between evaporation and precipitation as well as by freezing and thawing); and differences in water temperature. Figure 13 shows the circulation of the oceans. The flow pattern is sometimes referred to as a conveyor belt because the water follows a continuous looping path, taking thousands of years to make a full circuit.

North Atlantic is very cold, and it is salty because the fresh water forms icebergs, leaving the salty water behind. Thus, it is more dense than in other areas, and so it sinks. The water travels along the ocean bottom until it reaches the tropical Pacific or the Indian Ocean, where it upwells, bringing nutrients to the surface. From Jilan et al. (1996), p. 271.

Great Ocean Conveyor Belt

Sea-to-Air Heat Transfer

Atlantic Ocean Indian Ocean Pacific Ocean

rm Wa
Co ld a

w llo ha

Cu

nt rr e

nd S

alty De ep Current

12

The Carbon Cycle

difference in temperature (thermal forcing) pushes the tropical surface waters poleward. CO2 is more soluble in cold water than in warmer water, so surface water from the tropics continues to absorb additional CO2 as it moves toward higher latitudes and colder temperatures. In the rainy North Atlantic, the addition of fresh waterwhich weighs less than saltwaterpushes the more saline water (and the extra CO2) toward the ocean floor. Only the movement of carbon from shallow to deeper waters in the conveyor-belt circulation allows large amounts of CO2 to be stored in the oceans (Figure 14). The surface water has a limited capacity to absorb CO2. It will even emit CO2 to the atmosphere if the local atmospheric concentration is lower than the concentration in the surface water.

Could the Oceanic Circulation Change?


A process called haline forcing pushes oceanic circulation in the opposite direction from thermal forcing. This process is the result of differences in saltiness between the high latitudes, where more freshwater enters the ocean from precipitation or runoff, and the lower latitudes, where the water becomes saltier through evaporation. Under the current climatic conditions, thermal forcing is stronger than haline forcing, so ocean circulation follows the path shown in Figure 13. However, a warming climate increases precipitation, runoff, and the melting of ice, strengthening the haline forcing. At some point haline forcing might prevail over thermal forcing, so that oceanic circulation as we know it could slow down or even stop. That change would drastically affect climate in many regions, such as western Europe, which is currently warmed by the Gulf Stream. water, and, to a lesser extent, the waters salinity and alkalinity (opposite of acidity).

Chemical Processes
Chemical processes in the ocean also affect carbon storage. When CO2 enters the ocean, it reacts chemically with the water to form carbonate and bicarbonate ions. (Ions are chemical compounds of atoms that have gained or lost electrons, leaving them with a positive or negative charge.) As more of these ions are formed, more atmospheric CO2 can enter the ocean. This reaction depends on temperature, the amount of dissolved inorganic carbon already in the surface

Biological Processes
Biological processes in the oceans are driven by the marine biota (oceanic plants and animals). The marine biota are believed to play a minor role, if any at all, in the oceans ability to absorb the CO2 added to the carbon cycle by humans. In

Figure 14. Schematic circulation of the oceans, showing the sinking of cold, nutrient-rich, high-CO2 water near the poles and the rise of warm, nutrient-poor water near the equator. This rise of water at the equator releases CO2 to the atmosphere. From Garrels et al. (1975), p. 40.

CO2 Out

Dense Cold Water

CO2 In

Warm Light Water Low N utrien ients ts Nutr Low High N utrient nts s utrie N igh H
Equator

CO2 In

South Pole North Pole

13

Understanding Global Change: Earth Science and Human Impacts

fact, marine biota contribute only about 3 Gt C to the worlds total carbon balance, as Figure 7 shows. The marine biota play a crucial role, however, in keeping the level of atmospheric CO2 steady. Phytoplankton (microscopic ocean plants) take up dissolved inorganic carbon through photosynthesis, just as terrestrial plants take up atmospheric CO2 during photosynthesis. If that dissolved inorganic carbon remained in the ocean waters, the oceans could not absorb as much CO2 from the atmosphere as they do. In fact, if there were no oceanic photosynthesis, the atmospheric concentration of CO2 would be approximately 520 ppmmuch higher than the mid-1990s level of 350360 ppm. In the open ocean, there is more carbon available than the marine biota use for photosynthesis. Scientists think that phytoplankton populations do not use this extra carbon to grow and multiply either because they dont have enough other nutrients and/or light or because zooplankton (minute ocean animals) graze on them and keep their numbers down. If the nutrients in the ocean were evenly distributed in areas suitable for phytoplankton to live, the increase in photosynthesis would cause todays atmospheric CO2 concentration to drop to only 230 ppm. Besides using dissolved inorganic carbon for photosynthesis, the marine biota use it to manufacture both inorganic compounds (e.g., shells) and organic matter, such as cell tissue. As the organisms die and sink, the organic material falls into the deeper ocean, where it is decomposed and converted back to dissolved inorganic carbon. This mechanism of carbon transport to the deep ocean is called the biological pump (Figure 15). If the ocean takes up more atmospheric CO2, thus increasing the amount of dissolved inorganic carbon in the shallower waters, this increase wont be passed on to the deeper waters via the biological pump. Photosynthesis is the first step in the biological pump, and, as we just noted, there is already more carbon in the ocean that the marine biota can use for photosysthesis.

