Sie sind auf Seite 1von 205

SCOPE 21 -The Major Biogeochemical Cycles and Their

Interactions
1 Interactions of Biogeochemical Cycles
B. BOLIN, P. J. CRUTZEN, P. M. VITOUSEK, R. G. WOODMANSEE,
E. D. GOLDBERG AND R. B. COOK

An overview of contributions and discussions at the SCOPE


Workshop on the Interaction of Biogeochemical Cycles,
Örsundsbro, Sweden, 2530 May 1981.

1.1 Global Biogeochemical Cycles and Man


1.2 Biological Processes and Interactions of Biogeochemical Cycles
1.3 The Atmosphere
1.3.1 AtmosphereBiosphere Interaction
1.3.2 The pre-industrial and the present atmosphere
1.3.3 Looking towards the future
1.4 Terrestrial Ecosystems
1.4.1 Interactions of C, N, S and P in Natural Terrestrial Ecosystems
1.4.2 Natural Disturbances
1.4.3 Human Influences
1.4.4 The Impact of Fossil Fuel Combustion on the Global Ecosystem
1.5 The Freshwater System and Coastal Waters
1.5.1 Water as an Agent in Biogeochemical Cycles
1.5.2 The Natural System
1.5.3 Disturbances in the Weathering Cycle Created by Man
1.5.4 Human Waste and the Waters of the World
1.6 The Open Sea
1.6.1 Response Characteristics of the Oceans
1.6.2 Life in the Sea
1.7 Conclusions
References

1.1 GLOBAL BIOGEOCHEMICAL CYCLES AND MAN

The elements of carbon, nitrogen, phosphorus and sulphur, and of course hydrogen and
oxygen, are needed for the formation of the fundamental molecules on which life depends.
Their availability, circulation and interaction in nature has therefore been decisive for the
development of life on earth and also for the maintenance of the present global ecosystem.
On the other hand, the development of living organisms on earth and their ability to utilize
their immediate environment has modified the primordial distributions and circulation of
these fundamental elements. The hypothesis has even been advanced (Lovelock, 1979) that
the biosphere has developed in such a manner that close to optimum conditions for life on
earth prevail and are being maintained by the life processes themselves. Regardless of the
extent to which this may be the case, the biogeochemical cycles are crucial for maintenance
of life on earth in its present form and biological processes largely determine the main
features of these cycles.

Man today influences the major biogeochemical cycles on earth significantly. The annual
release of carbon dioxide to the atmosphere by burning fossil fuels is about ten percent of the
amount being used in primary production of organic matter by plants (Bolin et al., 1979).
During the last few hundred years about ten percent of the land surface, primarily forests and
grasslands, has been transformed into agricultural land. The formation of fixed nitrogen by
combustion and fertilizer manufacturing is currently about half of what is produced naturally
(Söderlund and Svensson, 1976; Rosswall, Chapter 2, this volume). The land transformations
that man has initiated also have resulted in a major translocation of nutrients, particularly
nitrogen compounds, from the soils to the rivers and lakes and ultimately to the sea.
Emissions of sulphur dioxide to the atmosphere, primarily by fossil fuel combustion,
probably exceed the natural emissions of gaseous sulphur from land and oceans; in industrial
regions S emissions by far exceed those due to natural processes (Cullis and Hirschler, 1980;
Ivanov, 1981; Freney et al., Chapter 2, this volume; Crutzen, Chapter 3, this volume). A
significant eutrophication of lakes and possibly coastal waters has been caused by the
increased release of phosphorus in waste material into rivers and estuaries; erosion of
agricultural lands adds to the problem (Richey, Chapter 13, this volume; Wollast, Chapter 14,
this volume).

The main features of the natural cycles of carbon, nitrogen, phosphorus and sulphur (C, N, P,
S) have been described during the last decade but many aspects of their dynamics are not well
known. It is therefore difficult to determine what the effects of man's intervention may have
been and may become in the future. For example, it is easy to predict that the amount of
carbon dioxide in the atmosphere will increase if man continues to emit this gas in significant
amounts by fossil fuel burning. It is more difficult to foresee how the overall carbon cycle
will change and how the natural quasi-equilibrium of the global ecosystem will be disturbed.
Additionally, cycles of the basic elements interact in the biological processes and a
disturbance of one of them will, in all likelihood, affect the others (Melillo and Gosz, Chapter
6, this volume; Cook, Chapter 12, this volume).

We have a limited knowledge of the character, magnitude and possible on-going changes of
the global cycles of C, N, P, and S. World wide measurements with good accuracy are
needed. A further difficulty is the mosaic structure of the biosphere, particularly on land. The
local variations usually obscure those on larger scales, which vary only slowly with time.
Also the state of the global ecosystem may be changing due to natural causes. Thus climate
variations on time scales greater than a decade probably affect the biogeochemical cycles and
ecosystems. It is not enough to establish that possibly large-scale and long term changes are
taking place. We must also try to understand whether they are due to natural processes or
brought about by man.

Another complication is our lack of familiarity with the scale of problems confronting us. We
can design extensive field experiments that help us in understanding the short term dynamics
of a limited, homogeneous community. It is more difficult to grasp the cumulative changes
that occur from one year or one decade to another. On the other extreme we can measure
global-scale pool sizes of CO2 in the relatively well-mixed troposphere, but we find it more
difficult to deal with heterogeneities in the distribution of more rapidly turning over chemical
species in the atmosphere, soil, or oceans. The kind of information required to build from
local studies to a regional-scale understanding of element interactions is largely lacking, and
it is only this information that will allow us to develop more realistic global models. The
importance of such regional-scale studies is well illustrated by the gradual anthropogenic
acidification of precipitation in industrial regions (Odén, 1968; Cook, Chapter 12, this
volume).

We are steadily increasing the use of the natural resources of the world. It is obvious that this
increase must not go beyond the level at which the global ecosystem will be seriously
damaged and a decrease in the renewable resources of the earth will occur. It is disturbing
that we do not yet adequately understand the global ecosystem and its dynamics to assess
what this statement implies. How wide-spread will the impact of acid deposition be if the use
of fossil fuels expands as energy-use scenarios predict and a corresponding increase of
sulphur emissions occurs? How will the quality of soils used extensively in agriculture
change? Will there be a significant change of the ozone layer due to increasing emissions of
nitrogen oxides and chlorine containing gases?

It is likely that the ultimate impact of man's interventions will not be adequately deduced
until the global processes as manifested in the steady flow of C, N, P, and S in the air, the
waters, and the soil are well understood. In such an effort to determine the main pathways
and rates of exchange for these major elements we will also learn about the flows of the
minor constituents in nature that are also essential for the life processes. Further, the fate of
some toxic substances emitted by man that find their way into the environment and are
globally dispersed (e.g. PCB), can be studied with similar methods. A readiness for action
can thereby be created that will be of use if and when new threats to the environment are
discovered in the future.

With the aim of furthering our basic knowledge of the global ecosystems the SCOPE
Workshop on Interactions of Biogeochemical Cycles was arranged. On the following pages
of this first chapter we shall give an overview of the papers that were presented to the
workshop and that will be found in the remainder of this report. We shall also summarize the
discussions held during the workshop. The most important conclusions will be given at the
end of this chapter. We hope that in this way a more comprehensive account of the full
workshop will be available.

1.2 BIOLOGICAL PROCESSES AND INTERACTIONS OF BIOGEOCHEMICAL


CYCLES

As emphasized in the introduction, the formation of the organic compounds in life processes
requires the availability of the elements C, N, P, and S and a number of trace elements in
distinct proportions. These biochemical processes occur on a spatial scale from the size of the
simple organic molecules to that of the individual cell, but they have far-reaching
consequences at all ecosystem levels both in marine and terrestrial systems.

Redfield (1958) synthesized a number of ideas regarding the chemistry of planktonic


organisms and the chemistry of the marine environment into a scheme that suggested
extensive control by the biota over the chemistry of some elements in sea-water and oxygen
in the atmosphere. These general relationships form a basic stoichiometric model for the
control of C, N, O, and P chemistry and a major part of the chemical control of Si
(Broecker,1974) and Al (Mackenzie et al., 1978; Stoffyn, 1979) in sea-water. The
biologically mediated stoichiometry of C, N, and P in organisms and sea-water is often
referred to as the 'Redfield Ratio'. Schindler (1977) extended the Redfield model to fresh-
water systems, thus enlarging the model's generality to a broader range of systems. The use of
this approach is fundamental to the modelling of the circulation of these elements in the
rivers, estuaries and the open ocean as described by Richey (Chapter 13, this volume),
Wollast (Chapter 14, this volume) and Fiadeiro (Chapter 17, this volume).

Because the cycles of C, N, O, P, and to a large extent S, are so strongly mediated and driven
by the biota, there has been a tendency for those studying biogeochemical cycles to adopt the
Redfield scheme too indiscriminately. There are deviations from this scheme in other
systems. As our knowledge increases, we will be able to consider the specific biochemical
processes in more detail and more complex models will evolve. Still the principle problem of
how biochemistry and physical transfer interact is well illustrated by the following simplified
overview. Table 1.1 shows typical C, N, S, and P ratios in emissions from fossil fuel
combustion, the burning of plant material and in various natural reservoirs of organic matter.

Table 1.1 Carbon, nitrogen, phosphorus and sulphur in several fluxes and natural reservoirs

C N S P

Fossil Fuel Emissions1 9300 36 130 1


Combustion of Biomass1 920 20 1 1
Land Plants1 790 7 .6 3 .1 1
Land Animals1 78 11 1 .1 1
Terrestrial Bacteria2 43 4 .3 0 .2 1
Terrestrial Fungi2 188 11 .7 0 .8 1
Soil Organic Matter1 54 3 1 .2 1
Ocean Plants1 129 12 2 .9 1
Ocean Animals1 93 12 0 .7 1
Ocean Sediments
212 16 1
(surface)3
1
From Delwiche and Likens (1977) and Likens et al. (1981).
2
From Ausmus et al. (1976) and assuming a P:S ratio of 5.7 for
bacteria and 1.25 for fungi (Bowen, 1979).
3
From Jørgensen (Chapter 18, this volume).

The oceanic system has a discrete water-air interface, a very deep water column and a
relatively thin sediment layer. The major producers of organic material are single-celled
algae, especially diatoms and dinoflagellates. The biomass of these organisms is largely
decomposed in the water column, but some variable fraction sinks to the sediments, which
may or may not be aerobic depending on the flux of organic matter and the circulation of
water which determines oxygen availability (Jørgensen Chapter 18, this volume). Where
sediments are anoxic, nutrient elements such as nitrogen and phosphorus are preferentially
liberated to return to the water column, and denitrification and sulfate reduction are usually
important processes. Also, in some marginal waters (eg. Baltic Sea, Black Sea and various
fjords and estuaries), anoxia may extend into the water column overlying anoxic sediments.

Soils in themselves do not have a carbon source so the equivalent terrestrial system must be
expanded to include plants and the atmosphere. To some extent plant biomass is consumed
and elements regenerated above ground, an equivalent to decomposition in the surface waters
of the ocean. Most plant products fall to the ground (shoots) or upon death are mixed in the
soil (roots). In most terrestrial systems the major part of decomposition occurs on and in the
soil, the equivalent to decomposition in marine sediments. As in oceans, the degradation of
organic material in soils is mainly accomplished by micro-organisms. In this sense there is a
direct analogy between the two systems. In soils and in the ocean the micro-organisms
generally have the same biochemical composition and thus the same resource needs as far as
synthesis of microbial biomass is concerned, although the predominance of fungi in soils shift
this stoichiometry somewhat (Table 1.1). The final decomposition products are, however,
basically identical. In addition, microbes are responsible for N-fixation, whether in root
nodules or free-living, and N-fixing organisms, like all other microbes, have a very definite
need for phosphorus and sulphur. So there is a tie between phosphorus availability and
accumulated nitrogen in soils as well as ocean water.

Three major chemical differences lead to pronounced contrasts between soil systems and the
marine systems. The first is that terrestrial plants are markedly enriched in carbon relative to
other elements (Table 1.1) by the extensive production of structural compounds of carbon
(cellulose, lignin, waxes, etc.). These structural components are relatively resistant to
degradation by micro-organisms. Thus the organic material entering a soil system is much
different from sedimenting marine material in the type of compounds present and the relative
proportion of carbon to other elements.

The second major difference is the strong role of the inorganic soil matrix and of inorganic
phosphorus chemistry in aerobic soils. The matrix provides some protection against the
degradation of organic matter. Most phosphorus in soils is present in a form that is
unavailable to plants or microbes at any given time. As a result, the ratio of total phosphorus
to C and especially N in soils is quite different from the ratios of these elements in marine
plankton and microbial biomass (Table 1.1).

The third difference between the terrestrial and marine (or aquatic) systems arises from the
first two. Due to the abundant flux of organic carbon to soils, C cycles more rapidly than N
and P and the soil retains N and P relative to C. Also, due to the importance of non-available
pools of phosphorus in soil, a large part of the P cycle within the soil is regulated by non-
biological processes. To a large extent its cycle is uncoupled from those of C and N.

Redfield did not incorporate S in his total scheme in terms of fixed ratios because sulphate is
so enormously abundant in sea-water. It is, however, related to the C, N, P, and O cycles in at
least two ways. First, S is an essential element for all life. Second, sulphate is the major
oxidizing agent for organic matter in anaerobic marine sediments, thereby regenerating
oxygen in the form of carbon dioxide (Jørgensen, Chapter 18, this volume). On geological
time scales reduced S is an oxygen sink that helps set oxygen levels in the water and
atmosphere (Holland, 1978). The semi-independent cycle of S in the sea is in some ways a
counterpart of the semi-independent cycle of P on land.
Even though on first consideration it may seem that soils and sea sediments are quite
different in terms of their redox potential, the sulphur cycle in the soil compares with that in
the sea in that all the above mentioned processes are possible. The contrast lies in a relative
scarcity of S on land (except for reduced sulphur in the form of pyrite) because sulphate is
highly soluble. S is largely stored in organic form in soils so that in fact there may be a more
direct relationship between S and C and N on land than in the sea. This coupling is
manifested in those cases in which S is a limiting nutrient, and additions of S via fertilizer,
dry or wet deposition produce a positive growth response.

The biological aspects of the ocean-based Redfield model are definitely central to cycles of
these same elements on land. All four cycles intersect where biological processes occur; each
element can limit the cycling of another. Comparison, however, cannot be made at the level
of stoichiometric equivalency. Nevertheless, from agricultural management studies we will
gain a better understanding of the types of relationships that can occur between C, N, S, and
P. So far we have merely begun to understand these key processes quantitatively and to
include them into regional or global models of the biosphere. It will be difficult to unravel the
conditions that prevailed before the industrial revolution. It is even more difficult to deal with
the evolution of the present biosphere and how man may modify it in the future.

The following four subsections will treat the specific processes in the global ecosystem and
man's interventions. We shall in order deal with
the atmosphere
the terrestrial vegetation and soils
the fresh-water system, rivers and coastal waters
the oceans and ocean sediments.

We will be following the fates and interactions of the key compounds from the atmosphere,
where changes usually are first detectable, to the sea, where in most cases the regulation of
man's activities ultimately will occur on time scales of centuries to millenia or even longer.
The main features of the C, N, P, and S cycles are presented separately in Chapter 2, which
will serve as a reference.

1.3 THE ATMOSPHERE

1.3.1 AtmosphereBiosphere Interaction

The atmosphere plays a decisive role in regulating the temperatures at the earth's surface.
Without the presence of the atmosphere, average surface temperatures would be about 18°C,
given an albedo of the earth-atmosphere system of 30%. In reality, the mean surface
temperature is about 35°C higher due to the trapping of a large fraction of the out-going
infrared surface radiation in the atmosphere. Clouds and the radiatively active gases H2O,
CO2 and O3 are of special importance in this respect. A doubling in the atmospheric carbon
dioxide content according to the best present knowledge is expected to result in a mean
surface temperature increase of 1.54°C (Manabe and Stouffer, 1980; Kellogg and Schware,
1981; Thompson and Schneider, 1981). Recently, attention has also been drawn to the
climatic importance of the minor constituents N2O, CH4 and the industrial gas CF2Cl2 at
volume mixing ratios as low as a few ppbv (Ramanathan, 1975; Wang et al., 1976).

The atmosphere is also the transport medium for many biologically or photochemically active
gases such as O2, N2, H2O, CO2, CO, CH4, O3, NO, NO2, HNO3, NH3, N2O, SO2, H2S,
CH3SCH3, and COS as well as airborne particulate matter containing C, N, S, P and other
nutrient elements. Several of the listed compounds are emitted or taken up by the biosphere
and the reduced compounds are oxidized in the atmosphere by photochemical reactions. A
key role in their oxidation is played by the OH radical, which is formed when ozone is
photolysed by ultraviolet solar radiation in the presence of water vapour (see Crutzen,
Chapter 3, this volume). Several compounds interact in the atmosphere to form particulate
matter. The nutrient elements C, N, and S are returned to terrestrial ecosystems and the
oceans by precipitation scavenging and by direct gas phase and particulate matter transfer at
the earth's surface (Reiners, Chapter 5, this volume; Liss, Chapter 15, this volume; Duce,
Chapter 16, this volume). The global ecosystem must have developed over geological time
scales in response to climate and to the supply of nutrient elements from the atmosphere and
the lithosphere. Agricultural and industrial activities have changed the balances and quantities
of inputs and outputs of nutrient elements in many ecosystems, which cause the ecosystems
to change according to their degree of resilience (see further Reiners, Chapter 5, this volume;
Houghton and Woodwell, Chapter 11, this volume; McGill and Christie, Chapter 9, this
volume).

1.3.2 The pre-industrial and the present atmosphere

During the industrial era the atmospheric transfer rates of many C, N and S containing gases
have changed greatly. Consequently, the atmospheric composition also has changed. What
was the atmospheric composition prior to man's impact on it? There has been remarkably
little thought given to this question.

By direct measurements we know that the amount of CO2 in the atmosphere has increased by
10% during the last 25 years. The abundances of other important trace gases also have been
modified both in response to climate changes and owing to man's emissions. No data on their
historical abundances are, however, available. Much of what will be given in the following is
therefore derived from extrapolations backwards using our limited knowledge of the
processes that govern the composition of the present atmosphere. In the future, analysis of air
bubbles in polar ice may provide important information on the history of atmospheric
composition. Certainly, changes have occurred during the past century in the atmospheric
concentration of the oxides of nitrogen (NO, NO2, N2O, HNO3), carbon monoxide, sulphur
dioxide, methane and other hydrocarbons. Also the organic chlorine content of the
atmosphere has changed dramatically during the past decades.

A. Carbon Dioxide

Extrapolating backwards from the known trends in the atmospheric CO2 concentrations
during the past decades and accepting the analysis that the terrestrial ecosystems have
contributed 125 ± 50 x 1015g C during the past 150 years, a pre-industrial CO2 volume
mixing ratio of 265 ± 30 ppmv (parts per million by volume) is derived, compared to 338
ppmv in 1981 (Brewer, 1978; Bolin, 1983). This lower value may roughly have represented a
natural, atmospheric equilibrium value maintained by the more pristine lands and oceans
before the agricultural and industrial eras. However, even this value of the atmospheric CO2
concentrations has probably not been constant. Analyses of the air trapped in Antarctic and
Greenland glaciers show that the atmospheric CO2 concentration seems to have varied
considerably in association with climatic conditions. For instance, during the last ice age,
between 15,000 and 30,000 years ago, atmospheric CO2 concentration seems to have been as
low as 200 ppmv (Berner et al., 1980; Delmas et al., 1980). Changes in atmospheric CO2
have possibly contributed to the climatic changes that have occurred during the last thirty
millenia. From the available data, however, it is more likely that atmospheric CO2 responded
to climate. The observed correlation between CO2 and mean surface temperatures cannot be
used to derive the likely climatic response to the on-going increase in the atmospheric CO2
concentrations (Thompson and Schneider, 1981).

B. Ozone and Nitrogen Oxides

The composition of the atmosphere is strongly influenced by the photochemistry of ozone. Its
photolysis by ultraviolet solar radiation leads to the production of the hydroxyl (OH) radical,
which reacts with many atmospheric compounds that would otherwise be much more inert
and abundant. Tropospheric reactions dominate the budgets of most gases. Tropospheric
ozone therefore influences the concentrations of many atmospheric trace compounds. The
oxides of nitrogen (NO and NO2) play an important role in these processes. Above a critical
NO concentration (>0.01 ppbv), ozone is created in the troposphere by the photochemical
oxidation of hydrocarbons to carbon monoxide and further to carbon dioxide. Since the
anthropogenic sources of oxides of nitrogen are estimated to be larger than the natural ones, it
may be inferred that the abundance of tropospheric ozone might have increased during the
past few centuries. Unfortunately, it is not possible to directly prove or disprove this
supposition due to the lack of sufficient ozone records. Despite larger surface destruction
rates in the northern hemisphere, the higher prevailing concentration of ozone there,
compared to that of the southern hemisphere could be explained by higher anthropogenic
inputs of NO in the northern hemisphere (Fishman et al., 1979). However, a larger downward
transfer of ozone from the stratosphere to the troposphere in the northern hemisphere may be
a contributing factor (Viezee and Singh, 1980; Husain et al., 1981). We note further that the
influence of NO on the ozone distribution leads to ozone depletion in most regions of the
stratosphere, but to an ozone increase in the troposphere (see Crutzen, Chapter 3, this
volume).

There is some observational evidence that the atmospheric loading of N2O has increased by
about 0.2% annually during the past 15 years (Weiss, 1981). This increase may be due to a
combination of the combustion of oil, coal and biomass and man's impact on the nitrogen
cycle, even though the role of artificial fertilizers seems less than originally estimated.
Assuming that the trend is mostly due to the production of N2O in fossil fuel combustion,
Weiss (1981) derived a pre-industrial atmospheric N2O volume mixing ratio of 270 ppbv,
which is 10% less than present levels. As the oxidation of N2O is the main source of NOx
(NO + NO2) in the stratosphere and as both NO and NO2 act as catalysts in the destruction of
stratospheric ozone, a 10% increase of N2O during the past century would have led to a 1.5%
reduction in total ozone, according to our present understanding of the photochemistry of the
atmosphere. No observational records are available to confirm this hypothetical depletion,
which furthermore may have been masked by the increase in tropospheric ozone due to the
industrial NO input described above.

C. Methane and Carbon Monoxide

Atmospheric methane and carbon monoxide play important roles in the photochemistry of the
atmosphere due to their influence in establishing ozone and hydroxyl concentrations (see
Crutzen, Chapter 3, this volume). The photochemical oxidation of methane also leads to the
formation of carbon monoxide. Measurements during the past years show an annual increase
in atmospheric concentrations of methane of about 2% (Rasmussen and Khalil, 1981; W.
Seiler, personal communication). There are many atmospheric sources in the environment,
such as release from wetlands, rice paddy fields, ruminants, termites, biomass burning, and
natural gas leaks. Several of these have been increasing with time due to human activities.
Their relative contributions to the global methane cycle are, however, rather uncertain.
Nevertheless, it is likely that pre-industrial methane concentrations were appreciably lower
than present ones.

The past temporal changes of carbon monoxide are more difficult to estimate. Its increased
atmospheric formation from methane oxidation and from growing industrial and biomass
burning activities has probably been countered by reduced atmospheric formation from
photochemical oxidation of hydrocarbon gases that are emitted by forest tree foliage. Expect
for the industrial carbon monoxide sources, the magnitudes of other sources of CO are not
well enough known to allow even a guess about whether CO has increased or decreased with
time.

D. Sulphur Dioxide and Carbonyl Sulphide

The atmospheric input of gaseous SO2 has increased much over the past centuries, especially
close to the main industrial centres of the world. According to present estimates (Granat et
al., 1976; Ivanov 1981; Freney and Rodhe, Chapter 2, this volume; Crutzen, Chapter 3, this
volume) the input of industrially produced SO2 to the atmosphere may well exceed the
natural formation rates of SO2 from volcanoes and from the oxidation of biogenic reduced
sulphur species, such as H2S, CH3SCH3, CH3SH, CH3SSCH and COS. SO2 is not produced
directly by the biosphere. For the earth as a whole the atmospheric circulation of sulphur
oxides and sulphuric acid has at least doubled, while in the industrialized parts of the world
the emission of SO2 and the deposition of SO2 and sulphuric acid has increased manyfold.
This has caused disturbances in the quantities and balance of nutrient supplies from the
atmosphere and the acidity of soils and lakes (Cook, Chapter 12, this volume). Detrimental
effects have occurred in the lakes of Scandinavia, Canada and the north-east of the U.S.A.,
where acidification of the waters has increased to levels that threaten lake life. These lakes
are examples of ecosystems with a low resistance against environmental changes.

Besides sulphuric acid, nitric acid has also contributed to the increased acidification of
precipitation. The increased input of ammonia, which has taken place simultaneously as a
result of growth in agriculture and increased use of nitrogen fertilizer, has only partially
balanced the increased acidification of clouds and rain resulting from sulphuric and nitric
acid. According to FAO statistics (FAO, 1979) present global use of nitrogen fertilizers
amounts to 5060 Tg N/year, of which possibly 10% is volatilized as NH3, while the industrial
production rates of SO2 and NOx amount to about 100 Tg S (Granat et al., 1976; Cullis and
Hirschler, 1980; Ivanov 1981), and 20 Tg N per year (Söderlund and Svensson, 1976). Due to
the growing acidification of aerosols and precipitation by HNO3 and H2SO4 that has occurred
during the industrial era, it is likely that the solubility of NH3 in water phases has increased
(Taylor et al., Chapter 4, this volume) and, therefore, the residence time decreased.

An important consequence of the increased atmospheric emissions of especially SO2 and the
resultant increase in the concentration of sulphate, its oxidation product, has been the
observed change in visibility in the industrial parts of the world. During stable meteorological
conditions in these regions the visual range is frequently reduced to less than 5 km for
hundreds of kilometers downwind of pollution centers, covering areas of several hundred
thousand square kilometers (Husar and Patterson, 1980). Together with the photochemical
smog produced from urban automotive emissions they constitute a drastic deterioration in
man's ability to enjoy his natural environment. In addition, the climatic implications of the
increased loading of the atmosphere by sulphate particles may be of significance (Bolin and
Charlson, 1976).

Carbonyl sulphide (COS), with a concentration of about 0.6 ppbv (Torres et al., 1980) and a
relatively long residence time, is the most abundant sulphur containing gas in the atmosphere.
COS is important because during periods with little volcanic activity its photo-oxidation in
the stratosphere is the main source of the sulphate that is contained in the stratospheric
sulphate aerosol layer at about 20 km altitude (Crutzen, 1976; Whitten et al., 1980; Meixner,
1981). In this way, the gas indirectly plays a role in the earth's radiation budget (Turco et al.,
1980). The sources and sinks of COS have not been established quantitatively, but as this gas
is also produced in the burning of coal and biomass, its atmospheric concentrations may be
increasing although much more observational work is needed to verify such a trend
(Hoffmann and Rosen, 1981). In contrast to the climatic effects of most gaseous emissions
(such as CO2, N2O, CH4 and CF2Cl2) to the atmosphere, increased atmospheric emissions of
COS tend to cool the earth's surface, because the enhanced sulphate aerosol layer in the lower
stratosphere increases scattering of incoming solar radiation back to space.

E. Organic Chlorine Compounds

From available information (Singh et al., 1979; Berg et al., 1980) we conclude that until a
few decades ago methyl chloride (CH3Cl) was the only significant chlorine containing
organic compound in the air, with a mixing ratio of about 0.6 ppbv (Lovelock, 1974). At
present the organic chlorine content of the atmosphere is almost five times higher due to the
presence of many additional industrial organic chlorine gases, especially carbon tetrachloride
(CCl4), the fluorocarbons F-11 (CFCl3) and F-12 (CF2Cl2) and methyl chloroform (CH3CCl3).
All these compounds are important because they are the only significant carriers of chlorine
into the stratosphere (Berg et al., 1980). Their photochemical breakdown in the stratosphere
leads to the production of photochemically active Cl and ClO radicals that catalytically
remove some ozone in the stratosphere (Molina and Rowland, 1974). While the chlorine
species did not play a dominant role in the ozone balance of the natural atmosphere (NO and
NO2 are much more important), their influence is now becoming increasingly significant,
affecting in particular the ozone concentrations between 30 and 45 km (Crutzen, Chapter 3,
this volume). Some reduction in the ozone concentrations at these levels has probably already
occurred. Significantly larger changes may be expected in the next half-century if the
production of industrial, organic chlorine compounds continues at about the present rates.
The total ozone content of the atmosphere may then decline by 310%, allowing more
ultraviolet radiation to affect the biosphere (WMO-NASA, 1982).

1.3.3 Looking towards the future

Many of the changes that occurred in the environment over the past 150 years will continue,
and additional ecosystems will be affected in the future, especially in the tropics. There is
little doubt that the concentrations of important constituents in the atmosphere will change
significantly during the next hundred years. The following summary emphasizes problem
areas that should be given particular attention in this regard. Man is expanding his activities
on earth in many ways and it is becoming increasingly clear that we must consider these
simultaneously to foresee the likely future changes in atmospheric composition. In particular
the expansions of agriculture and forestry will have to be considered carefully (cf. section
1.4.3). The need for a proper understanding of the interactions of biogeochemical cycles is
obvious from the preceding presentation.

A. The Tropics

The total area covered by tropical rain forest and seasonal forest is at present about 11 x 1012
and 5 x 1012 m2 respectively (Bolin et al., 1979). A significant fraction of these remaining
tropical forests may well disappear during the next 50 years (section 1.4.3). Table 1.2 shows
the present rate of deforestation in tropical America, Africa and Asia and the annual
emissions of carbon, nitrogen, sulphur and phosphorus (based on Lanly and Clement, 1979).
Burning of biomass in the process of deforestation and also in shifting cultivation and dry
grass fires in the savannas produces many important trace gases such as CO, CH4, N2O and
NOx (Crutzen et al., 1979). Emissions of particulate matter cause a large aerosol loading in
the atmospheric boundary layer above the tropical savanna. A substantial fraction consists of
organic matter and elemental carbon (soot) (Seiler and Crutzen, 1980) containing an
important submicron component. The extent to which it may influence the heat balance of the
atmospheric boundary layer is not well-known (Watson, 1979).

Deforestation in the tropics leads to smaller emissions of hydrocarbons to the atmosphere,


especially isoprene and terpenes. The natural production of carbon monoxide is also reduced.
On the other hand, the increased CO2 concentrations in the atmosphere might enhance the
productivity of the remaining forests (see section 1.4.4.A) and, furthermore, fast-growing
trees, such as eucalyptus or casuarina, which are prodigious terpene and isoprene producers,
will be planted for commercial use. Industrial activities emitting nitric oxide may have a
detrimental impact on local and regional air quality in the tropics in that photochemical smog
may be formed during stable meteorological conditions.

When more biomas is cleared in the tropics, more material will become available for
decomposition and for entrainment in run-off. This will probably result in an increase of
organic carbon in aquatic systems (see section 1.5.3). As eutrophication is increased in these
systems anaerobic conditions will become more prevalent and more methane will be
produced. The release of methane to the atmosphere will probably further increase due to
expansions in the sources of methane, such as livestock and rice paddy fields. The use of
nitrogen fertilizers has been shown to lead to enhanced CH4 production in rice paddy fields
(Cicerone and Shetter, 1981).

B. Pollution in the Arctic

There is growing observational evidence that industrial pollution is affecting the composition
of the air in the Arctic. During winter and spring, pollutants may travel thousands of
kilometres from the industrial areas of the Northern Hemisphere to form haze containing
sulphate, trace metals and soot of which about 80% is of anthropogenic origin (Rahn and
Heidam, 1981; Rahn 1981; Rosen et al., 1981). The build-up can be explained by the
inefficient scavenging of the pollutants due to infrequent precipitation in winter. In contrast
the Arctic is much less polluted during the summer. Black soot in the aerosol is a most
efficient absorber of solar radiation. A significant heating of the lower Arctic atmosphere in
the spring may occur (Rosen and Novakov, 1980).
Table 1.2 Present annual rate of deforestation in the tropics and the annual release of C, N, S,
P from forest clearing based on fixed proportionalities between these elements in vegetation
and soils (Lanly and Clement, 1979)

America Africa Asia Total

Rain Seasonal Rain Seasonal Rain Seasonal Rain Seasonal


forest forest forest forest forest forest forest forest

Area
cleared 2.7 0.2 0.6 0 1.7 0.1 5.0 0.3
10 2
(10 m )

C Veg 510 18 110 330 15 950 33


(Tg) Soil 110 7 20 70 6 200 13
Total 620 25 130 400 21 1150 46

N Veg 3.4 0.12 0.7 2.2 0.10 6.3 0.22


(Tg) Soil 8.5 0.56 1.8 5.4 0.46 15.7 1.02
Total 11.9 0.68 2.5 7.6 0.56 22.0 1.24

S Veg 0.34 0.01 0.07 0.22 0.01 0.63 0.02


(Tg) Soil 0.90 0.06 0.19 0.57 0.05 1.66 0.11
Total 1.24 0.07 0.26 0.79 0.06 2.29 0.13

P Veg 0.34 0.01 0.07 0.22 0.01 0.63 0.02


(Tg) Soil 0.90 0.06 0.19 0.57 0.05 1.66 0.11
Total 1.24 0.07 0.26 0.79 0.06 2.29 0.13

C. Expanding Agriculture

The demand for more food in the world will lead to a further increase in the demands on land
used for agriculture and the use of more fertilizers. According to some recent projections, the
emphasis will be on intensifying land use, but new land will also be brought under cultivation
(FAO, 1919). To bring out possible future implications of expanding and intensified
agriculture, let us consider the use of 25 x 1012 m2, which is close to the maximum possible,
rather than 15 x 1012 used today and application of nitrogen fertilizer at an annual rate of 10 g
N/m2 (100 kg N/ha) to such an increased area. The global use of nitrogen fertilizer would
reach 250 Tg N per year, which is about a five-fold increase. With the trend towards the use
of urea as the main form of fertilizer, the volatilization of ammonia to the atmosphere may be
increased by 25 Tg N annually, if typically 10% of the applied nitrogen is lost in this way.
The release of ammonia to the atmosphere may be further increased by 40 Tg N per year, if
the cattle population continues to grow by about 2% annually during the next three to four
decades, and by still by another 25 Tg N per year if the use of coal takes place according to
the projections of Rotty and Marland (1981) during this time. The total release of ammonia
from the land to the atmosphere may, under these conditions, increase by almost 100 Tg N.

 At this time we can only speculate about likely consequences of such a scenario.
Some possibilities are enumerated here.
 An enhanced formation of ammonium sulphate aerosol would occur with influence on
the radiative properties of the atmosphere (cf Bolin and Charlson, 1976).
 Ammonia emissions might partly neutralize the increased acidity of precipitation that
will result from the growth in sulphuric and nitric acid concentrations from
anthropogenic activities (see further below).
 The increased use of nitrogen fertilizer would probably lead to increased production
of N2O. The atmosphere contains about 1500 Tg N of N2O. In comparison, the direct
production of N2O in agricultural fields is probably rather small, of the order of a few
Tg N per year (Crutzen, Chapter 3, this volume). The subsequent fate of nitrogen that
is lost from agricultural fields by leaching, ammonia volatilization and crop removal
may supply additional, but unknown, quantities of N2O to the atmosphere. In
summary we may expect a probably slow, but steady increase in the atmospheric
nitrous oxide abundance. This might affect the distribution of ozone in the
stratosphere, as well as ground level temperatures, as N2O plays a role in the radiation
balance of the troposphere (Wang et al., 1976). Much further study is, however,
required to establish likely future changes.

Table 1.3 Expected input of S into atmosphere in the year 2025 based on fossil fuel
combustion as projected by Rotty and Marland (1981). A C:S ratio of 50:1 by weight is
assumed. Units are Tg yr-1

CO2-C Released SO2-S Released


Region
19751 2025 19751 2025

North America 1500 2300 30 45


Western Europe 1000 1100 21 22
U.S.S.R./Eastern
1100 3400 22 68
Europe
Other Developed
400 1000 8 20
Countries
China 500 2200 10 44
Latin America 200 1100 4 21
Middle East 50 800 1 16
Africa 50 400 1 9
South Asia 100 1300 2 27

Totals 4700 13600 100 270


1
The values in 1975 are calculated by assuming that S is emitted from each region in
proportion to energy use and from the total S emission data of Cullis and Hirschler
(1980).

D. Fossil Fuel Combustion.

A typical estimate of the future increase of fossil fuel combustion (Rotty and Marland, 1981),
implies that the atmospheric CO2 concentrations will be about 460 ppmv in the year 2025 and
680 ppmv soon after the middle of the next century (WMO, 1981). As mentioned before
(section 1.3.2.B), an increase of atmospheric N2O may also result. Significant climatic
changes may occur (Manabe and Stauffer, 1980) and have an impact on the global ecosystem.
This problem receives much attention at present and will not be discussed further here.

The emissions of SO2 to the atmosphere will undoubtedly increase due to the expected rise in
fossil fuel combustion, most of which will be coal combustion. The projection by Rotty and
Marland (1981) implies nearly a three-fold increase in the global anthropogenic sulphur
dioxide emissions to the atmosphere by the year 2025 (see table 1.3). As the SO2
concentrations in air increase, cloud water becomes more acidified causing the solubility of
SO2 in water and its water phase oxidation to decrease (Penkett et al., 1979; Taylor et al.,
Chapter 4, this volume). This would perhaps lead to an increasing possibility for SO2 to
penetrate through clouds to higher levels and may be important for long range transport of
sulphur compounds (Rodhe, 1981). The increased emissions might therefore affect such
ecosystems that so far have not been markedly influenced by atmospheric SO2 inputs, such as
Siberia and various regions in the tropics. If industrial emissions are not regulated more
efficiently than is now the case, environmental deterioration might occur in a manner similar
to what so far has taken place in and around the most heavily industralized parts of the world,
i.e. parts of North America and Europe. Continued industrialization in the tropics may
particularly affect savanna regions (e.g. the cerrados of Brazil). Little is known about the
natural atmospheric S deposition in the savannas, but it may amount to less than 1 g m-2 yr-1
(i.e. totally only a few Tg S as an average for all tropical savannas) due to the low S/C ratio
of vegetation, as compared to a possible 40 Tg S per year due to fossil fuel combustion (cf.
table 1.3). How these ecosystems would respond to such an increase in sulphur deposition
rates is unknown.

Through a variety of man's activities in the developed and the developing world, large
amounts of nutrients (N, P, S) will be transported through the atmosphere and waters and
accumulate in various ecosystems in amounts and ratios that differ markedly from pristine
conditions. Ecosystems respond to such perturbations in various ways: an imbalance of
minerals in the plants; higher nutrient content of senescent material returning as litter; and
higher rates of decomposition; and change in soil acidity, etc. Knowledge of the C, N, P and
S cycles of vegetation and soils must serve as a basis in assessing the effects of changes in
nutrient deposition patterns and from this knowledge we will possibly learn how to avoid
major deteriorations of the environment.

1.4 TERRESTRIAL ECOSYSTEMS

As already emphasized (section 1.2), living organisms and their metabolic products have an
important influence on the biogeochemistry of all but the most extreme terrestrial
ecosystems. For example, plants can decrease wind and water erosion of soil by a factor of 10
or more by adding litter to the soil and by binding the soil with roots. Plants and microbes
increase chemical weathering and the transport of soluble compounds to deeper soil layers
and to aquatic systems by raising the partial pressure of CO2 in the soil, by converting part of
the N and S they obtain from the atmosphere to mobile ions (i.e. nitrate and sulphate) and by
releasing hydrogen ions in exchange for cations in the soil solution. Micro-organisms, plants,
and animals produce many of the important trace constituents (or their precursors) in the
atmosphere (Crutzen, Chapter 3, this volume).

1.4.1 Interactions of C, N, S and P in Natural Terrestrial Ecosystems

Several interactions among the C, N, P, and S cycles under natural conditions can be
illustrated by examining the changes in element supply, availability, and mobility that occur
during the establishment and development of soil and vegetation on a previously unoccupied
site. If a new land surface is deposited following a volcanic eruption or exposed by the
melting of a glacier, the new substrate is at first nearly devoid of organic carbon and nitrogen.
The development of biota and soils on such sites requires: (i) C-fixation from CO2 by
photosynthesis; (ii) N-fixation by symbiotic or non-symbiotic micro-organisms, or slow
accumulation by the retention of small amounts of fixed N deposited from the atmosphere;
(iii) P which generally is derived from the weathering of the substrate, and therefore is
generally present in newly exposed material (Walker and Syers, 1976); and (iv) S weathered
from the substrate or accumulated from the atmosphere. Symbiotic nitrogen-fixing plants are
at a significant competitive advantage during the early stages of soil development because
they can obtain N and thus C much more rapidly than other plants if adequate amounts of P
and S are available. N-fixers add organic matter with a relatively low C to N ratio to the site;
the organic matter accumulates and increases the water- and cation-holding capacity of the
soil, and the level of available N in the site rises. The rates of N-fixation are eventually
reduced as the availability of N increases in the soil or when P or S become limiting to the N-
fixers. Once the availability of N in the soil is high, other plants without N-fixing capabilities
can grow. Competition for light, water, and other nutrients becomes more important, and
those plants that invest large fractions of their energy on structural tissue in stems or roots
generally have a competitive advantage. Consequently, N-fixation peaks relatively early in
soil development (Walker and Syers, 1976) and thereafter the relative availability of N is
more closely coupled to that of P and S.

Within terrestrial ecosystems, organisms have evolved the ability to operate over a fairly
wide range of C, N, P, and S contents (Melillo and Gosz, Chapter 6, this volume; Vitousek,
Chapter 7, this volume; Stewart et al., Chapter 8, this volume; McGill and Christie, Chapter
9, this volume; Hunt et al., Chapter 10 this volume). Much of the variation between
organisms can be explained by differential storage or retranslocation of C, N, S, P within the
organisms or by differential availability of these elements in the soil; this variation can in turn
cause differences in decomposition rates and nutrient availability. Consequently, the addition
of one element to a system may bring about a change in the form of storage rather than the
input or output of other elements.

Early in soil development, any change beyond the range of the metabolic flexibility of the
organisms can still be adjusted by additional N-fixation, or by continued mineralization of
soil P and some forms of S. With additional time and further soil development, however, the
availability of P gradually declines

as P is lost or transformed into unavailable forms. This very slow loss is irreversible under
natural conditions because there is relatively little replenishment of P from the atmosphere.
Eventually, this means that not enough C (and N) can be fixed to replenish losses of those
elements and the biomass and nutrient content of the system declines (Walker and Syers,
1976; Cole and Heil 1981; Melillo and Gosz, Chapter 6, this volume; McGill and Christie,
Chapter 9, this volume).

This directional change in P availability in soil development, and the attendant changes in C,
N, and S fluxes and storage, can be reversed by any process that brings unweathered material
to the surface. Since the cycle of soil development is generally slow relative to the cycle of
growth and wastage of continental glaciers, and slow relative to the frequency of volcanic
activity in tectonically active regions, the final state of P deficiency can only be observed in
geologically quiescent tropical and subtropical areas.

1.4.2 Natural Disturbances

The natural biogeochemical cycles of C, N, P, and S do not interact only through this slow
progression of soil development modified by changes in climate and landforms. Natural
disturbances are important features of most natural landscapes (White, 1979) and can
influence the natural biogeochemical cycling of C, N, P, and S. They are sudden in
comparison with the slow changes of soil development discussed in section 1.4.1. The most
important disturbances in grasslands and savannas are fire, grazing and severe droughts; in
forests they are fire, windstorms, and insect outbreaks. These natural disturbances can occur
with a frequency of one or more per year to one per several hundred years, and with a size
from less than one to many millions m2. They are integral components in determining the
long term quasi-steady state features of most grasslands and many forests. Even without such
catastrophic exogenous disturbances, tree falls occur frequently and can open large gaps in
the forest canopy.

Different disturbances affect C, N, P, and S cycling differently. Fire volatilizes organic C, N,


and S, producing a suite of major and trace atmospheric constituents including relatively
large amounts of CO, NO, and aerosols, but fire converts organic P to a more biologically
available form. Grazing by large herbivores concentrates nutrients into small patches of
faeces and urine, which can then be sites for volatilization losses of N as NH3, N2O, and some
amines. All of these disturbances can increase soil temperature and water content, thus
increasing the mineralization of N, P, and S at least for short periods of time.

The retention of these mineralized nutrients (particularly N) against losses by leaching or


volatilization after a disturbance is dependent in large part on the maintenance of the cycling
of nutrients between soils, micro-organisms and plants within terrestrial ecosystems
(Rosswall, 1976). When the cycles are interrupted by disturbance, losses of N are generally
somewhat increased. Organisms that decompose structural material with a high C:N ratio are
themselves likely to be N-limited as a consequence however (Melillo and Gosz, Chapter 6,
this volume; Vitousek, Chapter 7, this volume; Hunt et al., Chapter 10, this volume), and they
will take up available N from the soil solution, incorporate it in their metabolic machinery
and retain it against loss. This structural material thus plays a stabilizing role in retaining
nitrogen within disturbed systems.

The soilplantmicroorganism cycle is eventually re-established by rapidly growing pioneer


plants that invade disturbed sites and take advantage of the available light, water, N, and P to
grow. The site then returns towards its predisturbance state, drawing C and some N and S
from the atmosphere, and P and more N and S from the soil. This re-establishment is much
more rapid (tens to hundreds of years) than that observed during soil development, due in part
to the organic matter and nutrient stocks already present in the soil. Most natural terrestrial
ecosystems, then, are a patchwork of different aged vegetation resulting from natural
disturbance (Bormann and Likens, 1979). A few of the patches are likely to be losing C, N,
and S at any given time; many more are likely to be in various stages of C, N, and S
accumulation. Averaged over an ecosystem, the net result is no large loss or gain of C, N, P,
or S.

Interactions of the biogeochemical cycles of C, N, P, and S thus occur on several time scales
in terrestrial ecosystems: daily (or shorter) intervals involved in plant and microbial growth;
annual cycles of community-level growth and decay; several month to several hundred year
intervals of biological C, N, P, and S accumulation between natural disturbances; and several
thousand to several hundred thousand year intervals of soil development and degradation.

1.4.3 Human Influences

Man's influence on the C, N, P, and S cycles of terrestrial ecosystems is felt in most parts of
the world. The natural forests, semi-deserts and grasslands of the temperate regions have long
been disruptively exploited by humans for food and fibre. More recently, population
pressures, industrialization, and the search for ways to expand the world's natural resource
base have caused the tropical regions to become the foci of extreme and often destructive
exploitation (Melillo et al., Chapter 6, this volume; Vitousek, Chapter 7, this volume).

Man's impacts on natural terrestrial ecosystems range from alterations of the timing and
intensity of processes, such as fire and grazing, to the application of agricultural and
silvicultural treatments that create fundamentally different ecosystems, such as permanent
agriculture, shifting cultivation, timber plantations, or drainage and reclamation of wetlands.

The increased intensity of land use takes many forms, but historically almost all have the long
term consequence of decreasing the reserves of C, N, P, and S in soils. Even though
interactions between the cycles primarily are governed by biological processes, C, N, S, and
P appear capable of cycling somewhat independently in ecosystems (section 1.2; Stewart et
al., Chapter 8, this volume; McGill and Christie, Chapter 9, this volume).

The response of ecosystems to man's interventions occurs on time scales from days to
centuries. Simultaneously, natural variations are present and we are confronted with a major
difficulty in trying to separate these and to determine cause and effect (cf. sections 1.1 and
1.2). Examples of this basic difficulty will be given in the following discussion of the
pathways and magnitudes of C, N, P, and S losses from managed sites.

A. Fire

Fire is nearly universally employed in the conversion of forest land to other uses. Such fires
generally have the same effects as natural fires except that natural fires, like fires used in
shifting agriculture, are followed by relatively rapid regeneration. The increase in the amount
of organic material consumed as a consequence of human activities is substantial. According
to one estimate among several, an average of about 5 x 1010 m2 of tropical forest per year is
expected to be converted to other uses in the next 20 years (cf. Table 1.2 and Lanly and
Clement, 1979), equivalent to an annual amount of 0.3 % of the total area of tropical forests.
Tropical forest conversion includes a variety of alterations ranging from shortening the fallow
period in shifting agriculture to complete destruction of forest cover and replacement by
cultivated fields.

The inputs of gaseous C and N to the atmosphere from the burning of forests and savanna
have been discussed in section 1.4.4A. Frequent savanna and forest fires have occurred for
hundreds or thousands of years and it is not known if the amount of material consumed has
increased in recent years or if this burning causes a change in the net flux of C, N, and S from
terrestrial ecosystems.

B. Volatilization and Leaching

Upland cultivated soils in the temperate zone can lose up to 50% of their C and N, 3040% of
their S, and 1030% of their P in the 5070 years following land conversion (Stewart et al.,
Chapter 8, this volume); similar changes occur more rapidly in the tropics. A portion of the
C, N, and S and most of the P is lost to harvest and erosion (see below), but a larger fraction
of the C, N, and probably S is lost to the atmosphere in gaseous form or to aquatic
ecosystems in ionic form. Appropriate agricultural management can reduce these losses, but
it is difficult to eliminate them entirely.

These losses of C, N, and S occur because the decomposition rate is increased in cultivated
soils due to physical alterations and because less organic material is returned to the soil than
in natural systems. When organic C flux to the soil is reduced, microbes have a reduced
capacity to immobilize the increased levels of N and S released during decomposition
(McGill and Christie, Chapter 9, this volume). The N and S released exceeds the amount that
plants can take up (at least for part of the year). Some of the excess C, N, and S in cultivated
soils is leached to stream-water or ground-water as bicarbonate, dissolved organic C, nitrate,
dissolved organic N, and sulphate. The remainder is lost to the atmosphere as CO2, other C
gases, N2, NH3, N2O, and various sulphur gases (see section 1.3.2). The relative importance
of these pathways is largely unknown and probably variable among sites, except that most of
the C lost goes to the atmosphere as CO2.

After one to several years of cultivation, these losses reduce the ability of the soil to supply
adequate amounts of N, P, and S to crops, and fertilization is necessary to maintain high
levels of production. Much of the applied fertilizer is present in forms that can be easily
leached or volatilized. Present management procedures have transferred the almost closed
nutrient circulation of the pristine ecosystem (forest or grassland) into the open system of
modern farm land.

C. Erosion

The decline in soil organic matter content, the absence of a surface litter layer, and a decrease
in perennial root biomass in cultivated systems combine to reduce the cohesiveness of the
soil. Consequently, erosion by both wind and water is increased many-fold by human
activities. Pimentel et al., (1976) estimated that erosion has severely damaged much land in
the U.S. and high rates of erosion have been reported in other regions (Rapp, 1975). The
material that reaches rivers and lakes increases sedimentation, decreases the lifetime of lakes
and reservoirs and can decrease aquatic productivity by limiting light penetration.

In drier areas, increased grazing by domestic animals caused by increasing population


pressure has led to a decrease in plant cover and significant increases in both wind and water
erosion. This reduces infiltration rates of water, soil water holding capacity, and organic
carbon content and thus plays an important role in the deterioration of arid shrublands and
dry savannas. Even though the rate of anthropogenic land degradation is currently 0.05 x 1012
m2 yr-1, 36 x 1012 m2 of land, which supports one-sixth of the world's population, may
ultimately be at risk (United Nations, 1977).

D. Harvest

Organic C, N, P, and S are removed from croplands in harvested products and concentrated
into urban areas or into large-scale animal feedlots in developed countries. Organic material
from large land areas is transported into small urban areas and processed or consumed. A
large fraction of it is then discharged into waterways whether or not sewage systems are in
use; the rest is generally disposed of in highly concentrated form on land. Urban areas thus
represent a significant concentrated source of organic C and biologically available N, P, and
S. The major cities of the world are located near major rivers and coastal areas, and water
discharges into riverine and marine ecosystems will increase primary production and anoxia
(see section 1.5.4).

E. Grazing

Man has influenced C, N, P, and S interactions by managing grazing animals in various ways.
Management can be extensive (animals maintained in losely restricted areas by fencing or
herding) or intensive (pasture improvement by altering plant species composition,
fertilization, irrigation and supplemental feeding). With proper extensive management, the
only important alterations of any of the cycles is increased production of NH3 from the
animal waste products. Although such losses are of a low magnitude over large areas, they
have important regional atmospheric chemistry implications (see section 1.4.3 B).

With intensive management, animal numbers can be greatly increased in small areas.
Intensive management can cause increases in NH3 evolution in relatively small areas (section
1.4.3 B) and significant inputs in nitrate into surface or ground-water (section 1.5.4).

F Overview: Cultivation and Nutrient Balances

The cultivation of terrestrial ecosystems may cause substantial losses of C, N, P, and S


through all of the pathways discussed above. However, cultivated ecosystems can be
managed so that loss of soil fertility is minimized and even so that the organic matter,
nutrient content and productivity of once degraded sites are increased (Power, 1981). These
practices are now relatively energy intensive if they are to be effective on the time scale of
decades.

During the 50100 years of continuous cultivation in the North American Great Plains,
approximately half of the C and N, 3040% of the S and 1030% of the P in the soil has been
lost (Haas et al., 1957; Haas et al., 1961; and Stewart et al., Chapter 8, this volume). Similar
losses have taken place in less than 10 years in tropical areas (Nye and Greenland, 1964). The
loss of N, P, and S reduces the ability of the soil to supply these nutrients, and increasing
levels of fertilization are required to maintain crop production. Without perennial plants or
structural carbon in the soil, fertilizer nitrogen cannot be effectively retained from year to
year. Loss of soil organic matter and continued erosion decreases soil fertility and can
threaten the long term maintenance of productive agriculture on a site.
The fundamental concepts underlying alternative cultivation practices involve bringing into
balance the net inputs of organic C (by photosynthesis, residue return, and reduced
decomposition), N, P, and S, thus in some sense reconstructing the interacting nutrient cycles
of natural terrestrial ecosystems. After land abandonment, the cycles of these elements can
also be brought into balance by natural processes but these will often require tens or hundreds
of years, even provided enough P remains in available form in the soil (Cole and Heil, 1981).
In the light of the marked changes of the soils that are occurring throughout the world, the
further development of long term strategies for soil management is most urgent.

1.4.4 The Impact of Fossil Fuel Combustion on the Global Ecosystem

A. Changing atmospheric CO2 concentration

The CO2 released in fossil fuel combustion increases the atmospheric CO2 concentration on a
global scale (Bolin et al., 1979). The direct effects of elevated atmospheric CO2 on terrestrial
ecosystems are not well known, beyond the fact that both photosynthesis and growth increase
in individual plants that are not strongly light, water or nutrient limited. The water use
efficiency of plants (the ratio of net primary production to evapotranspiration) is also
increased because the CO2 concentration gradient between plants and leaves is increased. It is
not clear, however, that either natural or cultivated plants in the field are CO2 limited often
enough so that plant growth can be substantially increased in this way. Even if the growth of
individual plants was increased, competition would limit the possible increase in community-
level biomass. Some increase in terrestrial productivity (due to water use efficiency) and
some, but probably quite limited, increase in total biomass is the most likely result of
elevated atmospheric CO2 levels, and the latter will require considerable time.

B. Impact of N and S deposition

The majority of the S and N released is at present deposited in either gaseous (SO2, NOx,
HNO3) or ionic form (SO42, NO3, NH4+) within 1500 km of its origin (Rodhe, 1981). As the
emissions increase, the dispersion of these elements will very likely be more wide-spread (see
section 1.3.3D; Rodhe, 1981; Taylor et al., Chapter 4, this volume). Still, most of the N and S
released during fossil fuel combustion will be deposited within a few thousand kilometers of
where the fuel is consumed.

The overall effects of these changes on C, N, P, and S cycling in terrestrial ecosystems are
largely unknown. SO2 and other gases do reduce plant growth near major sites of fossil fuel
combustion. Most terrestrial ecosystems are relatively well buffered, however, and direct
toxic effects of increased acidity (comparable to the deleterious effects so clearly
demonstrated in many lakes) have only been conclusively established in a few areas (Ulrich
et al., 1980). Acid precipitation does cause increased leaching of sulphate and accompanying
cations and trace metals to stream-water and ground-water in many soils, but sulphate
adsorption in old, acid soils can reduce leaching to negligible levels (Johnson et al., 1980).
When increased leaching does occur, the loss of cations could eventually decrease terrestrial
productivity. Acid precipitation at current levels could also decrease rates of litter
decomposition by up to 10% (Cook, Chapter 12, this volume).

On the other hand, the N and S in precipitation could fertilize the impacted areas, increasing
production in N- or S-limited sites (Melillo and Gosz, Chapter 6, this volume). The effect
could be significant over considerable areas for N and in fewer sites for S, but there is no
conclusive evidence for any directional change in forest productivity attributable to changes
in atmospheric deposition. If forest biomass is increased by these N and S additions, some of
the CO2 released by fossil fuel emissions could accumulate in the increased biomass. The
maximum potential accumulation of C in this pool is estimated at 250500 Tg/yr, which is
510% of the C released by combustion (Melillo and Gosz, Chapter 6, this volume, and
discussions at the workshop).

The emissions to the atmosphere due to fossil fuel combustion means the addition of very
large amounts of C, N, and S to the natural circulation of these elements in nature. By
continuing this practice at a high and even increasing level we will change the chemical
climate on earth. It is urgent to establish more firmly what this will imply over the time scales
of one or a few centuries.

1.5 THE FRESHWATER SYSTEM AND COASTAL WATERS

1.5.1 Water as an Agent in Biogeochemical Cycles

The circulation of C, N, P, and S is largely due to the motions of air and water on earth. The
role of the atmosphere in this regard has been discussed (section 1.3). At this stage we are
concerned with the transfer of matter by water from land to sea.

The hydrologic cycle begins with the evaporation of water due to solar radiation and the
dispersion of water vapor about the atmosphere. Part of the water is deposited upon the
continents and returns to the oceans via rivers, coastal run-off and subterranean discharge;
part returns directly to the oceans. The atmospheric and terrestrial waters are continuously
altered chemically by processes along their way back to the oceans. Gas dissolution in the
water droplets of the atmosphere, biological activity in the rivers, and continental erosion are
examples of such processes. Many chemistries interact to give each river a unique
composition (Richey, Chapter 13, this volume).

Since the industrial revolution over a century ago, the combination of social, agricultural and
technological activities of man have distorted this cycle. Pollutants have entered the
atmospheric and the terrestrial waters. Rivers have been dammed and channelled; the natural
chemistries of weathering have been altered. As a consequence, the waters entering the
marine environment have many imprints of human society.

1.5.2 The Natural System

The world's rivers are the major transporting agents for continental weathering products to
the oceans. However, they are also a dynamic system, and have active biogeochemical cycles
that modify their dissolved and suspended loads (Richey, Chapter 13, this volume).

The sources of C, N, P, and S in rivers are precipitation, erosion, entrainment from flood
plains, chemical weathering, slope run-off and leaching. The processes involved have
generally produced C, N, P, and S species in oxidized states. The diversity of (1) various
weathering regimes, (2) the involvement of flood plains, (3) the character of the vegetation in
the drainage basin, (4) hydrographic conditions, and (5) the occurrence of episodic events
such as floods, determine river compositions. Aerobic respiration is the most important
process modifying the species, forms and amounts of C, N, P, and S transported by the rivers.
When lakes occur along a river path, flow is interrupted, leading to increased particulate
deposition and longer transit times favouring the occurrence of phytoplankton blooms. In the
hypolimnion (the cold, deep and undisturbed region of such lakes) there is a shift toward
reduced chemical species (e.g. less NO3 and more NH4+).

The waters of rivers mix with the waters of the sea in the estuarine zone. Chemical and
biological processes may profoundly change the speciation and the transfer of elements to the
adjacent coastal zone, depending to a large degree on the hydrodynamical properties of the
estuary. Precipitation, flocculation and sedimentation remove part of the riverborne organic
matter and some inorganic species essential for plant growth, such as iron. The decrease in
turbidity associated with these processes promotes high productivities when, as is usually the
case in estuaries, the nutrients are in high supply.

The nutrients are then transferred from solution to particulate organic matter and can be
trapped by sedimentation in deeper waters or in sediments. On the other hand, when high
productivity occurs, high turbidities can result from the large amount of phytoplankton in the
photic zone. As a consequence, there can be a severe limitation on the depth of
photosynthetic activity. This effect is enhanced by a thermal stabilization due to light
absorption and to associated warming in the uppermost layer. With limited nitrogen
availability, these stabilized conditions promote blue-green algae growth.

Another consequence may be increased sedimentation of organic matter that leads to high
biological activity in the surface sediments (benthos). Nutrient regeneration to the waters
overlying the sediments is controlled by the microbial degradation of organic matter
originating from river input and primary production in the surface waters. C, N, and P
compounds introduced into the pore waters from such processes re-enter the overlying waters
by molecular diffusion, bioturbation, sediment resuspension, and ground-water seepage.

Depending upon the intensity of the primary production and the hydrography of a coastal
area, there is a possibility that all of the oxygen in the sediments may be consumed through
the microbial degradation of organic matter. Anoxia can also occur during summer months in
the water column over these sediments. This condition can be expected in either a permanent
or seasonal mode depending upon whether the system has a permanent stratification (e.g. a
tropical river estuary) or a seasonal one (e.g. the Baltic Sea and some fjords).

Following the loss of oxygen from sedimentary strata, nitrate is utilized as the oxidizing
agent yielding nitrogen gas and ammonia as products. With the depletion of nitrate, which
generally occurs deeper in the sediment, sulphate becomes the oxidizing agent, and hydrogen
sulphide is produced. Finally, when all of the sulphate is utilized, fermentation and carbon
dioxide reduction are initiated with the evolution of methane (Jørgensen, Chapter 18, this
volume).

The coastal zone, which we define as that part of the marine environment bounded by the 200
metre isobath and the continents, receives the material transported by the rivers and estuaries.
This area produces about 25% of the primary productivity of the oceans and receives about
90% of the organic matter that reaches the ocean floor and is buried (Jørgensen, Chapter 18,
this volume). Further, about 50% of the fisheries that provide society with food are located
here.
The primary productivity of coastal waters can be limited by nitrogen deficiencies and in rare
cases by those of phosphorus. In such cases the waters are generally clear and the nutrients N
and P are rapidly taken up by the plankton.

In unpolluted rivers the dissolved organic carbon mixes with ocean water without
disintegration and most of it is exported to the open ocean. Particulate and dissolved organic
carbon in the estuary is formed from microbiologically labile material, which is extensively
recycled.

1.5.3 Disturbances in the Weathering Cycle Created by Man

The major perturbation in the weathering cycle brought about by man involves the increased
fluxes of C, N, P, and S throughout the hydrosphere and, from increases in N and P, a
consequent increase in photosynthetic activity. This tendency toward euthrophication is
especially evident in lakes and coastal marine waters and in some rivers.

The nutrient concentration of rivers on a global basis and over a 30 years' period has
increased by at least a factor of 2 for dissolved inorganic nitrogen and phosphorus (Wollast,
Chapter 14, this volume). This is a result of the wide-spread use of fertilizers, detergents, and
industrial chemicals containing N and P. In addition there is an increased transport of organic
carbon in the weathering cycle due to deforestation and land-use changes as well as through
the increased production of municipal, agricultural and industrial wastes. A part of these
materials is delivered to the coastal zone by the rivers. Much will be modified within the
rivers, especially near sources such as outfalls and manufacturing plants.

The increased biological activity due to the entry of N and P in rivers has led in some cases to
anoxia and consequent fish mortalities. Water impoundment for flood control, hydroelectric
power, or water storage has caused the retention of large amounts of particulate materials.
Additionally, earth moving activities have yielded large amounts of particulates to the rivers
(Richey, Chapter 13, this volume). Channelization has modified river transit by confining the
river to its bed and preventing plain flooding (entrainment) from influencing river chemistry.
Rivers may also carry elevated levels of substances antagonistic to primary productivity, such
as heavy metals, e.g. copper (Morel and Morel-Laurens, 1982).

The changes in the ratios of nutrients entering the coastal zone over the past decades have
brought about major changes as well. The entry of large amounts of P relative to N and Si
have lead to dominant plankton communities of blue-green and brown algae that replace the
silicon containing diatoms, which form the basis of many valuable food webs (Wollast,
Chapter 14, this volume). This situation develops in those coastal areas not subject to
extensive mixing with the open ocean.

Future changes in the food web are to be expected. Presently, a shortening of the food web,
high mortality of the phytoplankton and a subsequent rapid turnover of organic matter is
observed (Goldberg, 1979a; Wollast, Chapter 14, this volume).

The increased biological activity usually extends throughout the coastal region and probably
beyond it. The increased supply of organic detritus from plant production in the coastal zone
will shift the location of decomposition from the water column, which recycles organic
matter on time scales of days to weeks, to the sediments, which recycles it on time scales of
months to years (Martens and Jannasch, comment to Chapter 18, this volume).
The CaCO3 budget of the rivers and coastal ocean can be disturbed by the increased
productivity induced by man. Respiratory activity in some rivers has increased to such an
extent that high CO2 pressures, and the consequent decrease in pH, has led to undersaturation
of CaCO3 and to its dissolution. Such a situation has been observed in the coastal ocean with
the potential dissolution of the more readily soluble magnesium calcites.

Although these changes in the major weathering cycle have been observed, a complete
quantification of the changing nutrient budgets of C, N, P, and S for such major rivers as the
Amazon, Ganges, Yangzi, Zaire and Missisippi has not been made. Such data are essential
for understanding the carbon cycle, where large transient reservoirs might exist but at present
are not accounted for, as well as for understanding the present and future vitalities of food
chains with value for human society.

1.5.4 Human Waste and the Waters of the World

The fluxes of C, N, P, and S induced by man now exceed the natural fluxes in many parts of
the world. There are a number of effects of this on rivers and on the coastal ocean. Primarily,
there is the increased production of organic matter and its consequencesthe replacement of
the algae at the base of important food chains with less desirable species; increasing areas of
anoxia; and increased loadings of sediments with organic carbon.

This situation clearly has an impact upon the renewable resources of the coastal ocean.
Besides the food from the sea, recreational areas and aesthetic attractions of the coastal
environment can be threatened by extensions of anoxic areas. There is, in principle, a limit to
the use of the oceans as a receptor for a part of the wastes generated by society.

The world population is increasing and the utilization of materials that are to be disposed of
continues to rise. The coastal environment appears to many as the most convenient place for
disposal. Additional entries of organic materials and of nutrients, such as nitrogen and
phosphorus, would add to the existing problem of eutrophication.

It is thus essential that the assimilative capacities of the oceans be assessed with respect to
waste disposal. At the present time there have been only a few studies dedicated to this end
(see Goldberg, 1979b). These evaluations generally use any alteration of the structure of
communities of organisms as a criterion for determining whether the carrying capacity of a
coastal water has been exceeded. Such assessments are still in a primitive state.

Those responsible for maintaining the quality of the marine environment, at national or
international levels, can be guided by this concern. Optional methods of waste management
are to be assessed with the recognition of this problem of enhancing the eutrophication of
waters in the weathering cycle.

1.6 THE OPEN SEA

1.6.1 Response Characteristics of the Oceans

The open sea is by far the largest reservoir for the basic elements of our concern. On time
scales of many hundred years to millenia, the water and the sediments of the oceans will be
the main receptor of most of the human emissions regardless of whether these are made to the
air, the soils, the fresh-water system or directly into the oceans. The pathways of the
emissions from these various spheres to the sea are, however, long and the oceans respond
slowly to imposed disturbances. New quasi-equilibria between the various parts of the global
ecosystem will therefore only be established slowly. Due to the size of the ocean reservoir,
final equilibria may not be vastly different from those that existed during pre-industrial times.
The transient period on the other hand is and will continue to be characterized by marked
deviations from pristine conditions. The characteristic time scale of modern society is short
compared to the characteristic times of adjustment of the oceans. To provide a first order
picture of the role of ocean circulation in this regard, the interaction of the wind stress and the
heat and salt (thermohaline) transport processes must be considered.

The global winds generate a system of horizontally circulating gyres in the surface layers of
the oceans. Those occurring in the equatorial and subpolar regions are characterized by
upwelling and those in the subtropics by downwelling.

Atmospheric warming creates a warm surface layer in the tropics and subtropics, extending
north and south to 4045° latitude. In most of the subtropical anticyclonic gyres this layer is
thick due to the downwelling. In contrast, the surface layers in the equatorial regions are
shallow, merely to 100 m depth, and they extend well into the subtropical gyres in both
hemispheres. This shallow surface layer communicates with deeper layers by inclusion of
upwelling water in the equatorial region and in coastal upwelling zones, such as those off the
coasts of Peru, California and northwest Africa.

Moderately warm water occupies the layer down to around 1000 metres and equatorward of
the West wind maxima at 4045° latitude. This layer is the permanent thermocline, and
upwelling in low latitudes into the surface layer is limited to waters from this layer. The
poleward half of the subtropical gyres on the other hand are characterized by pronounced
seasonal variations and late winter convective activity. The volumes of the warm surface
waters and the moderately warm thermocline region are about 1 and 10%, respectively, of the
total ocean volume and have characteristic response times of a few years and a few decades
(Fiadeiro, Chapter 17, this volume).

The upwelling subpolar gyres that are subject to surface cooling have little surface
stratification and are sites of deep convective overturning. Although these high latitude
regions represent but a small fraction of the total ocean surface area, they are the only regions
where vertical communication by water motion occurs between the surface layers and the
deep water. This is limited to the North Atlantic and the Antarctic waters in our present
climate. From radiocarbon data, the gross residence time for the deep waters relative to
formation at the surface is 5001000 years (Fiadeiro, Chapter 17, this volume). The intrusion
of anthropogenic tracers as observed in the North Atlantic deep waters as far south as 30°N
latitude is consistent with this long response time due to the sluggish interior circulation of
the deep waters. The fraction of the deep water that carries these tracers is quite small.

The large volume of abyssal waters is not very well mixed. The microbial breakdown of
sedimenting organic debris within this water mass and the resultant release of nutrients can
significantly influence biogeochemical equilibria.

In studies of the role of the ocean in the global biogeochemical cycles rather simple models
have been developed. From these studies it is possible to successfully reproduce the gross
vertical distribution of, for example, radiocarbon and total dissolved inorganic carbon (e.g.
Oeschger et al., 1975). Fiadeiro (Chapter 17, this volume) emphasizes, however, that simple
models are incapable of depicting transient processes on time scales that are small compared
with the transfer times that are resolved by the model. Because the time scale of development
of modern industrial activity is a few decades to at most half a century, the traditional box-
models of the oceans must be replaced by models that can properly depict the motions on
time scales of less than 30 years. Such work is in progress (Bolin, 1983; Viecelli et al., 1981;
Fiadeiro, Chapter 17, this volume). Ultimately, detailed general circulation models will be
employed for such studies.

1.6.2 Life in the Sea

Photosynthesis in the sea is limited to the photic zone and therefore so is most life. Still some
remarkable communities have developed in the deep sea, of which those found around deep
sea vents are most spectacular (Jannasch, Chapter 19, this volume). The evolution of any
such features of the global ocean ecosystem is a very slow process, but their destruction may
occur much more quickly.

The turn-over of organic matter in the well mixed ocean surface layer (about 100 m deep) is
rapid and well governed by the Redfield hypothesis of distinct ratios of the elements involved
(see section 1.2). Concentrations of the major elements in the surface waters represent a
balance between biological processes, exchange with the atmosphere (Liss, Chapter 15, this
volume; Duce, Chapter 16, this volume) and with deeper layers of the ocean. Only a few
percent of the dead organic matter settles out of this layer as detritus. Still, all nutrient
inhomogeneities of the ocean below the surface layer stem from this material transfer into
deeper layers and its decomposition and dissolution.

Due to the ability of some marine organisms to fix nitrogen, it appears that of the major
nutrients, only phosphorus supply is truly critical. Thus, the removal of carbon in the form of
organic matter from the surface layers is, to first order, totally controlled by phosphorus.

Primary production, which requires light, is limited to a fairly thin layer near the ocean
surface. Photosynthetic organisms are very efficient in utilizing available nutrients and, in
fact, they recirculate much of the nutrient materials at a rapid rate. The result is that most of
the phosphorus is found in organic matter and very little inorganic phosphorus exists in
solution. Some small fraction of the organic matter does settle out, most in partially digested
faecal material and to some extent in dead settling organisms of a size large enough to
provide significant settling rates.

The decomposition of this organic material in the deep water or in the surface layers of the
sediments leads to a development of carbon dioxide supersaturation accompanied by nutrient
regeneration into dissolved form approximately in the Redfield ratios. The excess nutrients
and carbon dioxide are, on a long time scale, returned to the surface layers to partake again in
the biogeochemical cycles except for possible reactions with surface sediments and burial of
some materials in the sediment. The factors determining the difference in concentration of
these nutrients between surface and deep waters are the turn-over rate of the ocean as a
physical system and the internal dynamics of the biological system. Assuming the turn-over
rate is fixed, we have a basis for a nutrient controlled biological transport mechanism for
carbon dioxide. The system is likely to be nonlinear, since several trophic levels will be
involved. It is also complicated by the presence of silica, which seems to control the balance
between diatom and calcareous algae production. If, however, an increased phosphorus
supply were to cause a substantial increase in the photosynthesis in surface layers, growths in
grazer and predator populations would ultimately lead a new equilibrium in which practically
all the extra phosphate would be present in the deep waters along with a significantly larger
amount of excess dissolved CO2. Enhanced corrosiveness of the deep waters would then lead
to carbonate dissolution from sediments. Thus the oceanic PCO2 and hence, the carbon
dioxide concentration in the atmosphere would be controlled by the available phosphate
supply in the oceans. Such interactions might conceivably occur on time scales associated
with glacial and interglacial periods (Broecker, 1981). The importance of biogeochemical
interactions of this kind during the coming centuries due to increasing perturbations by man
needs to be carefully assessed.

Man's interference with the global ocean system is still small. Even though pollution of the
ocean surface layer can be detected all over the world, the direct impact of man on the C, N,
P, and S cycles has not yet been established by direct measurements in the open sea. The
required accuracy of measurements can be achieved but the natural variability implies that
long observation series are required. The expected change of total inorganic carbon in the
surface layer since the beginning of the industrial revolution due to CO2 transfer from the
atmosphere is 12% and the present rate of changes less than 0.4% or 0.01 mmol/1 per decade.
Similarly we can estimate that even if 20% of all phosphorus used by man reaches the open
sea (which is most unlikely) the photosynthesis could increase by only 12% if all of this
additional supply were used. Due to the size of the oceans, cumulative effects take place only
very gradually. It is, however, important to monitor the oceans carefully in these regards,
because small changes correspond to large net fluxes and are therefore of importance for the
overall budgets for the elements concerned.

1.7 CONCLUSIONS

In the course of the survey given in the preceding sections, a number of conclusions have
been reached; the contributed papers contain many additional specific results. An equally
important outcome of the workshop is, however, a more integrated view of how the global
ecosystem as a whole functions. Clearly, there are numerous ways whereby different parts of
this system interact. We have learned that when disturbances to the system becomes
significant in one region of the world, implications need to be carefully assessed for the globe
as a whole. Much further research is required to fully grasp this global interplay with regard
to both important detailed processes and the overall interactions. We shall briefly summarize
the outcome of the workshop in this regard in a series of concluding statements.

The participants of the workshop agreed that:

A. Man's activities on earth today induce fluxes of carbon, nitrogen, phosphorus and sulphur
that are of similar magnitude to those associated with the natural global cycles of these
elements; in limited areas man's influence dominates the cycles. The likely increase of man's
activities during the remainder of this and during the next century will undoubtedly mean
significant disturbances of the global ecosystem.

B. The most important ways whereby man is interfering with the global ecosystem are:
fossil fuel burning which may

a. double the atmospheric CO2 concentration by the middle of next century;


b. further increase the emissions of oxides of sulphur and nitrogen very
significantly;
expanding agriculture and forestry and the associated use of fertilizers (nitrogen and
phosphorus) significantly alter the natural circulation of these nutrients;
increased exploitation of the fresh-water system both for irrigation in agriculture and industry
and for waste disposal.

C. The present rate of change of the global ecosystem in response to man's activities will
increase further and this implies increased departures from the pre-industrial quasi-steady
state. No part of the globe will be untouched by these changes.

D. According to our present understanding, the most important impacts of these changes in
the long term perspective are:
a gradual change towards a warmer climate, the details and implications of which we know
very little about;
the concentration of ozone will decrease in the stratosphere, due to the increased release of
N2O and chlorine compounds and increase in the troposphere, due to the increased release of
NOx and hydrocarbons;
an increase of the areas affected by lake and stream acidification in mid-latitudes and
possibly also in the tropics; the ion balance of the soils may be significantly disturbed, as is
now being found with regard to aluminium;
a decrease of the extent of tropical forests, which will enhance the rate of increase in
atmospheric CO2 concentration and release other minor constituents to the atmosphere; this
may also contribute to soil degradation;
due to loss of organic matter and nutrients, soil deterioration will occur and this implies a
reduced possibility for the vegetation to return to pristine conditions; global mapping of
ongoing soil changes is urgently needed;
a trend toward the eutrophication of estuarine and coastal marine areas;
more frequent development of anoxic conditions in fresh-water and marine systems and
sediments.

E. The development and continuation of highly productive units in agriculture and forestry
means an increasing dependance on technological advances that, to be properly directed,
requires profound knowledge about long term modifications of the soil.

F. The long term implications of exploiting the natural resources of the earth are not well
understood, nor do we understand what is permissible in order to guarantee that present or
future (possibly higher) levels of productivity will not later decline. We need to develop a
strategy for how to assess the long term carrying capacity of the earth. It should be obvious
from the overview of this vast problem area that we are yet far from such an integrated view
of the global ecosystem.

1.8 REFERENCES

Ausmus, B. S., Edwards, N. T., and Witkamp, M. (1976) Microbial immobilization of


carbon, nitrogen, phosphorus and potassium: implications for forest ecosystem processes, in
Anderson, J. M. and MacFadyen, A. (eds) The Role of Terrestrial and Aquatic Organisms in
Decomposition Processes, London, Blackwell, 397-421.

Berg, W. E., Crutzen, P. J., Grahek, F. E., and Gitlin, S. N. (1980) First measurements of total
chlorine and bromine in the lower stratosphere, Geophys. Res. Lett., 7, 937-940.
Berner, W., Oeschger, H., and Stauffer, B. (1980) Information on the CO2 cycle from ice core
studies, Radiocarbon, 22, 227-235.

Bolin, B. (1983) Changing global biogeochemistry, in Brewer, P. (ed.) 50th Anniversary


Volume, Woods Hole Oceanographic Institution, The Future of Oceanography, New York,
Springer-Verlag, 305-326.

Bolin, B., and Charlson, R. J. (1976) On the role of the tropospheric sulphur cycle in the
shortwave radiative climate of the earth, Ambio, 5, 47-54.

Bolin, B., Degens, E. T., Duvigneaud, P., and Kempe, S., (1979) The Global Biogeochemical
Carbon Cycle, In Bolin, B., Degens, E. T., Kempe, S., and Ketner, P. (eds) The Global
Carbon Cycle. SCOPE Report No. 13, New York, Wiley,1-53.

Bowen, H. J. M. (1979) Environmental Chemistry of the Elements. New York, Academic


Press.

Bormann, F. H., and Likens, G. E. (1979) Pattern and Process in a Forested Ecosystem, New
York, Springer-Verlag.

Brewer, P. (1978) Direct observations of the oceanic CO2 increase, Geophys. Res. Lett., 5,
997-1000.

Broecker, W. S. (1974) Chemical Oceanography, New York, Harcourt Brace Jovanovich.

Broecker, W. S. (1981) Glacial to interglacial changes in ocean and atmosphere chemistry, in


Berger, A. (ed.) Climatic Variations and Variability: Facts and Theories, Dordrecht, D.
Reidel 111-121.

Cicerone, R. J., and Shetter, J. D. (1981) Sources of atmospheric methane: measurements in


rice paddies and a discussion, J. Geophys. Res., 86, 7203-7209.

Cole, C. V., and Heil, R. D. (1981) Phosphorus effects on terrestrial nitrogen cycling, in
Clark, F. E., and Rosswall, T. (eds) Terrestrial Nitrogen Cycles, Ecol. Bull. (Stockholm), 33,
363-374.

Cook, R. B. The impact of acid deposition on the cycles of C, N, P, and S. (Chapter 12, this
volume).

Crutzen, P. J. (1976) The possible importance of CSO for the sulphate layer of the
stratosphere, Geophys. Res. Lett., 3, 73-76.

Crutzen, P. J. Atmospheric interactionshomogeneous gas reactions of C, N, and S containing


compounds (Chapter 3, this volume).

Crutzen, P. J., Heidt, L. E., Krasnec, P. J., Pollock, W. H., and Seiler, W. (1979) Biomass
burning as a source of atmospheric gases CO, H2, NO, CH3Cl and COS, Nature, 282, 253-
256.
Cullis, C. F., and Hirschler, M. M. (1980) Atmospheric sulphur: natural and man-made
sources, Atmos. Environ., 14,1263-1278.

Delmas, R. J., Ascencio, J. M., and Legrand, M. (1980) Polar ice evidence that atmospheric
CO2 20,000 yr BP was 50% of present, Nature, 284,155-157.

Delwiche, C. C., and Likens, G. E. (1977) Biological response to fossil fuel combustion
products, in Stumm, W. (ed.) Global Chemical Cycles and their Alteration by Man. Berlin,
Dahlem Konferenzen, 73-88.

Duce, R. A. Biogeochemical cycles and the air-sea exchange of aerosols (Chapter 16, this
volume).

Fiadeiro, M. Physical-chemical processes in the open ocean (Chapter 17, this volume).

Fishman, J., Solomon, S., and Crutzen, P. J. (1979) Observational and theoretical evidence in
support of a significant in-situ photochemical source of tropospheric ozone, Tellus, 31, 432-
446.

Food and Agriculture Organization (FAO) of the U.N. (1979) Agriculture: Toward 2000,
Conference C/79/27, Rome, FAO, 30-31.

Freney, J., Ivanov, M. V., and Rodhe, H. Major global reservoirs and transfers of sulphur.
(Chapter 2, this volume).

Goldberg, E. D. (ed.) (1979a) Scientific Problems Relating to Ocean Pollution, Washington,


DC, U.S. Dept. of Commerce (NOAA).

Goldberg, E. D. (ed.) (1979b) Assimilative Capacity of U.S. Coastal Waters for Pollutants,
Washington, DC, U.S. Dept. of Commerce (NOAA).

Granat, L., Rodhe, H., and Hallberg, R. O. (1976) The global sulphur cycle, in Svensson, B.
H., and Söderlund, R. (eds) Nitrogen, Phosphorus and SulphurGlobal Cycles, SCOPE Report
No. 7. Ecol. Bull. (Stockholm), 22, 89-134.

Haas, H. J., Evans, C. E., and Miles, E. F. (1957) Nitrogen and carbon changes on Great
Plains soils as influenced by cropping and soil treatments, Washington, D.C., USDA
Technical Bulletin No. 1164.

Haas, H. J., Grunes, D. L., and Reichman, G. A. (1961) Phosphorus changes in Great Plains
soils as influenced by cropping and manual applications, Soil. Sci. Soc. Amer. Proc., 25, 214-
218.

Hoffmann, D. J., and Rosen, J. M. (1981) On the background stratospheric aerosol layer, J.
Atmos. Sci., 38,168-181.

Holland. H. D. (1978) The Chemistry of the Atmosphere and Oceans, New York, Wiley-
Interscience.
Houghton, R., and Woodwell, G. W. Effect of increased CO2 on the cycles of C, N, P and S
(Chapter 11, this volume).

Hunt, H. W., Stewart, J. W. B., and Cole, C. V. A conceptual model for the basis of
interactions among carbon, nitrogen, sulphur, and phosphorus transformations (Chapter 10,
this volume).

Husain, L., Dutkiewicz, V. A., and Rusheed, A. (1981) Stratospheric-tropospheric exchange


in the Northern and the Southern Hemisphere, EOS, 62, 292.

Husar, R. B., and Patterson, D. E. (1980) Regional scale air pollution: sources and effects, in
Kneip, T. J., and Lioy, P. J. (eds) Aerosols: Anthropogenic and Natural Sources and
Transport. Ann. N. Y. Acad. Sci., 338, 339-417.

Ivanov, M. V. (1981) The global biogeochemical sulphur cycle, in Likens, G. E. (ed.) Some
Perspectives of the Major Biogeochemical Cycles. SCOPE Report No. 17, Chichester, Wiley,
61-78.

Jannasch, H. W. Interactions between the carbon and sulphur cycles in marine environments
(Chapter 19, this volume).

Johnson, D. W., Hornbeck, J. W., Kelly, J. M., Swank, W. T., and Todd, D. E. (1980)
Regional patterns of sulphate soil accumulation: relevance to ecosystem sulphur budgets, in
Shriner, D. S., Richmond, C. R., and Lindberg, S. E. (eds) Atmospheric Sulphur Deposition:
Environmental Impact and Health Effects, Michigan, Ann Arbor Science, 507-520.

Jørgensen, B. B. Processes at the sedimentwater interface (Chapter 18, this volume).

Kellogg, W. W., and Schware, R. (1981) Climatic Change and Society, Boulder, Colorado,
Westview Press.

Lanly, J. P., and Clement, J. (1979) Present and future natural forest and plantation areas in
the tropics, Unasylva, 31(123),12-20.

Likens, G. E., Bormann, F. H., and Johnson, N. B. (1981) Interactions between major
biogeochemical cycles: terrestrial cycles, in Likens, G. E. (ed.) Some Perspectives of the
Major Biogeochemical Cycles, SCOPE Report No. 17, Chichester, Wiley, 93-112.

Liss, P. S. The exchange of biogeochemically-important gases across the air-sea interface


(Chapter 15, this volume).

Lovelock, J. E. (1974) Atmospheric halocarbons and stratospheric ozone, Nature, 252,292-


294.

Lovelock, J. E. (1979) Gaia, A New Look at Life on Earth, Oxford, Oxford University Press.

McGill, W. B., and Christie, E. K. Biogeochemical aspects of nutrient cycle interactions in


soil and organisms (Chapter 9, this volume).
MacKenzie, F. T., Stoffyn, M., and Wollast, R. (1978) Aluminium in seawater: control by
biological activity, Science, 199, 680-682.

Manabe, S., and Stauffer, R. J. (1980) Sensitivity of a global climate model to an increase of
CO2 concentration in the atmosphere, J. Geophys. Res., 85, 5529-5554.

Martens, C. S., and Jannasch, H. W. Cycling of metabolizable C, N, P and S in organic-rich


marine sediments. (Comment to Chapter 18, this volume).

Meixner, F. (1981) Die vertikale Verteilung des atmosphärischen Schwefel-dioxides in


Tropopausenbereich, Doktor-Dissertation, Frankfurt-am-Main Johann-Wolfgang-Goethe-
Universität, (176 pages).

Melillo, J. M., and Gosz, J. R. Interactions of biogeochemical cycles in forest ecosystems


(Chapter 6, this volume).

Molina, M. J., and Rowland, F. S. (1974) Stratospheric sink for chloro-fluoromethanes:


chlorine atom catalyzed destruction of ozone, Nature, 249, 810-812.

Morel, F. M. M., and Morel-Laurens, N. M. L. (1982) Trace metals and plankton in the
oceans, in Wong, C. S., Boyle, E., Bruland, K. W., and Burton, J. D. (eds.) Trace Metals in
Sea Water, Proc. NATO Adv. Res. Inst. New York, Plenum Press (in press).

Nye, P. H., and Greenland, D. J. (1964) Changes in the soil after clearing tropical forest,
Plant Soil, 21,101-112.

Odén, S. (1968) The acidification of air precipitation and its consequences in the natural
environment. Bulletin from the Ecological Research Committee Vol. 1. Swedish National
Science Research Council, Stockholm. Translation Consultants Ltd. Arlington, Virginia,
USA, 117 pages.

Oeschger, H., Siegenthaler, U., Shotterer, U., and Gugelmann, A. (1975) A box diffusion
model to study the carbon dioxide exchange in nature, Tellus, 27, 168-192.

Penkett, S. A., Jones, B. M. R., Brice, K. A. and Eggleton, A. E. J. (1979) The importance of
atmospheric ozone and hydrogen peroxide in oxidizing sulphur dioxide in cloud and
rainwater, Atmos. Environ., 13, 123-138.

Pimentel, D., Terhune, E. C., Dyson-Hudson, R., Rochereay, S., Samis, R., Smith, E. A.,
Denman, D., Reifschneider, D., and Shepard, M. (1976) Land degradation: effects on food
and energy resources, Science, 194, 149-155.

Power, J. F. (1981) Nitrogen in the cultivated ecosystem, in Clark, F. E., and Rosswall, T.
(eds) Terrestrial Nitrogen Cycles, Ecol. Bull. (Stockholm), 33, 529-546.