Biological CO2

Physical CO2

Particulate Dissolved Cooling Organic Inorganic Inorganic Gulf Flux Upwelling

Deep Ocean Circulation CaCO3 Organic Carbon CaCO3

Figure 15. The natural cycling of carbon in the ocean. Biological and physical pumps remove CO2 from surface waters and thereby provide the main driving forces for the ocean carbon cycle.

Exchanges between the Terrestrial Biosphere and the Oceans


Carbon moves from the terrestrial biosphere to the oceans both indirectly, via the global hydrologic cycle (through evaporation from land and rain over the oceans), and directly, via rivers. Global rivers are thought to have discharged an average of 1.21.4 Gt C per year into the oceans in the 1980s. A substantial fraction of this amount (up to 0.8 Gt C/yr) results from the natural cycling of carbon. The rest is the result of human activity, mostly soil erosion. Unlike the oceans, the rivers do not appear to be taking up additional atmospheric CO2.

The Anthropogenic Carbon Budget and Associated Uncertainties


As we have seen, human activities, including deforestation and release of fossil fuels, have significantly perturbed the natural short-term carbon cycle, causing an increase in the magnitude of the fluxes between the main carbon reservoirs. Table 3 lists the calculated sources and sinks arising from

14

The Carbon Cycle

changes in land use and the burning of fossil fuels. These values for the 1990s total 7.9 Gt C in sources (with an uncertainty of 1.6 Gt C) but only 6.3 Gt C in sinks (with an uncertainty of 2.0 Gt C). Since no carbon is leaving or entering the earth, the sources and sinks in the carbon budget must balance; carbon that leaves one reservoir has to enter another. Thus the difference between sources and sinks in Table 3 shows that there is probably a sink we have not yet identifiedoften called the missing sinkof around 1.6 Gt C/yr. This sink may be any or all of several processes Table 3 Average annual carbon budget for human-caused emissions to the atmosphere, 1990s (in Gt C uncertainty)
Activity Fossil-fuel burning and cement production Tropical deforestation Storage in the atmosphere Uptake by the ocean Uptake by land Total 7.9 1.6 Source strength 6.3 0.6 1.6 1.0 3.3 0.2 2.3 0.8 0.7 1.0 6.3 2.0 Sink strength

for which we do not yet know the rates, including forest regrowth, increased plant growth due to CO2 fertilization, increased plant growth due to N deposition, and interactions among these factors. The existence of the missing sink shows that we do not yet fully understand all the processes involved in the carbon cycle. Although the estimated uncertainties in Table 3 are large, the entire picture is less uncertain than are the individual terms, because what is known about each term in the budget constrains the possible range of the other budget terms. If one flux (e.g., uptake by the ocean) is assumed to have an extremely high or low value, this puts constraints on the other terms. Atmospheric accumulation and fossil emissions are directly measured and are known within approximately 10%. The value for ocean uptake is based on some measurements but primarily on an understanding of oceanic physics (water movement) and chemistry. In this budget uptake by land has not been measured but is instead calculated from the other terms in the budget.

15

Global Change Instruction Program

Chapter 3: Feedbacks in the Carbon Cycle


Sometimes when one process in a system affects another process, the effect itself leads to a change in the first process. This change is called a feedback. Feedbacks can be either positive, in which the original process or effect is eventually enhanced, or negative, in which the original process is eventually reduced. For example, if you work hard and get a bonus, youre motivated to work even hardera positive feedback. If you dont study and get a bad midterm grade, fear of failing may make you study and bring up your grade; this is a negative feedback, even though it has a postive effect on you! A number of processes in the carbon cycle can produce feedbacks. Lack of knowledge about feedbacks is a major factor limiting our ability to estimate future atmospheric CO2 concentrations. For example, an increase in atmospheric CO2 that leads to an increase in the earths surface temperature may be a positive feedback in some regions because warmer temperatures increase soil respiration, thereby further increasing atmospheric CO2. However, warmer temperatures can also extend the growing season of plants, and thus the length of time over which they draw CO2 out of the atmosphere. This effect would reduce the amount of atmospheric CO2, which then could reduce the amount of warming. Thus increased atmospheric CO2 has both positive and negative feedbacks, and it is hard to discern which would prevail globally. by which they lose it. Warming increases respiration (which, you may remember, releases CO2 to the atmosphere), especially of soils, because respiration is more sensitive to temperature than is photosynthesis (which takes up CO2 from the atmosphere). Conversely, cooling decreases respiration. The reduction in the atmospheric growth rate of CO2 during 199192 may have been the result of a decrease in respiration caused by cooling, because the 1991 eruption of Mt. Pinatubo in the Philippines produced a global atmospheric cooling by releasing large quantities of sulfur dioxide that blocked sunlight. A number of processes in the terrestrial biosphere could compensate for the reduced carbon storage. For example, in nutrient-limited forests, warming may increase carbon storage by releasing organic nutrients from the soil, which the trees could use to grow and thus store more carbon in their tissues. Also, some plants whose growing seasons are now limited by cold may take up more CO2 than at present if global warming increases their life spans. The records of past changes of climate (such as the warming trend of 18,000 years ago) all suggest that climate change leads to major changes in terrestrial carbon storage. Scientists project that warming could lead to large losses of terrestrial carbon (~200 Gt) over the next few hundred years if temperature changes are very rapid. However, a gradual rather than a sudden change to warmer temperatures could cause the terrestrial biosphere to eventually take up 100300 Gt C in response to warming, albeit after a significant short-term loss.

Terrestrial Feedbacks
Carbon Storage in Plants
Climate change affects carbon storage in terrestrial ecosystems because temperature, moisture, and radiation influence both photosynthesis, by which ecosystems gain carbon, and respiration,

Carbon Storage in Soils


Soils generally store more carbon as latitude decreases (that is, as you move farther from the equator) because dead plants decompose more slowly in colder environments. Soil carbon

16

The Carbon Cycle

storage also increases under conditions of high precipitation. Flooded soils, which have extremely low rates of decomposition, may accumulate large amounts of organic matter as peat. If high-latitude soils get warmer and wet soils get drier as a result of climate change, some researchers have suggested that these soils would release more CO2 to the atmosphere, creating a large positive feedback to the initial warming. This process could be under way already in the peatlands of the arctic. Estimates of how much carbon would be lost from soils globally as temperature rises range from 11 to 34 Gt C per degree of warming. It is possible that the response will be only 5 to 17 Gt because as temperature rises decomposition will speed up, releasing nutrients that are needed for plant growth. Increased plant growth will eventually lead to increased carbon storage in the soil.

8 Net Ecosystem CO2 Fertilization Effect

(gC/m2/day)

-2 1983 1984 Year 1985

Effects of Increased CO2 Concentration (CO2 Fertilization Feedback)


As discussed in Chapter 2, laboratory studies have shown that photosynthesis increases in plants exposed to elevated CO2. If this short-term gain translates to entire ecosystems, in the future more carbon may be stored in woody tissue or in organic matter in soils. Enhanced storage would serve as a sink for excess CO2 and thus act as a negative feedback, decreasing the rate of atmospheric CO2 increase. Whether enhanced fertilization can be sustained, and whether enhanced photosynthesis will translate to increased carbon storage, are currently hot topics of debate. Figure 16 shows the results of a study in which plants of the Alaskan tundra in their native setting were exposed to elevated CO2 (680 ppm, compared to the 1996 value of ~360 ppm) for three years. The study suggests that not all plants will continue to grow faster with continued higher levels of CO2. CO2-fertilized foliage is typically lower in nitrogen than is foliage grown under current concentrations of CO2. As this lower-N foliage dies and works its way through the decomposition process, soil decomposer organisms will require additional nitrogen. As a result, vegetation may become more nitrogen limited (a change that would weaken but not eliminate the effects of CO2 fertilization). Additional nitrogen released by

Figure 16. The CO2 fertilization effect on plants grown for three seasons at Toolik Lake, Alaska, under doubled atmospheric CO2 concentration (680 ppm). The effect is considered to be the difference between the flux measured at doubled and at ambient CO2 levels. Units are grams of carbon per square meter of tundra vegetation per day. From Oechel et al. (1994), p. 500.

warming, as described above, could alleviate the nitrogen stress induced by foliage high in CO2. Deposition of atmospheric nitrogen could also influence the effectiveness of CO2 fertilization, although the effect may be modest, especially where other nutrients or conditions limit plant growth. Empirical studies are few, but a recent analysis revealed that hay yield had not changed measurably over the past 100 years despite substantial increases in the amount of nitrogen in precipitation and a 21% increase in atmospheric CO2 concentration during the period of study.

Changes in the Geographical Distribution of Vegetation Types


Studies of the records of past climates (paleorecords) suggest that if they are able to do so, plant types (e.g., grasses, deciduous forests) will migrate to their preferred climatic conditions when changes in climate occur. For example, the ranges of several species of trees have changed in tandem with changes in climate since the last glaciation. However, if climate change is rapid, many species will not be able to migrate fast enough to keep up with the new conditions. Furthermore, some of those that can migrate

17

Understanding Global Change: Earth Science and Human Impacts

quickly still will not survive the change, because soil conditions in the new area will not be suitable for them. Thus climate change is likely to lead to loss of habitat and subsequent death for some plants, and some species may become extinct. These changes would cause not only a release of carbon to the atmosphere if the dead plants (especially trees) were burned, but also a loss of uptake via photosynthesis. Table 5 reflects the likely direction of changes in carbon storage caused by factors described above. Short-term includes effects that will likely take place in the next 10 years, and long-term includes changes that will not be observed for 50100 years or more. A plus sign means an increase in carbon storage; a minus sign means a decrease. A question mark indicates a high degree of uncertainty.

Table 5 Direction of feedback effects on carbon storage


Feedback factor Increased atmospheric CO2 Increased temperature
1

Short-term + +/+ -/+

Long-term + (?) + +/+/+

Increased precipitation2 Increased N deposition3 Vegetation redistribution4


1

Long-term increase is possible due to release of soil nutrients. Increased precipitation increases both net primary productivity and decomposition; either effect could dominate. Long-term effects could be negative if the amount of N deposited becomes toxic. Short-term effects could be negative if plants cannot migrate fast enough to areas of newly suitable climate. Effects would be positive if trees replaced grasses.