Rahn, K. A. (1981) The Mn/V ratio as a tracer of large-scale sources of pollution aerosol for
the Arctic, Atmos. Environ., 15, 1457-1464.

Rahn, K. A., and Heidam, N. Z. (1981) Progress in Arctic air chemistry, 1977-1980: a
comparison of the first and second symposia, Atmos. Environ., 15, 1345-1348.
Ramanathan, V. (1975) Greenhouse effect due to chlorofluorocarbons: climatic implications,
Science, 190, 50-52.

Rapp, A. (1975) Soil erosion and sedimentation in Tanzania and Lesotho, Ambio, 4, 154-163.

Rasmussen, R. A., and Khalil, M. A. L. (1981) Atmospheric methane (CH4): trends and
seasonal cycles. J. Geophys. Res., 86, 9826-9832.

Redfield, A. C. (1958) The biological control of chemical factors in the environment, Am.
Sci., 46,206-226.

Reiners, W. A. Transport processes within and between biomes (Chapter 5, this volume).

Richey, J. E. Interactions of biogeochemical cycles in fresh-water ecosystems (Chapter 13,


this volume).

Rodhe, H. (1981) Current problems related to the atmospheric part of the sulphur cycle, in
Likens, G. E. (ed.) Some Perspectives of the Major Biogeochemical Cycles, SCOPE Report
No. 17, Chichester, Wiley, 51-60.

Rosen, H., and Novakov, T. (1980) Soot in the Arctic, GCMN No. 8, Summary Report 1979,
Herbert, G. A. (ed.) Boulder, Colorado, USA, NOAA, ERL, 84-88.

Rosen, H., Novakov, T., and Bodhaine, B. A. (1981) Soot in the Arctic, Atmos. Environ., 15,
1371-1374.

Rosswall, T. (1976) The internal cycle between vegetation, micro-organisms, and soils, in
Svensson, B. H., and Söderlund, R. (eds) Nitrogen, Phosphorus and SulphurGlobal Cycles,
SCOPE Report No. 7, Ecol. Bull. (Stockholm), 22, 157-167.

Rosswall, T. The Nitrogen Cycle (Chapter 2, this volume).

Roddy, R. M., and Marland, G. (1981) Constraints on fossil fuel use, in Bach, W., Pankrath,
J. and Williams, J. (eds) Interactions of Energy and Climate, Dordrecht, D. Reidel, 191-212.

Schindler, D. W. (1977) Evolution of phosphorus limitation in lakes, Science, 195, 260-262.

Seiler, W., and Crutzen, P. J. (1980) Estimates of gross and net fluxes of carbon between the
biosphere and atmosphere from biomass burning, Climatic Change, 2, 207-248.

Singh, H. B., Salas, L. J., Shigeishi, H., Smith, A. J., Scribner, E., and Cavanaugh, L. A.
(1979) Atmospheric Distributions, Sources and Sinks of Selected Hydrocarbons, SF6 and
N2O. Report EPA-600/3-79-107, Springfield, Virginia, USA, NTIS.

Söderlund, R., and Svensson, B. H. (1976) The global nitrogen cycle, in Svensson, B. H., and
Söderlund, R. (eds) Nitrogen, Phosphorus and SulphurGlobal Cycles, SCOPE Report No. 7,
Ecol. Bull. (Stockholm), 22, 23-73.

Stewart, J. W. B., Cole, C. V., and Maynard, D. G. Interactions of biogeochemical cycles in


grassland ecosystems (Chapter 8, this volume).
Stoffyn, M. (1979) Biological control of dissolved aluminium in seawater: experimental
evidence, Science, 203, 651-653.

Taylor, G. S., Baker, M. B., and Charlson, R. J. Atmospheric interactions of C, N and S


Cycles: The role of aerosols and clouds (Chapter 4, this volume).

Thompson, S. L., and Schneider, S. H. (1981) Carbon dioxide and climate: ice and ocean,
Nature, 290, 9-10.

Torres, A. L., Maroulis, P. J., and Bandy, A. R. (1980) Atmospheric OCS measurements on
Project GAMETAG, J. Geophys. Res., 85, 7357-7360.

Turco, R. P., Whitten, R. C., Toon, O. B., Pollack, J. B., and Hamill, P. (1980) OCS,
stratospheric aerosols and climate, Nature, 283, 283-286.

Ulrich, B., Mayer, R., and Khanna, P. K. (1980) Chemical changes due to acid precipitation
in a loess-derived soil in central Europe, Soil Sci., 130, 193-199.

United Nations (1977) Desertification: its causes and consequences, U.N. Conference on
Desertification, Nairobi, Kenya, 29 Aug.9 Sept. 1977. Oxford, Pergamon Press.

Viecelli, J. A., Ellsaesser, H. W., and Burt, J. E. (1981) A carbon cycle model with latitude
dependence, Climatic Change, 3, 281-302.

Viezee, W., and Singh, H. B. (1980) The distribution of Beryllium-7 in the troposphere:
implication on stratospheric-tropospheric air exchange, Geophys. Res. Lett, 7, 805-808.

Vitousek, P. M. The effects on deforestation on air, soil, and water (Chapter 7, this volume).

Walker, T. W., and Syers, J. K. (1976) The fate of phosphorus during pedogenesis,
Geoderma, 15, 1-19.

Wang, W. C., Yung, Y. L., Lacis, A. A., Mo, T., and Hansen, J. E. (1976) Greenhouse effects
due to man-made perturbations of trace gases, Science, 194, 685-690.

Watson, J. G. (1979) Chemical element balance receptor model methodology for assessing
the sources of fine and total suspended particulate matter in Portland, Oregon, Ph.D.
Dissertation, Oregon Graduate Center, Beaverton, Oregon.

Weiss, R. F. (1981) The temporal and spatial distribution of tropospheric nitrous oxide, J.
Geophys. Res., 86, 7185-7195.

White, P. S. (1979) Pattern, process, and natural disturbance in vegetation, Bot. Rev., 45,229-
299.

Whitten, R. C., Toon, O. B., and Turco, R. P. (1980) The stratospheric sulphate aerosol layer:
processes, models, observations, and simulations, PAGEOPH, 118, 87-127.

Wollast, R. Interactions in estuaries and coastal waters (Chapter 14, this volume).
WMO (1981) On the Assessment of the Role of CO2 on Climate Variations and their Impact.
Geneva, WMO.

WMO/NASA (1982) The Stratosphere 1981: Theory and Measurements. Hudson, R. D.,
Reed, E. I, and Bojkov, R. D. (eds), Geneva, WMO.

Back to Table of Contents

The electronic version of this publication has been prepared at


the M S Swaminathan Research Foundation, Chennai, India.

SCOPE 21 -The Major Biogeochemical Cycles and Their


Interactions
2 C, N, P, and S Cycles: Major Reservoirs and Fluxes

2.1 Introduction
2.2 The Carbon Cycle
2.3 The Nitrogen Cycle
2.3.1 General
2.3.2 Distribution of Nitrogen
2.3.3 Fluxes of Nitrogen
2.4 The Phosphorus Cycle
2.5 The Sulphur Cycle
References

2.1 INTRODUCTION

The present chapter brings together the most recent information on reservoir sizes and fluxes
for the four major cycles with which we are concerned. Internal consistency between
different cycles has been aimed for to facilitate comparisons. Variability between different
estimates as quoted represents uncertainty in our knowledge about the relevant reservoirs and
processes.

The tables and diagrams given are not the result of re-assessment of the different cycles, but
should rather be considered as reference material that might facilitate the reading of the
remainder of the book. References should therefore be made to the original data sources
instead of the present chapter.

This overview was not available in its present form at the time of the workshop and some of
the authors of the following chapters do not explicitly refer to the present compilation but
rather to the original research on which it was based. It is judged that this synthesis still will
be of value.
2.2 THE CARBON CYCLE

BERT BOLIN

The global carbon cycle has been described in detail in SCOPE Report 13 (Bolin et al, 1979),
where references to earlier attempts to present a consistent picture of this cycle are also given.
Some new data have become available during the last few years, but they hardly warrant a
more critical re-analysis of earlier views. There are, however, uncertainties and
inconsistancies in that budgets do not always balance and further work is required to resolve
some of these problems. It should also be emphasized that we need to develop methods of
how to describe the carbon cycle in more detail, i.e. the way the transfer of carbon within the
ocean is accomplished or the cycling of carbon within the various main biomes, including the
soil. This is dealt with in several of the following chapters. The objective of this note is to
give a simple view of the overall transfers that govern the global carbon cycle. We shall thus
merely give a synthesis of the data given in SCOPE 13, with some minor modifications to
serve as background for the discussions of biogeochemical interactions.

Table 2.1 gives the distribution of carbon between the major reservoirs. For the atmosphere a
few different values are given, referring to the likely pre-industrial content of the atmosphere,
the accurately known values for 1959, when careful measurements began and the most recent
data from 1980.

Carbon reservoirs of the oceans are also reasonably well known, while there are considerable
uncertainties about the size of the carbon reservoirs in the terrestrial biosphere, the
pedosphere and the lithosphere. In some cases different estimates are quoted and a best
estimate finally chosen. Data for the lithosphere are not accurate to more than one significant
figure, even though sometimes two are given in the table. The values depend on the use of
proper values for the average amount of carbon in different rocks, estimates of their relative
occurrence and the thickness of the crust.

Table 2.2 summarizes in a similar manner the magnitude of the fluxes between the different
reservoirs. Again a best estimate is given based on the determinations quoted.

Figures 2.1 and 2.2 display in schematic form the carbon cycle. The exchanges between the
atmosphere, the terrestrial biomass, soil, rivers and the oceans including marine life as shown
in Figure 2.1 are relatively large compared with the reservoirs involved. Characteristic turn-
over times vary from a few weeks for marine biota to about one thousand years for the ocean.
The circulation in the rock cycle involves reservoirs that are one to ten million times larger.
Since the fluxes are one or two orders of magnitude less than those at work in air and ocean
transport and in photosynthesis, the characteristic turn-over times are hundreds of millions of
years.

Table 2.1 Major carbon reservoirs (when not otherwise indicated, in units of 1015 g C =
PgC). References are given after Table 2.2

Atmosphere Reference
Carbon dioxide
Pre-industrial 265 ppm 560 (1)
290 ppm 615 (1)
1959 316 ppm 670 (2)
1980 336 ppm 712 (2)
Methane 1.41 ppm 3 (3)
Carbon monoxide 0.11 0 .2 (3)
Other C-containing gases 0 .05 (3)

Atmosphere,
total (1980) 715

Oceans

Dissolved inorganic
37,400 (4)
carbon
Dissolved organic carbon 1,000 (5)
Particulate organic carbon 30 (5)
Biota 3 (6)

Ocean, total 38,500

Terrestrial biota and


pedosphere

Phytomass short-lived 130 830 (7)


long-lived 700 560 (1)
Animals 12 (3)
Man 0.03 (8)
Bacteria 2 (8)
Fungi 1 (8)
Standing dead organic
30 (8)
carbon
Litter 60 (8)
Peat (very uncertain) 160 (1),(8)
Soil organic carbon
1,500 (8),(9)
(excluding peat)
Organic carbon
in
biota and pedo-
sphere, total 2,3002,600

Lithosphere

Continental crust
26 x
Sediments, carbonate (10)
106
Sediments, non- 10 x
(10)
carbonate* 106
Igeneous rock, non- 79 x
(10)
carbonate 106
Igeneous rock, non- 1.1 x
(10)
carbonate 106
Oceanic crust
14 x
Sediments, carbonate (10)
106
Sediments, non- 6.0 x
(10)
carbonate* 106
0.3 x
Basalt, carbonate (10)
106
Basalt, non-carbonate 0.3 x (10)
106
65 x
Lithosphere, total
106

*These reservoirs contain the oil, gas and coal reserves of which 510 x 103 units (510 x
1018 Pg) can possibly be recovered for use.

Table 2.2 Major carbon fluxes (when not otherwise indicated in units of 1015 g C yr-1 = Pg
yr-1)

Atmosphere, terrestrial biosphere Reference

Carbon monoxide production 1.2 ± 0.6 (3)


Methane production 0.6+0.3 (3)
Photosynthesis 53 (7)
60 (8)
Litterfall 50 ± 10 (8)
Grass and forest fires (natural;
24 (4, 11)
gross)
Animal (herbitore) consumption 6 (8)
Decay of organic matter =
photosynthesis
Fluxes induced by man due to:
Fossil fuel combustion 1980 5.2 (12)
Fossil fuel combustion
165 (12)
18601980
1.0 ±
Deforestation 1970 (net)
0.5
2.5 (13)
1.5 ±
Agriculture 1970 (net)
1.0
Deforestation, agriculture
150 (13)
18601970

Atmospheresea

60 ± 15 mmole/m2
Air-sea exchange, average rate (4)
(ppm)yr
1860 CO2 concentration 290 ppm 75 ± 20 (4)
1980 CO2 concentration 336 ppm 90 ± 20 (4)
Net flux atmosphere to sea, 1980 2.5 (4)

Oceans, marine biota

Primary production 40 ± 10 (6)


Detritus fall out from surface
layer(~ 75 m) 4 (1),(6)
Catch of fish 0.006 (6)

Freshwater, oceans, lithosphere

River flux, inorganic, dissolved 0.5 (14)


River flux, inorganic, particulate 0.2 (14)
River flux, organic, dissolved 0.30.7 (15)
River flux, organic, particulate 0.20.5 (15)
Sedimentation in continental
0.05 (13)
basins
Weathering on land 0.3 (10)
Glacial erosion 0.03 (10)
Marine Erosion 0.005 (10)
Metamorphosis (sediments
igneous rock) 0.008 (10)
Subduction (marine sediments
igneous rock) 0.3 (10)
Sedimentation in sea, inorganic 0.15 (10)
Sedimentation in sea, organic 0.04 (10)

(1) Bolin et al. (1979) (8) Ajtay et al. (1979)


(2) Bacastow and (9) Schlesinger (1984)
Keeling (1981)
(3) Freyer (1979) (10) Kempe (1979a,b)
(4) Bolin et al. (1981) (11) Seiler and Crutzen (1980)
(5) Mopper and Degens (12) Rotty (1981)
(1979)
(6) De Vooys (1979) (13) Moore et al. (1981)
(7) Whittaker and (14) Kempe (1979a)
Likens (1975)
(15) Meybeck (1981)

Figure 2.1 Size of reservoirs (in 1015 g) and fluxes (in 1015 g yr-1) for the part of the carbon
cycle that is in a state of comparatively rapid turn-over, i.e. characteristic turn-over times less
than about 1000 years

Figure 2.2 Fluxes (in 1015 g yr-1) for the carbon cycle in the earth's crust, where the
characteristic turn-over times are of the order of 100 million years (based on Kempe, 1979a,
b)

2.3 THE NITROGEN CYCLE

THOMAS ROSSWALL

2.3.1 General

During the past decade there has been a rapid development towards a qualitative
understanding of the global nitrogen cycle. We are, however, still far from being able to
present a quantitative picture of the global nitrogen cycle. This inability is due to two main
reasons. The first one is inherent in all attempts to construct models of the annual transport of
elements between different reservoirs. In most cases the reported annual flows do not
represent truly integrated values but are based on extrapolations to a global, yearly basis of
determinations of flow rates at specific points in time and space. Secondly, there are still
lacunes in our knowledge of the qualitative aspects of the biogeochemical nitrogen cycle, and
certain previously neglected processes seem to be of major importance on a global scale.
Examples of this are the recent rediscovery of the importance of nitrification in the
production of nitrous oxide (Bremner and Blackmer, 1978; Cohen and Gordon, 1979;
Crutzen, Chapter 3, this volume) and the possible production of nitric oxide also during
nitrification (Lipschulz et al., 1981).

Figure 2.3 is a schematic representation of the biogeochemical nitrogen cycle. Unlike that of,
for example, phosphorus, the atmosphere plays an important role in the biogeochemical
nitrogen cycle on account of the importance of gaseous compounds (N2, N2O, NO, NH3) all
of which can be produced and consumed through biotic and abiotic processes. A more
accurate understanding of the biogeochemical nitrogen cycle has become of interest in recent
years as a result of the importance of processes leading to the production of nitrogen oxides
in view of the importance of such gases in the regulation of the chemical composition of the
atmosphere.
2.3.2 Distribution of Nitrogen

The global distribution of nitrogen is shown in Table 2.3. Although nitrogen is present in
relatively large concentrations in rocks, sediments, and the atmosphere, its availability in
compounds that can be utilized by most forms of life is severely restricted. This deficiency in
biologically available nitrogen in terrestrial and aquatic systems makes nitrogen one of the
most important limiting nutrients.

In the atmosphere a minute fraction of nitrogen occurs in forms other than N2. The
quantitatively most important form of combined nitrogen in the atmosphere is nitrous oxide
(N2O), which accounts for 99.5% of all combined nitrogen. The small amounts of nitrogen
compounds occuring in the atmosphere, do however, play a major role in regulating major
processes in the atmosphere (see Crutzen, Chapter 3, this volume).

Figure 2.3 A global nitrogen cycle. Units are in Tg (1012g) N yr-1. From Söderlund &
Rosswall (1982) based on Söderlund & Svensson (1976)

In the oceans, dimolecular nitrogen (in dissolved form) is also the most abundant form of
nitrogen. Nitrate and nitrogen in dead organic matter occur in approximately equal amounts.
Biomass nitrogen accounts for less than 0.001% of the total amounts of nitrogen in the
hydrosphere. The ratio of plant: animal: microbial biomass-N (Table 2.3) is 15:8.5:1, which
differs markedly from the ratios found in terrestrial biomass (25:0.4:1) in that animals
(mainly zooplankton) contain an appreciable reservoir of nitrogen as compared to what is
found in terrestrial systems.

The nitrogen in the lithosphere is inaccessible to living organisms except man, who, through
the burning of coal, will release some of the bound nitrogen into the atmosphere.

Nitrogen in terrestrial systems occurs mainly in soil organic matter, litter and soil inorganic
nitrogen (97% of total) with biomass accounting for only less than 3%. In the biomass, 95%
occurs in the plants.

Table 2.3 Major nitrogen reservoirs (in units of 1015 g N = Pg N)

Atmosphere % of total Reference

Dimolecular nitrogen 3 900 000 > 99.999 (1)


Nitrous oxide 1.4 < 0.0001 (2)
Ammonia 0.0017 < 0.0001 (2)
Ammonium 0.00004 < 0.0001 (2)
Nitric oxide + Nitrogen dioxide
0.0006 < 0.0001 (2)
(NOx)
Nitrate 0.0001 < 0.0001 (2)
Organic nitrogen 0.001 < 0.0001 (3)
Total 100

Ocean

Plant biomass 0.30 0.001 (3)


Animal biomass 0.17 0.0007 (4)
Microbial biomass 0.02 0.00006 (5)
Dead organic matter (dissolved) 530 2.3 (3)
Dead organic matter (particulate) 3240 0.010.1 (3)
Dimolecular nitrogen (dissolved) 22 000 95.2 (4)
Nitrous oxide 0.2 0.009 (3)
Nitrate 570 2.5 (6)
Nitrite 0.5 0.002 (3)
Ammonium 7 0.03 (3)
Total 100

Pedosphere including biota

Plant biomass 1114 2.6 (3)


Animal biomass 0.2 0.04 (4)
Microbial biomass 0.5 0.1 (3)
Litter 1.93.3 0.5 (3)
Soil: organic matter 300 63 (3)
inorganic 160 34 (7)

Total 100

Lithosphere

Rocks 190 000 000 99.8 (8)


Sediments 400 000 0.2 (8)
Coal deposits 120 0.00006 (9)
Total 100

(1) Robinson and Robbins (1970)


(2) Galbally and Roy (in
manuscript)
(3) Söderlund and Svensson (1976)
(4) Delwiche (1970)
(5) This compilation (Based on Mopper and Degens (1979) and a C/N
ratio in micro-organisms of 12.5.)
(6) Emery et al. (1955)
(7) Delwiche and Likens (1977)
(8) Stevenson (1965)
(9) Donald (1960)

Simpson et al. (1977) made a compilation of data on nitrogen reservoir sizes from four
different publications. Although the data in that compilation agree well with those in Table
2.3, there are estimates which differ by an order of magnitude. Galbally and Roy (in
manuscript) and Sweeney et al. (1977), for example, estimated the atmospheric content of
ammonia to be 1.7 and 1.5 Tg, respectively, while Delwiche and Likens (1977) gave a value
of 28 Tg. The similarity between many of the published data is, however, dependent on the
use of the same original source by many authors. In several instances the originally cited
sources do not present the basis for the calculations and the correctness cannot be assessed.
Unfortunately, there are few well documented reports on the distribution of nitrogen in
different reservoirs.

2.3.3 Fluxes of Nitrogen

There is considerably uncertainty with regard to most estimates of process rates in the global
biogeochemical nitrogen cycle (Table 2.4). Most estimates made in the past 10 years differ by
an order of magnitude between lowest and highest values. In addition, the uncertainty is
further increased because previously unidentified processes in the biogeochemical nitrogen
cycle probably exist. Examples of such new processes are the mesospheric source of N2O
from excited N2 as suggested by Zipf and Prasad (1982) and the possible production of NO
during nitrification (Lipschulz et al., 1981). A further quantification of natural fluxes of
nitrogen compounds between the oceanic-terrestrial systems and the atmosphere is needed in
order for reliable atmospheric models involving nitrogen compounds to be developed. Here
we only consider the exchanges of nitrogen compounds between the terrestrial, aquatic and
atmospheric systems, while the internal transfers are not discussed. It should be realized that,
for example, for the terrestrial systems, the internal transfers of nitrogen in the soilplant
subsystem are one to two orders of magnitude larger than the transfers to and from the
terrestrial systems (Rosswall, 1976).

Man is increasingly affecting the global biogeochemical nitrogen cycle. The industrial
production of nitrogen fertilizer will, towards the end of this century, probably become as
large as that produced through biological nitrogen fixation in the global terrestrial ecosystem
(Söderlund and Svensson, 1976). Increased combustion temperatures increase the production
of NOx, which contributes to the acidification of rain water. If we wish to determine the
possible implications of such increased amounts of fixed nitrogen for the global cycles, it is
essential that they are evaluated against background knowledge of the amounts of nitrogen
that are parts of the natural biogeochemical nitrogen cycle. It should be evident from the data
cited in Tables 2.3 and 2.4 that we are still far from having such a quantitative knowledge.

Table 2.4 Fluxes of nitrogen (Tg N yr-1) in the global biogeochemical nitrogen cycle. The
ranges summarize rates given by the following authors: Delwiche (1970), Burns and Hardy
(1975), Söderlund and Svensson (1976), McElroy et al. (1976), CAST (1976), Delwiche and
Likens (1977), Liu et al. (1977), Hahn and Junge (1977), Sweeney et al. (1977), NAS (1978)
and Bolin (1979). For a compilation of the individual estimates, see Rosswall (1981)
Tg
N

Biological nitrogen fixation: land 44200


Biological nitrogen fixation: ocean 1130
Atmospheric fixation (lightning) 0.530
Industrial fixation 60 (1)
Industrial combustion and fossil fuel burning
(NOx, N2O) 10-20 (2)
Fires 10200
Biogenic NOx production 090 (3)
Denitrification: land 43390
Denitrification: ocean 0330
N2O from denitrification: land 1669 (4)
N2O from denitrification: ocean 580 (5)
Nitrification N2O production: land ?
Nitrification N2O production: ocean 410 (6)
Ammonia volatilization 36250
+
Dry and wet deposition NH3/NH4 110240
Dry and wet deposition NOx 40116
Dry and wet deposition organic-N 10100
River run-off (total N) 1340

(1) Data for 1979/80 (UN, 1981)


(2) Crutzen (Chapter 3, this volume)
(3) The lowest value refers to the estimated of NOx production
from soils of
0.15 Tg given by Crutzen (Chapter 3, this volume).
(4) Söderlund and Svensson (1976).
(5) The lower estimate comes from Crutzen (Chapter 3, this
volume) and
the upper from Söderlund and Svensson (1976).
(6) Cohen and Gordon (1979).

2.4 THE PHOSPHORUS CYCLE JEFFREY E. RICHEY

Previous evaluations of the global P cycle have identified the key fluxes and reservoirs, and
have provided some estimates of their relative magnitudes (Stumm,1973; Lerman et al.,
1975; Pierrou, 1976). A revision is provided here (Table 2.5; Figure 2.4). The use of
phosphorus fertilizers in modern agriculture and the eutrophication of fresh waters from run-
off and effluent discharge are the most visible results of human intervention in the P cycle.

The intent of this work is to summarize our current understanding of the dynamics of the
phosphorus cycle by describing some of the chemical properties of P that affect its
distribution, reviewing the assumptions used in the calculation of earlier P budgets and,
where possible, up-dating them with new data (Table 2.5). It is becoming increasingly
evident that element cycles cannot only be viewed in their global aggregate, but must be
analysed on more relevant space scales. Therefore, the continental portion of the P cycle is
analysed with finer resolution for 10 major geopolitical zones (Table 2.6).

Phosphate is liberated into the environment by the weathering of the apatite rocks. Because
phosphate tends to precipitate to form materials of low solubility, sorb onto surfaces, and
form complexes with metal ions, much of the phosphate released by weathering is
immobilized (Van Wazer, 1973). As a result, of the total soil P, free phosphate is often
present in only trace quantities, with little leaching into fresh waters.

Phosphorus is present in the biota in a wide variety of organic compounds. These organic
compounds are characterized by either fairly weak POC ester bonds or stable PC bonds, and
undergo the hydrolytic degradation of esters and condensed polyphosphates. With these
properties, P is critical as a mobile entity of cell metabolism and as a basic structural element
of cell materials. Other than as an intermediate, phosphate does not participate in
reductionoxidation reactions, as do C, N, and S. Because the ambient concentrations are low
and the demands specific, phosphate becomes an important element for primary production in
both terrestrial and aquatic environments.

The terrestrial biota contains much less P than do the source rocks and soils, with the largest
reservoirs in the forests of North America, the U.S.S.R., Latin America, and Tropical Africa
(450560 Tg P). Although the total dissolved reservoir in the oceans is about 77,000 Tg P,
only about 50120 Tg P are contained in marine biota. This is due in large part to much of the
P being below the euphotic zone.

The primary reservoirs and fluxes of P involve dissolved phosphate ion (PO43), dissolved
organic P (DOP), and particulate inorganic P. The bulk of the P exists in the soil, marine
sediments, and, of course, crustal rocks as apatite. Where the concentration of apatite is great
enough, it can be commercially mined. The soil fraction is distributed approximately in
proportion to the area of a region, ranging from 5900 Tg P in Europe to 29,400 Tg P in
Tropical Africa, whereas the mineable rock is concentrated in North America (4,700 Tg P)
and Tropical Africa (5,900 Tg P). The particulate soil fractions can be mobilized by erosional
processes into rivers and subsequently to the oceans, and is estimated to be about 17 Tg P/yr.
A lesser amount, about 4 Tg P/yr, is carried by the wind into the atmosphere; however, the
residence time there is very short, and the atmospheric reservoir is only about 0.025 Tg P.
The P cycle does not have an important atmospheric gaseous component, unlike C, N, or S.

Table 2.5 The major reservoirs and fluxes of the global phosphorus cycle. Three previous
evaluations are compared, up-dated with new information, where possible, and a current
estimate derived (summarized in Figure 2.4). The means of calculations and sources of each
estimate are provided below. It must be remembered that all such calculations yield a value
with considerable uncertainty, not a precise number. It is important to examine the underlying
assumptions and sources of error for each term as given in the original reference

Lerman
Stumm Pierrou
RESERVOIRS (Tg P) et al. Up-date Reference
(1973) (1976)
(1975)
Atmosphere
Particulates over land 0.025 (1)
Particulates over oceans 0.003 (1)
Land
Biota 1950 3,000 1,805 2,600 (2)
Soil 200,000 160,000 96,000160,000 (3)
Mineable rock 31,000 9,920 3,1409,000 19,000 (4)
Fresh-water (dissolved) 90 90 (5)
Ocean
Biota 124 138 128 50120 (6)
Dissolved (inorganic) 124,000 92,600 120128,000 80,000 (7)
Detritus (particulates) 650 (8)
Sediments 8.4 x 108 4 x 108 840,000,000 (9)

FLUXES (Tg P/yr)

Atmosphere Atmosphere
2 1.0 (1)
(land) (ocean)
Atmosphere Atmosphere
0.3
(land) (ocean)
Atmosphere Land 3.79.3 3.2 (1)
Atmosphere Ocean 2.63.5 1.4 (1)
Land Atmosphere ? 4.3 (1)
Ocean Atmosphere ? 0.3 (1)
Marine
Biota 961 9921042 9901,300 6001,000 (10)
dissolved
Marine detritus Sediment 1.9 1.7 13 213 (11)
Terrestrial
Soils 229 63.6 136237 200 (12)
biota
Mineable rock Soil 12.4 12.4 12.6 14 (14)
Soil Fresh-water 2.512.3 47 (13)
Fresh-water
Oceans 1.9 1.7 1.54 (14)
(diss.)
Fresh-water
Oceans 17.4 17 (14)
(part.)

(1) The atmospheric reservoir and fluxes are directly from Graham and Duce (1979), who
summarized extensive measurements of atmospheric P concentrations and deposition
rates in marine and continental regions.
(2) Estimates of the terrestrial biota are generally calculated from estimates of C mass of
4.5 x 105 Tg C (Bolin, 1979) 8.3 x 105 Tg C (Whittaker and Likens, 1975), and C/P
atomic ratios of 500 (Stumm, 1973) to 833 (Deevey, 1970), though Pierrou (1976) used
dry-weight biomass conversions. These yield a most-likely value of 2,600 Tg P, with a
range of 1,4004,300 Tg P.
(3) Inorganic P in the soil has been computed from a total land area of 130 x 1012 m2 and a
soil depth of 0.6 m (Lerman et al. 1975) to 1.0 m (Pierrou, 1976), with a P content of
0.100.12 % (Taylor, 1964) and a soil density of 1 kg/dm3.
In their calculation of soil volume, Lerman et al. apparently divided by 0.6 m rather
than multiplying, yielding an over-estimate. Total soil P, including about 10% organic P
(Bohn, 1976), is thus 96,000160,000 Tg P, depending on soil depth.
(4) The amount of P in mineable rock, as 30% of P2O5, has been defined as that minimum
amount which is economically recoverable. As demand and technology increase, the P
content of mineable rock decreases and the reservoir `increases'. The estimates of
reservoir size and consumption rate are from Harris and Hare (1979) and FAO (1980).
Pierrou (1976) calculated the P in fresh-waters from a total volume of 7.2 x 1014 m3 and
(5) a mean concentration of 0.12 g P/m3. The concentration value is probably uncertain by a
factor of 2.
(6) Phosphorus in the marine biota is calculated generally from applying the Redfield ratio
of C/P = 106 and carbon estimates of 2,000 Tg C (Williams, 1975) to 5,000 Tg C
(Bolin, 1979). The upper estimate of Lerman et al. (1975) is based on out-dated
production data.
(7) The previous estimates of dissolved inorganic P in the oceans have used mean
concentrations of 0.080.10 g P/m3 and mean depths of 3,0003,500 m. More recent
GEOSECS data (Takahashi et al. 1981) suggest a concentration of 0.062 g P/m3.
(8) The inventory of marine detrital P can be calculated as the amount of particulate carbon
(3 x 104 Tg C; Mopper and Degens, 1978) times a detrital C/P atomic ratio of 120
(Broecker, 1974), or 650 Tg P.
(9) Stumm (1973) calculated the phosphorus in sediments from a geochemical mass
balance, which is based on more recent data than the value presented in Lerman et al.
(1975).
(10) Estimates of the photosynthetic uptake by marine biota have been obtained by applying
the Redfield ratio to productivity data, which range from 2.5 x 104 Tg C/yr to 4 x 104 Tg
C/yr (De Vooys, 1979), or 6001000 Tg P/yr. The release of P by decomposition is
assumed to be equal to photosynthetic uptake.
(11) Emery et al. (1955) assumed that the P content of sediments is 0.092% and that 1 cm of
solid sediment forms every 6000 years, for a rate of 13 Tg P/yr. Assumptions of steady
state, as calculated by the others, indicate that this figure might be high, and that a value
of 2 Tg/yr is more appropriate.
(12) The uptake and release of P by terrestrial biota has been estimated from productivity
estimates of 3 x 104 C/yr (Bolin, 1979) to 5 x 104 Tg C/yr (Whittaker and Likens, 1975)
and the C/P atomic ratios of 500822. A mean estimate of 200 Tg P/yr results.
(13) Phosphate is introduced into fresh-waters from natural leaching processes and from
various human activities it is difficult to differentiate between the sources.
Pierrou used leaching rates from respective land-use types to estimate the total input. A
similar analysis using the land-use types specified for the different geographic zones
and up-dated loss rates for the different land uses (forest, grass, desert-swamp, and
agriculture-pasture) from Rast and Lee (1978) gives the result 4 Tg P/yr.
(14) The river run-off of P to the oceans includes both natural and human-influenced
leaching and particulate erosion products, less that which is retained or consumed
within the river. The most recent calculation of dissolved export is described in Wollast
(Chapter 14, this volume). Particulate export, assuming a ratio of 0.075% for the P
content of total suspended sediments (e.g. Holland, 1978), is considerably greater.

Figure 2.4 A global phosphorus cycle. Fluxes are in Tg P yr-1 and reservoirs are in Tg P.
From Table 2.5

The uptake and release of phosphate by terrestrial plants is about 200 Tg P/yr, while the
equivalent marine flux is 6001000 Tg P/yr. Given the lesser marine biomass, the turnover rate
in the oceans is much greater than on land. The amount finally sedimenting in the oceans is
only a small fraction of the productionmineralization flux.

The natural cycle of P partly regulates the distribution of biomass because the supply and
levels of phosphate are low relative to the requirements for plant and animal nutrition.
Superimposed on the natural cycle is man's influence: the mining and consuming of
phosphates by society, and the release of P in domestic and industrial effluents.

Table 2.6 The distribution of the terrestrial and fluvial components of the global P cycle
(Table 2.5) according to major geopolitical zones. Means of calculation are described in the
footnotes

N. L. N. Tr.
Europ U.S.S. Pac. S. S.E. Referenc
Americ China Americ Afric Afric
e R Dev. Asia Asia e
a a a ME a
RESERVOIRS(Tg
P)
Biota 450 110 560 170 80 450 20 540 70 140 (1)
10,70 13,40 14,40 29,40 7,00 4,80
Soil 23,000 5,900 26,800 24,600 (2)
0 0 0 0 0 0
Mineable rock 4,700 50 980 280 1,600 1,260 2,900 5,900 800 530 (3)
FLUXES (TgP/yr)
Biota Soil 28 7 35 10 5 27 2 33 4 9 (1)
Rock Soil 2.5 4.0 2.3 1.6 0.8 1.0 0.4 0.2 0.6 0.2 (3)
Soil Freshwa 0.3- 0.4- 0.2- 0.6- 12- 0.4-
0.7 0.3 1.0-1.8 0.5-0.8 (4)
ter 0.6 0.6 0.3 1.0 0.3 0.7
0.2- 0.04- 0.04- 0.05- 0.1- 0.02
Freshwater 0.20.6 0.6-1.5 0.1-0.5 0.01 (5)
0.5 0. 0.1 0.2 0.4 -0.1
(1) The distribution of terrestrial biota and the uptake and release of P was calculated by
determining the percent vegetation type by region
and weighting that by kg C/m2 from Whittaker and Likens (1975).
(2) The soil P distribution by region was apportioned by the percent land surface in each
region.
(3) Mineable rock and fertilizer consumption were calculated by grouping the country-
specific and regional data of Harris and Hare (1979) and
FAO (1980) to the appropriate zones.
(4) The relative land cover in each zone was calculated, and the run-off ratios from Rast and
Lee (1978) applied.
(5) Calculated by river zone by establishing the range of P concentrations for the rivers of
that zone time their total water discharge.

The annual rate of fertilizer consumption in the industrialized regions is now about 10% of
the steady-state flux between the soil and biota, and approaches 50% in Europe. In the less-
developed regions the rate is lower. The total input of P to fresh-waters which also includes
wastes, is a considerable fraction of the fertilizer consumption. This suggests, that the
mobility of this `excess' P is sufficient to be transported away from its sites of application.
Most of this phosphorus is discharged into coastal waters.

Although adequate data on the pre-industrialization loads of P in rivers do not exist, there has
been at least a several-fold increase (Wollast, Chapter 14, this volume). The eutrophication of
lakes and coastal waters is well-known. This phenomenon also alters the cycles the C and N.
Given the above observations, the P cycle is thus significantly perturbed by the activities of
man. Figure 2.4 gives an overview of the reservoirs and fluxes of the global P cycle.

2.5 THE SULPHUR CYCLE

J. R. FRENEY, M. V. IVANOV AND H. RODHE

A number of attempts have been made over the last 20 years to assess the Earth's reservoirs
of sulphur and the natural and man-made transfers to determine how seriously the balance of
nature was being affected (Eriksson, 1960; Junge, 1963; Robinson and Robbins, 1970;
Kellogg et al., 1972; Friend, 1973; Cadle, 1975; Granat et al., 1976; Cullis and Hirschler,
1980; Ivanov and Freney, 1983). These reports show that man-made emissions are large and
seriously interfere with the natural sulphur cycle. As the sulphur cycle clearly interacts with
those of carbon, nitrogen and phosphorus, the main features of these reports are summarized
and discussed to serve as background information for the study on cycle interactions.

The most recent of these reports (Ivanov and Freney, 1983) includes the results from work
done in the Soviet Union that has not previously been readily available. This report provides
up-to-date information on many fluxes and attempts to separate, wherever feasible, the
natural and man-made contributions to these fluxes. The following summary is largely based
on it. In order to make the presentation short, no indications of the uncertainties appropriate
to the different values are included; this does not imply that the values are accurately known.
For a further discussion of this aspect the reader is referred to the original work.

Table 2.7 summarizes the current information on the distribution of sulphur between the
various spheres. The bulk of sulphur is contained in the lithosphere (24 x 109 Tg S) and the
hydrosphere (1.3 x 109 Tg S), moderate amounts occur in the pedosphere and only small
amounts (4.6 Tg S) are found in the atmosphere at any one time.

Table 2.7 Major sulphur reservoirs (Tg s)

Atmosphere(1) Tg S

Troposphere
Sulphate in aerosols 0.7
Sulphur dioxide 0.5
Carbonyl sulphide 2.3
Other reduced sulphur gases 0.8
Total in troposphere 4.3
Stratosphere 0.5

4.8

Hydrosphere (2)

Ocean water 1.3 x 109

1.3 x 109

Lithosphere (3)

Continental and subcontinental


Sedimentary 5.2 x 109
Granite 7.8 x 109
Basalt 8.8 x 109
Oceanic
Sedimentary (layer I) 0.3 x 109
Tholeitic and olivine basalt
0.6 x 109
(layer II)
Layer III 1.6 x 109

24.3 x 109

Pedosphere

Soil(4) 2.6 x 105


Soil organic matter (5) 1.1 x 104
Land plants(4) 760

2.7 x 105

(1) Ryaboshapko (1983)


(2) Volkov and Rozanov (1983)
(3) The lithosphere refers to the crust of the Earth. Migdisov et al. (1983)
(4) Krauskopf (1967)
(5) Freney and Williams (1983)

Within the lithosphere most of the sulphur occurs in rocks of the present continents and sub-
continents (22 x 109 Tg S) and the major forms are metal sulphides and sulphates. The
maximum concentration of reduced sulphur is found in argillaceous rocks of platform areas
(0.4% S) and the minimum concentration occurs in effusive rocks (0.04% S). Sulphate
sulphur varies from fractions of one per cent in humic pelagic formations to solid sulphate
layers in evaporites (Migdisov et al., 1983).

Figure 2.5 The major natural and anthropogenic (circled) fluxes of the global
biogeochemical sulphur cycle (Tg S/year) (from Ivanov and Freney 1983)

Roman figures denote:

I output of sulphur-containing minerals;


II industrial treatment of the sulphur-containing raw materials;
III inland water bodies;
IV volcanoes.

Fluxes:

P1 fuel combustion and metal smelting; P2 and P17 volcanic emission; P3 aeolian dust; P4
biogenic emission from land; P5 sea air-land air transport; P6deposition of large dust particles;
P7 wash-out and dry deposition; P8 land air-sea air transport; P9 weathering; P10 river runoff
to oceans; P1l transport to ocean in underground water; P12 runoff to inland water bodies; P13
input in fertilizers; P14 - leaching of fertilizers; P15 efflux from chemical industries; P16 efflux
of acid mine water; P18 abrasion of shores; P19 - sea spray; P20 wash-out and dry deposition;
P21 sedimentation of reduced sulphur; P22 sedimentation of sulphate; P23 biogenic emission
from oceans.

Table 2.8 Sulphur transfer between the atmosphere and the earth's surface according to
different authors(1) (Tg S/yr)

Robinson Granat Ivanov


Kellogg
Eriksson Junge and Friend et and
et al.
(1960) (1963) Robbins (1973) al. Freney
(1972)
Flux (1970) (1976) (1982)

Combustion, smelting,
39 40 70 50 65 65 113
etc.
Volcanic gases 1.5 2 3 28
Aeolian emission 0.2(2) 20
Biogenic gases from
77 70 68 58 5 16
land
Biogenic gases from
90
coastal regions 190 160 30 48 27 20
and the open ocean
Sea spray 44 44 43 44 44 140
Emission of long-lived
reduced 5
sulphur compounds
Uptake of SO2 by land
surfaces
and vegetation 77 70 26 15 15 28 17
Wash-out over land 55 70 86 86 43 51
Dry deposition of 57
15 20 10 20 16
sulphate over land
Uptake of SO2 by
70 70 25 25 10 11
ocean
Wash-out over ocean 60(3) 72 63 230
146 71 71
Dry deposition of
17
sulphate over ocean

(1) From Ryaboshapko (1983)


(3) Excluding sea salt

In the pedosphere the bulk of the sulphur occurs in organic compounds in soil and in living
plants (1.2 x 104 Tg S) while in the ocean, inorganic sulphate predominates. Carbonyl
sulphide appears to be the dominant form of sulphur in the atmosphere.

From Figure 2.5, the biggest transfers of sulphur occur between: the atmosphere and the
ocean due to wash-out and dry deposition (258 Tg S/yr); the pedosphere and the ocean in run-
off (208 Tg S/yr); the ocean and the atmosphere in sea spray (140 Tg S/yr); the lithosphere
and the hydrosphere by weathering (114 Tg S/yr); and the lithosphere and the atmosphere by
fuel combustion and metal processing (113 Tg S/yr). It should be pointed out that
accumulation of sulphur in the pedosphere has been neglected.

Table 2.8 shows a comparison between different estimates of the global sulphur fluxes
affecting the atmosphere. To a certain degree the differences are due to the better data
available during recent years and also reflect the large uncertainties still associated with many
of the fluxes, and the changes with time in some of the fluxes. For example, there has been a
very substantial rise in the anthropogenic emission of sulphur to the atmosphere between
1950 and the mid-seventies (Figure 2.6). During the latest years the increase has probably
halted. It should be noted that the emission estimates, on which Figure 2.6 is based, are
associated with considerable uncertainties. Recent estimates of sulphur emissions in Europe,
U.S.S.R. and North America (ECE, 1981) give a figure of about 50 Tg S/yr. If this figure
were correct, it seems unlikely that the global total at present would exceed 100 Tg S/yr.

At the present time the anthropogenic input most probably outweighs the natural input to the
atmosphere over land and is approximately equal to the natural transfer of sulphur from the
pedosphere to the oceans in river run-off. The total amount of sulphur in continental water
bodies has nearly doubled due to man's activity.

An even more dramatic anthropogenic impact on the sulphur cycle is evident if the
industrialized regions are considered separately. Within those regions (the eastern part of
North America, Europe and parts of eastern Asia) emissions into the atmosphere have
probably increased by at least one order of magnitude due to man's interference (Galloway
and Whelpdale, 1980).

Due to the limited atmospheric residence time of the emitted sulphur, much of it is deposited
within the same regions. As a result of these emissions, the chemical composition of air and
of precipitation in industrial regions has been greatly affected (Rodhe, 1981). The deposition
of sulphuric (and nitric) acid is now having a significant effect on water quality and maybe
also on soil chemistry in those areas (Cook, Chapter 12, this volume).

Figure 2.6 Estimates of the global emission of sulphur from anthropogenic sources between
1860 and 1980 (Ryaboshapko, 1983)

2.6 REFERENCES

Ajtay, G. L., Ketner, P., and Duvigneaud, P. (1979) Terrestrial primary production and
phytomass, in Bolin, B. et al. (eds) The Global Carbon Cycle, SCOPE Report No. 13,
Chichester, Wiley 129-182.

Bacastow, R. B., and Keeling, D. C. (1981) Atmospheric carbon dioxide concentration and
the observed airborne fraction, in Bolin, B. (ed.) Carbon Cycle Modelling, SCOPE Report
No. 16, Chichester, Wiley, 103-112.

Bohn, H. (1976) Estimates of organic carbon in world soils. Soil Sci. Am. J., 40, 468-470.

Bolin, B. (1979) On the role of the atmosphere in biogeochemical cycles. Quart. J. R. Met.
Soc., 105, 25-42.

Bolin, B., Degens, E. T., Duvigneaud, P., and Kempe, S. (1979) The global biogeochemical
carbon cycle, in Bolin, B., Degens, E. T., Kempe, S. and Ketner, P. (eds) The Global Carbon
Cycle, SCOPE Report No. 13, Chichester, Wiley, 1-56.

Bolin, B., Keeling, C. D., Bacastow, R. B., Björkström, A., and Siegenthaler, U. (1981)
Carbon cycle modelling, in Bolin, B. (ed.) Carbon Cycle Modelling, SCOPE Report No. 16,
Chichester, Wiley, 1-28.

Bremner, J. M., and Blackmer, A. M. (1978) Nitrous oxide: emission from soils during
nitrification of fertilizer nitrogen, Science, 199, 295-296.
Broecker, W. S. (1974) Chemical Oceanography, New York, Harcourt Brace Jovanovich.

Burns, R. C., and Hardy, R. W. F. (1975) Nitrogen Fixation in Bacteria and Higher Plants,
Berlin, Heidelberg, New York, Springer-Verlag.

Cadle, R. D. (1975) Volcanic emissions of halides and sulphur compounds to the troposphere
and stratosphere, J. Geophys. Res., 80, 1650-1652.

CAST (1976) Effect of Increased Nitrogen Fixation on Stratospheric Ozone. Council for
Agricultural Science and Technology Report No. 53, Ames, Iowa, Iowa State University.

Cohen, Y., and Gordon, L. I. (1979) Nitrous oxide production in the ocean, J. Geophys. Res.,
84, 347-354.

Cook, R. B. The impact of acid deposition on the cycles of C, N, P, and S. (Chapter 12, this
volume).

Crutzen, P. J. Atmospheric interactions homogeneous gas reactions of C, N, and S containing


compounds (Chapter 3, this volume).

Cullis, C. F., and Hirschler, M. M. (1980) Atmospheric sulphur: natural and man-made
sources, Atmos. Environ., 14, 1263-1278.

Deevey, E. S. (1970) Mineral cycles, Scient. Amer., 223, 149-158.

Delwiche, C. C. (1970) The nitrogen cycle, Sci. Amer., 223(3), 137-146.

Delwiche, C. C., and Likens, G. E. (1977) Biological response to fossil fuel combustion
products, in Stumm, W. (ed.) Global Chemical Cycles and their Alterations by Man, Berlin,
Dahlem Konferenzen, 73-88.

De Vooys, C. G. N. (1979) Primary production in aquatic environments, in Bolin, B.,


Degens, E. T., Kempe, S., and Ketner, P. (eds) The Global Carbon Cycle, Chichester, Wiley,
259-292.

Donald, C. M. (1960) The impact of cheap nitrogen, J. Austr. Inst. Agric. Sci., 26, 319-338.

ECE (1981) Situation energetique et pollution de l'air dans la CEE. Commission Economique
pour l'Europe, ENV/IEB/R. 13/Add. 1.

Emery, K. O., Orr, W. L., and Rittenberg, S. C. (1955) Nutrients in the ocean, in Essays in
the Honor of Captain Allan Hancock, Los Angeles, University of Southern California Press,
299-309.

Eriksson, E. (1960) The yearly circulation of chloride and sulphur in nature, Tellus, 12, 63-
109.

FAO (1980) FAO fertilizer yearbook statistics, Rome, FAO.


Freney, J. R., and Williams, C. H. (1983) The sulphur cycle in soil, in Ivanov, M. V., and
Freney, J. R. (eds), The Global Biogeochemical Sulphur Cycle, Chichester, Wiley, (In press).

Freyer, H. D. (1979) Atmospheric cycles of trace gases containing carbon, in Bolin, B. et al.
(eds), The Global Carbon Cycle, SCOPE Report No. 13, Chichester, Wiley, 101-128.

Friend, J. P. (1973) The global sulphur cycle, in Rasool, S. I. (ed.), Chemistry of the Lower
Atmosphere, New York, Plenum Press, 177-201.

Galbally, I. E., and Roy, C. R. (In manuscript) The atmospheric nitrogen cycle, in Rosswall,
T. (ed.) The Global Nitrogen Cycle Revisited.

Galloway, J. N., and Whelpdale, D. M. (1980) An atmospheric sulphur budget for eastern
North America, Atmos. Environ., 14, 409-418.

Graham, W. F., and Duce, R. A. (1979) Atmospheric pathways of the phosphorus cycle,
Geochim. Cosmochim. Acta., 43, 1195-1208.

Granat, L., Rodhe, H., and Hallberg, R. O. (1976) The global sulphur cycle, in Svensson, B.
H., and S6derlund, R. (eds) Nitrogen, Phosphorus and SulphurGlobal Cycles. SCOPE Report
No. 7, Ecol. Bull. (Stockholm), 22, 89-134.

Hahn, J., and Junge. C. (1977) Atmospheric nitrous oxide: A critical review, Z. Naturforsch.,
32A, 190-214.

Harris, G. T., and Hare, E. A. (1979) World Fertilizer Situation and Outlook 1978-1985,
Muscle Shoals, Ala., International Fertilizer Development Center.

Holland, H. H. (1978) The Chemistry of the Atmosphere and Ocean, New York, Wiley.

Ivanov, M. V., and Freney, J. R. (1983) The Global Biogeochemical Sulphur Cycle,
Chichester, Wiley (In press).

Junge, C. E. (1963) Air Chemistry and Radioactivity, New York and London, Academic
Press.

Katz, M. (1956) City planning, industrial plant location and air pollution, in Magill, P. L.,
Holden, F. R., and Ackley, C. (eds) Air Pollution Handbook, New York, McGraw Hill.

Kellogg, W. W., Cadle, R. D., Allen, E. R., Lazrus, A. L., and Martell, E. A. (1972) The
sulphur cycle, Science, 175, 587-596.

Kempe, S. (1979a) Carbon in the fresh-water cycle, in Bolin, B. et al. (eds) The Global
Carbon Cycle, SCOPE Report No. 13, Chichester, Wiley, 317-342.

Kempe, S. (1979b) Carbon in the rock cycle, in Bolin, B. et al. (eds) The Global Carbon
Cycle, SCOPE Report No. 13, Chichester, Wiley, 343-378.

Krauskopf, K. B. (1967) Introduction to Geochemistry, New York, McGraw Hill.


Lerman, A., Mackenzie, F. T., and Garrels, R. M. (1975) Modelling of geochemical cycles:
phosphorus as an example, Geol. Soc. Am. Mem., 142, 205-218.

Lipschulz, F., Safirion, O. C., Wofsy, S. C., McElroy, M. B., Valois, F. W., and Watson, S.
W. (1981) Production of NO and N2O by soil nitrifying bacteria, Nature, 294, 641-643.

Liu, S. C., Cicerone, R. J., Donahue, T. M., and Chameides, W. L. (1977) Sources and sinks
of atmospheric N2O and the possible ozone reduction due to industrial fixed nitrogen
fertilizers, Tellus, 29, 251-263.

McElroy, W. B., Elkins, J. W., Wofsy, S. C., and Yung, Y. L. (1976) Sources and sinks for
atmospheric N2O, Rev. Geophys. Space Phys., 14(2), 143-150.

Meybeck, M. (1981) River transport of organic carbon to the ocean. U.S. Dept. Energy.
Carbon Dioxide Effects Res. and Assessment Program, Conf. 8009140 UC-11.

Migdisov, A. A., Grinenko, V. A., and Ronov, A. B. (1982) Cycles of sulphur in the
lithosphere, in Ivanov, M. V., and Freney, J. R. (eds) The Global Biogeochemical Cycle of
Sulphur, Chichester, Wiley (In press).

Moore, B., Boone, R. D., Hobbie, J. E., Houghton, R. A., Melillo, J. M., Petersen, B. R.,
Shaver, G. R., Vörösmarty, C. J., and Woodwell, G. M. (1981) A simple model for analysis
of the role of terrestrial ecosystems in the global carbon budget, in Bolin, B. et al. (eds)
Carbon Cycle Modelling, SCOPE Report No. 16, Chichester, Wiley, 365-386.

Mopper, K., and Degens, E. T. (1979) Organic carbon in the ocean: Nature and cycling, in
Bolin, B., Degens, E. T., Kempe, S., and Ketner, P. (eds) The Global Carbon Cycle, SCOPE
Report No. 13, Chichester, Wiley, 293-316.

NAS (1978) Nitrates: An Environmental Assessment, Washington, D.C., NAS/NRC.

Pierrou, U. (1976) The global phosphorus cycle, in Svensson, B. H., and Söderlund, R. (eds)
Nitrogen, phosphorus and sulphurglobal cycles, SCOPE Report No. 7, Ecol. Bull.
(Stockholm), 22, 75-88.

Rast, W., and Lee, G. F. (1978) Summary analysis of the North American (U.S. Portion)
OECD Eutrophication Project: Nutrient loading - lake response relationships and trophic state
indices. EPA-6001378008.

Robinson, E., and Robbins, R. C. (1970) Gaseous sulphur pollutants from urban and natural
sources, J. Air Pollution Control Ass., 20, 233-235.

Robinson, E., and Robbins, R. C. (1971) Emissions, concentrations and fate of gaseous
atmospheric pollutants. Report of Stanford Research Institute, Menlo Park, California.

Robinson, E., and Robbins, R. C. (1979) Gaseous atmospheric pollutants from urban and
natural sources, in Singer, S. F. (ed.) Global Effects of Environmental Pollution, Dordrecht,
Reidel, 50-64.
odhe, H. (1981) Current problems related to the atmospheric part of the sulphur cycle, in
Likens, G. (ed.) Some Perspectives of the Major Biogeochemical Cycles, SCOPE Report No.
17, Chichester, Wiley.

Rosswall, T. (1976) The internal nitrogen cycle between micro-organisms, vegetation and
soil, in Svensson, B. H., and Söderlund, R. (eds) Nitrogen, Phosphorus and SulphurGlobal
Cycles, SCOPE Report No. 7, Ecol. Bull. (Stockholm), 22, 157-167.

Rosswall, T. (1981) The biogeochemical nitrogen cycle, in Likens, G. E. (ed.) Some


Perspectives of the Major Biogeochemical Cycles, SCOPE Report No. 17, Chichester, Wiley,
25-49.

Rotty, R. M. (1981) Data for global CO2 production from fossil fuels and cement, in Bolin, B.
(ed.) Carbon Cycle Modelling, SCOPE Report No. 16, Chichester, Wiley, 121-125.

Ryaboshapko, A. G. (1983) The atmospheric sulphur cycle, in Ivanov, M. V., and Freney, J.
R. (eds) The Global Biogeochemical Sulphur Cycle, Chichester, Wiley (In press).

Schlesinger, W. H. (1984) The world carbon pool in soil organic matter, in Woodwell, G. M.
(ed.) The Role of Terrestrial Vegetation in the Global Carbon Cycle. Methods for Appraising
Changes, SCOPE Report No. 23, Chichester, Wiley.

Seiler, W. , and Crutzen, P. J. (1980) Estimates of gross and net fluxes of carbon between the
biosphere and the atmosphere from biomass burning, Climatic Change, 2, 207-218.

Simpson, H. J. (rapporteur), Broecker, W. S., Garrels, R. M., Gessel, S. P., Holland, H. D.,
Holser, W. T., Junge, C., Kaplan, I. R., McElroy, M. B., Michaelis, W., Mopper, K.,
Schidlowski, M., Seiler, W., Steele, J. H., Wofsy, S. C., and Wollast, R. F. (1977) Man and
the global nitrogen cyclegroup report, in Stumm, W. (ed.) Global Chemical Cycles and Their
Alterations by Man, Berlin, Dahlem Konferenzen, 253-274.

Söderlund, R., and Svensson, B. H. (1976) The global nitrogen cycle, in Svensson, B. H., and
Söderlund, R. (eds) Nitrogen, Phosphorus and SulphurGlobal Cycles, SCOPE Report No. 7,
Ecol. Bull. (Stockholm), 22, 23-73.

Söderlund, R., and Rosswall, T. (1982) The nitrogen cycles, in Hutzinger, O. (ed.) The
Handbook of Environmental Chemistry Vol. 1 Part B. Berlin, Heidelberg, New York,
Springer-Verlag 61-81.

Stevenson, F. J. (1965) Origin and distribution of nitrogen in soil, in Bartholomew, W. V.,


and Clark. F. E. (eds) Soil Nitrogen, Agronomy, 10, 1-42, Madison, Wisc., American Society
for Agronomy Inc., Publishers.

Stumm, W. (1973) The acceleration of the hydrogeochemical cycling of phosphorus, Water


Res., 7, 131-144.

Sweeney, R. E., Liu, K. K., and Kaplan, I. R. (1977) Oceanic nitrogen isotopes and their uses
in determining the sources of sedimentary nitrogen, in Robinson, B. W. (ed.) Stable Isotopes
in the Earth Sciences, DSIR Bulletin, 220, 9-26, Wellington, New Zealand Dept. of Scientific
and Industrial Research.
Takahashi, T., Broecker, W. S., and Bainbridge, A. E. (1981) Supplement to the alkalinity
and total carbon dioxide concentration in the world oceans, in Bolin, B. (ed.) Carbon Cycle
Modelling, SCOPE Report No. 16, Chichester, Wiley, 159-200.

Taylor, S. R. (1964) Abundance of chemical elements in the continental crust: a new table,
Geochim. Cosmochim. Acta, 28, 1273-1285.

UN (1981) Statistical Yearbook 1979/80, New York, United Nations.

Van Wazer, R. (1973) The compounds of phosphorus, in Griffith, E., Beeton, A., Spencer, J.,
and Mitchell, D. (eds) Environmental Phosphorus Handbook, New York, Wiley, 169-178.

Volkov, I. I., and Rozanov, A. G. (1983) The sulphur cycle in oceans. Part 1. Reservoirs and
Fluxes of Sulphur, in Ivanov, M. V., and Freney, J. R. (eds) The Global Biogeochemical
Sulphur Cycle. Chichester, Wiley (In press).

Whittaker, R. H., and Likens, G. E. (1975) The biosphere and man, in Lieth, H. et al. (eds)
Primary Productivity of the Biosphere, Ecol. Stud., 14, 305-328, Berlin, Springer-Verlag.

Williams, P. J. Le B. (1975) Biological and chemical aspects of dissolved organic material in


sea-water, in Riley, J. P., and Skirrow, G. (eds) Chemical Oceanography, 2nd edn, Vol. II,
London, Academic Press, 301-363.

Wollast, R. Interactions in estuaries and coastal waters (Chapter 14, this volume).

Zipf, E. C., and Prassad, S. S. (1982) A mesospheric source of nitrous oxide, Nature, 295,
133-134

Back to Table of Contents

The electronic version of this publication has been prepared at


the M S Swaminathan Research Foundation, Chennai, India.

SCOPE 21 -The Major Biogeochemical Cycles and Their


Interactions
3 Atmospheric InteractionsHomogeneous Gas Reactions of C N, and S
Containing Compounds
P. J. CRUTZEN

Abstract
3.1 Introduction
3.2 Tropospheric Photochemistry
3.3 Stratospheric Photochemistry
3.4 The Most Important Carbon Compounds
3.4.1 Carbon Dioxide
3.4.2 Carbon Monoxide
3.4.3 Methane
3.4.4 Isoprene and Terpenes
3.5 The Most Important Nitrogen Compounds
3.5.1 Nitric Oxide, Nitrogen Dioxide and Nitric Acid
3.5.2 Ammonia
3.5.3 Nitrous Oxide
3.6 The Most Important Sulphur Compounds
3.6.1 Sulphur Dioxide
3.6.2 The Reduced Sulphur Gases H2S, (CH3)2S, CH3SH, COS and CS2
3.7 Conclusions
Acknowledgements
References
Comment to Chapter 3
J. E. Lovelock
References

ABSTRACT

This article is a review of the most important photochemical processes that take place in the
atmosphere and of the cycles of many C, N, and S containing atmospheric constituents. The
emphasis is on the essential role of tropospheric ozone and how the distribution of this gas is
influenced by photochemical interactions between carbon and nitrogen containing
compounds, which are changing substantially due to man's activities. There is a large
potential for tropospheric ozone formation, of which at most only 10% is realized due to
generally low NOx concentrations. The input of NOx in the troposphere may be dominated by
anthropogenic sources so that tropospheric ozone may have increased in the industrial era.

The budgets (sources and sinks) of many compounds are derived and it is shown that these
are mostly dominated by tropospheric reactions. However, several gases are rather inert in the
troposphere and their photochemical breakdown in the stratosphere leads to products that
influence stratospheric ozone and sulphate aerosol. There often remain large uncertainties in
the quantitative aspects of the atmospheric cycles of trace constituents.

3.1 INTRODUCTION

Of the elements carbon, nitrogen, sulphur, and phosphorus, gas phase reactions are by far the
least important for phosphorus. There are some indications that phosphine (PH3) is released
from tropical forest soils (R. Herrera, personal communication), but the residence time of PH3
is probably so short that little atmospheric transfer can occur. Phosphorus enters the
atmosphere mainly on soil dust (Duce, Chapter 17, this volume). On the other hand, several
C, N, and S compounds have considerable atmospheric concentrations, mainly in
combination with the elements O and H. These compounds play an essential role in the gas
phase photochemistry and the cycling of many volatile elements in the atmosphere, not only
those containing C, N, and S, but also, for example, the photochemically important halogen
compounds.
The composition of the natural atmosphere is to a considerable extent determined by
biological processes. With the exception of CO2, the typical natural cycle of an element
consists of release at the earth's surface as a reduced, often hydrogen-bound gas, its
subsequent oxidation in the atmosphere by gas phase photochemical reactions, and finally the
removal from the atmosphere by precipitation scavenging, and deposition on the earth's
surface. Here, such compounds may serve as electron acceptors in anoxic environments
making possible the many microbiological reduction-oxidation processes in the soils and
waters of the earth. As two examples we mention methane (CH4), which is photochemically
converted to carbon monoxide (CO), and hydrogen sulphide (H2S), which is converted to
sulphur dioxide (SO2) in the atmosphere. Further oxidation leads next to carbon dioxide
(CO2) and sulphuric acid (H2SO4). Carbon dioxide is removed from the atmosphere mainly
through photosynthesis in vegetation, while sulphuric acid is removed mainly through wet
and dry deposition. I will note in the following that the natural sources of many biogenic
gases are not well known. It should also be mentioned here that CO2 does not play a
significant direct role in the photochemistry of the earth's atmosphere, so that the
photochemistry of CO2 will not be discussed in the following. The reason for this is that CO2
absorbs the photochemically active ultraviolet radiation in the same wavelength regions as
molecular oxygen, which is a thousand times more abundant. The role of CO2 in atmospheric
chemistry is indirect by its influence on the temperature structure of the atmosphere and
climate.

In contrast to natural, biological processes, man's activities cause the emissions of increasing
amounts of oxidized gases into the atmosphere. The total global supply of anthropogenic
fixed nitrogen, as NO, and of sulphur, as SO2, are now an appreciable fraction of the total
supply of these nutrient elements. In contrast to the biogenic emissions, industrial emissions
occur mainly as point or regional sources, so that atmospheric concentrations of the various
gaseous pollutants are non-uniformly distributed and their concentrations often strongly
correlated. This leads to strong photochemical and chemical interactions, that may result in
very different photochemical behavior to that which occurs in clean air masses.

In the following discussion I will review the homogeneous, atmospheric reactions and cycles
of the most important gases that contain carbon, nitrogen, and sulphur in combination with
hydrogen and oxygen. The gases that will be considered are important because of their role in
biological, climatic and photochemical processes. I will emphasize the important role of
tropospheric ozone in atmospheric photochemistry, how it influences the cycles of many
trace elements and how man's activities affect tropospheric ozone. The experience of the last
decade has already shown that there are many ways in which human activities may affect the
balance between ozone formation and destruction in the stratosphere and thereby influence
the intensity of ultraviolet radiation at the ground, and maybe climate, with possible
repercussions for the biosphere.

In the following sections, brief reviews of the homogeneous photochemistry of the


troposphere and stratosphere will be presented. Next, the budgets of the photochemically
most important carbon, nitrogen and sulphur containing gases will be taken up for review. I
will emphasize new developments since the SCOPE 7 study (Svensson and Söderlund, 1976),
and interactions between cycles whenever they can be recognized.

3.2 TROPOSPHERIC PHOTOCHEMISTRY


Although only about 10% of all atmospheric ozone (O3) is located in the troposphere, the
lowest 1017 km of the atmosphere, this small amount of ozone is nevertheless of fundamental
importance for the composition of the earth's atmosphere. The reason for this is the
production of the highly reactive hydroxyl (OH) radical by the two reactions:

O(1D) +O2 (<310 nm;1 nm = l0-


(R1) O3 + hv 9
m)
(R2) O(1D) + H2O 2OH

where O(1D) denotes an electronically excited oxygen atom. It is the attack by OH that
initiates the oxidation of many trace gases in the atmosphere, the most important examples of
which are shown in Table 3.1 and Figure 3.1 (Levy,1971, 1974; McConnell et al., 1971). In
the background troposphere, about 60% of the OH radicals react with CO, and about 40%
with CH4. Smaller fractions react with the other gases listed in Table 3.1.

The average concentration of hydroxyl in the atmosphere is now estimated to be about 6 x


105 molecules cm-3, with an uncertainty of about a factor of two (Crutzen et al., 1978;
Derwent and Eggleton, 1981; Gidel et al., 1982). This estimated range is consistent with the
global observations of methylchloroform, which is removed from the atmosphere by
reactions with OH and which does not appear to have any atmospheric sources other than the
well known industrial emissions. Most hydroxyl is located in equatorial regions, where the
intensity of ultraviolet radiation is at a maximum and where absolute water vapour densities
are highest (Crutzen et al., 1978; Derwent and Eggleton, 1981). The calculated OH
distribution also explains most of the features of the C14O observations in the troposphere
(Volt et al., 1979).

Although OH reacts overwhelmingly with CO and CH4, these reactions do not necessarily
lead to the removal of hydroxyl from the atmosphere, because hydroxyl will mainly act as a
catalyst. For instance, in the presence of sufficient concentrations of another catalyst, nitric
oxide, the oxidation of carbon monoxide will lead to the formation of tropospheric ozone,
without loss of OH and NO, as follows:

(R3) CO + OH H + CO2
(R4) H + O2 + M HO2 + M
(R5) HO2 + NO OH + NO2
(R6) NO2 + hv NO + O (<400 nm)
(R7) O+O2+M O3+M

net: CO + 2 O2 CO2 + O3

A competing chain of reactions, leading to ozone destruction:


(R3) CO + OH H + CO2
(R4) H + O2 + M HO2 + M
(R8) HO2 + O3 OH + 2O2

net: CO + O3 CO2 + O2
again does not lead to loss of hydroxyl, and takes place whenever the ratio of the atmospheric
concentrations of NO and O3 is less than 2 x 10-4. With ozone volume mixing ratios
increasing from about 20 x 10-9 (20 ppbv) at ground level to 100 ppbv at the tropopause, the
break-even point between reaction chains (R3R7) and (R3, R4, R8) is attained at nitric oxide
volume mixing ratios of 4 x 10-12 (4 pptv) at ground level and 20 pptv at the tropopause.
These are indeed very low concentrations, but they may nevertheless not be exceeded in
extensive regions of the troposphere because of the very short residence times of the oxides
of nitrogen, NO and NO2, which are a result of the rapid formation of highly water-soluble
nitric acid by the reactions:

(R9) NO + O3 NO2 + O2
NO2 +
(R10) HNO3 (+M)
OH(+M)

Through similar reactions ozone is also produced in the oxidation of CH4 and other
hydrocarbons if the mixing ratios of NO are sufficiently large. McFarland et al. (1979) have
indeed indicated the possibility of very low background concentrations of NO by the
measurements of volume mixing ratios of less than 10 ppt in the marine boundary layer of the
tropical Pacific. Noxon (1981) measured profiles of NOx and reported a mixing ratio for NOx
(NO + NO2) of 30 pptv at 3 km altitude at Hawaii. There are, unfortunately, too few
measurements to derive a typical distribution of NO in the troposphere, and to make a good
estimate of the sources and sinks of tropospheric ozone. The only term that can be estimated
with reasonable reliability is the global ozone loss by the reactions (R1) and (R2), amounting
to about 8 x 1010 molecules cm-2s-1 (Fishman et al., 1979a). This ozone loss may be compared
with the estimated downward transport of about 6 x 1010 molecules cm-2s-1 stratospheric
ozone by meteorological processes and an ozone loss rate at the ground of the same
magnitude (Fabian and Pruchniewicz, 1977). In order to balance the ozone budget of the
troposphere, it is, therefore, quite feasible that photochemical interactions between carbon
and nitrogen containing gases lead to tropospheric ozone production. This is particularly
important as the atmospheric sources of NO, CO, CH4 and other hydrocarbons are greatly
influenced by anthropogenic activities, especially in the Northern Hemisphere.

Table 3.1a Budgets of carbon species; atmospheric lifetimes in hours, months, or years;
diffusion distances in EW, SN and vertical directions (in km) over which concentrations are
reduced to 30% by chemical reactions. Lifetimes and removal rates calculated with (OH) = 6
x 105 molecules cm-3. 1 ppmv = 10-6; 1 ppbv = 10-9; 1 pptv = 10-12

Secondary Atmospheric Transport


Gas Direct source/year Removal by
source/year life- distances
Source
Source identification times x, y, z (km);
identification
volume mixing
ratios
in unpolluted
tropo-
sphere

CO 416 x 1014g CO 3.7 9.3 x 1014g 30 x 1014g 2 months 4000, 2500, 10


CO CO
methane
Biomass burning OH 50200 ppbv
oxidation
4.5 x 1014g
6.4 x 1014g CO 413 x 1014g CO
CO
Uptake by
Industry C5H8, Cl0H16
soils
0.22 x 1014g CO oxidation
Vegetation

CH4 0.30.6 x 1014g CH4 4 x 1014g CH4 7 yrs Global


Rice paddy fields OH 1.5 2.0 ppmv
0.32.2 x 1014g CH4
Natural wetlands
0.6 x 1014g CH4
Ruminants
< 1.5 x 1014g CH4
Termites
0.31.1 x 1014g CH4
Biomass burning
0.2 x 1014g CH4
Gas leakage

C5H8 8.3 x 1014g C 8.3 x 1014g C 10 hrs 400, 200, 1


Cl0H16 Trees OH 010 ppbv

Table 3.1b Budgets of nitrogen species. For explanation see Table 3.1a

Direct Secondary Atmospheric Transport


Gas Removal by
source/year source/year life- distances
Source Source
times x, y, z (km);
identification identification
volume mixing
ratios
in unpolluted
tropo-
sphere

NOx 1220 1012g N 1.01.5 1012g N 2585 1012g N 1.5 d 1500,400, 10


Oxidation of
(NO+NO2) Industry OH 1100 pptv
N2O
1040 1012g N Deposition
Biomass burning on soils and
110 1012g N oceans
Lightning
115 1012g N
Soils
0.15 1012g N
Ocean
0.25 1012g N
Jet aircraft

HNO3 1585 1012g N Rain 3d 3000, 600,1.5


NO2 + OH 10-300 pptv

N2O 1.8 1012g N 611 1012g N 100200 yrs global


Fossil fuel
Stratospheric 300 ppbv
burning
12 1012g N photolysis
Biomass burning
12 1012g N
Oceans, estuaries
13 1012g N
Cultivation
natural soils
< 3 1012g N
Fertilized fields
?
Natural soils

NH3 1020 1012g N Rain <9d < 9000, 1000, 3


Domestic animals 03 ppbv
26 1012g N
Wild animals
< 3 1012g N
Fertilized fields
<30 1012gN
Natural fields
412 1012g N
Coal burning
<60 1012g N
Biomass burning
Table 3.1c Budgets of sulphur species. For explanation see Table 3.1a. COS and CS2 are not
listed here, because too little is known about their sources and sinks. The industrial source of
CS2 is about 2 x 1011g S per year

Secondary Transport
Gas Direct source/year Removal by Atmospheric
source/year distances
Source Source
Life-times x, y, z (km);
identification identification
volume mixing
ratios
in unpolluted
tropos-
sphere

SO2 64 x 1012g S 40100 x 1012g S OH 5d 5000, 700, 2.5


oxidation H2S,
Coal burning Rain 10200 ppty
DMS
26 x 1012g S
Petroleum burning
11 x 1012g S
Non-ferrous ores
1030 x 1012g S
Volcanoes

H2S < 4 x 1012g S OH 2d 2000, 500, 1.5


(CH3)2S Agricultural fields 0100 ppty
CH3SH 3142 x 1012g S
Open ocean
10 x 1012g S
Coastal waters
16 x 1012g S (?)
Tropical forests
24 x 1012g S (?)
Wetlands

There are tropospheric ozone data available that indicate a possible increase of 20% in the
middle troposphere (28 km) at north temperate latitudes over the time period 19671979
(Angell and Korshover, 1979), although such an increase seems not to be present in the
Mauna Loa data of the past 6 years (Komhyr, personal communication).

To emphasize further the importance of tropospheric ozone I also point out its role in the
earth's radiation budget, because of its strong absorption band located in the atmospheric
`window' region near 9.6 µm. As a result of this, ozone absorbs, and radiates back to the
earth's surface, terrestrial radiation that otherwise would escape to space. The particular
efficiency of the tropospheric ozone fraction is caused by the pressure broadening of the
absorption lines in the troposphere. A hypothetical doubling of tropospheric ozone has been
calculated to lead to a surface temperature increase of about 0.7°C (Fishman et al., 1979b).

An approximate upper limit to the potential ozone production rates in the troposphere can be
derived by assuming that there is enough NO present in the atmosphere that during the
oxidation of a hydrocarbon molecule to CO, as with methane, there are 23 ozone molecules
produced per carbon atom (Crutzen, 1973; Fishman et al., 1979a). The further oxidation of
CO to CO2 subsequently adds another ozone molecule to the atmosphere. Making use of the
information on the sources of CO, CH4 and non-methane hydrocarbons (Table 3.1), an
average global, tropospheric ozone production potential of 2 x 1012 molecules cm-2s-1 can be
calculated. The actual tropospheric ozone production is probably at most 10% of this
potential, because ozone loss at the ground or by photochemical reactions is about 1011
molecules cm-2s-1 (Fishman et al., 1979a). This indicates that much oxidation of
hydrocarbons, which occurs mainly in the tropics, must take place without the production of
ozone. One may guess that this is so because the air above the tropical forests is at present
practically devoid of nitric oxide. Future expansions of industrial and biomass burning
activities in the tropics may, however, have large implications for the tropospheric ozone
distribution, because of the enhanced supply of nitric oxide to the boundary layer of the
atmosphere in which the oxidation of isoprene and terpenes to CO occurs. It should be
mentioned here that most of the nitric oxide produced by lightning is deposited above the
tropical boundary layer where the concentrations of these highly reactive hydrocarbons
should be rather low. I will return to the possible implications of human activities on the air
chemistry of the tropics when I discuss the role of nitric oxide in the production of carbon
monoxide.

Figure 3.1 Compilation of the most important photochemical processes in the atmosphere,
including estimates of flux rates expressed in moles per year between the earth's surface and
the atmosphere and within the atmosphere. The processes are numbered and explained in the
figure texts. The symbol 'E' means: 10y where y is the number that follows E, e.g. 2E13 is 2
1013

As carbon monoxide is the main reactant of hydroxyl, an increase in the atmospheric carbon
monoxide content could lead to lower tropospheric concentrations of OH and thereby cause
an increase in the tropospheric abundances of those gases that are mainly removed by
reactions with OH (Wofsy, 1976). This could have further chemical and climatological
consequences, as for instance a larger transfer of gases to the stratosphere with possible
effects on ozone. It is likely, however, that a rise in the tropospheric CO abundance by
anthropogenic activities will be accompanied by increases in the tropospheric NO abundance.
This would lead to higher ozone concentrations. The increases in NO and O3 would tend to
increase the OH concentrations through reactions (R1), (R2), and (R5), thereby counteracting
the effect of CO.

Methane oxidation is not only a potentially significant source of CO, but the atmospheric
concentrations of OH may largely be determined by the particular reaction paths that are
followed during methane oxidation. Again, nitric oxide plays an important role in this. With
enough NO present (>10 pptv) the reaction path, leading to formaldehyde (CH2O), is as
follows:
(R11) CH4 + OH CH3 + H2O
CH3 + O2 +
(R12) CH3O2 + M
M
CH3O2 +
(R13) CH3O + NO2
NO
(R14) CH3O + O2 CH2O + HO2
(R5) HO2 + NO OH + NO2
(R6) NO2 + hv NO + O (<400 nm), 2x
(R7) O + O2 + M O3 + M, 2x

net: CH4 + 4 O2 CH2O + H2O + 2 O3


Further reactions of CH2O towards CO lead then to some net gain of hydroxyl (Levy, 1971).

With little NO present, CH4 oxidation may follow the pathway:

(R11) CH4 + OH CH3 + H2O


CH3 + O2 +
(R12) CH3O2 + M
M
CH3O2 +
(R15) CH3O2H + O2
HO2
CH3O2H +
(R16) CH3O + OH
hv
(R14) CH3O + O2 CH2O + HO2

net: CH4 + O2 CH2O + H2O

However, the photolysis of CH3O2H is slow, which results in a residence time of 1 week for
this compound. Thus the methyl hydroperoxide may be rained out of the atmosphere or react
with the earth's surface or aerosol particles. If this is the case, the oxidation of CH4 will lead
to a loss of two odd hydrogen radicals (OH and HO2 and no formation of CO would occur in
the atmosphere. It follows that the pathways of methane oxidation in the atmosphere are not
yet satisfactorily resolved. More intricate questions will have to be addressed regarding the
oxidation of higher hydrocarbons.

An interesting, and potentially important, interaction between carbon and nitrogen gases in
the atmosphere occurs through the formation of certain organic nitrate molecules during the
oxidation of hydrocarbons. The most important of these are probably the peroxy-acyl nitrates,
especially peroxy-acetyl nitrate (PAN) with the chemical formula CH3C(=O)O2NO2. PAN
thermally decomposes in the atmosphere according to the reactions

(R17) CH3(C=O)O2NO2 CH3(C=O)O2 + NO2


CH3(C=O)O2 +
(R18) CH3 + CO2 + NO2
NO

or,

(R19) CH3(C=O)O2 + CH3 (C=O)O2H + O2


HO2
CH3(C=O)O2H +
(R20) CH3 + CO2 + OH
hv

depending on the concentration of nitric oxide. Based on information by Hendry and Kenly
(1977) and Cox and Roffey (1977), the following values are calculated for the atmospheric
residence time of PAN against photochemical destruction for several heights in the U.S.
standard atmosphere: at z = 0 km, T = 288K, (PAN) = 3 days; z = 4 km, T= 262K, (PAN) = 1
month; z = 6 km, T = 249K, (PAN) = 1 yr; z = 8 km,

T = 235K, (PAN) = 15 yrs.

Because of the strong temperature dependence of reaction (R17), the lifetime of PAN against
thermal decomposition increases dramatically with height. Likewise, higher PAN
concentrations in the atmosphere may be favoured during the winter season (Hendry and
Kenley, 1977). Although photolysis of PAN should be rather slow, the only ultraviolet
spectrum of PAN published so far (Stephens, 1969) does not contain enough information to
exclude significant photolysis in the middle and upper troposphere and stratosphere. Reaction
of PAN with OH may likewise not entirely be neglected in these regions (Atkinson et al.,
1979), but should not result in an atmospheric residence time shorter than several months.

Surprisingly, PAN is slowly lost from the atmosphere by wet and dry removal. A laboratory
study by Garland and Penkett (1976) gave a deposition rate of PAN on water surfaces that
was slower than that of ozone. Like ozone, PAN is not removed by rainfall, so that in the free
troposphere above 3 km, the atmospheric lifetime of this gas is clearly much larger than that
of NOx. Long range transport of NOx may, therefore, occur in the middle and upper
troposphere with PAN as the vehicle. When PAN reaches the boundary layer it is thermally
decomposed and NOx is released to the atmosphere (Crutzen, 1979).

PAN is formed in polluted air by photochemical reactions, involving non-methane


hydrocarbons and NOx and has been observed in many urban environments (e.g. Stephens,
1969; Lonneman et al, 1976; Nieboer and van Ham, 1976; Penkett et al., 1977; Spicer, 1977;
Tuazon et al., 1981). At high concentrations this gas is a major phytotoxicant and it affects
health by causing eye irritation. In polluted air close to urban centres it is often present in
concentrations similar to those of nitric acid and typically 1020% those of NOx (Spicer, 1977;
Tuazon et al., 1981). PAN is formed from acetaldehyde (CH3CHO), which is a
photochemical intermediate in the photochemical decay of many non-methane hydrocarbons
(Demerjian et al., 1974). The simplest chain of reactions leading to the formation of PAN
occurs, following the oxidation of ethane, as follows:

(R21) C2H6 + OH C2H5 + H2O


(R22) C2H5 + O2 + M C2H5O2 + M
(R23) C2H5O2 + NO C2H5O + NO2
(R24) C2H5O + O2 CH3CHO + HO2
(R25) CH3CHO + OH CH3CO + H2O
(R26) CH3CO + O2 + M CH3(C=O)O2 + M
CH3(C=O)O2 + NO2
(R17) CH3(C=O)O2NO2 + M
+M
C2H6 is present in the troposphere at sufficiently high volume mixing ratios (12 ppbv) (Singh
et al., 1979; Rudolph et al., 1981) so that appreciable production of PAN from NOx may take
place anywhere in the atmosphere (Singh and Hanst, 1981); this makes the possible role of
PAN in global atmospheric photochemistry especially worthwhile to investigate. As
observations at ground level in rural areas show (Penkett et al., 1977; Singh et al., 1979)
PAN is not restricted to urban environments. In fact, PAN has been observed on several
cruises in the subtropical and tropical Atlantic, whenever colder and more polluted air masses
could reach the lower latitudes. Typical sea level concentrations during such episodes were of
the order of 50 pptv (Guicherit, private communication). With more PAN expected in the
background middle and upper troposphere (Crutzen,1979; Singh and Hanst, 1981), PAN
could be an important reservoir species for the long range transport of NOx.

The previously mentioned reactions are of critical importance in tropospheric


photochemistry, because of their role in determining the concentrations of OH in background
air. Most photochemistry of the gas phase in the troposphere is derived from this. For
example, it is normally assumed that after the initial attack by hydroxyl, the reduced sulphur
gases H2S, (CH3)2S and CH3SH are rapidly converted to SO2. However, this has only been
demonstrated in the laboratory for H2S, while SO2 formation did not take place for (CH3)2S
(Cox and Sandalls, 1974), so that atmospheric reaction chains that do not produce SO2,
cannot be ruled out. In the presence of enough NO, we may even speculate about the
possibility of formation of carbonyl sulphide (COS) via the reaction sequence:

(R27) CH3SCH3 + OH CH3SCH2 + H2O


CH3SCH2 + O2 +
(R28) CH3SCH2O2 + M
M
CH3SCH2O2 +
(R29) CH3SCH2O + NO2
NO
CH3S + CH2O
(R30a) CH3SCH2O
(probable)
CH3SCHO + HO2 (not
(R30b) CH3SCH2O + O2
impossible)
(R31) CH3SCHO + OH H2O + CH3 + COS

Although less likely, it may not be ruled out that the initial attack of OH on (CH3)2S occurs
by addition of OH to S (Atkinson et al., 1978; Kurylo, 1978). Also, in the absence of enough
NO, peroxides and many other compounds may be formed that can be removed
heterogeneously from the atmosphere (Bentley et al., 1972; Panther and Penzhorn, 1980), so
that only a fraction of the (CH3)2S may be converted to COS.

Because COS is a sulphur compound with a much longer atmospheric lifetime ( years) than
(CH3)2S (days), it will be transported globally and reach the stratosphere (Hanst et al., 1975;
Sandalls and Penkett, 1977; Inn et al., 1979; Torres et al., 1980). The photolysis of COS
leads to the SO2 and H2SO4 needed to explain the stratospheric sulphate layer during periods
of low volvanic activity (Crutzen, 1976). Through its influence on the stratospheric aerosol
layer, COS is of some significance for the earth's climate; the existence of any long term
trends in the atmospheric abundance of COS should be ascertained (Turco et al., 1980;
Hofmann and Rosen,1981).
Carbonyl sulphide, with an atmospheric volume mixing ratio of about 0.5 ppbv, is probably
the most abundant sulphur species of the atmosphere. Recently, there have been some
speculations that this gas (or CS2) would also be the precursors of the SO2 that has been
measured in the middle and upper troposphere over the Pacific (Maroulis et al., 1980).
However, the proposed reaction mechanisms via reaction with hydroxyl (Logan et al., 1979;
Sze and Ko, 1980) have been shown to be incorrect (Atkinson et al., 1978; Ravishankara et
al., 1980). The hypothesis that COS oxidation in the stratosphere is an important source of
SO2 and sulphate during periods with little volcanic activity has, however, probably been
shown to be correct from measurements of COS and SO2 in the stratosphere (Inn et al., 1979;
Meixner, 1981).

A particularly interesting kinetic observation has recently been made by Jones et al. (1982),
who showed that CS2 reacts with OH only in the presence of O2 via

(R32a) CS2 + OH + O2 COS + SO2 + H


with an effective rate coefficient of k32a(O2) = 2 x 10-12cm3 molecule-1s-1.

These results have been confirmed by Niki and co-workers (Niki, personal communication).

Likewise, Jones et al. (1982) have reported that the photodissociation reactions occur with a
quantum yield of 10-3 in the 320 nm absorption band, making the two listed pathways to COS
and SO2 almost of equal importance. These reactions may be particularly important as
sources of atmospheric COS. Recently, this gas has also been identified as a potential
corrosion agent (Graedel et al., 1981).

(R32b) CS2 + hv CS2*


(R32c) CS2*+ O2 COS + O

The oxidation of SO2 to H2SO4 in the presence of NOx also involves initiation by reaction
with hydroxyl, leading, for example, to the reaction sequence (Calvert et al., 1978):

(R33) SO2 + OH + O2 HSO5


(R34) HSO5 + NO HSO4 + NO2
(R35) HSO4 + HO2 H2SO4 + O2

At low NO concentrations the oxidation steps are less certain (Davis and Klauber, 1975;
Davis et al., 1979).

An interesting possible influence that NOx might exert on the SO2 oxidation cycle in
industrial air masses has recently been pointed out (Rodhe et al., 1981). Because reaction
(R10) between OH and NO2 is about ten times faster than reaction (R33) it may act as a sink
for OH when the mixing ratio of NOx is of the order of a few ppb or larger. Under situations
when photochemical oxidation is important, the oxidation of SO2 may, therefore be delayed
until the concentrations of NO2 become smaller than about 1 ppbv. At higher NOx
concentrations, the production of H2O2 from will be suppressed because of reaction (R5). As
H2O2 may be an important oxidant of SO2 in water (Penkett et al., 1979), the oxidation of
SO2 in clouds likewise may be delayed.
(R36) HO2 + HO2 H2O2 + O2

Although mainly gas phase reactions leading to SO2 oxidation have been discussed, it is clear
that heterogeneous and aqueous phase reactions may be at least as important as homogeneous
reactions for the oxidation of SO2 to H2SO4, especially during winter time at middle and high
altitude (see e.g. Shaw and Rodhe, 1981).

Under most atmospheric conditions the conversion of NOx to HNO3 via reaction (R10) occurs
within a few days. The atmospheric residence time of highly soluble HNO3 must be shorter
than that of water vapour, which is about 9 days (Levine and Schwartz, 1981). As this is
much less than the time scale of 2 months needed for HNO3 to convert back to OH and NO2,
particularly through the photolysis reaction the formation of nitric acid by reaction (R10)
provides an efficient sink for NOx. In this way, an appreciable upward transport of NO and
NO2 to the stratosphere is made difficult.