Oceanic Feedbacks
Feedbacks Related to Chemical and Biological Processes
Since CO2 is more soluble in colder water than in warmer water, warmer sea surface temperatures will affect the oceans ability to dissolve CO2 and their carbon chemistry. A warmer ocean might cause dissolved organic carbon to decompose faster and convert to CO2, reducing the amount of atmospheric CO2 that can be absorbed by the oceans (a positive feedback). However, researchers do not think that this accelerated decomposition would significantly change the amount of atmospheric CO2 because the quantity of dissolved organic carbon in the surface waters of the ocean is less than 700 Gt; the atmosphere currently stores more than this amount (773 Gt). Warming is also expected to cause a decrease in the extent of sea ice, which could increase plankton and other marine growth in high-latitude regions. This would result in a greater uptake of atmospheric CO2, thereby acting as a negative feedback. assess. Warming of the sea surface is predicted to cause a strong reduction of the deep, salty current (see Figure 13, p. 12), which is driven by the cooling and sinking of surface water at high latitudes. This suggests that the oceans ability to serve as a sink for excess atmospheric CO2 would decrease with warming. Supporting this theory are studies of ocean behavior during El NioSouthern Oscillation (ENSO) events, which are characterized by reduced upwelling of cold bottom waters in the equatorial Pacific. These studies indicate that fluxes of CO2 to the atmosphere are enhanced during ENSO events. A reduction of the processes by which shallow and deep water are exchanged would impact the oceanic carbon cycle in several ways. The downward movement of surface waters, which have been in contact with the atmosphere and thus are enriched with excess CO2, would be reduced, thereby reducing the oceans capacity to take up excess CO2. Any decrease in the amount of water moving upward within the ocean would affect the biological pump by reducing the flow of nutrients to the ocean surface. In regions where nutrients are the limiting factor, this would result in reduced plankton and other marine growth.

Feedbacks Related to Ocean Circulation


Changes in oceanic circulation and their effects on the oceanic carbon cycle are difficult to

18

The Carbon Cycle

With less marine growth at the surface, less carbon would be exported to depth as organisms sink and decompose, which could lead to an increase in CO2 concentration at the oceans surface.

UV-B Radiation
The depletion of stratospheric ozone is an example of a human-caused change that impacts the oceanic carbon system even though it is not directly related to global warming. Ozone depletion could cause increased ultraviolet radiation (UV-B, corresponding to light wavelengths of 280320 nanometers, or billionths of a meter), which might affect marine life and thus influence marine carbon storage. However, even a complete cessation of marine productivity in high latitudes, where increases in UV-B are expected to occur, would result in an atmospheric CO2 increase of less than 40 ppm.

19

Global Change Instruction Program

Chapter 4: Implications for Global Climate


The schedule selection was relatively arbitrary many other choices could have been madebut all the schedules involved a smooth transition To prepare for a 1995 international meeting from 1990 emission levels to the levels that would regarding the state of the global environment (led be required to attain the target concentration. The by the United Nations Environment and project was designed to illustrate the relationship Development Program), the Intergovernmental between emissions and resulting CO2 concentraPanel on Climate Change (IPCC) requested an intions. This relationship is not straightforward, depth study of the carbon cycle. The participating because as atmospheric CO2 concentrations scientists were given a set of atmospheric CO2 increase, the percentage taken up by the ocean concentrations ranging from 350 to 1,000 ppm and the terrestrial biosphere decreases. (that is, from todays level to three times as great Figure 17 shows the emissions schedules that as todays). Their job was to produce schedules of would lead to stabilization at three levels: 450 total anthropogenic emissions that would allow ppm, 550 ppm, and 650 ppm. In all cases emisatmospheric CO2 to stabilize at the chosen levels. sions continue to rise initially and then drop sharply and rapidly. The IPCC parFigure 17. Emissions of CO2 (from fossil-fuel burning, deforestation, and cement proticipants allowed the initial rise duction) that would lead to stabilization of atmospheric concentrations at 450, 550, and because they considered it unlikely 650 ppm. After about the year 2050 emissions are assumed to drop drastically that governments would be able to because otherwise it would be very difficult to reach the target concentrations. Data enforce strict reductions before techfrom Schimel et al. (1995). nology improved and fossil-fuel use 12 became more efficient. The large decreases that follow do not represent predictions about changing poli10 650 ppm cy or technology; they are simply the patterns that would be necessary to 8 stabilize CO2 at the chosen levels. 550 ppm Stabilization at any of these concentration levels (all of which are higher 6 than todays level) is only possible if 450 ppm emissions are eventually reduced 4 well below 1990 levels. For comparison, Figure 18 shows the atmospheric CO2 concentrations 2 resulting from three scenarios about future increases in human popula0 tion, economic growth, and energy 1990 2050 2110 2170 2230 2290 needs. (These scenarios, labeled

Future Concentrations of CO2 in the Atmosphere

Anthropogenic Emissions (Gt C/yr)

Year

20

The Carbon Cycle

950 850 Concentration (ppm) 750 650 550

is92e

is92a

is92c 450 350 250 1990 2000

disturbance of the carbon cycle and the atmospheric concentration of CO2, rather than as definite predictions of future atmospheric concentrations. As one group of scientists noted, An improved understanding of the carbon cycle is essential to predict the future rate of any atmospheric CO2 increase and to plan eventually for an international CO2 management strategy (Tans et al., 1990).