(R37) HNO3 + hv OH + NO2

It remains, however, of interest to explore the possibility of some leakage of NOx and NO2
(and other pollutant gases, such as SO2) through the tropospheric water vapour and cloud
filter to higher altitudes (middle and upper troposphere, lower stratosphere). One mechanism
for such transport may be the rapid upward transport in frontal zones and especially
thunderstorms so that reaction (R10) cannot be completed. Thunderstorms, which penetrate
the tropopause, may directly transfer lightning-produced and boundary-layer NOx into the
lower stratosphere. As a consequence of this, outside polluted areas, such rapid transfer could
bring about an increase of NOx mixing ratios with altitude, providing a complementary
interpretation to the NOx profiles to the downward transport from the stratosphere proposed
by Liu et al. (1980). This may be especially important for the tropical band and during
summertime at mid-latitudes. Another mechanism may be transport at high latitudes during
winter, when little or no OH is produced by reaction (R2) and when, furthermore,
precipitation occurs in rather small amounts and more as snow, which may be less efficient
than rain in scavenging trace gases. This possibility of global NOx transport is speculative,
though not entirely unreasonable, judging from observations of considerable pollution levels
in the Arctic regions during winter (Rahn and Heidam, 1981; Shaw, 1981).

Observations of HNO3 by Lazrus and co-workers (Lazrus and Gandrud, 1974; Huebert and
Lazrus, 1978) and of NO2 by Noxon (1978, 1979) in areas remote from pollution sources
have shown volume mixing ratios of NO2 and HNO3 that are substantially larger in the
stratosphere than in the troposphere. This indicates a net transfer of odd nitrogen from the
stratosphere to the troposphere, but does not rule out the possibility of a significant transport
of some tropospheric NOx to the lower stratosphere at preferred locations. One-dimensional
models are totally inadequate to address this issue and far too few data are available to make
a reliable judgment on this interesting matter.

Finally, I shall also briefly discuss whether photochemical reactions may affect ammonia. For
this gas, both uptake at the earth's surface and release by microbiological processes should be
considered in order to establish the overall net source of NH3 to the atmosphere. The only
homogeneous gas phase reaction of ammonia is again reaction with OH:

(R38) OH + NH3 NH2 + H2O


This reaction is, however, rather slow (k38 1.5 x 10-13cm3 molecule-1s-1), so that with average
(OH) 6 x 105 molecules cm-3, NH3 would remain in the atmosphere for a season, if reaction
(R38) represents the only atmospheric loss process. In reality, ammonia will be removed
from the atmosphere on the average during a period of about 9 days, which represents the
average residence time of water vapour in the atmosphere. Thus, at most only a fraction of
atmospheric NH3, about 10%, could react with OH. In addition, the NH2 radical that is
formed can react with HO2:

(R39) NH2 + HO2 NH3 + O2

providing a pathway back to NH3. Such reactions with radicals or minor atmospheric
constituents are important because the reaction

(R40) NH2 + O2 + M NH2O2 + M

is slow. It remains to be seen whether NH3 oxidation would represent a source or a sink for
atmospheric NOx (McConnell, 1973) through reactions such as (Lesclaux and Demissy, 1977;
Hack et al., 1978; Kurasawa and Lesclaux, 1980):

(R41) NH2 + NOx H2O + N2Ox-1(x = 1,2)


(R42) NH2 + O3 NH2O + O2

where further oxidation of NH2O may lead to NOx, probably via HNO formation. The break-
even point between reactions (R41) and (R42) occurs at about 60 pptv of NO. Such mixing
ratios of NOx can indeed be reached in continental, agricultural areas where we expect
substantial emissions of atmospheric NH3 to take place. It may be that NH3 oxidation implies
a sink for some atmospheric NOx and a source for N2O. It is, however, not possible to make a
numerical assessment because of insufficient information on the distributions of NOx, OH
and NH3.

The great importance of NH3 in atmospheric chemistry is due to its role in establishing the
pH of rain and cloud water. A discussion of this may be found in the overview paper of
Taylor et al. (Chapter 4, this volume).

3.3 STRATOSPHERIC PHOTOCHEMISTRY

The temperature structure and the dynamic processes in the stratosphere are to a major degree
determined by the absorption of solar ultraviolet energy by ozone. The total amount of ozone
in the stratosphere is nevertheless rather small; it corresponds to the number of air molecules
that are contained in a 3 mm thick layer at standard temperature and pressure. As I have
noted, the troposphere contains only 10% of the atmospheric ozone, so that most ozone is
located in the stratosphere, which is between 10 and 50 km. Atmospheric ozone also plays an
important ecological role. Most ultraviolet solar radiation between about 210 and 310 nm is
filtered out by atmospheric ozone. However, penetration of UV radiation to ground level
starts at about 300 nm and increases by orders of magnitude within the next 20 nm. It is this
radiation that may be biologically harmful. The penetration: of this radiation to the ground is
enhanced by reduction of the ozone column.
Since the beginning of the 1970s it has become increasingly clear that a number of human
activities can lead to global changes in the amount of stratospheric ozone. Following
suggestions by Johnston (1971) and Crutzen (1971), initial attention was directed to pollution
of the stratosphere by direct injections of NO from high-flying aircraft. Earlier, Crutzen
(1970) had proposed that NOx (NO + NO2) would catalyse the destruction of ozone and limit
its stratospheric abundance by a simple set of photochemical reactions:

(R43) O3 + hv O + O2 (< 1140 nm)


(R44) O + NO2 NO + O2
(R45) NO + O3 NO2 + O2

net: 2 O3 3 O2
This catalytic chain of reactions largely balances the formation of ozone in the stratosphere
through the reaction sequence:
(R46) O2 + hv 2 O (< 240 nm)
(R7) O + O2 + M O3 + M(2x)

net: 3 O2 2 O3
The main source of NOx in the stratosphere is most probably the oxidation of nitrous oxide
(N2O) via (Crutzen, 1971; McElroy and McConnell, 1971; Nicolet and Vergison, 1971):
(R1) O3 + hv O(1D) + O2 (>310 nm)
(R47) O(1D) + N2O 2 NO

Stratospheric ozone may especially be affected by compounds that are relatively inert in the
troposphere, because of a low solubility in water, a slow photolysis and a slow reaction with
OH (e.g. nitrous oxide and several chlorocarbon gases, such as natural CH3Cl and industrially
produced CFCl3, CF2Cl2, CCl4, and CH3CCl3). Another way to influence stratospheric
chemistry is by direct injection of material in the upper troposphere and stratosphere well
above most atmospheric water and water vapour (e.g., volcanic eruptions, meteoritic impacts,
nuclear weapons testing, and emissions from aircraft in the stratosphere).

The oxides of nitrogen, NOx, play a remarkable catalytic role in the ozone balance of the
atmosphere. Above about 25 km the net effect of NOx additions to the stratosphere will be a
lowering of ozone concentrations by the set of reactions present earlier (R43 + R44 + R45).
However, below about 25 km in the stratosphere, NOx protects ozone from destruction. An
important reason for this is a set of reactions:

(R5) HO2 + NO OH + NO2


(R6) NO2 + hv NO + O
(R7) O+O2+M O3+M

net: HO2 + O2 OH + O3

As we have shown before, this reaction set is also basically the cause of ozone production
that takes place in the troposphere. In the lower stratosphere, the chain of reactions (R5 + R6
+ R7) tends to counteract the destruction of ozone by the catalytic reaction pair:
(R48) OH + O3 HO2 + O2
(R8) HO2 + O3 OH + 2 O2

net: 2 O3 3 O2
by deferring it into the chain:
(R48) OH + O3 HO2 + O2
(R5) HO2 + NO OH + NO2
(R6) NO2 + hv NO + O
(R7) O+O2+M O3+M

no net chemical effect.

An additional role of NOx in the stratosphere occurs through interactions with Cl and ClO. As
with NOx, chlorine atoms and chlorine monoxide molecules participate in an effective
catalytic chain of reactions that converts ozone back to molecular oxygen. This cycle goes as
follows:

(R43) O3 + hv O + O2
(R49) O + CIO Cl + O2
(R50) Cl + O3 ClO + O2

net: 2 O3 3 O2
The presence of NO in the stratosphere makes this cycle less effective, as the reactions
(R51) NO + ClO Cl + NO2
(R52) Cl + CH4 HCl + CH3

transform the catalysts Cl and ClO into HCl, which does not react photochemically with
ozone. Furthermore, since the reaction

(R53) ClO + NO2 + M ClNO3 + M

ties up both some ClO and some NO2 as non-reactive ClNO3, it is clear that ozone removal
by NOx additions to the stratosphere is mitigated by the NOx interference in the chlorine
cycle. The photochemical chains of reactions required to explain the distribution of
stratospheric ozone and to predict the further effects of human activities are thus quite
complicated. However, additions of NO to the low stratosphere may tend to increase local
ozone concentrations by reducing some of the ozone loss that would otherwise occur because
of the catalytic action of OH and HO2, and Cl and ClO.

The importance of the ozone production and protection by NOx in the stratosphere was
dramatically emphasized by discoveries of Howard and Evenson (1977) and Zahniser and
Howard (1979), who found reaction (R8) and especially (R5) to be much faster than
previously estimated. This finding resulted in substantial downward revisions of estimated
total ozone-column reductions due to stratospheric NOx additions from aircraft (Duewer et
al., 1977; Crutzen and Howard, 1978; Logan et al., 1978). As NOx is produced by the
oxidation of N2O via reaction (R47), the same conclusions are also partially valid regarding
the possible effects of a future rise in the atmospheric content of nitrous oxide. In this case,
however, more of the additional NOx produced in reaction (R47) will reach the upper regions
of the stratosphere, where the reaction set (R43R45) is more important.

An increase in NOx may be caused by man's intervention in the nitrogen cycle through the
increasing use of fixed nitrogen in agriculture, through changes in the composition of the
earth's soil and vegetation, and through fossil fuel and biomass combustion processes. I will
review the current state of knowledge regarding atmospheric N2O sources and sinks later in
this article. Here I briefly review the sensitivity of stratospheric ozone to N2O changes and
the latest developments in ozone depletion predictions.

In 1974, theoretical predictions (Crutzen, 1974) of total ozone reductions that used the then
recommended rate constants yielded the following approximate relationship between total
ozone change (V3) and a hypothetical increase in the volume mixing ratio of atmospheric
nitrous oxide, µ(N2O),

A doubling in the atmospheric abundance of N2O was therefore expected to yield a 20%
decrease in total ozone. New rate constants for reactions (R5) and (R8), determined by
advanced laboratory techniques first led to substantial downward revisions in the ozone-
reduction estimates. Actually, it was consequently estimated that an increase in the
atmospheric N2O abundance could lead to an increase in total ozone. Presently, however, it is
calculated that an increase in the atmospheric N2O abundance will lead to a decrease in total
ozone. This author's own calculations indicate a loss in total stratospheric ozone of about
12% for a doubling of N2O, keeping all other factors that affect ozone constant. The reason
for this finding is that substantially smaller concentrations of OH in the lower stratosphere
below about 30 km are now predicted through new calculations of the rate coefficients related
to the formation and photolysis of HO2, NO2, and HNO3. The role of NO2 is that of a catalyst
in the reaction cycles:

(R54) HO2 + NO2(+M) HNO4 (+M)


(R55) OH + HNO4 H2O + NO2 + O2

net: OH + HO2 H2O + O2


and
(R10) OH + NO2 (+M) HNO3 (+M)
(R56) OH + HNO3 H2O + NO3
(R57) NO3 + hv NO2 + O

net: 2 OH H2O + O

These reactions enhance the loss of HOx from the lower stratosphere, so that much less
HNO3. and more NOx, is calculated than was the case before. The newer calculations are in
much better agreement with observations (Coffey et al., 1981a). As a consequence, the
normal ozone balance in the lower stratosphere is again much more affected by NOx
catalysis, i.e. reactions (R43R45), and not so much by HOx, and ClOx,. Thus the
compensation effects of NOx, in the HOx, and ClO2 catalytic cycles, as discussed before, can
not make up for the enhanced loss of ozone due to NOx, catalysis in the entire stratosphere.
The importance of COS as a source of sulphur in the stratosphere, where this gas is
efficiently photolysed by ultraviolet radiation has already been discussed. More detailed
discussions on the many aspects of stratospheric photochemistry are available in recent
reviews (e.g. WMO/NASA,1982).

3.4 THE MOST IMPORTANT CARBON COMPOUNDS

3.4.1 Carbon Dioxide

With volume mixing ratios of about 3.3 x 10-4, carbon dioxide is by far the most abundant
carbon-containing gas in the atmosphere. Its atmospheric concentration is increasing by about
0.5% or roughly 1.6 ppm per year (Freyer, 1979). This increase is mainly due to the burning
of fossil fuels, which amounts to about 5.3 x 1015g C/yr (Rotty, 1981). Only about 58% of
this input remains airborne, but a better estimate can not be made because the source or sink
of CO2, connected with changes in the global biomass is not well known (see, e.g., Seiler and
Crutzen, 1980; Melillo and Gosz, Chapter 6, this volume).

Carbon dioxide is important for the radiative heat balance of the earth. A doubling of
atmospheric CO2, which may become established by the middle of the next century, can
cause a climatic warming by 1.56°C (Hansen et al., 1981). Contrary to the effects in the
troposphere, an increase in atmospheric CO2 leads to a cooling of the stratosphere, which
leads to increased ozone concentrations, partially offsetting the ozone depletions caused by
other anthropogenic activities, such as chlorofluorocarbon release. CO2 is not a
photochemically active gas, except insofar as it affects atmospheric temperatures. We will
restrict the discussion of it to what is said above and refer to the extensive reports of the past
years, e.g. the SCOPE 13 report (Bolin et al.; 1979).

3.4.2 Carbon Monoxide

This gas has somewhat variable concentrations in the atmosphere. High concentrations, about
150200 ppbv, are found near the earth's surface at the middle and high latitudes of the
Northern Hemisphere, which clearly suggests an anthropogenic source (Seiler, 1974). In the
Southern Hemisphere the CO volume mixing ratios are near 5060 ppbv southwards of 20°S
(Seiler, 1974; Heidt et al., 1980). Model calculations show, however, that the anthropogenic
source of carbon monoxide cannot be the dominant one. Because of reaction with OH, a very
large source of CO must be located in the equatorial regions that cannot be supplied from
local industrial emissions or transported from mid-latitudes. Large natural, mainly tropical,
sources of CO with a total strength of 23 x 1015g CO/yr must exist. The oxidation of methane
can only contribute about 25% of this.

Two important mechanisms of CO production in the tropics have, therefore, been postulated.
Zimmerman et al. (1978) proposed that the oxidation of isoprene (C5H8) and of terpenes
(C10H16), which are emitted by trees, constitute the required source. They estimated
worldwide isoprene and terpene emission rates of 3.5 x 1014 and 4.8 x 1014 g C/yr
respectively, and derived a range of global CO production rates of between 4 and 13 x
1014g/yr from their oxidation. The other important source of carbon monoxide in the tropics
comes from the burning of biomass, mainly due to a variety of land clearing operations, grass
fires in savannas, agricultural waste and firewood burning (Seiler and Crutzen, 1980). The
total estimated CO source from these fires was estimated at about 8 x 1014 g C/yr with an
uncertainty of at least a factor of 2 (Crutzen et al., 1979). Although the carbon monoxide
budget of the atmosphere may be explained in these terms, there are large uncertainties
connected with these numerical estimates, especially concerning the tropical emissions.

Recent work by Seiler and Fishman (1981) has identified seasonal variations in mid-latitudes
with CO maxima occurring during the winter and CO minima during the summer months.
This behaviour can be explained by the much more pronounced photochemical destruction of
CO by reaction with OH in summer, when this radical is calculated to be much more
abundant than during the winter season. This seasonal behavior may prove to be very useful
in deriving better information on the natural contributions of CO to the atmosphere from the
oxidation of the non-methane hydrocarbons that are also emitted by trees at maximum rates
during summer.

3.4.3 Methane

The annual destruction rate of methane in the atmosphere is calculated to be equal to 4 x 1014
g CH4 with an uncertainty range of maybe 26 x 1014 g CH4. This calculation is based on an
average tropospheric OH concentration of about 6 x 1015 molecules cm-3, which has an
uncertainty of a factor of two (Lovelock, 1977; Singh et al., 1979; Derwent and Eggleton,
1981; Gidel et al., 1982) in accordance with the methylchloroform balance of the atmosphere.
As uptake of methane by microbiological processes in soils and waters has never been found
to be significant (Ehhalt, 1974; Seiler, private communication), this rules against the
substantially larger sources of methane that have been proposed by Ehhalt (1974) and
recently by Sheppard et al. (1981). From this I derived the tentative budget of the biological
sources of methane, which is presented in Table 3.1. A substantial portion of the methane will
be oxidized in the atmosphere to CO, but in the absence of enough NO there will be a rain-
out of about 1014 g C as CH3O2H. It appears that several sources contribute comparable
quantities of CH4 to the atmosphere, and some of these sources are expanding world-wide.
For the period 19721978 (FAO, 1975, 1977, 1980), the cattle population increased by 1.2%
annually, the rice paddy area by 1.7% and the rice production by 4.6% per year. Of particular
importance as an interaction between biogeochemical cycles are the observations by Cicerone
and Shetter (1981) who observed a more than fourfold enhancement in CH4 yields when rice
fields were supplied with nitrogen fertilizer. Nevertheless, their global estimate of CH4
emissions from rice fields is less than 6 x 1013 g per year. It is interesting to note that biomass
burning is also a significant source of methane to the atmosphere, supplying between 30 and
110 Tg annually (Crutzen et al., 1979). It is quite likely that the extent of burning has also
grown substantially during the past decade, but no reliable statistics are available. An
additional production of methane of maybe 1.5 x 1014 g CH4 per year takes place in the
digestive systems of termites during the decay of wood (Zimmerman et al., 1982). This must
still be shown by field measurements.

An important, abiological source of CH4 is venting and leakage of natural gas. Current world-
wide natural gas consumption amounts to about 1015 g CH4/yr. A world-wide average
leakage rate of only 2% would supply 2 x 1013 g CH4 annually to the atmosphere. A good
statistical evaluation of the leakage rate has not been published, and the 2% number based on
unofficial information from gas supply companies is rather uncertain.

Observations by Rasmussen and Khalil (1981) have shown, an annual global increase in
methane by about 2% during 19781980. This increase has also been discovered by Seiler
(private communication) from data extending back to 1976. Spectroscopic measurements
from the Jungfraujoch in Switzerland seem to rule out an increase by more than 10% during
the time period 19551977 (Zander, private communication), so that the increase in
atmospheric methane may only have been of more recent dates. Carbon isotope ratio
measurements of CH4 and its potential atmospheric sources may be of great value in
identifying the source of the increase in atmospheric CH4 over the past years (Rust, 1981;
Stevens and Rust, 1981). Because of the important role of methane in both stratospheric and
tropospheric photochemistry, an increase in CH4 will significantly affect air chemistry.
Furthermore, methane plays a role in the earth's radiation budget (Wang et al., 1976).

3.4.4 Isoprene and Terpenes

Isoprene and terpenes seem to be the main volatile organic compounds that are emitted by
trees (Went, 1960; Zimmerman et al., 1978). They are important because they may be
converted to CO. Their residence time in the atmosphere is only a few hours. Extrapolating
from field measurements of emissions made in the U.S.A., Zimmerman et al. (1978)
estimated isoprene and terpene emission rates of 3.5 1014 and 4.8 1014 g C/yr respectively.
These are uncertain by a factor of two and it may not be excluded that other hydrocarbon
gases are also emitted in important quantities in the tropics.

Virtually nothing is known about the oxidation paths of these compounds in the atmosphere.
As discussed for methane, it appears that the availability of nitric oxide is an important factor
that determines the routes and time constants of their oxidation in the atmosphere. If too little
NO is present in the air, it is possible that many photochemically rather long-lived and water-
soluble gaseous intermediates (e.g. organic peroxides and alcohols) are formed. These
intermediates may be removed by precipitation so that there is no guarantee that the
efficiency of CO formation from each emitted carbon atom will be close to unity. Conversely,
any future additions of NO from industry and fires to the boundary layer of the tropical
forests could substantially shorten the time needed to break down C5H8 and C10H16 via
aldehydes to carbon monoxide. This may have implications for the future atmospheric CO
distribution. The accompanying effects on ozone formation have already been discussed.
There may even exist a potential for photochemical smog formation under stable
meteorological conditions during the dry season in the trophics.

3.5 THE MOST IMPORTANT NITROGEN COMPOUNDS

3.5.1 Nitric Oxide, Nitrogen Dioxide and Nitric Acid

I group these gases together because photochemical reactions establish an equilibrium


between NOx and NO2, after which HNO3 is formed via reaction (R10). The source of nitric
oxide to the troposphere will be discussed in the following sections.

A. High Temperature Combustion

In the SCOPE 7 report the global source of NO was estimated at 23 Tg N in 1975 (Söderlund
and Svensson, 1976). These estimates were obtained by an extrapolation of earlier 1965
emission estimates of 15 Tg N by Robinson and Robbins (1968). On the other hand, an
analysis of NOx emissions recently made by Böttger et al. (1980), estimates the global NOx
production from combustion engines to range between 8.2 and 18.5 Tg N/yr. For specific
continents and regions more detailed analyses have been compiled by Smith (1980). For
Western Europe, Smith reports an annual atmospheric emission rate of about 2 Tg N. For the
U.S.A. and Japan, production rates were estimated at 6.6 and 0.7 Tg N/yr respectively. A
consideration of these more recent figures casts some doubt on the emission estimates
by Söderlund and Svensson (1976) and demonstrates clearly an unsatisfactory state of affairs
that calls for more complete and up-dated analyses of the industrial NOx production rates.
Besides, because of the short lifetime of NOx in the atmosphere, regional budgets are
important for pollution studies. Globally, we will adopt here a range of 1220 Tg N per year as
the NOx production rate estimate from industrial high temperature combustion processes.

Considerable attention has recently also been given to the emissions of NOx by subsonic
aircraft above 9 km in the Northern Hemisphere, which is estimated at 0.25 Tg N/yr. (Liu et
al., 1980). Although this source of NO is much smaller than that produced by all other fossil
fuel combustion processes, the NOx will reside in the atmosphere much longer than the nitric
oxide that is produced at low altitudes, because of a lower probability of uptake of HNO3 in
cloud water. Liu et al. (1980) postulated, therefore, that the aircraft nitric oxide would
contribute significantly to the NOx concentrations in the upper troposphere together with the
nitric oxide that is of stratospheric origin. They also postulated that these sources of NOx
would exert a considerable effect on the tropospheric ozone. The validity of this argument
also depends, however, on the efficiency of transfer of the industrial NO and NO2 from the
ground to the upper troposphere. As NO and NO2 are not very soluble and do not react in
water (Lee and Schwartz, 1981), such transport may occur by occasional vigorous
overturning of the troposphere during frontal passages and thunderstorms. This cannot be
estimated by the current 1-D and 2-D models of the atmosphere that consider only averaged
motions and average deviations thereof, and that thereby may grossly under-estimate the
transfer of photochemically reactive gases to the middle and upper troposphere.

B. Biomass Burning in the Tropics

The nitric oxide resulting from this process is mainly a product of the oxidation of the fixed
nitrogen that was already bound in the biomas. This source of nitric oxide is hard to estimate,
but based on the studies by Crutzen et al. (1979) and by Seiler and Crutzen (1980), it should
be of the order of 20 Tg N/yr with about a factor of two uncertainty. This may only represent
30% of all fixed nitrogen that is contained in the burned biomass, so that there remains the
possibility of large atmospheric sources of other nitrogenous gases. One of these gases is
N2O; other possibilities are ammonia, hydrogen cyanide (HCN), and other nitriles. The
presence of HCN in the atmosphere was predicted by Crutzen et al. (1979), because of a long
atmospheric residence time and the likelihood of a substantial atmospheric source from
biomass burning (Schmeltz and Hoffmann, 1977). Its presence in the atmosphere above 12
km at an average volume mixing ratio of 0.17 ppbv has recently been reported (Coffey et al.,
1981b). The reaction coefficient of HCN with OH varies from about 3 x 10-14cm3 molecule-1
s-1 at ground level to slower than 7 x 10-15cm3 molecule-1 s-1 above 12 km (Fritz et al., 1981).
This means that the oxidation of HCN in the atmosphere can at most yield 0.1 Tg N/yr of
NOx, which is quite insignificant. Rain-out of HCN in the atmosphere may likewise be
neglected because of insufficient solubility of HCN in water (Landolt-Börnstein, 1962), so
that only uptake at the earth's surface can be a significant sink for atmospheric HCN. With a
maximum deposition velocity of about 0.51 cm/s on the ocean and at land (Liss and Slater,
1974), the transfer of fixed nitrogen from the atmosphere is at most about 10 Tg N/yr, which
may be of ecological significance provided HCN can be used efficiently as a fixed nitrogen
source.

C. Lightning
For this source of NO widely different estimates ranging from 4.2 TgN/yr (Tuck, 1976) to 40
TgN/yr (Chameides et al., 1977) were published several years ago. Most recent estimates are
about 34 TgN/yr (Dawson, 1980; Hill et al., 1980), or even 1.8 TgN/yr (Levine et al., 1980).
I estimate a wide range of possible values of 110 Tg N/yr, for which the more recent
estimates, being based on better statistics on lightning stroke energy dissipation and lightning
stroke frequency, are more heavily weighted.

D. Soil Exhalation

This source still awaits careful evaluation. Laboratory studies have shown that emission of
NOx is much dependent on soil acidity (Nelson and Bremner, 1970). The SCOPE 7 report
(Söderlund and Svensson, 1976) gives a range of 114 Tg N/yr for global emission, based on
two Russian studies, while Galbally and Roy (1978) propose 10 Tg N/year as a global rate
extrapolating from the results of their box measurements on some Australian soils. With so
little new information available, I propose a range of 115 Tg N/yr for the NOx release from
soils, which includes the possibility of no appreciable emission at all.

E. Ocean Release

Nitric oxide is also released from the ocean surface as a result of nitrite photolysis. The
source strength has been estimated at 0.15 Tg N/yr (Zafiriou and McFarland, 1981)

F. Stratospheric Production of NO from the Oxidation of N2O

There are many estimates available for this source of NO, depending on the assumptions
which are made about the stratospheric distribution of N2O, of which there are too few
measurements. The assumptions about the tropical distributions of N2O are especially critical
in making these estimates. From the analysis of one balloon flight at 9°N and several
observations at other latitudes, Schmeltekopf et al. (1977) derived an average global,
vertically integrated, production rate of NO of 1.6 Tg N/yr. The measured profiles of N2O
with high stratospheric N2O concentrations have been confirmed by three other equatorial
flights by the same research group (Goldan et al., 1980). However, the estimate of
Schmeltekopf et al. (1977) is too high by a factor of two, because these workers considered
the photochemical process to last for 24 hours each day. The analysis by Johnston et al.
(1979) avoided this error and led to a production of 1 Tg N/yr. This may also be somewhat on
the high side, however, as no proper account could be taken of the temperature dependence of
O(1D) quantum yield in O3 photolysis near 310 nm. If the estimate by Johnston et al. (1979)
has an uncertainty of about 40%, the production of NO from N2O oxidation may range
between 0.7 and 1.4 Tg N/yr. This source determines the abundance of NO in the stratosphere
and is critical for stratospheric ozone chemistry. In polar regions, galactic cosmic rays
produce NO at rates between 0.024 and 0.036 Tg N annually, depending of the phase of the
solar cycle. Sporadic solar cosmic ray events may occasionally produce almost ten times as
much NO per event as galactic cosmic rays in a whole year, but on the average this source of
NO is also small compared to that derived from N2O oxidation. The possibility of some
significant contribution to the stratospheric NOx budget from higher layers in the atmosphere
can, however, not be totally ruled out. In the ionosphere 200400 Tg N/yr of NOx are produced
by the ionizing action of short wave ultraviolet radiation and 20 Tg N/yr by auroral activity at
solar maximum (Crutzen, 1979). By far most of the fixed nitrogen produced in this way is
removed by the reactions
(R58) NO + hv N + O (<191 nm)
(R59) NO + N N2 + O

net: 2 NO N2+2ON2+O2

It is difficult to estimate how much fixed nitrogen may nevertheless escape this sink because
it depends on vertical exchange between the stratosphere and the layers above 100 km, about
which little is known. A rather small leakage would be sufficient to establish a significant
effect, and this may especially take place during the winter at high latitudes when reaction
(R58) does not take place.

3.5.2 Ammonia

Although ammonia may not participate in any significant homogeneous gas phase reactions
in the atmosphere, it is nevertheless of much importance in the chemistry of the troposphere
through its influence on rain-water chemistry (Taylor et al., Chapter 4, this volume). Most
fixed nitrogen is circulated in the atmosphere in the form of ammonia. Except during the
winter, ammonia is often present at ground levels over land with volume mixing ratios of 110
ppbv (Breeding et al., 1973; Shrendikar and Lodge, 1975; Georgii and Lenhard, 1978; Hoell
et al., 1980). Over the oceans much lower NH3 volume mixing ratios seem to be far more
common (Ayers and Grass, 1980) . Normally, a very marked drop of NH3 mixing ratios with
height is observed, so that typically volume mixing ratios below 1 ppbv are observed above 3
km (Kaplan, 1973; Georgii and Müller, 1974; Georgii and Lenhard, 1978). Exceptions are the
higher values which were measured in the eastern U.S.A. in March 1979 (Hoell et al., 1980).

The role and cycling of NH3 in the atmosphere are complex. Plants can both absorb
(Hutchinson et al., 1972) and emit NH3 (Farquhar et al., 1979). A study by Denmead et al.
(1976) showed emission of NH3 from the soil, but re-absorption of this NH3 within the plant
canopy. As a consequence of this, an appreciable loss of ammonia to the atmosphere was
detected from a grazed pasture, but only a small flux from an ungrazed one. Because of this
role of vegetation it is particularly hard to estimate the atmospheric budget of NH3. In the
following we can, therefore, only show the gross inputs of NH3 to the atmosphere, and even
for these the uncertainties are very substantial:

A. NH3 Release from Animal Excretions

Healey et al. (1970) listed animal urea excretion rates for most important domestic animals in
the U.K. From this information and statistics on the population of domestic animals (FAO,
1975), one can extrapolate to an urea production of about 50 Tg/yr in the developed world.

In the developing world the food intake per animal is, however, only 3060% of that in the
developed world, so that urea production per head should be about half as large as in the
developed countries. From the available statistics in the developing world one may calculate
an additional urea production of 20 Tg/yr, so that the total world-wide urea production by
domestic animals adds up to about 70 Tg or 35 Tg N/yr. Healey et al. (1970) assumed that
10% of the nitrogen in the urea would volatilize to NH3. This release estimate is, however,
very low in comparison to the results of field studies obtained by Denmead et al. (1974) and
especially Porter (1975) that indicated a relative volatilization loss of 30%. Adopting this
ratio as more representative, the global production of NH3 from urea would be about 10 Tg
N/yr. To this must be added the contribution from faeces, which may double the release of
ammonia to 20 Tg N/yr (Böttger et al., 1980). This estimate is somewhat smaller than those
derived by Söderlund and Svensson (1976) and Böttger et al. (1980). For the release of
ammonia from wild animals we will accept the relatively small source of 26 Tg N as given by
Söderlund and Svensson (1976).

B. Release of NH3 from Fertilized Fields

The loss of NH3 from fertilized agricultural fields is estimated to be about 5% (Bolin and
Arrhenius, 1977) . With an annual fertilizer application rate of almost 60 Tg N, the release of
NH3 is equal to about 3 Tg N/yr.

C. Ammonia Release from Uncultivated Fields

This is an unknown quantity. Dawson (1977) made a detailed analysis of biomass


decomposition and from a soilatmosphere exchange model proposed an NH3 source of 27 Tg
N/yr, with most emission taking place at temperate latitudes. Even if this emission was
roughly correct, substantial re-absorption of ammonia is probably going to take place as in
the study by Denmead et al. (1976) in the ungrazed pasture. Böttger et al. (1980) actually rule
out a significant source of this type on the basis of a few quoted measurements on NH3
release from soils, and Söderlund and Svensson (1976) do not even discuss it. Here I will
consider the value of Dawson (1977) as an upper limit. In particular, I do not rule out a
significant release of NH3 especially during early spring when fresh vegetation is not well
developed and dead material on the ground is abundant. When the soil is warmed by solar
radiation much release of NH3 to the atmosphere may indeed occur.

D. NH3 Release from Biomass Burning

Referring to the discussion given earlier regarding NO release, a volatilization of ammonia of


up to 60 Tg N/year that takes place especially in the tropics may not be ruled out. This
requires of course that much of the ammonia is not oxidized in the fires. Söderlund and
Svensson (1976) mention, however, that ammonia is a surprisingly stable gas in combustion
systems.

E. NH3 from Coal Burning

I will use the estimate of Söderlund and Svensson (1976), who proposed 412 Tg N as the
global source.

3.5.3 Nitrous Oxide

During the years that have passed since the publication of the SCOPE 7 report, much
knowledge has been collected about the questions related to N2O. Atmospheric measurements
and measurements of soil release now strongly indicate that the total global source of N2O
must be much smaller than some of the values that were listed by Söderlund and Svensson
(1976). Photochemical reactions in the stratosphere may, therefore, represent the main sink
for atmospheric nitrous oxide. For reasons stated before, the loss rate of 15 Tg N2O-N/yr of
Schmeltekopf et al. (1977) is probably too high by a factor of 2. Instead, Johnston et al.
(1979) estimated a loss rate of 8.4 Tg N/yr. With this loss rate the atmospheric lifetime of
N2O is about 180 years.
The processes that add significant and similar quantities of N2O to the atmosphere will be
discussed below.

A. Production by Fossil Fuel Combustion

Weiss and Craig (1976) estimated that 1.6 Tg N would be produced annually by coal and fuel
oil burning. To this must be added about 0.15 Tg N from natural gas burning as estimated by
Pierotti and Rasmussen (1976). The fossil fuel combustion source of N2O is increasing at a
rate of 3.5% per year. The increase in N2O that results from this source almost satisfies the
observed world-wide increase in N2O concentrations by about 0.2% per year (Weiss, 1981).
Additional, potentially important production of N2O may take place by the use of reducing
catalysts in automobile engines to cut down on NO emissions (Pierotti and Rasmussen, 1976;
Weiss and Craig, 1976). Weiss and Craig (1976) estimated a potential global rise in N2O
emissions rates by 2.1 Tg N/yr, if all 1976 year cars were to be equipped with Pt catalysts.

B. N2O Production from Biomass Burning

This source was estimated by Crutzen et al. (1979) at 8 Tg N/yr. More recent data (to be
published) indicate, however, a lower emission rate of 12 Tg N/yr.

C. N2O Production in the Oceans and Estuaries

According to Cohen and Gordon (1979) this source of N2O may range between 4 and 10 Tg
N per year, whereby the production of N2O would take place mainly during nitrification.
Elkins et al. (1978) estimate an N2O release of 7 Tg N/yr for the latitude belt 20°S to 20°N,
and in a recent review Hahn (1981) derives about 14 Tg N/yr for the entire ocean. Seiler
(1981) and Weiss (1981) from their extensive oceanic observations conclude, however, that
the oceanic source of atmospheric N2O is small compared to the stratospheric sink. Although
significant quantities of N2O may be released from estuaries (Kaplan et al., 1978), the N2O
production from oceans and estuaries is probably much lower than 10 Tg N/yr.

D. N2O production from the Cultivation of Natural Soils

According to estimates by Bohn (1977) and Wilson (1978), during the past century the
average annual, world-wide loss of organic carbon due to the ploughing of agricultural soils
has amounted to 12 x 1015sg C. With an average C/N ratio of 20 (Delwiche and Likens,
1977), the annual loss of fixed nitrogen from these soils has amounted to 50100 Tg N. In a
recent study Terry et al. (1980) reported that 2.7% of the fixed nitrogen that is lost in the
drained, cultivated soils of South Florida did appear as N2O. If similar ratios applied for other
ecosystems that lose organic nitrogen due to agricultural activities, the annual global source
of N2O would be equal to 13 Tg N. This source of N2O may, therefore, have been significant
during the past century and may have led to an increase in atmospheric N2O.

Recent studies generally show relative emissions of N2O nitrogen compared to fertilizer
nitrogen input typically of the order of 1% or less (McKenney et al., 1978; Breitenbeck et al.,
1980; Conrad and Seiler, 1980; Mosier and Hutchinson, 1981; Mosier et al., 1981a, b; Ryden,
1981; Seiler and Conrad, 1981). There is a tendency for the relative N2O release to be higher
for ammonium containing fertilizer, an indication for the importance of the nitrification
process as a source of atmospheric N2O. High yields of N2O (46.8%) were measured by
Bremner et al. (1981) after application of anhydrous ammonia. Altogether, the global source
of N2O from nitrogen fertilization is probably smaller than a few Tg N/yr. Thus the impact of
the increased nitrogen fertilization on stratospheric ozone is probably not a very urgent
matter. This view is supported by the observations by Weiss (1981) of a global upward trend
in atmospheric N2O of only 0.2% per year. In fact, Weiss explains that all the atmospheric
N2O increase is due to the increased rate of fossil fuel combustion.

E. N2O Release from Natural Soils

About this very little information is available. Seiler and Conrad (1981) normally found even
smaller N2O release rates (0.14g N/ha/day) in natural soils than in fertilized soils near Mainz,
Germany. Mosier et al. (1981b) derived an average N2O flux of 2.3 g N/ha/day from a native
prairie in N.E. Colorado during the summer months. In the latter case the loss of N2O
accounted for 10% of the N inputs from atmospheric deposition and N2 fixation. Earlier
studies quoted by Hahn and Junge (1977) gave release rates of N2O from unfertilized or
natural soils in the range 0.134 g N/ha/day. With such limited and highly variable information
it is clearly possible only to guess at global N2O releases from terrestrial ecosystems.

3.6 THE MOST IMPORTANT SULPHUR COMPOUNDS

3.6.1 Sulphur Dioxide

A. Anthropogenic Sources of SO2

A detailed compilation of the industrial atmospheric sources of SO2 was made by Cullis and
Hirschler (1980). The man-made sources for 1976 add up to 104 Tg S/yr, of which 62%
comes from coal, 25% from petroleum, 11% from non-ferrous ores, and 2% from a variety of
other industries.

B. SO2 Production from Biomass Burning

Another potential source of sulphur dioxide could be biomass burning. With typical S/C
atomic ratios of about 0.002 in biomass, the global emission of sulphur from this source
could be about 6 Tg S/yr with an uncertainty of a factor of two. SO2 may not be the dominant
sulphur gas produced in fires, because an appreciable fraction of more reduced gases, such as
H2S and COS, may also be emitted.

C. SO2 Release from Volcanoes

An important natural source of SO2 to the atmosphere is provided by volcanic eruptions.


Starting from the very low production estimates of 1.5 Tg SO2/yr by Kellogg et al. (1972),
the estimated SO2 inputs have been going up: 7.5 Tg SO2/yr (Cadle, 1975), 47 TG SO2/yr
(Naughton et al., 1975) and maybe 100 Tg SO2/yr (Cadle, 1980). An earlier estimate by
Bartels (1972), based on elemental analysis of the Greenland icecap, was 34 Tg SO2/yr.
Therefore, volcanic emissions apparently contribute 1030% of the anthropogenic SO2 input to
the atmosphere. During periods with intense volcanic activity the contributions may be
substantially larger.

3.6.2 The Reduced Sulphur Gases H2S, (CH3)2S, CH3SH, COS and CS2
The SCOPE 7 compilation by Granat et al. (1976) gives a total annual emission rate of
biogenic, reduced sulphur of 37 Tg S/yr, about 90% of which emanates from the oceans or
coastal waters. Cullis and Hirschler (1980) give, however, appreciably higher biogenic
sulphur fluxes of almost 100 Tg S annually, about half of which comes from the land. The
latter result seems to have been obtained by a questionable extrapolation of some measured
sulphur fluxes over fertilized agricultural fields in Sweden (Siman and Jansson,1976) to all
ecosystems of the world. At this stage it is, therefore, hard to assign any significance to the
H2S release rates given by Cullis and Hirschler, except their value for agricultural soils in the
developed world of about 4 Tg S annually. Biogenic sulphur flux estimates over non-saline
soils in the eastern U.S.A. by Adams et al. (1980) show large variabilities by orders of
magnitude, but average emission rates were estimated at 0.2 gm-2yr-1 for coastal wetlands, 0.3
gm-2yr-1 for inland soils high in organic matter and 0.015 g-2yr-1 for more `typical' inland soils.
For all soils in the study area an average emission rate of 0.03 gm-2yr-1 was calculated. This is
about a factor of ten smaller than the sulphur emission of 0.20.3 kg m-2yr-1 obtained by Siman
and Jansson (1976) for fertilized Swedish fields, so that the contribution of agricultural
sulphur emissions by Cullis and Hirschler (1980) should be considered as an upper limit.

Airborne observations by Crawford and Reisinger (1980) in the South-eastern United States
have indicated the possibility of substantial biogenic sulphur fluxes from wetland areas of the
order of 12 g Sm-2yr-1. If such a flux were characteristic for most terrestrial wetlands of the
world with a total estimated area of 2 x 1012m2, an annual flux of 24 Tg S would result.
Considering the much lower emissions given below, this estimate may be on the high side.

There exists a large variability in reported sulphur emissions from salt marshes and tidal flats,
with emission values quoted as 0.51940 gm-2 yr-1 (Adams et al., 1980), 39 gm-2yr-1 (Steudler
and Peterson, 1980) and <0.20.6 gm-2yr-1 (Aneja et al., 1979a). To extrapolate from these
values to global estimates may be hopeless. More promising are methods that average over
emissions from larger areas of an ecosystem. For a salt marsh at Wallops Island, Goldberg et
al. (1981) estimated average H2S emission rates of 0.4 g S m-2yr-1 in December and 6 g Sm-
2 -1
yr in July, maybe due to the expected temperature dependence. The value of these
estimates is, clearly dependent on good interpretation of the meteorology during the
particular observation period. Only a limited analysis of this could be made by Goldberg et
al. (1981). If I derive from their work average emission rates of 3 and 6 gm-2yr-1 for the
temperate and the tropical latitudes respectively, considering that these fluxes would
represent sulphur emissions in estuaries, in algal bed and reef areas and oceanic upwelling
zones with a combined area of 2.4 x 1012m2 (Whittaker and Likens, 1975), the production of
sulphur from these ecosystems could almost reach 10 Tg S/yr.

Sulphur emission rates from the open oceans have been estimated by Bonsang (1980) to be in
the range 3142 Tg S annually. This range is larger than the estimate of 27 Tg S/yr of Granat
et al. (1976). As only emissions of dimethylsulphide could be measured, this range may have
to be shifted upwards.

Relatively large H2S emissions may occur in the tropics, but much of the emitted sulphur may
be recycled within the tropical forest stands (Brinkmann and Santos, 1974). Relatively large
H2S emissions of 16 Tg S/yr have, nevertheless, been proposed to occur in wet tropical
forests (Delmas et al., 1980). These emissions give rise to precipitation with a pH range of
4.75.7 (Stallard and Edmond, 1981).
Clearly, the world-wide biogenic and volcanic emissions are poorly known, but together they
could be in the range 70100 Tg S/yr. Thus, on a global average, they may turn out to be quite
comparable to the industrial emissions. In the heavily industrialized areas of the Northern
Hemisphere, the industrial emissions should, however, be much larger than the natural
emissions, (e.g. Granat et al., 1976; Galloway and Whelpdale, 1980).

The importance of COS for the stratospheric sulphate layer has been discussed before.
Besides an estimated loss rate of about 5 x 1010 S/yr by photolysis in the stratosphere, little is
known about the sources and sinks of this compound. Possible sources include emissions
from wetland areas (Aneja et al., 1979a, b; Adams et al., 1980), coal processing, volcanic and
fumarolic emissions (Crutzen, 1976), biomass burning (Crutzen et al., 1979) and oxidation of
CS2. Besides through photochemical breakdown in the stratosphere, COS may also be
removed from the atmosphere by deposition on soils and vegetation and on the ocean surface.
Decomposition and subsequent hydrolysis in the ocean may limit the lifetime of COS to 10
years (Johnson, 1981), which is much shorter than the lifetime of about 30 years due to
stratospheric photolysis. It appears, however, that the ocean surface waters are always
supersaturated with COS, so that they supply about 0.4 Tg S annually to the atmosphere
(Rasmussen, private communication). Altogether, the global source may exceed 1 Tg S per
year, so that COS must be taken up at the earth's surface, most probably by plants. Graedel et
al. (1981) have pointed to the potential role of COS as a corrosive gas.