2020

2040 Year

2060

2080

2100

Options for Mitigating CO2 Releases

In an ideal world, mitigation options would not be necessary because people Figure 18. Future atmospheric CO2 concentrations for three scenarios of would realize the potential damage caused anthropogenic emissions. IS92a: world population 11.3 billion by 2100; ecoby perturbation of the natural carbon cycle nomic growth 2.9% per year from 1990 to 2025; energy supplies mostly oil and natural gas, some solar and biofuels. IS92c: world population 6.4 billion and would work to curtail emissions from by 2100; economic growth 2.0% per year for 19902025; less oil and natural land use and use of fossil fuels. This not gas than IS92a, more nuclear power. IS92e: world population 11.3 billion by being a perfect world, there are several 2100; economic growth 3.5% from 1990 to 2025; more oil, same gas as options for mitigating human-induced carIS92a, phaseout of nuclear power. Data from Schimel et al. (1995). bon cycle perturbation. The options center on three basic themes: (1) reduce emissions, (2) switch to a renewable energy source, IS92a, c, and e, were developed in a 1992 IPCC and (3) increase the amount of carbon stored in report.) IS92c is considered optimistic in terms the ocean and in the terrestrial biosphere. of emissions because it assumes an eventual Obviously, these themes overlap. For example, decrease in human population. Scenario IS92e is switching to a renewable energy source, such as considered pessimistic by some because it solar power, reduces emissions from fossil fuels. assumes a phase-out of nuclear power (which Likewise, planting forests to store carbon creates most people believe would lead to increased fosa renewable energy source. sil-fuel emissions), high economic growth, and Emissions will be reduced only if new techmoderate increases in human population. nology becomes available that makes conservaScenario IS92a is considered by some to be realtion economically sound or if governments force istic in its projection of human population the reduction, through either subsidies, incengrowth, fossil-fuel use, and economic growth. tives, or taxes (disincentives). Keep in mind that In fact, the relative realism of all the scenarany major transformation of the energy, industry, ios is hotly debated, since for a variety of reasons and transportation sectors of the economy is not they do not take into account all of the factors likely to happen overnight, so emissions will conknown to affect emissions. For example, they do tinue to grow in the immediate future. not include the effects of changes in land use, We are probably all familiar with some of the changes in the nitrogen cycle, or climate feedbacks. options for reducing emissions, such as taking The effects of changing land use alone could add public transportation instead of driving a car and hundreds of Gt C to the atmosphere over the next keeping buildings cooler in the winter and one to several centuries, or could eventually store warmer in the summer. This section will concensimilar amounts over many centuries. trate on the other types of mitigation options, Thus these projections serve only as illustrathose involving alternative energy sources and tions of the relationship between human

21

Understanding Global Change: Earth Science and Human Impacts

carbon storage. One issue that must be remembered when assessing any of these options is whether it contributes to or undermines the goal of supporting a growing human population and simultaneously providing the required amounts of food, fodder, fiber, and biomass. An analysis of the socioeconomic costs associated with various mitigation options is presented in IPCC (1997).

Table 6 Carbon stored in trees and forest litter as a percentage of national fossil-fuel emissions After Brown et al. (1996).
Carbon Stored Fossil-Fuel in Trees Carbon and Litter Emissions (Mt)* (Mt/yr) >87 113 978 1,5002,000 12,000 10,000 1,113 18,585 164 8 18 268 136 137 131 1,300 Percentage of Fossil-Fuel Emissions Removed by Forests 1.5 44 28 2 37.5 3.6 6 6

Developing Renewable Energy Sources


Development of renewable energy sources would reduce emissions of traditional fuels. However, this transformation would have to be economically feasible, either through technological developments or via incentives. For example, vegetable oil crops are already used to produce biodiesel, and this fuel can be used in current diesel engines. The cost of biodiesel, however, is slightly higher than the cost of petroleum diesel, and so its use is not common. Planting biofuel crops not only promotes alternative energy use, it also reduces emissions of traditional fuels and temporarily increases carbon storage. It is estimated that dedicated energy plants could be grown sustainably on 811% of marginal to good cropland in the temperate zone. One drawback is that biofuel crops compete with food crops for limited resources (e.g., water and nutrients). Use of recycled wood and paper products and industrial timber and paper industry wastes as biofuels is an interesting option, but one that is not yet feasible.

Country Great Britain New Zealand Finland Germany Canada India Poland USA

*Mt is millions of tons.

depicts the amount of carbon that can be either stored or saved (not released to the atmosphere) by forests in several latitudinal bands. Note that the forests of the tropics have the greatest potential to affect future atmospheric CO2 levels. In some parts of the ocean, lack of iron (rather than lack of nutrients) appears to limit marine plant growth. It has been suggested that
Figure 19. Average annual rate of carbon uptake and storage per decade through forest management practices by latitudinal region. Note that the forests of the tropics have the greatest potential to reduce future atmospheric CO2 concentrations. From Brown et al. (1996), p. 786.
C Sequestered or Conserved (Gt/yr) 2.5 2.0 Tropics Temperate Boreal Total

Increasing Carbon Storage


Where sufficient land is available, planting forests can result in a fairly long-term (50 to hundreds of years) storage of carbon. Forestry currently offsets 90% of carbon released in Sweden through fossil-fuel burning. Table 6 shows the percentage of fossil-fuel emissions that is offset by forestry in eight other countries. In addition to serving as a storehouse for carbon, forests, like agricultural crops, can be used as an alternative source of energy. Wood grown in plantations can be used to generate electricity instead of using coal; this substitution can dramatically reduce CO2 emissions to the atmosphere. Figure 19

1.5

1.0

0.5

1995

2005

2015 Year

2025

2035

2045

22

The Carbon Cycle

increasing oceanic iron concentrations in these areas could speed up the biological pump and thus reduce atmospheric CO2 levels. However, experiments in adding iron to ocean waters (called iron fertilization) have shown that although productivity was greatly increased at first, the final effect was negligible because the additional plant material simply joined the food chain. In other words, iron fertilization led to more sharks rather than to less atmospheric CO2. Most scientists now believe that even excessive iron fertilization of selected areas of the ocean would have a relatively small impact on atmospheric CO2 levels. Several carbon storage options also not only reduce climate change but increase or restore ecosystem health. One such option is to convert marginal or surplus cropland and pasture to forest or natural (ungrazed) grassland. Another is to

improve land-management practices, such as returning crop residues to fields, decreasing periods of fallow (when erosion danger is high), and reducing tillage. These changes would both increase carbon storage and reduce its loss in agricultural fields. Because soils and vegetation have a finite capacity to store carbon, mitigation options that increase carbon storage are not sustainable over many centuries. To stabilize atmospheric CO2 concentrations, eventually we will have to cut emissions, preferably via a combination of alternative fuel sources and increased efficiency.

23

Global Change Instruction Program

Questions and Discussion Topics


Chapter 1
1. What is the smallest reservoir of carbon? State two reasons why this reservoir is important. 2. What is the largest reservoir of carbon? 3. What causes the zigzag pattern in the atmospheric CO2 record? 4. The amplitude (distance between peaks and valleys) of the seasonal cycle of CO2 is greater at Barrow, Alaska, than it is at Mauna Loa, Hawaii. Discuss why. 13. How would a warmer, drier climate affect soil carbon storage? Would this effect be a positive or a negative feedback to the global carbon cycle? 14.How are nitrogen levels in vegetation affected by increased atmospheric CO2? Could altered nitrogen levels affect future atmospheric CO2? If so, how? 15.What regions of the globe are most likely to store additional carbon as a result of N deposition? 16.What scientific evidence suggests that plants may relocate under an altered future climate? Is there any reason we should worry about plants if the future climate is very different from todays?

Chapter 2
5. What are the major sources of CO2 to the atmosphere? What are the main sinks that can take up CO2 released by anthropogenic activities? 6. Why did growth in emissions of fossil fuels decrease after 1973? Are there any other periods of lower emissions in the record? What might have caused these? 7. What is the Suess effect, and what does it tell us? 8. What is radiocarbon, and what is it used for? 9. What is the biological pump, and how does it affect atmospheric levels of CO2? 10. What is the missing sink? Does it involve carbon that is missing from the carbon cycle?

Chapter 4
17. In Figure 17, what is the approximate year in which emissions begin to level off for each of the three scenarios shown? 18. List two ways forests can affect future atmospheric CO2 levels. Do both have the same lifetimethat is, will the effects of each last the same amount of time? 19. Some actions to reduce atmospheric CO2 levels are called no regrets options, meaning that they have positive effects in addition to their potential to reduce atmospheric CO2 levels. List three no regrets actions and describe their positive environmental effects.

Chapter 3
11. State how an increase in temperature could increase or decrease carbon storage (a) on land and (b) in the ocean. 12.How can an increased level of atmospheric CO2 act as a feedback to the global carbon cycle through its influence on plant growth?

24

Global Change Instruction Program

Glossary
Anthropogenicinvolving the impact of humans on nature; induced or altered by the presence or activities of humans Biofuela nonfossil, natural, carbon-based fuel. Biological pumpthe export of newly produced organic carbon from the surface ocean (mostly via sinking) and regeneration of dissolved inorganic carbon at depth. Biosphereall areas on earth where living organisms are found; includes oceans, land, and lower atmosphere. Biotathe animal and plant life of a region. Carbon cyclethe complex array of chemical, physical, and biological processes by which carbon flows through the living and nonliving environment. CO2 fertilization effectthe enhancement of plant growth as a result of elevated atmospheric CO2 levels. Decompositionthe decay of dead organic material; decomposition is carried out by microorganisms and other decomposers that break down complex molecules and release nutrients. Dissolved inorganic carboninorganic carbon molecules (chiefly carbonate and bicarbonate ions) that are in solution in water; accounts for 95% of ocean carbon. Dissolved organic carbonorganic carbon compounds (e.g., from dead marine plants and animals) that have dissolved in water. El NioSouthern OscillationEl Nio is an oceanic event that historically has occurred roughly every three to seven years, when warm surface seawater moves eastward along the equator from the western Pacific toward South America. The Southern Oscillation is a variation in air pressure over the South Pacific between Tahiti and Australia. The two phenomena are closely linked. Fallowleft unseeded after being plowed; uncultivated. Feedbackthe return of a portion of the output from any process or system to the systems input; the return may either add to the initial input (positive feedback) or subtract from it (negative feedback). Flow/fluxthe rate at which a variable enters or leaves a reservoir. Fluctuations/oscillationsvariations in the value of a variable, usually around the variables average value. Ice corea tube of ice taken from a glacier (e.g., in Greenland or Antarctica), used by scientists to study the properties of samples of trapped air for which the age can be determined. Ionan atom that carries a positive or negative charge as a result of having lost or gained an electron. Isotopeone of two or more species of atoms of the same chemical element that have the same atomic number and nearly identical chemical behavior, but differ in mass.