The industrial release of CS2 to the atmosphere in the U.S.A. is equal to 1.35 x 1011g/yr
(Peyton et al., 1976), which may be extrapolated to a world-wide, annual release of about 4 x
1011g. The resulting production of COS via reactions (R32a-c) is equal to 1.7 x 1011g S/yr,
emphasizing that COS may be influenced by anthropogenic activities. It is easy to show that
the average measured volume mixing ratio of 30 pptv in the continental U.S.A. boundary
layer (02 km) in May and June is compatible with a mean OH concentration of about 1.5 x
106 molecules cm-3, so that anthropogenic emissions may well explain much of the
observations of Bandy et al. (1981).

3.7 CONCLUSIONS

There are many photochemical interactions between the atmospheric cycles of carbon,
nitrogen, and sulphur containing species, so that these species cannot be studied
independently of each other. In the troposphere the interactions between carbon and nitrogen
compounds contribute significantly to the abundance of tropospheric ozone. The photolysis
of ozone by ultraviolet solar radiation leads to the formation of the highly reactive OH
radical, which is the only tropospheric compound capable of attacking such important gases
as CO, NOx H2S, and the hydrocarbons, including those containing sulphur and halogens.
Without the presence of ozone in the troposphere, the atmospheric abundance of all these
gases would have been much higher than observed, in several cases by orders of magnitude.
This is clearly of profound importance in air chemistry. Because of man's activities the
factors determining the photochemistry of tropospheric ozone are changing. Thus, it is
important to improve our understanding of the distribution of this gas in the troposphere,
which also includes better knowledge about the flux of stratospheric ozone to the
troposphere.

Using the currently available knowledge of tropospheric chemistry I have discussed


important photochemical reactions in the atmosphere and have also derived estimates of the
sources and sinks of several important atmospheric trace gases as compiled in Table 3.1 and
Figure 3.1. The influence of man's activities on most of the processes is substantial. The
fluxes of compounds to the stratosphere are relatively small, but they are, in several cases, of
fundamental importance for the photochemistry of the stratosphere. Some compounds that are
inert in the troposphere, such as N2O, CFCl3, CF2Cl2 and CCl4, are of special importance as
their photochemical breakdown in the stratosphere leads to the reactive radicals NO, NO2, Cl
and ClO that attack ozone through catalytic chains of reactions. Therefore, even small inputs
of these and other stable compounds at the earth's surface may have profound effects on
upper atmospheric chemistry.

Finally, due to expansions in agriculture during the past century the soils have lost about 12 x
1015g C/yr (Bohn, 1977; Wilson, 1978). With an N/C ratio of 0.05 (Delwiche and Likens,
1977), this implies a loss of 50100 Tg N each year from agricultural soils, which is
comparable to the present use of artificial nitrogen fertilizer. The fate of these two sources of
fixed nitrogen in the environment may well have similarities. In the same process also about
3550 Tg S have been affected annually, a substantial fraction of which has moved from the
agricultural soils into different environments (see McGill and Christie, Chapter 9, this
volume).

ACKNOWLEDGEMENTS

For helpful discussions and comments I thank Drs. R. Conrad, J. Lovelock, H. Niki, W.
Seiler, Mr. R. Chatfield and the participants at this SCOPE meeting. Special thanks go to Dr.
A. F. Tuck for this thorough and generously constructive review of this paper.

3.8 REFERENCES

Adams, D. F., Farwell, S. O., Pack, M. R., and Robinson, E. (1980) Biogenic Sulphur gas
emissions from soils in Eastern and Southeastern United States. Paper 80-40.5, 16 pp, 73rd
Annual Meeting of the Air Pollution Control Association, Montreal, Quebec, June 22-27.

Aneja, V. P., Overton, J. H., Cupitt, L. T., Durham, J. L., and Wilson, W. E. (1979a) Direct
measurement of emission rates of some atmospheric biogenic sulphur compounds, Tellus, 31,
174-178.

Aneja, V. P., Overton, J. H., Cupitt, L. T., Durham, J. L., and Wilson, W. E. (1979b) Carbon
disulphide from biogenic sources and their contribution to the global sulphur cycle, Nature,
282, 493-496.

Angell, J. K., and Korshover, J. (1979) Comparison of ozone variations derived from
ozonosondes and Umkehr measurements for the period 1969-1976, Mon. Weather Rev., 107,
599-607.

Atkinson, R., Perry, R. A., and Pitts, J. N., Jr. (1978) Rate constants for the reaction of OH
radicals with COS, CS2 and CH3SCH3 over the temperature range 299-430K, Chem. Phys.
Lett., 54, 14-18.

Atkinson, R., Darnall, K. R., Lloyd, A. C., Winer, A. M., and Pitts, J. N., Jr. (1979) Kinetics
and Mechanisms of the reaction of the hydroxyl radical with organic compounds in the gas
phase, Adv. Photochem., 11, 375-488.
Ayers, G. P. and Grass, J. L. (1980) Ammonia gas concentrations over the Southern Ocean,
Nature, 284, 539-540.

Bandy, A. R., Maroulis, P. J., Shalaby, L. and Wilner, L. A. (1981) Evidence for a short
tropospheric residence time for carbon disulphide, Geophys. Res. Lett., 8, 1180-1183.

Bartels, O. G. (1972) An estimate of volcanic contributions to the atmosphere and volcanic


gases and sublimates as the source of the radio isotopes 10Be, 35S, 32P and 22Na, Health Phys.,
22, 387-392.

Bentley, M. D., Douglass, I. B., Lacadie, J. A., and Whittier, D. R. (1972) The photolysis of
Dimethyl sulphide in air, J. Air Poll. Control Assoc., 22, 359-363.

Bohn, H. L. (1977) On organic soil carbon and CO, Tellus, 30, 472-475.

Bolin, B., and Arrhenius, E. (1977) Nitrogenan essential life factor and a growing
environmental hazard, Ambio, 6, 96-105.

Bolin, B., Degens, E. T., Duvigneaud, P., and Kempe, S. (eds) (1979) The Global Carbon
Cycle, SCOPE Report No. 16, New York, Wiley.

Bonsang, B. (1980) Cycle Atmosphérique du soufre d'origine marine. Thèse de Doctorat


D'Etat es Sciences Physiques, Université de Picardie.

Böttger, A., Ehhalt, D. H., and Gravenhorst, G. (1980) Atmosphärische Kreisläufe von
Stickoxyden and Ammoniak, Report Jül-1558, Institut für Chemie 3, Kernforschungsanlage
Jülich (FRG).

Breeding, R. J., Lodge, Jr., J. P., Pate, J. B., Sheesley, D. C., Klonis, H. B., Fogle, B.,
Anderson, J. A., Englert, R. T., Haagenson, P. L., McBeth, R. B., Morris, A. L., Pogue, R.,
and Wartburg, A. F. (1973) Background trace gas concentrations in the Central United States,
J. Geophys. Res., 78, 7057-7064.

Breitenbeck, G. A., Blackmer, A. M. and Bremner, J. M. (1980) Effects of different nitrogen


fertilizers on emission of nitrous oxide from soil, Geophys. Res. Lett. 7, 85-88.

Bremner, J. M., Breitenbeck, G. A. and Blackmer, A. M. (1981) Effect of anhydrous


ammonia fertilization on emission of nitrous oxide from soil, J. Environ. Qual., 10, 77-80.

Brinkmann, W. L. F. and Santos, V. de M. (1974) The emission of biogenic hydrogen


sulphide from Amazonian floodplain Lakes, Tellus, 26, 262-267.

Cadle, R. D. (1975) Volcanic emissions of halides and sulphur compounds to the troposphere
and stratosphere, J. Geophys. Res., 80, 1650-1652.

Cadle, R. D. (1980) A comparison of volcanic with other fluxes of atmospheric trace gas
constituents, Rev. Geophys. Space Phys.,18, 746-752.

Calvert, J. G., Su, F., Bottenheim, J. W., and Strausz, O. P. (1978) Mechanism of the
homogeneous oxidation of sulphur dioxide in the troposphere, Atmos. Environ., 12, 197-226.
Cicerone, R. J., and Shetter, J. D. (1981) Sources of atmospheric methane: measurements in
rice paddies and a discussion, J. Geophys. Res., 86, 7203-7209.

Chameides, W. L., Stedman, D. H., Dickerson, R. R., Rusch, D. W., and Cicerone, R. J.
(1977) NOx production in lightning, J. Atmos. Sci., 34, 143-149.

Coffey, M. T., Mankin, W. G., and Goldman, A. (1981a) Simultaneous spectroscopic


determination of the latitudinal, seasonal and diurnal variability of stratospheric N2O, NO,
NO2 and HNO3, J. Geophys. Res., 86, 7331-7341.

Coffey, M. T., Mankin, W. G. and Cicerone, R. J. (1981b) Spectroscopic detection of


stratospheric hydrogen cyanide (HCN), Science (submitted).

Cohen, Y., and Gordon, L. I. (1979) Nitrous oxide production in the ocean, J. Geophys. Res.,
84, 347-353.

Conrad, R. and Seiler, W. (1980) Field measurements of the loss of fertilizer nitrogen into the
atmosphere as nitrous oxide, Atmos. Environ., 14, 555-558.

Cox, R. A., and Roffey, M. J. (1977) Thermal decomposition of peroxyacetyl nitrate in the
presence of nitric oxide, Environ. Sci. Technol., 11, 900-906.

Cox, R. A. and Sandalls, F. J. (1974) The photo-oxidation of hydrogen sulphide and dimethyl
sulphide in air, Atmos. Environ., 8, 1269-1281.

Crawford, T. L., and Reisinger, L. M. (1980) Transport and transformation of sulphur oxides
through the Tennessee Valley Region, EPA600/780126, June 1980.

Crutzen, P. J. (1970) The influence of nitrogen oxides on the atmospheric ozone content,
Q.J.R. Meteorol. Soc., 96, 320-325.

Crutzen, P. J. (1971) Ozone production rates in an oxygenhydrogennitrogen oxide


atmosphere, J. Geophys. Res., 76, 7311-7327.

Crutzen, P. J. (1973). A discussion of the chemistry of some minor constituents in the


stratosphere and troposphere, PAGEOPH, 106-108, 1385-1399.

Crutzen, P. J. (1974) Estimates of possible variations in total ozone due to natural causes and
human activities, Ambio, 3, 201-210.

Crutzen, P. J. (1976) The possible importance of CSO for the sulphate layer of the
stratosphere, Geophys. Res. Lett., 3, 73-76.

Crutzen, P. J. (1979) The role of NO and NO2 in the chemistry of the troposphere and
stratosphere, Ann. Rev. Earth Planet. Sci., 7, 443-472.

Crutzen, P. J., Fishman, J., Gidel, L. T., and Chatfield, R. B. (1978) Numerical investigations
of the photochemical and transport processes which affect halocarbons and ozone in the
atmosphere. Annual Summary of Research, Department of Atmospheric Sciences, Colorado
State University, Fort Collins, Colorado.
Crutzen, P. J., Heidt, L. E., Krasnec, J. P., Pollock, W. H., and Seiler, W. (1979) Biomass
burning as a source of atmospheric gases CO, H2, N2O, NO, CH3Cl and COS, Nature, 282,
253-256.

Crutzen, P. J., and Howard, C. J. (1978) The effect of the HO2 + NO reaction rate constant on
one-dimensional model calculations of stratospheric ozone perturbations, PAGEOPH, 116,
497-510.

Cullis, C. F., and Hirschler, M. M. (1980) Atmospheric sulphur: natural and man-made
sources, Atmos. Environ., 14, 1263-1278.

Davis, D. D., and Klauber, G. (1975) Atmospheric gas phase oxidation mechanisms for the
molecule SO2, Int. J. Chem. Kin. Symp., Series 1, 543-556.

Davis, D. D., Ravishankara, A. R., and Fischer, S. (1979) SO2 oxidation via the hydroxyl
radical: atmospheric fate of HSOx radicals, Geophys. Res. Lett., 6, 113-116.

Dawson, G. A. (1977) Atmospheric Ammonia from undisturbed land, J. Geophys. Res. 82,
3125-3133.

Dawson, G. A. (1980) Nitrogen fixation by lightning, J. Atmos. Sci., 37, 174-178.

Delmas, R., Baudet, J., Servant, J., and Baziard, Y. (1980) Emissions and concentrations of
hydrogen sulphide in the air of the tropical forest of the Ivory Coast and temperature regions
in France, J. Geophys. Res., 85, 4468-4474.

Delwiche, C. C., and Likens, G. E. (1977) Biological response to fossil fuel combustion
products, in Global Chemical Cycles and their Alteration by Man, W. Stumm, (ed.), Berlin,
Dahlem Konferenzen.

Demerjian, K. L., Kerr, J. A., and Calvert, J. G. (1974) The mechanism of photochemical
smog formation, Adv. Environ. Sci. Technol., 4, 1-262.

Denmead, O. T., Freney, J. R., and Simpson, J. R. (1976) A closed ammonia cycle within a
plant canopy, Soil Biol. Biochem., 8, 161-164.

Denmead, O. T., Simpson, J. R., and Freney, J. R. (1974) Ammonia flux into the atmosphere
from a grazed pasture, Science, 185, 609-610.

Derwent, R. G., and Eggleton, A. E. J. (1981) Two-dimensional model studies of methyl


chloroform in the troposphere, Q.J.R. Meteorol. Soc., 107, 231-242.

Duce, R. A. (1982) Biochemical cycles and air/sea exchanges of aerosols, Chapter 16 this
volume.

Duewer, W. H., Wuebbles, D. J., Elsaeeser, H. W., and Chang, J. S. (1977) NOx catalytic
ozone destruction: sensitivity to rate coefficients, J. Geophys. Res., 82, 935-942.

Ehhalt, D. H. (1974) The atmospheric cycle of methane, Tellus, 26, 58-70.


Elkins, J. W., Wofsy, S. C., McElroy, M. B., Kolb, C. E., and Kaplan, W. A. (1978) Aquatic
sources and sinks of nitrous oxide, Nature, 275, 602-604.

Fabian, P., and Pruchniewicz, P. G. (1977) Meridional distribution of ozone in the


troposphere and its seasonal variations, J. Geophys. Res., 82, 2063-2073.

FAO (Food and Agricultural Organisation) (1975) Production Yearbook 1974, 28, Rome,
FAO.

FAO (Food and Agricultural Organisation) (1977) Production Yearbook 1976, 30, Rome,
FAO.

FAO (Food and Agricultural Organisation) (1980) Production Yearbook 1979, 33, Rome,
FAO.

Farquhar, G. O., Wetselaar, R., and Firth, P. M. (1979) Ammonia volatilization from
senescing leaves of maize, Nature, 203, 1257-1258.

Fishman, J., Solomon, S., and Crutzen, P. J. (1979a) Observational and theoretical evidence
in support of a significant in-situ photochemical source of tropospheric ozone, Tellus, 31,
432-446.

Fishman, J., Ramanathan, V., Crutzen, P. J., and Liu, S. C. (1979b) Tropospheric ozone and
climate, Nature, 282, 818-820.

Freyer, H. D. (1979) Variations in the atmospheric CO2 content, in: Bolin, B., Degens, E. T.,
Kempe, S., and Ketner, P., (eds) The Global Carbon Cycle, SCOPE Report No. 13,
Chichester, Wiley, 79-99.

Fritz, B., Lorenz, K., Steinert, W., and Zellner, R. (1981) Rate and probable mechanism of
the tropospheric oxidation of HCN by OH radicals, Presented at the Second Meeting
(working party 2) of COST 61 a bis, Leuven, Belgium, Feb. 11-12,1981.

Galbally, I. E., and Roy, C. R. (1978) Loss of fixed nitrogen from soils by nitric oxide
exhalation, Nature, 275, 734-735.

Galloway, J. N., and Whelpdale, D. M. (1980) An atmospheric sulphur budget for eastern
North America, Atmos. Environ., 14, 409-417.

Garland, J. A. and Penkett, S. A. (1976) Absorption of peroxy acetyl nitrate and ozone by
natural surfaces, Atmos. Environ., 10, 1127-1131.

Georgii, H. W., and Lenhard, U. (1978) Contribution to the atmospheric NH3 budget,
PAGEOPH, 116, 385-392.

Georgii, H. W., and Müller, W. J. (1974) On the distribution of ammonia in the middle and
lower troposphere, Tellus, 26, 180-184.
Gidel, L. T., Crutzen, P. J., and Fishman, J. (1982) A two-dimensional photochemical model
of the atmosphere. I. Chlorocarbon emissions and their effect on stratospheric ozone, J.
Geophys. Res. (submitted).

Goldan, P. D., Kuster, W. C., Albritton, D. L., and Schmeltekopf, A. L. (1980) Stratospheric
CFCl3, CF2Cl2 and N2O height profile measurements at several latitudes, J. Geophys. Res.,
85, 413-423.

Goldberg, A. B., Maroulis, P. J., Wilner, L. A., and Brandy, A. R. (1981) Study of H2S
emissions from a salt water marsh. Atmos. Environ., 15, 11-18.

Graedel, T. E., Kammlott, G. W., and Franey, J. P. (1981) Carbonyl sulphide: potential agent
of atmospheric sulphur corrosion, Science, 212, 663-665.

Granat, L., Rodhe, H., and Hallberg, R. O. (1976) The global sulphur cycle. in Svensson, B.
H., and Söderlund, R. (eds) Nitrogen, Phosphorus and SulphurGlobal Cycles. SCOPE Report
No. 7, Ecol. Bull. (Stockholm), 22, 89-134.

Hack, W., Schacke, H., Schröter, M., and Wagner, H. Gg. (1978) Reaction of NH2-radicals
with NO, NO2, C2H2, C2H4 and other hydrocarbons, 17th Symposium (International) on
Combustion, 505-513.

Hahn, J. (1981) Nitrous oxide in the oceans, in Delwiche, C. C. (ed.) Denitrification,


Nitrification, and Atmospheric Nitrous Oxide, New York, Wiley, 191-277.

Hahn, J., and Junge, C. (1977) Atmospheric nitrous oxide: a critical review, Z. Naturforsch.,
32a, 190-214.

Hansen, J. E., Johnson, D., Lacis, A., Lebedeff, S., Lee, P., Rind, D., and Russell, G. (1981)
Climate impact of increasing atmospheric CO2, Science (submitted).

Hanst, P. L., Spiller, L. L., Watts, D. M., Spence, J. W., and Miller, M. F. (1975) Infrared
measurement off fluorocarbons, carbon tetrachloride, carbonyl sulphide and other
atmospheric trace gases, J. Air Pollut. Control Ass. 25, 1220-1226.

Healey, T. G., McKay, H. A. C., Pilbeam, A., and Scargill, N. D. (1970) Ammonia and
ammonium sulphate in the troposphere over the United Kingdom, J. Geophys. Res., 75, 2317-
2321.

Heidt, L. E., Krasnec, J. P., Lueb, R. A., Pollock, W. H., Henry, B. E., and Crutzen, P. J.
(1980) Latitudenal distributions of CO and CH4 over the Pacific, J. Geophys. Res., 85, 7329-
7365.

Hendry, D. G., and Kenley, R. A. (1977) Generation of peroxy radicals from peroxynitrates
(RO2NO2). Decomposition of peroxyl nitrates, J. Am. Chem. Soc., 99, 3198-3199.

Hill, R. D., Rinker, R. G., and Wilson, H. D. (1980) Atmospheric nitrogen fixation by
lightning, J. Atmos. Sci., 37, 179-192.
Hoell, J. M., Harward, C. N., and Williams, B. S. (1980) Remote infrared heterodyne
radiometer measurements of atmospheric ammonia profiles, Geophys. Res. Lett. 7, 313-316.

Hofmann, D. J., and Rosen, J. M. (1981) On the background stratospheric aerosol layer, J.
Atmos. Sci., 38, 168-181.

Howard, C. J., and Evenson, D. K. (1977) Kinetics of the reaction of HO2 with NO, Geophys.
Res. Lett., 4, 437-440.

Huebert, B. J., and Lazrus (1978) Global tropospheric measurements of nitric acid vapor and
particulate nitrate, Geophys. Res. Lett., 5, 577-580.

Hutchinson, G. L., Millington, R. J., and Peters, D. B. (1972) Atmospheric ammonia:


absorption by plant leaves, Science, 175, 771-772.

Inn, E. C. Y., Vedder, J. F., and Tyson, B. J. (1979) COS in the stratosphere, Geophys. Res.
Lett., 6, 191-193.

Johnston, H. (1971) Reduction of stratospheric ozone by nitrogen oxide catalysis from SST
exhaust, Science, 173, 517-522.

Johnston, H. S., Serang, O., and Podolske, J. (1979) Instantaneous global nitrous oxide
photochemical rates, J. Geophys. Res., 84, 5077-5082.

Johnson, J. E. (1981) The lifetime of carboxyl sulphide in the troposphere, Geophys. Res.
Lett., 8, 938-940.

Jones, B. M. R., Burrows, J. P., Cox, R. A., and Penkett, S. A. (1982) The atmospheric photo-
oxidation of CS2, Atmos. Environ. (submitted).

Kaplan, L. D. (1973) Background concentrations of photochemically active trace constituents


in the stratosphere and upper atmosphere, PAGEOPH, 106-108, 1342-1345.

Kaplan, W. A., Elkins, J. W., Kolb, C. E., McElroy, M. B., Wofsy, S. C., and Duran, A. P.
(1978) Nitrous oxide in fresh water systems: an estimate for the yield of atmospheric N2O
associated with disposal of human waste, PAGEOPH, 116, 423-438.

Kellogg, W. W., Cadle, R. D., Allen, E. R., Lazrus, A. L., and Martell, E. A. (1972) The
sulphur cycle, Science, 175, 587-596.

Kurasawa, H., and Lesclaux, R. (1980) Rate constant for the reaction of NH2 with ozone in
relation to atmospheric processes, Chem. Phys. Lett. 72, 437-442.

Kurylo, M. J. (1978) Flash photolysis resonance fluorescence investigation of the reaction of


OH radicals with dimethyl sulphide, Chem. Phys. Lett. 58, 233-237.

Landolt-Börnstein (1962) Zahlenwerte und Funktionen, Vol. 5, Berlin, Springer-Verlag, 629.

Lazrus, A. L., and Gandrud, B. W. (1974) Distribution of stratospheric nitric acid vapor, J.
Atmos. Sci., 31,1102-1108.
Lee, Y. N and Schwartz, S. E. (1981) Evaluation of the rate of uptake of nitrogen dioxide by
atmospheric and surface liquid water, J. Geophys. Res. (submitted).

Lesclaux, R., and Demissy, M. (1977) On the reaction of NH2 radicals with oxygen, Nouveau
J. Chim., 1, 443-444.

Levine, J. S., Rogowski, R. S., Gregory, G. L., Howell, W. E., and Fishmann, J. (1980)
Simultaneous measurements of NOx, NO and O3 production in a laboratory discharge:
atmospheric implications, Geophys. Res. Lett., 8, 357-360.

Levine, S. Z., and Schwartz, S. E. (1981) In-cloud and below-cloud scavenging of nitric acid
vapor, Atmos. Environ. (in press).

Levy, II, H. (1971) Normal atmosphere: Large radical and formaldehyde concentrations
predicted, Science, 173, 141-143.

Levy, II, H. (1974) Photochemistry of the troposphere, Adv. Photochem., 9, 5325-5332.

Liss, P. S., and Slater, P. G. (1974) Flux of gases across the airsea interface, Nature, 247,181-
184.

Liu, S. C., Kley, D., McFarland, M., Mahlman, J. D., and Levy II, H. (1980) On the origin of
tropospheric ozone, J. Geophys. Res., 85, 7546-7552.

Logan, J. A., McElroy, M. B., Wofsy, S. C., and Prather, M. J. (1979) Oxidation of CS2 and
COS: sources for atmospheric SO2, Nature, 281, 185-188.

Logan, J. A., Prather, M. J., Wofsy, S. C., and McElroy, M. B. (1978). Atmospheric
Chemistry: response to human influence, Phil. Trans. R. Soc. (London), 290, 187-234.

Lonnemann, W. A., Bufalini, J. J., and Seila, R. L. (1976) PAN and oxidant measurements in
ambient atmospheres, Env. Sci. Technol., 10, 374-380.

Lovelock, J. E. (1977) Methyl chloroform in the troposphere as an indicator of OH radical


abundance, Nature, 267, 32-34.

McConnell, J. C., McElroy, M. B., and Wofsy, S. C. (1971) Natural sources of atmospheric
CO, Nature, 233, 187-188.

McConnell, J. C. (1973) Atmospheric ammonia, J. Geophys. Res., 78, 7112-7821.

McElroy, M. B., and McConnell, J. C. (1971) Nitrous oxide: a natural source of stratospheric
NO, J. Atmos. Sci., 28,1095-1098.

McFarland, M., Kley, D., Drummond, J. W., Schmeltekopf, A. L., and Winkler, R. H. (1979)
Nitric oxide measurements in the equatorial Pacific region, Geophys. Res. Lett., 6, 605-008.

McGill, W. B., and Christie E. K. Biogeochemical aspects of nutrient cycle interactions in


soil and organisms, Chapter 9, this volume.
McKenney, O. J., Wade, D. L., and Findlay, W. I. (1978) Rates of N20 evolution from N-
fertilized soil, Geophys. Res. Lett., 5, 777-780.

Maroulis, P. J., Torres, A. L., Goldberg, A. B., and Bandy, A. R. (1980) Atmospheric SO2
measurements on project Gametag, J. Geophys. Res., 85, 7345-7349.

Meixner, F. (1981) Die vertikale Verteilung des atmosphärischen Schwefel-dioxides im


Tropopausenbereich. Doktor-Dissertation, Johann-Wolfgang-Goethe-Universität, Frankfurt-
am-Main.

Mosier, A. R., and Hutchinson, G. L. (1981) Nitrous oxide emissions from cropped fields, J.
Environ. Qual., 10, 169-173.

Mosier, A. R., Hutchinson, G. L., Sabey, B. R., and Baxter, J. (1981a) Nitrous oxide
emissions from barley plots treated with ammonium nitrate or sewage sludge, J. Environ.
Qual. (in press).

Mosier, A. R., Stillwell, M., Parton, W. J., and Woodmansee, R. G. (1981b) Nitrous oxide
emissions form a native shortgrass prairie, Soil Sci. Soc. Am. J. (submitted).

Naughton, J. J., Lewis, V., Thomas, D., and Finlayson, J. B. (1975). Fume compositions
found at varies stages of activity at Kilauea volcano, Hawaii, J. Geophys. Res., 80, 2963-
2966.

Nelson, D. W., and Bremner, J. M. (1970) Gaseous products of nitrate decomposition in soils,
Soil Biol. Biochem., 2, 203-215.

Nicolet, M., and Vergison, E. (1971) L'Oxyde azoteux dans la stratosphere, Aeronomica
Acta, A 91.

Nieboer, H., and van Ham, J. ((1976) Peroxyacetyl nitrate (PAN) in relation to ozone and
some meteorological parameters at Delft in the Netherlands, Atmos. Environ., 10, 115-117.

Noxon, J. F. (1978) Tropospheric NO2, J. Geophys. Res., 83, 3051-3057, Correction: J.


Geophys. Res., 85, 4560-4561 (1980).

Noxon, J. F. (1979) Stratospheric NO2, II. Global behavior, J. Geophys. Res., 84, 5067-5076.

Noxon, J. F. (1981) NOx in the mid-Pacific troposphere. Geophys. Res. Lett., 8, 1223-1226.

Panther, R., and Penzhorn, R. D. (1980) Alkyl sulphonic acids in the atmosphere, Atmos.
Environ., 14, 149-151.

Penkett, S. A., Sandalls, F. J., and Jones, B. M. R. (1977) PAN measurements in England -
Analytical methods and results, VDI-Berichte, 270, 47-54.

Penkett, S. A., Jones, R. M. R., Brice, K. A., and Eggleton, A. E. J. (1979) The importance of
atmospheric ozone and hydrogen peroxide in oxidizing sulphur dioxide in cloud and rain
water, Atmos. Environ.,13, 123-137.
Peyton, T. O., Steele, R. V., and Mabey, W. R. (1976) Carbon disulphide, carbonyl sulphide,
Final Report EPA Contract No. 68-01-2940, Stanford Research Institute, Menlo Park, CA,
USA.

Pierotti, D., and Rasmussen, R. A. (1976) Combustion as a source of nitrous oxide, Geophys.
Res. Lett., 3, 265-267.

Porter, K. S. (1975) Nitrogen and Phosphorus. Food production, Waste and the Environment,
Ann Arbor, Michigan, Ann Arbor Science Publishers Inc.

Rahn, K. A., and Heidam, N. Z. (1981) Progress in arctic air chemistry, 1977-1980: a
comparison of the first and second symposia, Atmos. Environ., 15, 1345-1348.

Rasmussen, R. A., and Khalil, M. A. K. (1981) Atmospheric methane (CH4): trends and
seasonal cycles, J. Geophys. Res., 86, 9826-9832.

Ravishankara, A. R., Kreutter, N. M., Shah, R. C., and Wine, P. H. (1980) Rate of reaction of
OH with COS, Geophys. Res. Lett., 7, 861-864.

Robinson, E., and Robbins, R. C. (1968) Sources, abundance and fate of gaseous atmospheric
pollutants, Final Report, Project P.R.-6755, Stanford Research Institute, Menlo Park, CA,
USA.

Rodhe, H., Crutzen, P. J., and Vanderpol. A. (1981) Formation of sulphuric and nitric acid in
the atmosphere during long-range transport, Tellus, 33,132-141.

Rotty, R. M. (1981) Distribution and changes in industrial carbon dioxide production,


Reference IV-l, WMO/ICSU/UNEP Scientific Conference on Analysis and Interpretation of
Atmospheric CO2 Data, Report WCP-14, Geneva, Switzerland, September 1981.

Rudolph, J., Ehhalt, D. H., and Tönnissen, A. (1981) Vertical profiles of ethane and propane,
J. Geophys. Res. (in press).

Rust, F. (1981) Ruminant methane (13C/12C) values: relation to atmospheric methane,


Science, 211,1044-1046.

Ryden, J. C. (1981) N2O exchange between a grassland soil and the atmosphere, Nature, 292,
235-237.

Sandalls, F. J., and Penkett, S. A. (1977) Measurements of carbonyl sulphide and carbon
disulphide in the atmosphere, Atmos. Environ., 11, 197-199.

Schmeltekopf, A. L., Albritton, D. L., Crutzen, P. J., Goldan, P. D., Harrop, W. J.,
Henderson, W. R., McAfee, J. R., McFarland, M., Schiff, H. I. Thompson, T. L., Hoffman,
D. J., and Kjome, N. T. (1977) Stratospheric nitrous oxide altitude profiles at various
latitudes, J. Atmos. Sci., 34, 729-76.

Schmeltz, I., and Hoffmann, D. (1977) Nitrogen-containing compounds in tobacco and


tobacco smoke, Chem. Rev., 77, 295-311.
Seiler, W. (1974) The cycle of atmospheric CO, Tellus, 26,117-135.

Seiler, W. (1981) The ocean as a source and sink of atmospheric trace gases, IAMAP Conf.
Proceedings, Hamburg, 1981 (submitted).

Seiler, W., and Conrad, R. (1981) Field measurements of natural and fertilizer-induced N2O
release rates from soils. J. Air Pollution Control Ass. (in press).

Seiler, W., and Crutzen, P. J. (1980) Estimates of gross and net fluxes of carbon between the
biosphere and the atmosphere from biomass burning, Climatic Change, 2, 207-247.

Seiler, W., and Fishman, J. (1981) The distribution of carbon monoxide and ozone in the free
troposphere, J. Geophys. Res. (submitted).

Shaw, G. E. (1981) Eddy diffusion transport of arctic pollution from the mid-latitudes: a
preliminary model, Atmos. Environ., 15, 1483-1490.

Shaw, R. W., and Rodhe, H. (1981) Non-photochemical oxidation of SO2 in regionally


polluted air during winter, Report CM-53. Dept. of Meteorology, University of Stockholm.

Sheppard, J. C., Westberg, H., Hopper, J. F., Ganesan, K., and Zimmerman, P. (1981)
Inventory of global methane sources and their production rates, J. Geophys. Res. (submitted).

Shrendikar, A. D., and Lodge Jr, J. P. (1975) Micro-determination of ammonia by the ring
oven technique and its application to air pollution studies, Atmos. Environ., 9, 431-435.

Siman, G., and Jansson, S. L. (1976) Sulphur exchange between soil and atmosphere with
special attention to sulphur release directly to the atmosphere, Swed. J. Agr. Res., 6,135-144.

Singh, H. B., and Hanst, P. L. (1981) Peroxyacetyl nitrate (PAN) in the unpolluted
atmosphere: an important reservoir for nitrogen oxides, Geophys. Res. Lett., 8, 941-944.

Singh, H. B., Sales, L. J., Shigeishi, H., Smith, A. J., Scribner, E., and Cavanagh, L. A.
(1979) Atmospheric distributions, sources and sinks of selected halocarbons, hydrobcarbons,
SF6, and N2O, Environmental Protection Agency Report EPA600/379107.

Smith, I. (1980) Nitrogen oxides from coal combustion-environmental effects, IEA Coal
Research, London, U.K., Report Nr. ICTIS/TR 10.

Söderlund, R., and Svensson, B. H. (1976) The global nitrogen cycle, in, Svensson, B. H.,
and S6derlund, R. (eds) Nitrogen, Phosphorus, and SulphurGlobal Cycles, SCOPE Report
No. 7, Ecol. Bull. (Stockholm), 22, 23-73.

Spicer, C. W. (1977) The fate of nitrogen oxide in the atmosphere, Adv. Environ. Sci.
Technol., 7,163-261.

Stallard, R. F., and Edmond, J. M. (1981) Geochemistry of the Amazon. I: Precipitation


chemistry and the marine contribution to the dissolved load at the time of peak discharge, J.
Geophys. Res., 86, 9844-9858.
Stephens, E. R. (1969) The formation, reaction, and properties of Peroxyacyl Nitrates (PANs)
in photochemical air pollution, in, Pitts J. and Metcalf R. (eds) Advances in Environmental
Sciences, New York, Wiley, Vol. 1, 119-146.

Stevens, C. M., and Rust, F. E. (1981) The carbon isotope composition of atmospheric
methane, J. Geophys. Res., (submitted).

Steudler, P. A., and Peterson, B. J. (1980) Gaseous sulphur release from a salt marsh, Paper
80-40.3, 73rd Annual Meeting of the Air Pollution Control Assoc., Montreal, Quebec, June
22-27.

Svensson, B. H., and Söderlund, R. (eds) (1976) Nitrogen, Phosphorus and SulphurGlobal
Cycles, SCOPE Report NO. 7, Ecol. Bull. (Stockholm), 22.

Sze, N. D., and Ko, M. K. W. (1980) Photochemistry of COS, CS2, CH3, SCH3 and H2S:
implications for the atmospheric sulphur cycle, Atmos. Environ., 14, 1223-1229.

Taylor, G. S., Baker, M. B., and Charlson, R. J. Atmospheric interactions of C, N and S


cycles: the role of aerosols and clouds, Chapter 3, this volume.

Terry, R. E., Tate III, R. L. and Duxbury, A. (1980) Nitrous oxide emissions from drained,
cultivated organic soils of South Florida, presented at 73rd Annual Meeting of the Air
Pollution Control Assoc., Montreal, Quebec, June 22-27.

Torres, A. L., Maroulis, P. J., Goldberg, A. B., and Bandy, A. R. (1980) Atmospheric OCS
measurements on Project GAMETAG, J. Geophys. Res., 85, 7357-7360.

Tuazon, E. C., Winer, A. M., and Pitts, Jr, J. N. (1981) Trace pollutant concentrations in a
multi-day smog episode in the California south coast air basin by long pathlength FT-IR
spectroscopy, Environ. Sci. Technol. (in press).

Tuck, A. F. (1976) Production of nitrogen oxides by lightning discharges, Quart. J. R. Met.


Soc., 102, 749-755.

Turco, R. P., Whitten, R. C., Toon, O. B., Pollack, J. B., and Hamill, P. (1980) OCS,
stratospheric aerosols and climate, Nature, 283, 283-286.

Volz, A., Ehhalt, D. H., Derwent, R. G., and Khedim, A. (1979) Messung von
atmosphärischem 14CO: eine Methode zur Bestimmung der troposphärischen OH
Radikalkonzentration. Berichte KFA Jülich No. 1604.

Wang, W. C., Yung, Y. L., Lacis, A. A., Mo, T., and Hansen, J. E. (1976) Greenhouse effects
due to man-made perturbations of trace gases, Science, 194, 685-690.

Weiss, R. F., and Craig, H. (1976) Production of atmospheric nitrous oxide by combustion,
Geophys. Res. Lett., 3, 751-753.

Weiss, R. F. (1981) The temporal and spatial distribution of tropospheric nitrous oxide, J.
Geophys. Res., 86, 7185-7195.
Went, F. W. (1960) Organic matter in the atmosphere and its possible relation to petroleum
formation, Proc. Nat. Acad. Sci., 46, 212-221.

Whittaker, R. H., and Likens, G. E. (1975) in, Primary Productivity of the Biosphere Lieth,
H., and Whittacker, R. H., (eds), New York, Springer-Verlag, 305-328.

Wilson, A. T. (1978) Pioneer agriculture explosion and CO2 levels in the atmosphere, Nature,
273, 40-41.

WMO/NASA (1982) The Stratosphere 1981: Theory and Measurements, Hudson, R. D.,
Reed, E. I., and Bojkov, R. D. (eds), Geneva, WMO.

Wofsy, S. C. (1976) Interactions of CH4 and CO in the Earth's atmosphere, Ann. Rev. Earth
Planet. Sci., 4, 441-469.

Zafiriou, O. C., and McFarland, M. (1981) Nitric oxide from nitrite photolysis in the central
equatorial Pacific, J. Geophys. Res., 86, 3173-3182.

Zahniser, M., and Howard, C. J. (1979) The reaction of HO2 with O3. J. Chem. Phys.,
73,1620-1626.

Zimmerman, P. R., Chatfield, R. B., Fishman, J., Crutzen, P. J., and Hanst, P. L. (1978)
Estimates on the production of CO and H2 from the oxidation of hydrocarbon emissions from
vegetation, Geophys. Res. Lett., 5, 679-682.

Zimmerman, P. R., Greenberg, J. P., and Crutzen, P. J. (1982) Termites: a potentially large
source of atmospheric trace gases, Science 218, 563-565.

COMMENT TO CHAPTER 3

J. E. LOVELOCK

During the past decade the scientific understanding of the great cycles of the elements at the
Earth's surface seems to have matured. The biosphere is now accepted as a participant and the
stratosphere, once the exclusive domain of elitist aeronomers, is now the concern of us all.
These are, I think, all changes for the good and it was therefore a pleasure to read Dr.
Crutzen's paper on atmospheric interactions and have them confirmed.

Indeed I am so much in accord with the spirit of the paper that criticism would be difficult
were it not for a few disagreements over matters of fact. There follows my comments on
these disagreements and also on things we agree about.

Can we be sure that there are no gaseous phosphorus compounds in the atmosphere (section
3.1)? Among the compounds we have sought but not yet found are P(CH3)3 and NPO. They
would be interesting even at the 10-12 ppv level in view of their probable brief residence
times. The existence of As(CH3)3 and N(CH3)3 makes P(CH3)3 worth looking for.

One of the key steps in the better understanding of atmospheric chemistry was the discovery
of the role of the hydroxyl radical (section 3.2). I would like to ask Dr. Crutzen if he is
confident that ozone is the principal precursor of OH in the natural troposphere. There is
some compelling although circumstantial evidence for active photochemistry involving
oxides of nitrogen over tropical waters and in the sea surface of these regions.

The probable climatic consequences of the accumulation of CO2 are well discussed. To many
it must come as a surprise to hear that tropospheric ozone has a similar warming effect
(section 3.2). Since the abundance of both of these gases is connected with combustion and
since their effects are additive, we shall presumably hear more about tropospheric ozone in
the near future.

I agree that the natural and man induced burning of tropical vegetation is an important source
of atmospheric nitrogen compounds and indirectly as a source of tropospheric ozone (section
3.2 and section 3.5). Among the interesting candidate compounds known to be present in
smoke but not discussed, are methyl nitrite and methyl nitrate.

Oceanic sources of NO and associated compounds seem to be dismissed in a single reference


(McFarland, 1979). Yet Zafiriou, McFarland and Bromund (1980) propose that the tropical
oceans are a source of tropospheric NO. They found that the photolysis of the nitrite ion in
sea water by solar radiation in the near ultraviolet generates both OH radicals and NO. The
NO concentration in equilibrium with the tropical ocean was up to 1000 times greater than its
observed aerial concentration (10-11 ppv).

During an expedition on the Meteor in 1973 to the tropical regions of the North Atlantic I
found the production of high concentrations of PAN and other alkylperoxy nitrates. This
observation, which was at the time a mere curiosity, takes on a new significance in the light
of the Zafiriou and McFarland (1981) findings and also when the unexpected atmospheric
stability of these nitrate esters is taken into account.

An interesting sulphur compound not here discussed is CH3SO3H. According to Cox


(personal communication) methyl sulphonic acid is a significant product in the oxidation of
CH3SCH3 and CH3SH by OH radicals. CH3SO3H is physically and chemically rather like
H2SO4 and may not be distinguished from it in the normal routine of sulphate aerosol
measurement. The measurement of the ratio of CH3SO3H to H2SO4 in the tropospheric
aerosol might provide a means of distinguishing the relative importance of biological and
industrial inputs.

It is usually assumed that volcanos vent SO2. This may be true of emissions from fumaroles
at comparatively quiescent periods, but no one yet has been able to sample directly the gases
coming from a volcano in full voice. This is the important time to sample, when the gases go
straight to the stratosphere and in substantial quantity. There are some reasons to believe that
the direct emissions are of reduced gases rather than SO2.

The coexistence of methane (section 3.4.3) at 1.5x10-6 ppv and oxygen in the current
atmosphere is a chemical anomaly so profound that it would reveal to an observer with near
certainty the presence of life on this planet. It is the item of evidence that most clearly
justifies the new orthodoxy about the contemporary atmosphere: that which takes in the
biosphere as an active participant.

The residence time assigned to methane has fluctuated with the estimates of the OH
abundance over the range 2 to 20 years. Now at last we seem to have a fairly firm mid-range
value of 7 years.
The potential significance of the claim that methane is rising in concentration by 2% per year
(Rasmussen and Khali, 1981) demands confirmation. It may indeed be rising in concentration
but to prove it requires an absolute accuracy of analysis, including the preparation of
standards, of at least 0.5% and a great deal of faith in the constancy of the atmosphere being
sampled.

Like methane, N2O (section 3.5.3) at 0.33 x 10-6 ppv is a chemical anomaly characteristic of
the Earth and a consequence of its biosphere. It is good to know that it is not now regarded as
a serious threat to stratospheric ozone.

I have been obliged professionally to follow the rapid unfolding of stratospheric chemistry as
observations refined the various hypotheses about the chlorine catalysed depletion of ozone. I
disagreed with some of the conclusions aeronomers drew about the consequences of ozone
changes on our future health. This did not stop me from being deeply impressed with their
professional competence. It is good to know that these powerful talents are now applied to the
more important problems of tropospheric chemistry.

The stratospheric experience has revealed how important trace molecular species and radicals
can be. Before the chlorineozone affair who would have thought of looking for chlorine
nitrate or peroxynitric acid in the stratosphere? I know for certain that there are substantial
quantities of as yet unidentified substances in the troposphere. It remains to be seen how
much they may enlighten our views on the cycles of the elements.

REFERENCES

Zafiriou, O., McFarland, M., and Bromund, R. H. (1980) Nitric oxide in sea water. Science
207, 637-639

Back to Table of Contents

The electronic version of this publication has been prepared at


the M S Swaminathan Research Foundation, Chennai, India.

SCOPE 21 -The Major Biogeochemical Cycles and Their


Interactions
4 Heterogeneous Interactions of the C, N, and S Cycles in the Atmosphere:
The Role of Aerosols and Clouds
G. S. TAYLOR, M. B. BAKER AND R. J. CHARLSON

Abstract
4.1 Introduction
4.2 General Nature of Interactions in Aerosols and Clouds
4.3 Possible Interactions
4.4 The System SO2, CO2, NH3, H2SO4, and H2O (liquid)
4.5 Oxidation of S(IV) to S(VI) in Wet Particles
4.6Conclusions: Flux Relationships
Acknowledgements
References

ABSTRACT

The cycles of C, N and S interact in the atmosphere within the condensed phases of aerosol
particles and clouds. Much of this interaction occurs in the aqueous phase due to the
hygroscopic nature of compounds like NH4, HSO4 and the water solubility of trace gases,
such as SO2 and NH3. In order to explore the nature of these interactions, the equilibria of the
system H2O(1), SO2, NH3, H2SO4 and CO2 are examined. Subsequently, oxidation of S(IV) to
S(VI) within the aqueous phase is considered as a function of NH3, H2O(1), and the source
strength of S(IV).