25

Understanding Global Change: Earth Science and Human Impacts

Net primary productionthe accumulation of organic matter in plants, calculated as the difference between photosynthesis and respiration. Paleoecologythe branch of ecology concerned with identifying and interpreting the relationships of ancient plants and animals to their environment. Paleorecordrecord of fossil data, including trapped and preserved tree pollen (from animals nests, feces) and growth records (tree rings), air bubbles (ice cores), and marine plants and animals (sea sediments). Peata substance consisting of partially carbonized vegetable material, chiefly mosses, usually found in bogs. Photosynthesisthe process by which light, carbon dioxide, and water are used by plants to make carbohydrate products, essentially converting light to chemical energy. Phytoplanktontiny photosynthetic plants that live in oceans and lakes; phytoplankton form the base of the aquatic food chain. Renewable resourcea resource that can be replaced as fast as it is exploited. Respirationan oxygen-consuming metabolic process used by plants and animals to break down organic substances, yielding energy and releasing carbon dioxide. Sinka reservoir that takes up excess of a variable from a system; e.g., the ocean is a sink for excess CO2. Sourcea reservoir that supplies a variable to a system; e.g., fossil fuels are a source of CO2 to the actively cycling carbon pool. Speciesa fundamental category in the classification of living organisms; organisms of the same species share common characteristics and appearance. Subalpinean ecosystem type usually characterized by spruce and fir trees, this area is immediately below timberline.

Temperate zonethe areas between the Tropic of Cancer and the Arctic Circle and between the Tropic of Capricorn and the Antarctic Circle. Thermohaline circulationa conveyor belt movement of water in the ocean thought to originate with the sinking of cold water in the North Atlantic and driven by temperature and salinity differences. Tillagethe turning over of soil in agricultural fields. Upwellingthe upward movement of water from depths of typically 50150 m at speeds of approximately 13 m per day, resulting from the lateral movement of surface water. Zooplanktonanimal life, often microscopic, that drifts in oceans or lakes.

26

Global Change Instruction Program

Suggested Reading
Bolin, B., E.T. Degens, S. Kempe, and P. Ketner, 1979: The Global Carbon Cycle. John Wiley & Sons, New York, 491 pp. Brown, S., J. Sathaye, M. Cannell, and P. Kauppi, 1996: Management of forests for mitigation of greenhouse gas emissions. In Climate Change 1995: Impacts, Adaptations and Mitigation of Climate Change; Scientific-Technical Analyses; Contribution of Working Group II to the Second Assessment Report of the Intergovernmental Panel on Climate Change (R.T. Watson, M.C. Zinyowera, and R.H. Moss, eds.), Cambridge University Press, Cambridge, U.K., 773779. Cole, V., C. Cerri, K. Minami, A. Mosier, N. Rosenberg, D. Sauerbeck, J. Dumanski, J. Duxbury, J. Freney, R. Gupta, O. Heinemeyer, T. Kolchugina, J. Lee, K. Paustian, D. Powlson, N. Sampson, H. Tiessen, M. van Noordwijk, and Q. Zhao, 1996: Agricultural options for mitigation of greenhouse gas emissions. In Climate Change 1995: Impacts, Adaptations and Mitigation of Climate Change; Scientific-Technical Analyses; Contribution of Working Group II to the Second Assessment Report of the Intergovernmental Panel on Climate Change (R.T. Watson, M.C. Zinyowera, and R.H. Moss, eds.), Cambridge University Press, Cambridge, U.K., 745771. Ennis, C.A., and N.H. Marcus, 1994: Biological Consequences of Global Climate Change. Global Change Instruction Program, University Corporation for Atmospheric Research, Boulder, Colorado, 49 pp. Heimann, M., 1993: The Global Carbon Cycle. Springer-Verlag, New York, 599 pp. Houghton, R.A., and J.L. Hackler, 1994: The net flux of carbon from deforestation and degradation in South and Southeast Asia. In Effects of Land Use Change on Atmospheric CO2 Concentrations: Southeast Asia as a Case Study (V.H. Dale, ed.), Springer-Verlag, New York, 301327. Houghton, R.A., and D.L. Skole, 1990: Carbon. In The Earth as Transformed by Human Action (B.L. Turner II, W.C. Clark, R.W. Kates, J.F. Richards, J.T. Mathews, and W.B. Meyer, eds.), Cambridge University Press, New York, 393408. Intergovernmental Panel on Climate Change (IPCC), 1990: Climate Change: The IPCC Scientific Assessment. Cambridge University Press, Cambridge, U.K., 365 pp. IPCC, 1996: Climate Change 1995: The Science of Climate Change; Contribution of Working Group I to the Second Assessment Report of the Intergovernmental Panel on Climate Change (J.T. Houghton, L.G. Meira Filho, B.A. Callander, N. Harris, A. Kattenberg, and K. Maskell, eds.), Cambridge University Press, Cambridge, U.K., 572 pp. IPCC, 1997: Stabilisation of Atmospheric Greenhouse Gases: Physical, Biological, and Socioeconomic Implications (J.T. Houghton, L.G. Meira Filho, D.J. Griggs, and K. Maskell, eds.). United Nations Environment Programme/World Meteorological Organization, Geneva, Switzerland, 48 pp. Keeling, R.F., and R. Shertz, 1992: Seasonal and interannual variations in atmospheric oxygen and implications for the global carbon cycle. Nature 358, 723727. Koch, G.W., and H.A. Mooney, 1996: Carbon Dioxide and Terrestrial Ecosystems. Academic Press, San Diego, California, 443 pp.