The results are:


 Almost all S(IV) is in the gas phase when clouds are acidified below pH5.5, while most
ammonia is present as aqueous NH4+. Gas phase ammonia concentrations are probably often
in the range of 0.5 to 5.0 pptv.
 The amount of S(IV) available for oxidation in the water phase is a strong function of pH,
with more S(IV) available at higher pH, and is increased by adding NH3.
 CO2 is unimportant to the pH of atmospheric water in most cases.
 S(IV) slowly oxidizes to S(VI) in aerosols and clouds, but can produce significant amounts
of SO42- after a number of passages through clouds.
 The products of oxidation of S(IV) in cloud water decrease the reaction rate such that the
production of SO42- is a nonlinear function of time and of source strength of SO2.
 The overall effect of increased SO2 emissions could be a slowing of its oxidation allowing
geographical regions under the influence of acid precipitation to increase in area.

4.1 INTRODUCTION

The heterogeneous interactions of C, N, and S cycles within the atmosphere depend on the
existence of aerosols and water clouds. * These particles provide sites for rapid reaction of
key molecules that otherwise react slowly (or not at all) in the gas phase. Such reactions in or
on particles are expected to be particularly important in situations where photochemistry is
not effective (Shaw and Rodhe, 1981). In addition, many of the reaction products of gaseous
N and S compounds are themselves liquids or solids of low vapour pressure under normal
conditions and are found as aerosol particles. The water solubility of these reactions products
allows their incorporation into cloud droplets and removal from the atmosphere by
precipitation. The nature of the interactions of C, N, and S cycles thus will depend on the
physical and chemical characteristics of both aerosols and clouds. In addition, these
interactions cause shifts in chemical equilibria that in turn cause nonlinear responses of
atmospheric concentrations and fluxes to changes in the source strength of materials injected
into the air.

Atmospheric aerosols are chemically heterogeneous and are composed of large numbers of
compounds. While it has been customary and useful to treat these simply as physically mixed
systems (Junge, 1963), more recent results demonstrate clearly that both the chemical
composition and chemical interactions of the tropospheric aerosol are strong and regular
functions of particle size (Stevens et al., 1978). Most of the available data are from low
altitudes in industrialized continental regions, but similar results are found in marine areas
(Duce, Chapter 16, this volume).

*Aerosols are defined as suspensions in a gas of particles of liquid or solid material that are stable to either
gravitational sedimentation or Brownian coagulation over some period of time. In the case of atmospheric
aerosols, the relevant time scale is a fraction of an hour or longer. Thus, ordinary water clouds are aerosols;
however, it is customary to designate water clouds separately as having a relative humidity over 100%. Fog may
be defined as cloud in contact with the ground.

The mass or volume distributions of aerosols are consistent with the composition-size
dependence and show that the particles have distinctly different chemical composition and
morphology above and below 1 µm (Figures 4.1 and 4.2). The mass mode below 1 µm is
often called the accumulation mode, because mass accumulates there due to condensation and
coagulation. These same processes tend to mix internally the accumulation mode particles
(i.e. each particle has some of each constituent). In contrast, the lack of condensation and
coagulation of coarse mode particles, along with their mechanical origins, tends to leave
these aerosols as an external mixture (i.e., individual particles may have different
compositions). In addition, the coarse and accumulation modes are externally mixed with
each other, and do not interact chemically to any known extent. Because of their chemical
composition the sub-µm aerosols usually contain some liquid water in an aqueous phase.

Figure 4.1 The volume distributions of aerosols plotted as dV/d (log Dp) (linear) versus
diameter (Dp) on a log scale. (Whitby and Sverdrup, 1980). Reproduced by permission of
John Wiley & Sons, Inc.

The accumulation mode aerosols are to a large degree the result of interactions of C, N and S
cycles, both within a condensed phase and via preceding gas-phase production reactions.
From Crutzen's review of homogeneous reactions (Chapter 3, this volume), gas phase
production of H2SO4 (reactions R3335) results either in nucleation with H2O of new particles
of hydrated H2SO4 or in the condensation of H2SO4 and H2O on the surfaces of pre-existing
particles. Both of these possibilities are important. The former is a necessity for the creation
of new particles in the absence of direct injection of particles into the atmosphere (e.g. from
combustion). The latter results in a build-up of the mass of sulphate compounds in the size
class between 0.1 and 1 µm. Chemical reaction of NH3 with H2SO4 results in an aerosol of
composition ranging from NH4HSO4 to neutralized (NH4)2SO4. The reaction of NO2 with OH
produces HNO3, some of which may attach to or be dissolved in pre-existing aerosol
particles. HNO3 can be neutralized by NH3 forming NH4NO3 in the condensed phase of
accumulation mode aerosols. Sulphate and nitrate combined with ammonium usually are the
dominant inorganic species formed in the accumulation mode by direct gas-particle
conversion. In addition, a wide variety of organic materials are often found in this same size
range, although their total mass is usually considerably less than the mass of sulphates and
nitrates, The gas-phase interactions discussed by Crutzen (Chapter 3, this volume) thus have
direct implications for the formation and composition of submicrometer aerosols.

Figure 4.2 A conceptual representation of the volume or mass distributions of aerosols as


related to source, transformation and sink processes. Fine particulate matter also contains
some primary emissions such as C (0) (elemental carbon) and Pb. Ordinate is as in Figure 4.1
(Shaw and Stevens, 1980), Reproduced by permission of New York Academy of Sciences

Besides such gas phase interactions, C, N, and S compounds interact within the condensed,
aqueous phase of aerosols and clouds, with chemical reactions resulting in an increased mass
of dissolved substances. In particular, the oxidation of SO2 in the liquid phase of clouds
results in increased amounts of dissolved SO42-. Since most clouds evaporate and do not
result in precipitation, the dissolved material is released back to the atmosphere in the
accumulation mode of the aerosol. There are several known reaction mechanisms of SO2 and
its dissociation products in the aqueous phase. Among the more likely are oxidation by O3
and H2O2 (Penkett et al., 1979) although oxidation by O2 may be of some importance. No
single mechanism can be identified yet as the most important one.

The lifetimes in air of the two mass modes are also distinctly different, largely due to the
different mechanisms of removal, Particles greater than several µm tend to be removed by
sedimentation and usually have lifetimes of much less than a day, Particles below 1 µm
diameter have very small sedimentation velocities (<10-2 cm s-1) and are removed from the air
mainly by incorporation into cloud droplets and precipitation, As a result, their lifetime is
dictated by the period between precipitation events, and typically is a few days to perhaps a
few weeks in mid-latitudes. As a result of little, if any, extensive interaction of the gas phase
with the coarse mode, there is little interaction of C, N and S cycles with particles much
above ca, 1 µm in size. Although we will emphasize sub-µm particles, we note the
importance to precipitation chemistry of below-cloud scavenging by rain of basic soil dust.
We also leave open the question of coarse-particle interaction with NOx and HNO3.

Clouds of liquid water have droplets from ca 5 to 50 µm diameter, typically yielding liquid
water contents of perhaps 0.10.5 g m-3; some vigorously precipitating clouds may have water
contents in excess of 0.5 g m-3. Fogs have somewhat lower liquid water contents, typically
below 0.1 g m-3, as well as somewhat smaller droplets. The water in clouds and fogs contains
soluble salts from the remains of condensation nuclei, water soluble gases (e.g. CO2, SO2,
NH3, etc.) as well as suspended but insoluble materials (e.g. mineral grains and soot). The
lifetime of an individual cloud droplet is often a fraction of an hour though several hours may
be possible in fog. Since only a small fraction of clouds result in precipitation, aerosol
particles take part in the condensationevaporation cycle many times before precipitation can
remove them. In the physical process of cloud nucleation, the chemical property of water
solubility of the nuclei is a dominant factor. Due to the solubility of the key compounds
(ammonium sulphates and nitrates) the products of interaction of N and S cycles in the
atmosphere are probably the dominant cloud condensation nuclei. To illustrate the ways in
which the cycles of C, N, and S influence each other within aerosols and clouds, we will
select a limited family of interactions as useful examples. After choosing the system,
NH3CO2SO2-H2SO4H2O(1), we will examine the characteristics of the equilibria that control
the interactions of the dissolved species. We can then investigate chemical reactions that
occur within this system in the presence of O2, O3, and H2O2. This allows consideration of the
response of sulphate production in aerosols and clouds to changes in the source strengths of
SO2 and NH3. The results allow an understanding of the nonlinearity of sulphate production
in response to varying SO2 strengths.

4.2 GENERAL NATURE OF INTERACTIONS IN AEROSOLS AND CLOUDS


That the atmospheric cycles of C, N, and S interact within both aerosols and clouds is
indicated by several observations:

 Submicrometer aerosols usually contain NH4+, SO42-, C(O) (graphitic carbon as soot) and
organic carbon (see Table 4.1).
 Sulphur dioxide is known to oxidize to form H2SO4. In turn, H2SO4 has a low vapour
pressure and condenses as particles, NH3 is known to be emitted from the ground and reacts
with H2SO4 particles to form the salts (NH4)2SO4, NH4HSO4, etc.
 Clouds and fog form on soluble nuclei of condensation such as sulphates, and cloud and
fog water usually contain dissolved ammonium and sulphate ions (see Table 4.2).
 Precipitation usually contains measurable amounts of SO42-, NH4+, NO3-, C(O) and HCO3-
and may be an important source of these compounds to terrestrial ecosystems (see Table 4.3).
 NH3, SO2, HNO3 and NO2 are water soluble, are present in air and to some degree dissolve
in cloud and fog water as well as in hydrated aerosol particles (see Table 4.4).
These examples help to define the specific modes of interaction of C, N, and S compounds
with and within the condensed phases of aerosol particles and clouds. As far as is known, the
atmospheric interactions are exclusively abiotic and do not involve any phosphorus
compounds. Six modes of atmospheric interactions are:

1. Chemical reaction (e.g. 2NH3 + H2SO4 (NH4)2SO4)


2. Gas exchange and equilibria with hydrated, wet particles or droplets (e.g. SO2 +
H2O(1) H2SO3 H+ + HSO3-, etc.)
3. Condensation of low vapour pressure material to yield either new particles or to add
to the mass of pre-existing ones (e.g. condensation of H2SO4 that was produced
photochemically)
4. Adsorption of materials into the surfaces or pre-existing particles (e.g. on the surface
of carbon particles where specific areas of 100 m2 g-1 are possible)
5. Surface reactions (e.g. OH + organic film products)
6. Capture of a basic particle (e.g. CaCO3) by a cloud or rain drop with subsequent
reaction.

Table 4.1 Typical composition of atmospheric aerosols, concentrations in µg m-3

a) Industrial Region*
Specie or element Concentration

Diameter less than 2.5 µm


SO42- 1.070.0
Al 0.0080.05
As 0.04
Br 0.013.0
C (graphitic) 010.0
Cd 0.020.4
Cl 0.0055.0
Cr 00.002
Cu 0.0020.14
Fe 0.021.0
K 0.010.2
Mg 0.0060.15
Mn 0.0020.03
NH4+ 0.55.0
Ni 0.0020.1
NO3- 3.060.0
Pb 0.56.0
Si 0.060.6
Ti 0.030.06
V 0.0040.007
Zn 0.003-0.06
Diameter greater than 2.5
µm
Al 0.3 2.0 or more (from soil)
Si 1.0 10.0 or more (from soil)
P 0.030.1
S 0.20.5
Cl 0.2 5.0 or more (from oceans)
Ca 0.4 1.0 or more (from soil)
Fe 0.3 1.0 or more (from soil)

*Adapted from Charlson et al. (1978) and Stevens et al., (1978).

b) Background Aerosol†

Concentrations (µg m-3)

Continental Marine
boundary boundary Free
Specie layer layer troposphere

SO42- (excess)‡ < 0.20.4 0.33.0 < 0.1 0.6


< 0.01
NH4+ 0.040.1 0.2 < 0.02

†from Huebert and Lazrus (1980).


‡SO41- (excess) is the amount of SO42- remaining after the
amount from sea-salt is subtracted
based on either Na+ or Cl data.

Table 4.2 Example cloud water composition, µeq litre-1

Mean of Mean of
Petrenchuk and Scott and
Specie Drozdova (1966) Laulainen (1979)

H+ 100 150
NH4+ 800 430
Na+ 120 53
K+ 100 44
Ca2+ 400 90
Mg2+ 100 4
SO42- 1500 460
NO3- 90 330
Cl- 300 35

Table 4.3 Inorganic composition of rain and snow water, Forshult Station, Sweden 19751979
(incl)*

Quantity ± 25% range

21
Amount of precipitation (mm) 42 57

Specie (µeq litre-1)

49.8
2-
SO4 72.8 100.0

8.6
-
Cl 13.5 19.8

12.6
-
NO3 21.7 40.3

3.3
+
NH4 13.4 24.2

1.7
+
K 3.8 9.6

6.0
+
Na 11.8 18.0

5.8
Mg2+ 9.6 14.0

16.1
2+
Ca 22.1 39.5

25.1
+
H (from pH) 50.1 79.4

*Data courtesy of R. Söderlund, IMI Network, Department of Meteorology, University


of Stockholm, S-106 91 Stockholm, Sweden.

Table 4.4 Gas Phase SO2, NH3, CO2*

Background
SO2 Northern Hemisphere
Boundary Layer 89 ± 69 ppt
Free Troposphere 122 ± 85
Southern Hemisphere
Boundary Layer 57 ± 18
Free Troposphere 90 ± 21

NH3 Very uncertain


1001000 ppt (ocean surface equilibrium)
1-100 ppt (equilibrium with acidic, wet sulphate
aerosols)

CO2 340 ppm


Industrial Region, Boundary Layer (estimates)
Away from SO2 0.13 ppb
sources NH3 10-31 ppb
CO2 340400 ppm

Close to SO2 up to 1 or a few ppm


sources NH3 10-31 ppb or more (uncertain)
CO2 3401000 ppm or more

*Adapted from Maroulis et al. (1980) and Lau and Charlson (1977).,

Collectively, these interactions play significant roles in controlling the molecular form of the
aerosol particles as well as the ionic species present in rain and snow water.

The sub-µm sulphate aerosol often exhibits the chemical features of a single compound. For
example, some particle samples show deliquescence and X-ray diffraction of pure (NH4)2SO4
(Charlson et al., 1978) or the volatility of H2SO4 at ca. 125°C (Coburn et al., 1980) Since
there are no known direct sources of fine-particle (NH4)2SO4, its existence also demonstrates
the simplest interactions of the S and N cycles in the atmosphere. This also indicates the sub-
µm aerosol particles containing deliquescent salts (e.g. NH4NO3, NH4HSO4, (NH4)2SO4) or
hygroscopic liquid (e.g. H2SO4) understandably also contain liquid water. At 80% R.H.,
taking Raoult's law as a means of approximating the composition, such particles would be 80
mole percent water and 20 mole percent dissolved species. Hence at R.H. < 90 or 95%, the
aqueous phase of such particles is sufficiently concentrated that non-ideal behaviour will
occur.

The calculations presented in this paper are for water solutions that behave ideally. Extension
to non-ideal solutions is not included, but the general tendency will be for the concentration
of dissolved gaseous species to decrease (salting-out effect) and for ionic forms to increase
since their activity coefficients will be substantially less than unity.

Figure 4.3 Matrix of potential interactions of C, N and S in air defined by oxidation state.
Elemental sulphur is omitted because it is not known to occur in air. The shaded corner
(SO42-, SO2, CO2, NH3, and RNH2) will be treated in section 4.4

4.3 POSSIBLE INTERACTIONS

In order to explore objectively the range of interactions, we consider the various C, N and S
compounds that exist in tropospheric air. Figure 4.3 depicts a three-dimensional matrix of the
oxidation states and species that are involved. Because COS reacts on time scales of years
(Johnson, 1981) and because H2S and RSH (mercaptans) are expected to react fairly rapidly
to SO2, oxidation states IV and VI of sulphur are of primary interest. Neither N2 nor N2O
react on or in particles, nor does CO.

Thus, we can focus on interactions of S(IV) and S(VI) with organic C, CO2, NH3, NO, NO2,
HNO2, HNO3 and H2O(1). These are represented qualitatively in Figures 4.4 and 4.5 for
S(IV) and S(VI) respectively. Here we assume that NO oxidizes to NO2 before any
interactions with C or S in particles, and we further assume that NO2 produces HNO3 and
HNO2 upon incorporation in hydrated particles. For convenience we list all organic
compounds as C(IV).

The most common interactions are acid equilibria, gasliquid exchanges and the influence of
NH3 or amines as bases neutralizing weak or strong acids. A potential source of gas phase
HNO3 is its volatilization that may be caused if sufficient H2SO4 is added to a particle, e. g.
by photochemical production. Since S(IV) and S(VI) coexist in the same particles, acidbase
equilibria are clearly sensitive to the proportions of the weak and strong acids that are
present. Because the rate of oxidation of S(IV) to S(VI) depends on which dissolved species
are present, it is necessary to consider the equilibria that are involved.

Figure 4.4 Interactions of S(IV) with C and N in wet particles. Note: (1) Acid equilibria
occur in all nine cases. (2) Catalytic oxidation of S(IV) by C(O) may occur independently of
the presence of any N species
Figure 4.5 Interactions of S(VI) with C and N in wet particles. Note: (1) Volatilization of
HNO3 probably occurs under acidic conditions independent of the presence of some C
species (Tang, 1980). (2) Neutralization of H2SO4 by NH3 is probably independent of C

In order to approach a complete list of factors controlling the equilibrium, we might include
the components used by Liljestrand and Morgan (1981): SO2, NO, NO2, HNO2, HNO3 NH3,
HCl, H2SO4, CO2, kaolinite, Al(OH)3 and H2O(1). However, it is instructive at this stage to
limit this discussion to the simpler system SO2, CO2, NH3, H2SO4 and H2O(1) (Figure 4.3,
shaded corner). This may be justified for understanding the chemistry of these components
alone, and may be further justified to the extent that there are atmospheric situations in which
the other components are present in concentrations that are low enough to be neglected.

4.4 THE SYSTEM SO2, CO2, NH3, H2SO4, AND H2O (LIQUID)

Chemical equilibrium of gaseous SO2, CO2 and NH3 with wet particles of less than 5 µm
diameter occurs in less than a few seconds (Freiberg and Schwartz, 1981). Thus, multiple
equilibrium expressions can be used to approximate the overall composition that results from
the weak acid/weak base interactions that occur along with the existence of strong acids such
as H2SO4 (Liljestrand and Morgan, 1981). This step is a necessary prelude to examining the
oxidation of S(IV) in solution, a discussion of which follows in section 4.5. The equilibrium
concentrations can be calculated numerically; however, they also may be presented
graphically, a form which reveals more clearly the interrelations of the many constituents.

Table 4.5 lists the relevant expressions and constants that were used to construct the log
concentration versus pH diagrams (Sillén, 1967) in Figures 4.64.8 and further used for the
oxidation computations in section 4.5. In addition, Table 4.6 gives the partitioning of SO2,
and NH3 between liquid and gas phases in clouds and fog.

Figures 4.6 4.8 are the simplest representation of a master variable diagram (Sillén, 1967) in
which the concentration of each species is given as a function of a master variable, in this
case, pH. The nine equilibrium expressions of Table 4.5 first can be put into logarithmic
form, resulting in nine linear equations. The three Henry's law expressions (Equations (2),
(5), and (7)) yield aqueous concentrations of SO2 • H2O, NH3•H2O and CO2•H2O that are
independent of pH and are governed by the gas phase equilibrium pressure of each gaseous
species. These result in values represented by three lines parallel to the pH axis.

The remaining six equations can be solved to yield linear expressions for the concentration of
each dissociated species as a function of pH, giving six more lines. A seventh line is log[H+]
= pH, and [SO42-] is included as a pH-independent quantity. The equilibrium condition is
depicted by the pH at which the sum of the positive ions is equal to the sum of the negative
ions. At this pH, the equilibrium concentrations of the other species are also graphically
depicted. We caution that these plots do not include equations of mass balance and hence
apply only at the equilibrium condition. Each plot expresses one solution to the equations for
a particular set of data and therefore must not be interpreted literally as dynamic diagrams in
which processes can be studied.

The three example conditions depicted in Figures 4.64.8 were selected to represent a range of
concentrations of SO2, NH3, [SO42-] and liquid H2O as might be found in actual cases; CO2 is
constant at 340 ppmv. The equilibrium conditions in Figure 4.6 are intended to represent an
industrial region and have been selected as the initial condition for some of the oxidation
calculations of section 4.5. The initial (SO42-) level of 3 ppbv and SO42- of 5 µg m-3 represents
a polluted industrial location that is well removed from the immediate influence of sources.
The bulk of the (NH4+) was initially contained in nuclei of condensation ((NH4)2SO4) and the
liquid water content reached in this activated cloud is 0.5 g m-3. The resulting equilibrium
partial pressure of NH3 is only 7 pptv, due to the acidity of the solution.

Table 4.5 Values of constants used

Equilibrium
Equilibrium Value at 5°C Reference
constant

1.82 x 10-15 mol2


(1) H2O H+ + OH Kw = [H+][OH] Yui (1940)
litre-2
Robinson and Stokes
(1959)
PCO2 Johnstone and Leppla
SO2(g) + H2O SO2 . KHS 0.379 atm
(2) (1934)
H2O = . litre mol-1
[SO2 H2O] Lowell et al. (1970)
[H+][HSO3]
. + Lowell et al. (1970)
(3) SO2 H2O H + HSO3 K1S= 0.0206 mol litre-1
Yui (1940)
[SO2 . H2O]
[H+][SO32 ]
K2S 8.88 x 10-8 Lowell et al. (1970)
(4) HSO3H+ + SO32
= mol litre-1 Yui (1940)
[HSO3]
PNH3 Morgan and Maass
NH3(g) + H2O NH3 . KHN 7.11 x 10-3 atm litre
(5) -1 (1931)
H2O = . mole
[NH3 H2O] Freiberg (1974)
[NH4][OH]
1.5 x 10-5 mol litre- Yui (1940)
(6) NH3 + H2O NH4+ + OH Kb = 1
Freiberg (1974)
[NH3 . H2O]
Robinson and Stokes
PCO2
CO2(g) + H2OCO2 . KHC (1959)
(7) 16.6 atm litre mol-1
H2O = Morgan and Maass
[CO2 . H2O]
(1931)
[H+][CO3] Yui (1940)
. + 2.94 x 10-7 mol
(8) CO2 H2O H + HCO3 K1C Robinson and Stokes
litre-1
[CO2 . H2O] (1959)
2
[H+][CO3 ] Yui (1940)
K2C 2.74 x 10-11
(9) HCO3 H+ + CO32- -1 Robinson and Stokes
= mol litre
[HCO3] (1959)
Figure 4.6 Industrial region. The equilibrium pH of the system is determined at the point of
charge balance (O). At that pH, the concentrations of other species are also given for the
equilibrium conditions. We caution that these plots do not include equations of mass
conservation and thus strictly apply only at the equilibrium pH. Assumptions:

aerosol
= 5.0 µg m-3
(SO24)
liquid
= 0.5 g m-3
water
aerosol
= 1.7 µg m-3
(NH4+)
PSO2 = 3 ppbv; 7.86 µg m-3
Calculated:
= 7.1 x 10-12 atm = 5 x 10-3 µg m-3
PNH3
pH = 4.6

For comparison to the hypothetical case of Figure 4.6, we can consider the composition of
rain-water from an actual site downwind of an industrial region. Here we assume that rain-
water has a composition similar to and governed by the composition of the clouds from
which it fell. Figure 4.7 is based on 5 years of precipitation chemistry data from Forshult,
Sweden (Table 4.3) with the assumption of an SO2 concentration of 1 ppbv. This case might
be termed a cloud in well-aged industrial or continental air.

Figure 4.7 Forshult (Sweden) precipitation 19751979. Assumptions:

PCO2 = 330 ppmv


PSO2 = 1 ppbv
2[SO2-] + [Cl-] + [NO3-] = 108 µeq litre-1
2[Ca2+] + 2[Mg2+] + [K+] + [Na+] = 47 µeq litre-1
pH = 4.3
calculated:
PNH3 = 0.23 pptv
Note: [SO32-] and [CO32-] are omitted for simplicity

Finally, Figure 4.8 represents a clean background case, either in the boundary layer or free
troposphere. The total SO2 concentrations, i.e. SO2 (gas) plus SO2 (dissolved), are
representative of remote locations in the northern hemisphere (Maroulis et al., 1980). Recent
data for sulphate aerosol in remote locations indicate sulphate concentrations of 0.1 to 1 µg
per standard cubic meter with molar ratios of (NH4+) to SO42- between 1 and 2. The
[NH4+]/[H+] of 2 was picked arbitrarily for this example.

Several general features are evident in these figures, the equilibrium expressions in Table 4.5,
and the partition expressions in Table 4.6:
a. Almost all the S(IV) is present as SO2 gas, for realistic liquid water contents and pH <
5.5. In contrast, almost all the ammonia is in solution as NH4+.
b. At any realistic pH (3.5 or higher) SO2.H2O is less than one-tenth of the other
dissolved forms of S(IV), largely due to the high value of K1S. As a result, the total
amount of dissolved S(IV) is always a strong function of pH. This implies that the
production rate of SO42 in a cloud decreases as hydrogen ion in the cloud water
increases, regardless of oxidation mechanism. (This presumes reaction rates for HSO3
or SO32 comparable to or greater than for SO2.H2O).
c. Carbon dioxide is important for the pH of atmospheric liquid water only in cases of
low (SO2). and low (SO42) as in Figure 4.8. Even then it is not dominant.
d. The charge balance of cloud water may be significantly effected at times by HSO3,
even in the presence of realistic amounts of (SO42-) (Hales and Dana, 1979).
e. pH below ca. 5 also requires the presence of strong acids, except in the instance of
high concentrations of SO2, such as near sources.

Figure 4.8 Background cloud. Assumptions: Liquid water content = 0.5 g m-3

T = 5C
[NH4 ]/[H+] =2
+

Mass concentration (SO42) = 0.1 to 1.0 µg m-3


PNH3 = 1.6 X 10-12 atm (10-3 µg m-3)
PSO2= 0.1 ppbv (0.26 µg m-3)
At (SO42) = 0.1 µg m-3 pH = 5.6 NH4+ (aq) = 0.05 µg m-3
S (IV) (aq) = 0.06 µg m-3
At (SO42) = 1 µg m-3 pH = 4.95 NH4+ (aq) = 0.2 µg m-3
S(IV) (aq) = 0.013 µg m-3
Note: [SO32] and [CO32] are omitted for simplicity

Given this set of equilibrium conditions, it is now possible to consider chemical reactions that
occur within the liquid phase. In doing so, we want to emphasize the ways in which ammonia
and pH influence the reaction rates. A major goal in doing this is to study the overall response
of the amounts of reaction products (e.g. (SO42-) to changes in the inputs of reactants (SO2).

Table 4.6 Partitioning* of SO2 and NH3 between liquid and gas phases as functions of liquid
water content, L, and [H+]

(SO2)

L (gm-
3 0.1 0.01
) = 0.5

[H+] SO2† SO2 SO2

10-6 0.6 0.1 0.01


10-5 0.06 0.01 10-3
10-4 6 X 10-3 10-3 10-4

NH3 L (gm-3) = 0.5 0.1 0.01

[H+] NH3‡ NH3 NH3

10-6 10 3 0.3
10-5 102 30 3
10-4 103 300 30

*x is the ratio of moles of a soluble gas in the liquid phase to the moles in the gas phase
for a given cloud volume. In general it is given as:
L[x]RT
x=
Px
where L is liquid water content in units appropriate to R, the gas constant, and Px, the partial
pressure of molecules x. [x] is the molar concentration of all species of x in solution and T is
absolute temperature.

LK1SRT
†SO2 for pH <6
[H+]KHS

LKb[H+]RT
‡NH3=
KHNKw

4.5 OXIDATION OF S(IV) TO S(VI) IN WET PARTICLES

One of the outstanding practical problems of atmospheric chemistry is to be able to


understand and to predict the response of chemical processes in the atmosphere to increases
or decreases of input materials. In the case of the system in section 4.4, we would like to
know what happens to the composition of cloud and rain-water when SO2. sources are
increased (e.g. due to industrial development) or decreased (e.g. when controls are
instigated). Not only must we consider the changes in equilibria, we also must include the
effects of chemical reaction.

We have already seen in Chapter 3 (Crutzen, this volume) the possibility of delays in the
oxidation of SO2. These may be due to effective competition by NO2 for OH radicals in the
gas phase (reaction R10) or by decreases in H2O2 and subsequent decrease in oxidation of
SO2 in cloud-water caused by competition for HO2 radicals by NO (reaction R5 followed by
R36). Such interactions of NOx with the oxidation of SO2 would cause some SO2 to be
transported farther before being oxidized and deposited in rain-water, possibly leading to
increases in the area of regions influenced by acid precipitation. Besides such gas phase
interactions, shifts of equilibria in the liquid phase can lead to similar effects.

It is well established that S(IV) is easily oxidized in aqueous solutions. Models of the
oxidation of SO2 in clouds have been developed that include oxidation by O2 (Scott and
Hobbs, 1967; Easter and Hobbs, 1974; Hegg and Hobbs, 1979). Other models simply treat
the oxidation rate as an input parameter, typically in the range of a few to several percent per
hour (see e.g. Scott, 1981). Considerable effort has been expended on studies of various
liquid-phase reaction mechanisms, including oxidation by O2, O3, and H2O2 and catalysis by
metal ions as well as the role of mass transport of gases within clouds (Freiberg and
Schwartz, 1981). Table 4.7 lists the reaction mechanisms we include in our calculations.

As noted numerous times in the literature, the oxidation rates show dependences on pH and
on NH3. The sense of this dependence is that the rate of conversion of S(IV) to S(VI)
decreases as the aqueous phase becomes acidic. Thus a reaction product, H2SO4, tends to
decrease the rate of reaction indicated earlier for the case of the effect of gas phase reactions
of NOx, acidification of cloud water by the production of H2SO4 should also delay the
oxidation of SO2. To explore the role of this pH dependence, we will calculate the rate of
production of SO42- in a cloud as a function of the SO42- source strength. Although they are
clearly important, we neglect entirely any gas phase reactions. The purpose of this section is
not to study liquid-phase oxidation mechanisms, but rather to use the mechanisms that have
been proposed to address the question of the relationship of SO42- production rates in cloud-
water to SO42- source strengths. We are particularly interested in determining the response of
SO42- production to SO42- source strengths, the nature of the responses (e.g., linear versus
nonlinear) and to what degree this depends on NH3. We herein include a brief description of
our model and qualitative results; the model will be presented in detail elsewhere (Taylor et
al., in preparation).

The presence of ambient ammonia or other basic gases or particles increases the liquid-phase
oxidation rate (Junge and Ryan, 1958; Scott and Hobbs, 1967) by neutralizing the pH as acid
sulphates are formed. The concentration of ammonia in the gas phase has such a large effect
on the rate and total amount of oxidation (Easter and Hobbs, 1974; Adamowicz, 1979;
Overton et al., 1979) that ammonia concentrations and source rates must be studied along
with SO42- concentrations and source rates in any attempt to understand the effect of varying
these parameters on SO42- production.

Table 4.7 Oxidation mechanisms

Rate Equation Constants at 5°C Reference

d[SO4= ] kl = 2.2 x 10-2 min-1


(O2)= k1 + k2 Larson et al.
(1) k2 =4.3 x 102 litre0.5 mol-0.5
[H+]0.5[SO32-] (1978)
dt min-1
d[SO4= ] kO3 = 1.2 x 106 litre0.62 min-1
(O3) = kO3KHO3PO3[HSO3- Penkett et al.
(2) kHO3=2.00 x 10-2 mol litre-1
][H+]-0.38 (1979)
dt atm-1
d[SO4= ]
(H2O2) = kH2O2 kH2O2= 4.0 x 104 litre mol-1 Penkett et al.
(3)
[H2O2][S(IV)] min-1(pH = 4.3) (1979)
dt

There is no general acceptance as to the ambient air concentration of NH3 (Lau and Charlson,
1977; Hales and Drewes, 1979), and there are also major questions as to source strengths of
NH3. Söderlund and Svensson (1976) could only account for approximately one-quarter of
the NH3 sources that are needed to balance the global NH3 cycle. We therefore include a
source of ammonia as a variable parameter in our model and make no attempt to analyse or
justify the relationship between NH3 and SO2 emission rates.

Our calculations were done using a simple box model of uniform wet haze at 5°C. The box is
a closed system except for sources of SO2 and NH3 which are assumed to be instantaneously
and uniformly mixed. The box has dimensions 1 km x 1 km x 1 km and there are no
advection or sink terms. The liquid water content of the haze is determined by the choices of
initial sulphate aerosol concentration, size and sulphate composition of the condensation
nuclei (CN), the fraction of total sulphate used as CN, and the supersaturation. To simplify
the calculation only one form of sulphate ((NH4)2SO4) and one composition and size of CN
are used, resulting in only one size droplets. All droplets have the same concentration of
SO42- and associated cations. Also for simplicity all SO42- in the initial calculations is
assumed to be in the CN.

The droplet equilibrium size is determined by the standard Köhler curve equation. We chose
to use the equilibrium size before activation and to keep the droplet the same size throughout
the oxidation process (this is an approximation that is true only if the increase in mass of the
droplet remains small). The chemical composition of the CN is impure (NH4)2SO4; however,
its effect on the equilibrium droplet size is approximated as if it were pure (NH4)2SO4. The
impurities are assumed to be chemically inert. The parameters remaining that must be defined
are the supersaturation and the mass of the CN, and they in turn determine the resulting liquid
water content, L, and the initial concentration of NH4+ and SO42- in the liquid water.

Our values of 3.0 µm for the droplet radii and 0.01 g m-3 for the liquid water content seem
reasonable when compared to measurements made by Garland et al. (1973). In choosing one
size droplet to comprise the entire liquid water content of the system, 3.0 µm gives a
reasonable number of droplets and liquid water content for a high humidity haze. The liquid
water content measured by Garland et al. (1973) was 0.05 g m-3 40 minutes after the onset of
the fog, so 0.01 g m-3 would seem reasonable for earlier in a developing fog or haze. Clearly,
our assumption that the droplets do not grow must be considered only a first approximation,
taken to avoid the complexity of the cloud droplet growth equations. We also consider higher
liquid water content later in order to assess the sensitivity of the model to this parameter.

The pressures of gas that initially come into contact with the above droplets are 3 ppb SO2,
330 ppm CO2 and a variable amount of ammonia (0.110 ppb). For any given condition the
total amount of NH3 and NH4+ is constant. The resulting equilibrium concentrations are found
by simultaneously solving the equations for PSO2,PNH3 and [H+] from Table 4.5, and taking
into account that total nitrogen will include the [NH4+] from the CN as well as from NH3(g).

Oxidation rate calculations were carried out using the mechanisms listed in Table 4.7 and
starting with the above initial conditions (time = 0). The oxidation proceeded for 30 minutes,
a typical lifetime for a cloud droplet, after which the composition changes were assessed. The
oxidation of S(IV) to S(VI) is assumed to be the rate determining step and considered to be
irreversible.

The concentration of ozone was held constant at 50 ppb during the oxidation. The
concentration of hydrogen peroxide was 0.3 ppb, all of which dissolved in the water to give a
concentration of 1.3 x 10-3 mole litre-1. H2O2 was depleted during oxidation and was solved
simultaneously with the other differential equations.
d[H2O2
PH2O2
] =
where [H2O2] =
kH2O2[H2O2][S(IV)]
LRT
dt

The amount of sulphate produced per cubic meter of fog or cloud by these oxidation
mechanisms was evaluated for nine different SO2 source rates (QSO2). Initial levels of NH3
(0.1, 1.0 and 10 ppb) and a source strength of 4.5 10-4 kg km-2 min-1 were used for three
separate cases. An ammonia source strength of 4.5 10-5 kg km-2 min-1 and a source strength of
NH3 proportional to that of SO2 provide two more cases. These first five cases had liquid
water contents of 0.01 g m-3, while the sixth case had 0.5 g m-3.

Figure 4.9 shows the influence of the initial amounts of NH3, with increasing amounts of
sulphate produced with increasing initial NH3. As expected, the amount of sulphate produced
is a nonlinear function of QSO2. Figure 4.10 compares the cases of fixed QNH3 with one where
NH3 sources are proportional to sources of SO2. This might be the case, for instance, where
NH3 is used as a fertilizer in the same region where SO2 is produced. Here, the amount of
SO42- produced is a nonlinear function of QSO2 in the opposite sensethat is positive as opposed
to negative feedback occurs. Figure 4.11 compares production rates of SO42- for liquid water
contents of 0.01 and 0.5 g m-3 with the initial values and source strength of NH3 as used
previously (Case 1). Here, more SO42- is produced at the higher liquid water content while the
nonlinear behaviour is preserved.

Figure 4.9 Amount of sulphate produced (µg m-3) in model fog of 0.01 g
m-3 liquid water in 30 minutes versus source strength of SO2 (QSO2) kg km-
2
min-1. Other parameters:
Case 1: PNH3 (t=0) = 0.1 ppb
QNH3 = 4.5 10-4 kg km-2 min-1
PSO2 (t=0) = 3 ppb
Case 2: PNH3 (t=0) = 1.0 ppb
QNH3 = 4.5 10-4 kg km-2 min-1
Case 3: PNH3 (t=0) = 10 ppb
QNH3 = 4.5 10-4 kg km-2 min-1

4.6 CONCLUSIONS: FLUX RELATIONSHIPS

The major goals of this approach are to quantify the relationships between fluxes and reaction
rates and to explore departures from linearity. However, before this can be done, it is clearly
necessary to identify key pathways and to establish the sensitivities of the cycles to key
variables. The use of simple, diagnostic models make it possible to ascertain at least some of
the controlling factors. However, we caution that this kind of model is only diagnostic and
not well suited for predictive applications. Perhaps the most important outcome at this stage
is that the results from the model will guide further modelling exercises and will suggest the
important variables that need to be measured. Among the conclusions at this stage are:
Figure 4.10 The same as Figure 4.9, except:
Case
PNH3 (t=0) = 0.1 ppb
4:
QNH3 = 4.5 x 10-5 kg km-2
min-1
Case
PNH3 (t=0) = 0.1 ppb
5:
QNH3 = 0.27 QSO2

Figure 4.11 The same as Figure 4.9, except:


Case L (liquid water content) =
6: 0.5g m-3
PNH3 (t=0) = 0.1 ppb
QNH3 = 4.5 10-4 kg km-2 min-1

1. As expected, QNH3 is important to the pH of atmospheric water and to the rate of


conversion of S(IV) to S(VI). The high liquid-phase solubility at pH < 5 predicates
low (ppt) gas-phase concentrations in many instances. While actual environmental
values for PNH3 and QNH3 are needed to improve the model, the low concentrations of
NH3 make determining these parameters difficult.
2. Increased QSO2 causes increased production of SO42- in wet aerosols, an effect that is
enhanced by increasing QNH3. Decreased pH of atmospheric water causes a decrease
in the rate of oxidation of S(IV) so that the induced production rate of SO42- is a
nonlinear function of QSO2. This could conceivably cause SO2 to be transported
farther prior to oxidation or dry deposition and an increase in the geographical area
influenced by acid deposition.
3. S(IV) in wet aerosols and clouds is apparently an effective sink for low levels of
H2O2. Future models might consider the relationship of QSO2 and QH2O2 in an air
parcel. Since 0.3 ppb of H2O2 is small compared to the initial 3 ppb SO2, it is quickly
consumed. While 0.3 ppb is on the low side of usual atmospheric concentrations
(Kok, 1980) it is reasonable value for areas not influenced by much photochemistry.

These results are consistent with the observations of Altshuller (1976, 1980) that reductions
in local SO2 emissions in the Eastern and Midwestern U.S.A. between 1963 and 1972 caused
only modest reductions in sulphate aerosol concentration. Data collected in 1974 in Europe
(LRTAP) similarly showed a possible nonlinear relationship of sulphate in rain-water as a
function of total sulphur (SO2 and SO42-) (OECD, 1977). In their cloud model, Easter and
Hobbs (1974) found, using several oxidation mechanisms, that sulphate production decreased
as initial SO2 rose above 10 ppb. This is also consistent with the lack of a clear-cut trend of
decreasing pH in rain in recent years in Scandinavia (Granat, 1978).

Since oxidation of sulphur dioxide is considered to be the major source of sulphate aerosols
(Gartrell and Friedlander,1975; Yuen et al., 1979) as well as the major source of acid
precipitation and since reaction in aerosol and cloud particles seems to be possible or
probable (Hegg and Hobbs, 1978, 1979) it is worthwhile to develop a more complete and
quantitative understanding of these interactions. We fully can expect these interactions to be
nonlinear due to both homogeneous gas-phase reactions and heterogeneous cloud and aerosol
reactions. The expected nonlinearity in both cases is a slowing-down or delay of the
oxidation of SO2 as SO2 and NO sources increase. However, the degree of nonlinearity is not
known at this time, resulting in uncertainties in the effectiveness of various control strategies.

ACKNOWLEDGEMENTS

Financial support was provided by the International Meteorological Institute in Stockholm


and also by the National Science Foundation, under Grant ATM 7808163. Special thanks are
due to Professor Bert Bolin for suggesting the topic, and Professors Henning Rodhe, Peter
Liss, Robert Duce, Dr. Rolf Söderlund and Dr. Robert Cook for helpful discussions.

4.7 REFERENCES

Adamowicz, R. F. (1979) A model for the reversible washout of sulphur dioxide, ammonia
and carbon dioxide from a polluted atmosphere and the production of sulphates in raindrops,
Atmos. Environ.,13, 105-121.

Altshuller, A. P. (1976) Regional transport and transformation of sulphur dioxide to sulphates


in the U.S., J. Air Pollut. Cont. Ass., 26, 318-324.

Altshuller, A. P. (1980) Seasonal and episodic trends in sulphate concentrations (19631978)


in the eastern United States, Environ. Sci. Technol., 14, 1337-1348.

Charlson, R. J., Covert, D. S., Larson, T. V., and Waggoner, A. P. (1978) Chemical
properties of tropospheric sulphur aerosols, Atmos. Environ.,12, 39-53.

Coborn, W. G., Djukic-Husar, J., and Husar, R. B. (1980) Monitoring of sulphuric acid
episodes in St. Louis, Missouri, J. Geophys. Res., 85(C8), 4487-4494.

Crutzen, P., Atmospheric interactionshomogeneous gas phase reactions, Chapter 3, this


volume.

Duce, R., Biogeochemical cycles and the airsea exchange of aerosols, Chapter 16, this
volume.

Easter, R. C., and Hobbs, P. V. (1974) The formation of sulphates and the enhancement of
clouds condensation nuclei in clouds, J. Atmos. Sci., 31,1586-1594.

Freiberg, J. E. (1974) Effects of relative humidity and temperature on iron catalyzed


oxidation of SO2 in atmospheric aerosols, Environ. Sci. Technol, 8, 731-734.

Freiberg, J. E., and Schwartz, S. E. (1981) Oxidation of SO2 in aqueous droplets: mass-
transport limitation in laboratory studies and the ambient atmosphere, Atmos. Environ., 15,
1145-1154.

Garland, J. A., Branson, J. R., and Cox, L. C. (1973) A study of the contribution of pollution
to visibility in a radiative fog, Atmos. Environ., 7, 1079-1092.
Gartrell, G. Jr, and Friedlander, S. K. (1975) Relating particulate pollution sources: the 1972
California Aerosol Characterization Study, Atmos. Environ., 9, 279-299.

Granat, L. (1978) Sulphate in precipitation as observed by the European Atmospheric


Chemistry Network, Atmos. Environ., 12, 413-424.

Hales, J. M., and Dana, M. T. (1979) Regional-scale deposition of sulphur dioxide by


precipitation scavenging, Atmos. Environ., 13, 1121-1132.

Hales, J. M., and Drewes, D. R. (1979) Solubility of ammonia in water at low concentrations,
Atmos Environ., 13, 1133-1147.