27

Understanding Global Change: Earth Science and Human Impacts

Leemans, R., and G.J. van den Born, 1994: Determining the potential distribution of vegetation, crops and agricultural productivity. Water, Air, and Soil Pollution 76, 133161. Mackenzie, F.T., 1997: Global Biogeochemical Cycles and the Physical Climate System. Global Change Instruction Program. University Corporation for Atmospheric Research, Boulder, Colorado, 75 pp. Post, W.M., T.-H. Peng, W.R. Emanuel, A.W. King, V.H. Dale, and D.L. DeAngelis, 1990: The global carbon cycle. American Scientist 78, 310326. Schimel, D.S., 1995: Terrestrial ecosystems and the carbon cycle. Global Change Biology 1, 7791. Schimel, D., I.G. Enting, M. Heimann, T.M.L. Wigley, D. Raynaud, D. Alves, and U. Siegenthaler, 1995: CO2 and the carbon cycle. In 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, and K. Maskell, eds.), Cambridge University Press, Cambridge, U.K., 3571. Schlesinger, W.H., 1991: Biogeochemistry: An Analysis of Global Change. Academic Press, San Diego, California, 443 pp. Tans, P.P., I.Y. Fung, and T. Takahashi, 1990: Observational constraints on the global atmospheric CO2 budget. Science 247, 14311438. Townsend, A.R., B.H. Braswell, E.A. Holland, and J.E. Penner, 1996: Spatial and temporal patterns in terrestrial carbon storage due to deposition of anthropogenic nitrogen. Ecological Applications 6, 806814. Trabalka, J.R., 1985: Atmospheric Carbon Dioxide and the Global Carbon Cycle. Office of Energy Research, U.S. Department of Energy, Washington, D.C., 316 pp. Trenberth, K.E., ed., 1992: Climate System Modeling. Cambridge University Press, Cambridge, U.K., 788 pp. United Nations, 1992: Earth Summit Convention on Climate Change, 314 June 1992. United Nations Conference on Environment and Development, Rio de Janeiro, Brazil, 21.

Wisniewski, J., and A.E. Lugo, 1992: Natural Sinks of CO2. Kluwer Academic Publishers, Boston, Massachusetts.

Other References
Barnola, J.-M., D. Raynaud, Y.S. Korotkevitch, and C. Lorius, 1987: Vostok ice core provides 160,000 year record of atmospheric CO2. Nature 329, p. 410. Bazzaz, F., 1990: The response of natural ecosystems to the rising global CO2 levels. Annual Review of Ecology and Systematics 21, p. 170. Garrels, R.M., F.T. Mackenzie, and C. Hunt, eds., 1975: Chemical Cycles and the Global Environment: Assessing Human Influences. William Kaufmann, Inc., Los Altos, California. Holland, E.A., B.H. Braswell, J.-F. Lamarque, A. Townsend, J.M. Sulzman, J.-F. Mller, F. Dentener, G. Brasseur, H. Levy II, J.E. Penner, and G. Roelofs, 1997: Variations in the predicted spatial distribution of atmospheric nitrogen deposition and their impact on carbon uptake by terrestrial ecosystems. Journal of Geophysical Research 102, p. 15,861. Houghton, R.A., 1994: The worldwide extent of land-use change. BioScience 44, p. 309. Jilan, S., E. Miles, E. Desa, B.N. Desai, J.T. Everett, J.J. Magnuson, A. Tsyban, and S. Zuta, 1996: Oceans. In Climate Change 1995: Impacts, Adaptations and Mitigation of Climate Change: Scientific-Technical Analyses; Contribution of Working Group II to the Second Assessment Report of the Intergovernmental Panel on Climate Change (R.T. Watson, M.C. Zinyowera, and R.H. Moss, eds.), Cambridge University Press, Cambridge, U.K., p. 271. Oechel, W.C., S. Cowies, N. Grulke, S.J. Hastings, B. Lawrence, T. Prudhomme, G. Riechers, B. Strain, D. Tissue, and G. Vourlitis, 1994: Transient nature of CO2 fertilization in Arctic tundra. Nature 371.

28

Potrebbero piacerti anche