Hegg, D. A., and Hobbs, P. V. (1978) Oxidation of sulphur dioxide in aqueous systems with
particular reference to the atmosphere, Atmos. Environ., 12, 241-253.

Hegg, D. A., and Hobbs, P. V. (1979) The homogeneous oxidation of sulphur dioxide in
cloud droplets, Atmos. Environ., 13, 981-988.

Huebert, B. J., and Lazrus, A. L. (1980) Bulk composition of aerosols in the remote
troposphere, J. Geophys. Res., 85, 7337-7344.

Johnson, J. E. (1981) The lifetime of carbonyl sulphide in the troposphere, Geophys. Res.
Lett., 8, 938-940.

Johnstone, H. F., and Leppla, P. W. (1934) The solubility of sulphur dioxide at low pressures,
J. Am. Chem. Soc., 56, 2233-2238.

Junge, C. E., and Ryan, T. G. (1958) Study of SO2 oxidation in solution and its role in
atmospheric chemistry, Q. J. R. Met. Soc., 54, 46-55.

Junge, C. E. (1963) Air Chemistry and Radioactivity, New York, Academic Press.

Kok, G. L. (1980) Measurements of hydrogen peroxide in rainwater, Atmos. Environ., 14,


653-656.

Larson, T. V., Horike, N. R., and Harrison, H. (1978) Oxidation of sulphur dioxide by
oxygen and ozone in aqueous solution: a kinetic study with significance to atmospheric rate
processes, Atmos. Environ., 12, 1597-1611.

Lau, N., and Charlson, R. J. (1977) On the discrepancy between background atmospheric
ammonia measurements and the existence of acid sulphate as a dominant atmospheric
aerosol, Atmos. Environ., 11, 475-478.

Liljestrand, H. M., and Morgan, J. J. (1981) Spatial variations of acid precipitation in


Southern California, Environ. Sci. Technol. 15, 333-339.

Lowell, D. S. et al. (1970) A theoretical description of the limestone injection-wet scrubbing


process. National Technical Information Service Publication, Vol. I, No. Pb 193-029.
Maroulis, P. J., Torres, A. L., Goldberg, A. B., and Bandy, A. R. (1980) Atmospheric SO2
measurements on Project Gametag, J. Geophys. Res., 85, 7345-7349.

Morgan, O. M., and Maass, O. (1931) An investigation of the equilibrium existing in


gaswater systems forming electrolytes, Can. J. Res., 5,162-199.

OECD (1977) The OECD Programme on long range transport of air pollutants,
measurements and findings. Organization for Economic Cooperation and Development,
Paris, France.

Overton, J. H. Jr., Viney, P. A., and Durham, J. L. (1979) Production of sulphate in rain and
raindrops in polluted atmospheres, Atmos. Environ., 13, 355-367.

Penkett, S. A., Jones, B. M., Brice, K. A., and Eggleton, A. E. (1979) The importance of
atmospheric ozone and hydrogen peroxide in oxidizing sulphur dioxide in clouds and
rainwater, Atmos. Environ., 13, 132-137.

Petrenchuk, O., and Drozdova, V. M. (1966) On the chemical composition of cloud water,
Tellus, 18, 280-286.

Robinson, R. A., and Stokes, R. H. (1959) Electrolytic Solutions, Butterworths, London, 559.

Scott, B. C. (1981) Predictions of in-cloud conversion rates of SO2 to SO4 based upon a
simple chemical and kinematic storm model. Battelle Pacific Northwest Laboratory, PNL-S
80538, Richland, WA.

Scott, B. C., and Laulainen, N. S. (1979) On the concentration of sulphate in precipitation, J.


Appl. Meteorol., 18, 138-147.

Scott, W. D., and Hobbs, P. V. (1967) The formation of sulphate in water droplets, J. Atmos.
Sci., 24, 54-57.

Shaw, R. W., and Stevens, R. K. (1980) Trace element abundances and chemistry of
atmospheric aerosols: current techniques and future possibilities. Ann. New York Acad. Sci.,
338, 13-25.

Shaw R. W., and Rodhe, H. (1981) Non-photochemical oxidation of SO2 in regionally


polluted air during winter. Report CM53, Department of Meteorology, University of
Stockholm.

Sillén, L. G. (1967) Master Variables and Activity Scales, in Gould, R. F. (ed.) Equilibrium
Concepts in Natural Water Systems, Advances in Chemistry Series No. 67, Washington,
American Chemical Society, 45-56.

Söderlund, R., and Svensson, B. H. (1976) The global nitrogen cycle, in Svensson, B. H. and
Söderlund, R. (eds) Nitrogen, Phosphorus, and SulphurGlobal Cycles, SCOPE Report No. 7,
Ecol. Bull. (Stockholm), 22, 23-73.

Stevens, R. K., Dzubay, T. G., Russworm, G., and Rickel, D. (1978) Sampling and analysis
of atmospheric sulphate and related species, Atmos. Environ.,12, 56-68.
Tang, I. M. (1980) On the equilibrium partial pressures of nitric acid and ammonia in the
atmosphere, Atmos. Environ., 14, 819-828.

Taylor, G. S., Baker, M. B., and Charlson, R. J. (19XX) Non-linearity of Sulphate Production
in Cloud as a Function of SO2 Source Strength. Manuscript in preparation, University of
Washington, FC-05, Seattle, WA 98195, USA.

Whitby, K. T., and Sverdrup, G. (1980) California Aerosols: Their Physical and Chemical
Characteristics, in Hidy, G. M., et al. (eds) The Character and Origins of Smog Aerosols,
New York, Wiley Interscience, 499.

Yuen, T. S., Harrison, H., Charlson, R. J., and Baker, M. B. (1979) The global sulphur box
model, Atmos. Environ., 13, 1351-1360.

Yui, T. (1940) On the electrolytic dissociation constant of sulphurous acid, Tokyo Inst. Phys.
Chem. Res. Bull., 19, 1229-1236.

Back to Table of Contents

The electronic version of this publication has been prepared at


the M S Swaminathan Research Foundation, Chennai, India

SCOPE 21 -The Major Biogeochemical Cycles and Their


Interactions
5 Transport Processes in the Biogeochemical Cycles of Carbon, Nitrogen,
Phosphorus, and Sulphur
W. A. REINERS

Abstract
5.1 Introduction
5.2 Transport Processes
5.2.1.Mobilization Processes
5.2.2 Transport Processes
5.2.3 Deposition Processes
5.3 Inter-Biome Transport Via the Atmosphere
5.3.1. Residence Times and Travel Distances
5.4 Summarizing Conclusions.
Acknowledgements
References

ABSTRACT
Material transport occurs over very wide ranges of space and time, and is an intrinsic part of
biogeochemical cycles. The transport aspects of the carbon, nitrogen, phosphorus, and
sulphur cycles are examined in this paper. This examination has three principal sections. The
first is a broad review of the many processes underlying transport of these elements, ranging
from intercellular translocation within organisms to global atmospheric circulation. The
processes are divided into mobilization, transport itself, and deposition of materials. The
second section is a more detailed analysis of atmospheric transport of the several forms of
these four elements. The third section summarizes the conclusions.

Throughout this paper, special attention is given to interactions between C, N, P and S cycles.
Such interactions are most pronounced in five areas. (1) biological mechanisms of
mobilization; (2) human-caused mobilization employing energy derived from fossil fuels; (3)
photochemical reactions in the atmosphere; (4) biological mechanisms of deposition; and (5)
influences of anthropogenically derived acids.

5.1 INTRODUCTION

Material transport is, by definition, an intrinsic part of biogeochemical cycles. There can be
no cycles without movement of matter. Movement occurs over a huge range of spatial scales.
It can range from microns, in the circulation of ions within unicellular organisms, to tens of
thousands of kilometers in global atmospheric circulations. As units of matter cross critical
boundaries, the scale of their movement can shift abruptly. An ion or molecule may escape
from a very small-scale cycle like an intraplant system and become incorporated in a global-
scale transport process, and thereby a larger-scale cycle. The reverse must also happen at
approximately equal rates. These shifts in scale of movement underlie the loss and input rates
of specific cycling systems. At the same time, these shifts represent the interconnectedness of
cycles of all scales.

Movement also occurs over a very wide range of rates. Rate of transport varies enormously in
both an absolute sense, such as km/yr, and in a relative sense, such as percent of the cycling
system traversed/yr. Both scales are essential considerations for evaluating the behaviour of
all cycling systems, whether they are single organisms, isolated ponds, or the entire planet.

In this paper, transport of matter is examined in the context of the cycles of carbon, nitrogen,
phosphorus, and sulphur. The objectives of this paper are to: (1) broadly review transport
processes for these elements; (2) highlight critical interactions between these elements in the
transport aspects of their cycles; and (3) illustrate and suggest approaches that may be useful
in modeling and predicting transport of these elements on a regional or larger space scale.

This paper is organized into three sections. In the first and main section, transport process is
defined and described. The second section is devoted to atmospheric transport in particular.
The last section is a brief list of summarizing conclusions.

5.2 TRANSPORT PROCESSES

Transport can be separated into three phases: mobilization, transport perse, and deposition.
Mobilization is a necessary part of moving a substance from one place to another; it involves
freeing the substance from an immobile, fixed form so that it can be moved by a carrier force.
Physical and chemical weathering, so-called biological mineralization, soil deflation, and gas
production through combustion are examples of mobilization.
The transport phase entails the actual movement of a substance from one point to another.
The deposition phase includes those processes that cause immobilization of a transported
substance. These include, for example, chemical precipitation, gas absorption, and inertial
impaction.

With this broad view of transport, a very large list of processes can be generated. I have
attempted to arrange many of the possible processes into categories in order to more easily
make comparisons of their importance for specific elements, spatial scales, or geographic
areas (Table 5.1).

5.2.1.Mobilization Processes

I have divided mobilization processes into three major classes: (A) natural biological-caused,
(B) natural physical-caused, and (C) human-caused processes. As with all classifications, this
approach is arbitrary to some degree, so that some overlaps between some classes occur and
important interactions between processes are obscured.

Table 5.1 Major processes involved in the transport of carbon, nitrogen, phosphorus, and
sulphur. Substances listed parenthetically in Sections A and C are merely representative of
families of products; e.g., CO32- includes its equilibration members in aqueous solution.
Relative ranges of mobility in Section B are the following: very local, < 1 km; local, 110 km;
mesoscale, 10100 km; regional, 1001000 km; continental, 10003000 km; hemispheric, >
3000 km < global; global, > 40,000 km.

A. Mobilization Processes and Products


Biological Processes
Plant processes
`Mineralization' of carbon compounds by respiration (CO2)
Production of volatile organics (terpenes)
Movement of chemical substances from internal tissues to surfaces (HCO3- ,
NH4+, NO3-, PO43-, SO42-)

Animal processes
`Mineralization' of organic materials (CO2, CH4, urea, NH4+)
Incorporation of food and ions into mobile animals

Microbial processes
`Mineralization' (CO2, CO, NH3, PO43-, SO42-)
Nitrification (NO2-, NO3- )
Denitrification (NO, N2O, N2)
Sulphate reduction (H2S, dimethylsulphide)
Organic carbon reduction (CH4)
Phosphate reduction (P2H4)

Natural Physical Processes


Mineral weathering
Physical weathering (inorganic and fossil organic particulates)
Chemical weathering (S2-, SO42-, PO33-)

Soil processes
Cation exchange (NH4+)
Entrainment by wind and moving water (particulate organic and inorganic
forms)
Ammonia volatilization (NH3)
Glacial acquisition (organics and inorganics)
Marine aerosol formation from evaporated spray (HCO3- , PO43-, S2-, SO42-)
Volcanic exhalations (CO2, SO2, NH3)
Atmospheric electrical discharges (NO3)
Wildfires (hydrocarbons, CO2, CO, NH3, NOx, PO43-, SO2, COS)
Suspension of particulates on leaf surfaces (HCO3-, NH4+, SO42-)

Human-caused Processes
Wood fuel combustion (see above)
Land management fires (see above)
Fossil fuel combustion (CO2, CO, NOx, N2, SO2)
Agricultural land tillage (organic and inorganic soil and crop particulates)
Food and fibre processing (dissolved and particulate organics, soil)
Fertilizer manufacture and application (CO2, NH3, NO3-, PO43-, SO42-)
Mining (organic and inorganic particulates, S2-, SO42-, PO43-)
Smelting (S2-, SO2, CO2)
Petrochemical extraction, refining and transport (volatile hydrocarbons)
Industrial processing (miscellaneous)

B. Transport Mechanisms and Relative Ranges of Transport


Biological Transport Mechanisms
Plant uptake from soil and translocation (very local)
Canopy leaching to throughfall-stemflow (very local)
Plant litterfall to ground (very local)
Animal uptake and subsequent movement, excretion, death (very local to
continental)

Mass-wasing Transport Mechanisms (very local to local) Glacial Transport (very


local to continental)
Fluvial Transport Mechanisms (particulate and dissolved, organic and inorganic)
Saturated and unsaturated flow in soil (very local)
Ground-water flow (very local to regional)
Surface-water flow (very local to continental)
Oceanic currents (hemispheric to global)

Atmospheric Transport Mechanisms (gases and aerosols)


Saltation (very local to mesoscale)
Suspension (mesoscale to global)
Solution (mesoscale to global)
Human Transport (food and fibre products, ores, industrial products (mesoscale
to global))

C. Deposition Processes and Materials


From Mass-wasting Transport
Deposition at base of slope (massive organics, inorganics)

From Glacial Transport


Deposition as till, kames, moraines, out-wash, etc. (organics, and especially
inorganics of all sizes)

From Fluvial Transport


Uptake of dissolved substances of sessile organismse.g., vascular aquatic plants,
corals (HCO3-, NO3- , PO43-)
Nitrogen fixation by sessile organisms (N2)
Physical sedimentation (organic and inorganic particulates)
Chemical sedimentation (CO32-)

From Atmospheric Transport


Wet deposition
Incident precipitation via rainout and washout (HCO3-, NH4+, NO3-, PO43-,
SO42-)

Dry deposition
Sedimentation (large particles)
Gaseous absorption
Active: photosynthesis (CO2), N fixation (N2)
Passive absorption to chemical sinks by foliage, soil, water (CH4,
hydrocarbons, NH3, NOx, SOx, H2S)
Inertial impaction (soil particulates, HNO3, NH4Cl, H2SO4, (NH4)2SO4,
MPO4 aerosols > 1 µm)
Molecular diffusion (HNO)3, NH4C1, H2SO4, (NH4)2SO4 aerosols < 0.1 µm)

A. Biological Mobilization Processes

Biological mobilization processes are subdivided into those promulgated by plants, animals,
and micro-organisms. Four general comments can be made about these biological processes.
First, all of them are globally ubiquitous, but between ecosystem types there are very definite
differences in the strength of their activities. Fresh-water wetlands, coastal marshes, and
shallow shelf areas are particularly important for the critical reduction activities listed under
microbial processes. With closer analysis, at least a graded series of importance by ecosystem
type could be developed.

The second comment is that microbial processes, in particular, give rise to mobile products
that can become involved in long range transport as well as local, intrasystem cycles. This is
particularly true for the microbiologically mediated reduction processes leading to formation
of highly mobile gaseous forms of these elements.

The third point is that the strength of all of these processes is roughly proportional to the
cycling rate or resource flow-through rates of ecosystems. Ultimately, all of these functions
depend on cell numbers and cellular metabolic rates. These, in turn, are dependent on
resource supplies. Cells require C, N, P, and S in particular ratios for accumulation of mass
and numbers and in different ratios for maintenance. The required proportions of these
elements can be highly variable, depending on the specific system; but, in the last analysis, all
of these four elements, and other elements, can be rate-limiting resources. This is a
fundamental, underlying interaction between C, N, S, and P for all biologically driven
portions of biogeochemical cycles.

The fourth point regarding biological processes is that virtually all of them are sensitive to
pH. Thus, there are important interactions between biological rates and the concentrations of
acids of C, N, S, and P. The most interesting of these from a perturbation viewpoint involves
the effects, both direct and indirect, of anthropogenically derived sulphuric and nitric acids.
Cook (Chapter 12, this volume) reviews many of these effects.

B. Natural Physical Mobilization Processes

The second class of mobilization processes are physical-chemical phenomena occurring


whether or not man is a dominating influence. This class is further subdivided into mineral
weathering, soil-centred, and miscellaneous other processes.

Mineral weathering as a direct, liberating process is much more pertinent for C, P, and S than
for N, because rocks generally do not contain much N. Nevertheless, there are important
relationships between N and weathering, as will be discussed later. Rates of mineral
weathering are highly variable over the face of the earth. Weathering potential definitely has
predictable, geographic patterns in that it is roughly correlated with temperature and
precipitation (Strakhov, 1967; Loughnan, 1969; Birkeland, 1974). On the other hand, this
potential can be approached only to the extent that weatherable minerals occur in a
weathering zone. In large part, this availability is related to rock type, relief, and youthful
terrain (Gibbs, 1967; Johnson et al., 1968; Reynolds and Johnson, 1972). All of these
geological factors have complex geographic patterns and are not usually characteristic of
ecosystem types. For example, deglaciated terrainyouthful in terms of exposed, weatherable
mineralsis located under arctic tundra, boreal forest, deciduous forest, grasslands, and
mountain complexes. On the other hand, one ecosystem type, such as the tropical forest
biome, covers a wide range of relief (Gibbs, 1967). Thus, a geographic model of weathering
rate as a mobilization process for transport is not likely to be feasible using climatic or
ecological criteria for geographic units. Rather, such a model would have to be based on
specific conditions, especially geological characteristic, including relief and rock type, in
conjunction with climatic and biotic factors.

Some of the most interesting interactions between C, N, P, and S relevant to weathering occur
through the effects of biota. Plants can contribute to weathering through mechanical effects,
but the most important influence of all the biota is probably through the excretion of
hydrogen ion and the generation of acids (carbonic, organic, nitric, and sulphuric) (Gorham et
al., 1979). Thus, weathering is influenced by acid production, which is a product of the
amount and kind of plant cover, organic detritus, and microbiota present. These, in turn, are
influenced by the supply of limiting resources, including C, N, P, and S. Another interaction
involves the reciprocal effects of rock-weathering rate and N-fixation rate in primary
succession or on impoverished substrates (Reiners, 1981). Nitrogen fixation is apparently
dependent on an adequate P supply and to a lesser extent on pH, especially for rhizobia
(Granhall, 1981). Thus, the weathering rate of apatite and metallic cations may influence the
N-fixation rate. To the extent that N supply limits biological activity, it can, in turn, influence
weathering rates by limiting acid production, so that there is a mutual feedback between
weathering and N fixation under these circumstances.

Mineral acids of anthropogenic origin that are not neutralized by the soil should contribute to
mineral weathering (Gorham and McFee,1980; Turk and Peters, 1979) or be transferred to
stream ecosystems. Enhanced weathering effects caused by such mineral acids have been
undetectable or are of very local extent (Johnson et al., 1972; Johnson et al., 1981) or are
confused by soil leaching effects. In some ecosystems, disturbance may lead to enhanced
nitric acid production via nitrification (Vitousek et al., 1979), possibly enhancing weathering
potential temporarily.

Glacial acquisition of materials from overridden plants, soils, and rocks can represent
enormous mobilization rates for local environments or particular periods of history. While
most of the burden of glaciers is mineral, and therefore mainly of importance for inorganic
forms of C, P, and S, some of the burden can be organic and thus include N.

Three processes are included under the soil category in Table 5.1. Ammonium is the only
cationic form of all four elements, and thus only nitrogen can be mobilized by cation
exchange. It can be mobilized for leaching transport through replacement by other cations,
especially hydrogen. Hydrogen can be supplied by airborne nitric and sulphuric acids, as
described above, and by organic, carbonic, and nitric acids formed by natural processes in the
soil.

Entrainment is the process by which loosened material is acquired by wind and water as
transport agent. Natural wind and water erosion are associated principally with arid and semi-
arid climates, but these zones can be enlarged by human cultivation of former grasslands. The
relationship between mean annual precipitation and wind and water erosion is shown in
Figure 5.1 (from Marshall, 1973). Curve A indicates that under natural vegetation, the peak
erosion potential occurs in areas that receive about 300500 mm precipitation per yearsemi-
arid environments. Under these conditions precipitation is inadequate to allow formation of
complete canopy cover. The relationship between canopy cover and wind and water erosion
is shown in Figure 5.2. Curve B of Figure 5.1 shows the erosion potential from bare grounda
condition that exists at least for short periods of time following cultivation or some fires.
Curves C and D of Figure 5.1 show the relationship between precipitation and wind erosion
potential under natural vegetation and bare ground. Wind erosion decreases as vegetation
cover increases due to increased water available for plant growth. Wind erosion from bare
ground decreases with increased precipitation, presumably because moist soils are less
susceptible to wind erosion. Clearly, when soils are dry and plant cover is low, potential
erosion is high from both wind and water. Excellent reviews of these processes are found in
Statham (1977), Foster (1977), and Branson et al. (1981).

Atmospheric transport of deflated soil material is of great importance at local, regional, and
global scales. Idso (1976) reviewed the meteorology and world geography of dust storms.
Most aspects of wind erosion in the Sahara are presented in more detail by Morales (1979).
The geography of deflation in the United States can be seen in Fletcher et al. (1978). For this
process, a geographic model for dust source areas might well be associated with the desert
and grassland biomes, but better resolution could be gained from other geographic criteria, as
demonstrated by the wind erosion map in Fletcher et al. (1978).

Figure 5.1 Relationships of water erosion (continuous lines) and wind erosion (broken lines)
with increasing mean annual precipitation. The curves for water erosion indicate the
relationships with mean annual precipitation for (A) areas of natural vegetation cover and (B)
bare ground (after Schumm, 1969). The curves for wind erosion indicate the relationships
with mean annual precipitation for (C) areas of natural vegetation cover and (D) bare ground.
These curves are based on what would be expected from the relationship of wind erosion to
vegetation cover and to moist soil (from Marshall, 1973). Reproduced from Drought, edited
by J. V. Lovett by permission of Angus & Robertson Publishers, Sydney, Australia

There are only weak interactions between elements involved in the mobilization of detached
particulates, but there are definite chemical associations for the material transported. Wind-
and water-transported material, especially that transported long distances, is enriched in
lighter fractions of the surface soil horizons (Branson et al., 1981). These fractions are often
rich in organic matter and thus richer in C, N, S, and P than deeper soil horizons or soil parent
material. This enrichment may be minimal when the transported material is derived from
desert soils that have not accumulated much organic matter (Rahn et al., 1979) and may be
maximal for former grassland soils that have been turned to the plough. Areas undergoing
desertification might produce dust and sediments richer in C, N, S, and P than traditional
desert sources do. Dust derived from recently ploughed grasslands of Oklahoma, U.S.A., had
a loss-on-ignition percentage of 7.26, indicating a carbon percentage of 3.6; nitrogen,
phosphorus, and sulphur percentages were 0.19, 0.08, and 0.07, respectively (Kilmer,1979).

Figure 5.2 Interrelationships between drought, plant cover, and soil erosion by wind and by
water. As drought severity increases, plant cover decreases. The markedly nonlinear nature of
the relationship of wind and of water erosion to plant cover is noteworthy (from Marshall,
1973). Reproduced from Drought, edited by J. V. Lovett by permission of Angus &
Robertson Publishers, Sydney, Australia

The process of ammonia volatilization is mainly related to ammonia supply and soil pH.
Thus, interactions between the biogenic elements are not particularly critical. Söderlund and
Svensson (1976) indicate that most naturally derived ammonia comes from animal
excrement, but broad geographic areas with alkaline soils, such as grasslands (Woodmansee
et al., 1981), may be important sources of ammonia. Fertilization of circumneutral soils with
urea and anhydrous ammonia, coupled with extensive periods of little or no crop cover,
would seem to be a very important ammonia source (J. Freney and R. Woodmansee, this
workshop, personal communication). Atmospheric ammonia concentrations over the ocean,
tropical land areas, and `other' land areas (Table 5, Söderlund and Svensson, 1976; Figure
5.3a) suggest that the land is a more important source than the oceans and that tropical
regions may be especially strong sources. The same regions producing high levels of
ammonia may also be major sinks, as indicated in Table 14 of the same reference. Junge's
map (1958) of ammonia in precipitation in the United States strongly suggests a Great Plains
origin for much of the ammonia returned in wet deposition further east (Figure 5.3b).

Figure 5.3 Maps indicating distributions of NH3 in the United States and north-western
Europe. (Above) Average NH4+ concentration (mg/litre) in rain over the United States in
JulySeptember 1956 (from Junge, 1958). Reproduced by permission of American
Geophysical Union. (Below) Mean annual surface concentrations of NH4 compounds and
NH3 gas for the years 19681972 (from Söderlund, 1977). Reproduced by permission of the
Royal Swedish Academy of Sciences

The increased use of ammonia fertilizer and development of feedlots in that region may have
significantly increased the emission levels since Junge's data were collected in 1955. Most of
the ammonia that combines with sulphate further downwind over the more industrialized
parts of the country may be derived from natural soil processes augmented by agricultural
practice. In the North-western Europe case (Figure 5.3b), the largest contributory source is
believed to be domestic animal waste, followed by coal combustion and fertilizers
(Söderlund, 1977). A good potential may exist for producing geographic models of ammonia
sources for global modelling. Grasslands, croplands, and tropical regions may compose large
natural sources that might be overlain with specific industrial sources. There are no obvious
chemical interactions between C, N, P, and S aside from those general ones mentioned earlier
concerning nutrient limitations for biological processes and mineral acid effects.

Aerosols created by ocean spray jets contain organic C, organic N, organic and inorganic
phosphate, nitrate, and especially sulphate (Söderlund and Svensson, 1976; Pierrou,1976;
Granat et al., 1976; Graham et al., 1979; Duce, Chapter 16, this volume). Six to ten percent
of the aerosols so generated fall on land (Granat et al., 1976), so there is a net transport by
this means from the sea margin to several hundred kilometers over land. Such transport
obviously represents a transfer from oceans to terrestrial ecosystems but does not
discriminate between terrestrial ecosystem types.

Volcanic emissions have been extremely important in the geochemical development of the
earth and its atmosphere. Estimates of contemporary contributions vary widely; most recently
they have been raised for sulphur to 29 Tg/yr (Ivanov, 1981; Freney and Rodhe, Chapter 2,
this volume). These emissions are major sources for remote areas and primary sources of
temporal variation over larger areas. The distribution and activity of volcanoes are readily
mapped for geographic modelling, but the distribution of active volcanoes does not coincide
with biome distributions.

Wildfires are very important sources of various compounds of C, N, S, and even P. Fires have
consistently been considered in the major SCOPE papers of chemical cycles (Svensson and
Söderlund, 1976; Bolin et al., 1979a), and details on emissions have been reviewed (e.g.,
Raison, 1979; Crutzen et al., 1979; Crutzen, Chapter 3, this volume). A particularly good set
of reviews on the geography of fire is in Mooney et al. (1981), and a good review on
secondary effects of burning on chemical cycles is provided by Woodmansee and Wallach
(1981). Wildfires are more wide-spread than commonly is thought, but a mapping pattern of
emissions in time and space from biomes having seasonal climates is definitely feasible.
Chemical interactions are especially interesting for this process, both in increasing the
probability of fire (e.g., N limitations in soil), in chemical conversions in the atmosphere, and
in post-fire effects (Woodmansee and Wallach, 1981).

The physical process of suspending particles from plant canopies into the air is directly linked
to the mobilization of materials from the interior of plant tissues to the surface. This
suspension, and resuspension, is very difficult to measure in the field and to treat in a
theoretical way (Fish, 1972; Slinn, 1976; Beauford et al., 1977). The chemical nature of such
suspended particulates is not well known, but C, S, and P are certainly involved. Tropical
forests may be especially strong sources of such particulates (Crozat, 1978; Hogan and
Mohnen, 1979; Lawson and Winchester, 1979). Much of this suspended material may be
resuspended locally, but some may escape for intercontinental transport (Hogan and Mohnen,
1979). Such local transport also greatly complicates the measurement of atmospheric inputs
for watershed budget studies because of contamination of bulk precipitation collections.

C. Human-caused Mobilization Processes

All the aspects of human-caused mobilization are familiar to most readers and are covered in
one form or other in the major SCOPE publications on the global cycles of C, N, P, and C
(Svensson and Söderlund, 1976; Bolin et al., 1979b). Human activities can largely be viewed
as local accelerations of natural processes, and involve the same chemical interactions. The
effects of man on erosion can be seen in Figures 5.1 and 5.2. Where natural vegetation occurs
and is not seriously overgrazed, erosion losses are generally low compared with sites that
have been denuded by cultivation or other land-clearing activities. Wood fuel and land
management fires will differ in seasonality and secondary effects from wildfires, but will
produce similar emission products. Fossil fuel combustion, fertilizer applications, mining,
and smelting are very much like intensive weathering processes. Agricultural tillage and food
and fibre processing are exaggerated versions of soil processing by meso- and micro-fauna
and flora. Petrochemical extraction and processing is a vast enhancement of reactions
resulting from natural petroleum or asphalt seepage losses. Industrial processing, on the other
hand, includes syntheses of materials that are totally exotic in the natural environment, such
as some halogen-containing organic substances.

The major difference between human-caused and natural mobilization processes is in the
geographic distribution, extreme localization, and very high intensity of the human-caused
processes. All three of these properties are well known from economic and marketing data.
Just as the rate of fossil fuel combustion is the best-established datum for any flux in the C
cycle, so should these other human-caused mobilization processes be easily estimated
compared with natural processes. In order to establish a geographic model of sources and
sinks, these estimates can be set as overlays on more diffuse natural processes.

One very important chemical interaction underlying these human-caused processes is the
intense relationship between hydrocarbons and other mobilized substances. Except for fires
and non-industrialized agriculture, the bulk of these processes is driven by fossil-fuel-
(hydrocarbon) derived energy.

5.2.2 Transport Processes

Following mobilization, substances can be transported by a variety of agencies. Remaining as


inclusive as possible, I have listed four classes of transport mechanisms representing all
scales of transport distance (Table 5.1).

A. Biological Transport Mechanisms

The mechanisms listed in this category are significant almost exclusively on a short range
basis. The three plant processes listed are of little consequence on even an inter-ecosystem
spatial scale, but they are essential in the maintenance of local, or `internal', cycles
(Rosswall,1976). Animal transport is occasionally an important medium range phenomenon,
as in the cases of marine birds creating guano deposits, anadromous fish fertilizing small
streams, or , migratory herds of ungulates moving across vast areas.

B. Mass-wasting Transport Mechanisms

Mass-wasting is the gravitational transfer of material. It includes a number of typescreep,


solifluction, mudflows, landslides, avalanches, and subsidences. Naturally, such movements
are usually associated with substantial relief in the terrain, whether terrestrial or submarine.
Mobilization for mass-wasting can arise from weathering processes, but other factors may be
the causes, such as volcanism or tectonic movements. In any case, all four elements are
usually involved, and the range of movement generally is less than 10 km. For an excellent
estimate of the role of mass-wasting in material transfer on mountainous watersheds, see
Swanson et al. (1981).

C. Fluvial Transport Mechanisms

Water movement over the earth's surface is an obvious and primary mechanism for the
transport of nearly all substances. This process might actually be considered to begin with
atmospheric wash-out, but that is classified later as a depositional mechanism. Mechanisms in
Table 5.1 begin with transport through the solumwhat is commonly termed leaching. This is a
very critical process and includes a number of interesting chemical interactions. For example,
the leachability of cations like ammonium is partly a function of the concentration of other
cations like hydrogen. Hydrogen can be derived from acids of C (carbonic), N (nitric), and S
(sulphuric). The last two can be of natural or anthropogenic origin. Leachability is also a
function of the concentration of soluble anions such as bicarbonate, nitrate, and sulphate.
Although anions are not retained against the mass flow of water by exchange reactions to the
extent demonstrated by cations, there can be a variable amount of anionic absorption,
depending on the mineralogy and other chemical attributes of the soil (Nye and Tinker,
1977). Soil leaching need not lead to total removal from the soil profile. Much of the transfer
can be over short distances, simply to lower horizons. Soil solutions can even become diluted
as they pass through B, C, and D horizons (Cronan and Schofield, 1979).

Ground water is 18.8% of all the water on earth and 95.7% of all fresh water (Garrels et al.,
1975). The flux rate through this reservoir is highly variable, ranging from very fast in karst
regions, to exceedingly slow in the more massive aquifers, to functionally zero in cases of
connate water. The chemistry of ground water in a wide variety of materials in the United
States is thoroughly reviewed in White et al. (1963). The importance of ground water as a
transport medium is as variable as the world's geology. Certainly, fluxes could not be
generalized on a biome or even ecosystem basis; transport would have to be estimated on an
individual region basis.

The movement of dissolved and solid materials from land to lakes, reservoirs, and the sea by
surface run-off is one of the most important fluxes for all biogeochemical cycles. The
hydrologic cycle has the net effect of collecting a diffuse source of energy, precipitation, into
a concentrated form, running water (Statham, 1977). During this concentrating process, soil
material is transported from the upland positions of watersheds to lowland positions as
dissolved ions, suspended substances, and bedload. In addition, the greater energy of flowing
water in the lower reaches of drainages can cause channel cutting, thereby greatly increasing
the entrained material in the water. The importance of these entrained materials in streams
and rivers is addressed by Richey (Chapter 13, this volume). As the energy of water is
reduced in downstream channels, deposition occurs, as discussed below. The importance of
such fluxes for individual element cycles is calculated in Chapter 2 (this volume) and the
references contained therein. For more detailed geographic analysis, hydrological data are
available from world atlases, while world river water chemistry is reviewed by Livingstone
(1963). While in transit, waterborne substances may undergo degassing, biological uptake,
and sedimentation (Hynes, 1970) that will lead to temporary re-assortment in sediments or
permanent distillation to the atmosphere. These processes are particularly important at the
mouths of rivers, where deposition can be prominent and freshwater meets saltwater (Wiley,
1977; Wollast, Chapter 14, this volume).

Once in the ocean, water mixes according to fairly well-known and understood patterns
(Neumann, 1968; Stern, 1975; Fiadeiro, Chapter 17, this volume), and re-assortment of water
contents will take place according to the many biogeochemical processes of recirculation,
degassing, and differential sedimentation (Broecker,1974; Gibbs, 1977; Wollast, this
volume). Transport in the ocean is treated by Fiadeiro (Chapter 17, this volume).

D. Atmospheric Transport Mechanisms

The atmosphere is the principal medium for fast, long range transport. For carbon and
nitrogen, the atmosphere is the principal source for land ecosystems and many aquatic
systems. The atmosphere is also the principal source of sulphur for many unpolluted land
ecosystems (e.g. maritime zones). This importance has expanded over wider regions with the
increase in emissions of anthropogenic sulphur.

Saltation is only of local importance in moving soil materials. In contrast, suspension of


particles is extremely important as a long-distance mode of transport. Truly suspended
particles range from 10-3 µm to 1 µm in diameter. Below 0.1 µm, particles are subject to
deposition by Brownian movement and are little affected by gravitational settling. Between
0.1 and 1 µm, particles are little affected by either factor and have the longest survival time in
the atmosphere. Particles from 1 to 10 µm diameter have a significant sedimentation rate but
are maintained in the atmosphere by turbulence (Twomey, 1977). These large particles are
readily deposited to surfaces through inertial impaction.

Aerosol concentrations range from very low values in hundreds/cm3 in the cleanest air of the
troposphere to tens of thousands/cm 3 in urban areas. This is a large number, but it generally
represents a relatively small mass, comparable with the mass of trace gases. The total mass of
suspended particulate matter in the entire atmosphere is estimated by Twomey (1977) to be
about 1011 g. This may be low by about an order of magnitude; an estimate of about 1012 g
seems more consistent with emission rates and turn-over times. This can be compared with 2
x 1015 g N for N2O and 1.1 x 1012 g N for NH3 (Söderlund and Svensson, 1976).

Aerosols are derived from soil deflation, sea spray jets, volcanoes, suspension of particles on
canopies, and condensation from various gases. Table 5.2 gives an estimate of the magnitudes
of production by several natural and human sources. Clearly, C, N, and S are prominent
among the chemicals involved in aerosol formation. Interactions between compounds of these
elements are extremely important in the formation of aerosols (Taylor et al., Chapter 4, this
volume). Phosphorus is of trivial importance from the viewpoint of total mass of aerosols, but
atmospheric transport of phosphate aerosol is a major part of the short term phosphorus cycle
(Richey, Chapter 2, this volume) and may, in the long run, be a critical source for
oligotrophic land and aquatic ecosystems.

Table 5.2 Estimates of particles smaller than 20 µm radius emitted into or formed in the
atmosphere (106 metric tons/yr) (from SMIC, 1971). Reproduced by permission of MIT Press

Source Size

Natural
Soil and rock debris* 100500
Forest fires and slash-bruning debris* 3150
Sea salt (300)
Volcanic debris 25150
Particles formed from gaseous emissions:
Sulphate from H2S 130200
Ammonium salts from NH3 80270
Nitrate from NOx 60430
Hydrocarbons from plant exudations 75200
Subtotal 7732200

Man-made
Particles (direct emissions) 1090
Particles formed from gaseous emissions:
Sulphate from SO2 130200
Nitrate from NOx 3035
Hydrocarbons 1590
Subtotal 185415
Total 9582615

*Includes unknown amounts of indirect man-made contributions.

The mixing and subsequent movement of gases in the atmosphere is the most important
means of transport for carbon, nitrogen, and sulphur in the atmosphere. Gaseous contents of
the atmosphere are given in Svensson and Söderlund (1976) and Freyer (1979), and sources
and lifetimes in the atmosphere are given by Crutzen (Chapter 3, this volume). The chemical
interactions between these gases are described by Crutzen (Chapter 3, this volume), and the
relationships between the gases and particulates is exceedingly intimate (Taylor et al.,
Chapter 4 this volume). In terms of the transport aspect of biogeochemical cycles, the most
important chemical interactions between carbon, nitrogen, and sulphur occur in the biological
aspects of mobilization and deposition and in atmospheric chemistry occurring while in
transit.

The geography of atmospheric transport is a function of emission source areas, lifetimes of


chemical species and physical forms in the atmosphere, movement of air masses, and
distribution of factors contributing to deposition. Further discussion of atmospheric transport
will follow later in this paper.

E. Human Transport

The spatial scale and variety of material transfer affected by human activities is immense. An
analysis of such transport would require a major review in economic geography. Man
undoubtedly accomplishes more transfer by releasing waste materials into the atmosphere or
into rivers and the sea than he does by actually moving material itself. Such activity is
covered indirectly as a result of human-caused mobilization (Table 5.1).

5.2.3 Deposition Processes

A. Deposition from Mass-wasting Transport

Deposition of material moved downslope through the force of gravity simply occurs when the
transported mass reaches a stable position, generally at the base of the slope, although
momentum can carry material some distance up opposing slopes or across flat ground. This is
almost entirely a physical process, involving no interactions between critical elements and is
entirely of local significance.

B. Deposition by Glacial Transport

The manner of deposition by glaciers is quite variable in every respect. Usually deposition is
most prominent around glacial margins, where glacial advance is balanced by ablation, or by
iceberg calving into lakes or the sea. Deposition can also cover broad areas where down-
wasting creates vast till plains or out-wash carries debris over extensive drainage basins.
Persistent deposits are, of course, mainly terrestrial and have very long term effects on the
nature of ecosystems developing on such deposits (Jenny, 1980). Material delivered to rivers
or directly to the sea will undergo the same sorting and sedimentation as other fluvial solids.

C. Deposition from Fluvial Transport

Eroded material from upland positions of watersheds is progressively deposited downhill as


gradient changes occur (Branson et al., 1981). For example, some eroded soil material from
hill slopes may be deposited on lower-slope positions or in low-order drainage basins, and
some may be carried further downstream, depending on the energy of the flowing water. As a
consequence of this progressive deposition, the total amount of material eroded from the
upland positions of watersheds during storms or snow melting is greater than the amount
measured as entrained substances in water flowing out of the watershed.

The ecological impacts of the deposition of sediment on toe slopes and low-order drainages
are increased soil depth, increased soil water-holding capacity caused by each, greater
proportion of fine-textured soil material and organic matter, and, consequently, often
increased storage of C, N, P, and S and in cycling rates of N, P, and S. The impact of
deposition on high-order drainages is often the formation of fertile floodplains (i.e., many of
the important agricultural centres of the world, especially in the semi-arid and arid regions).
However, deposition of sediment, especially that resulting from erosion caused by man's
activities, can have serious economic and environmental impact: (1) detrimental deposition
on land and crops; (2) alteration of water quality of fresh-water, estuarian, and marine
communities (see Ritchie, Chapter 13, this volume); and (3) aggradation of river channels,
increasing flood hazards and reducing reservoir storage capacity.

As noted earlier, some material in water is transferred to the atmosphere through gas escape
or aerosol suspension (see Liss, Chapter 15, this volume and Duce, Chapter 16, this volume).
The reverse is true, as well. Nitrogen gas and carbon dioxide for example, do dissolve from
the atmosphere into water and become biologically fixed. Deposition processes in this
category include those for which dissolved or suspended substances are temporarily, at least,
removed from the moving water column and fixed in place. The first two processes include
biological fixation by sessile organisms on benthic or reef substrates. Floating organisms can
also fix nitrogen and carbon; take up nitrate, phosphate, and sulphate; and, by their death or
excretion of faecal pellets, lead to sedimentation of these elements in a position below this
zone of synthesis (Broecker, 1974). Besides organic matter, calcium carbonate and many
other minerals can be sedimented in this manner, as well (Lowenstam, 1981). Much of the
organic matter and calcium carbonate thus sedimented is redissolved and recirculated in deep
ocean currents (Jørgensen, Chapter 19 this volume). However, some of it will ultimately be
deposited for the long term and either up-thrust as new sedimentary rock on land or, less
likely, lost to metamorphosis in crustal subduction (Kempe, 1979).

D. Deposition from Atmospheric Transport

The mode and rate of deposition of substances from the atmosphere depends on the form of
the substance, meteorological conditions, and the nature of the surface. The various modes of
deposition have been described in many places; one of the most succinct has been the
description by Fowler (1980), which will be followed here.
It is simplest to first consider wet deposition. Rain and snow are responsible for
approximately half of the deposition of nitrogen and sulphur in well-watered industrial
countries (Galloway and Whelpdale, 1980; Grennfelt et al., 1980). The rate of deposition on a
particular locale depends on the chemistry of the air in which droplets are formed and on the
annual rainfall over the locale. Fowler (1980) calculated that 6070% of the sulphur and
nitrogen in wet deposition comes from the cloud condensation nucleus pathway, and about
20% comes from the solution and oxidation of gaseous species in droplets that will become
raindrops. This oxidation may not go to completion (Dana, 1980). Again, heterogeneous
reactions between chemical species of carbon, nitrogen, and sulphur and with other elements
are critical for controlling rates at which gas in the atmosphere can be swept out by wet
deposition. While aerosols impacted on falling drops comprise only a small part of the
chemical load of raindrops, this is a very important mechanism for removing small aerosols
from the atmosphere. Butcher and Charlson (1972) reiterate Junge's (1963) approximation
that 8090% of particulate removal occurs by incorporation of aerosols into hydrometeors and
delivery by wet deposition.

Strictly speaking, all other forms of deposition are classified as dry deposition, although one
formsedimentationis generally measured together with wet deposition in `bulk precipitation'
collectors. The other anomaly is that cloud droplet impaction is classified as dry deposition
even though it is quite a `wet' process. Sedimentation is the deposition of larger-sized
particles (> 10 µm diameter). Soil dust, volcanic ash, and locally transported organic
fragments probably make up most of this category, which is usually a minor source, except
for downwind of deflation areas. Kilmer (1979) gives an example of dust deposition from a
single storm that originated in the TexasOklahoma Panhandle in 1937. Dust was carried over
800 km, leaving deposits of 10 g/m2 in Ames, Iowa; 5 g/m2 in Marquette, Michigan; and 4
g/m2 in New Hampshire.

I have divided gaseous absorption into active and passive processes. This seems appropriate
in that CO2 and N2 fixationtwo of the major driving forces in the biosphereare rate limited by
distinct physiological processes and not usually by physical ones. Wherever physiological
processes come into consideration, the basic requirement for a proper balance of limiting
elements becomes a germane chemical interaction for C, N, P, and S in the manner described
by Redfield (1958).

Passive gaseous absorption is not entirely passive, either. Deposition rates will ultimately
depend on partial pressure differentials in the absorbing material. All of the substances listed
in this category in Table 5.1 must undergo a chemical reaction to reduce this back pressure.
Some reactions are at least partly biologicale.g., microbial decomposition of methane and
stomatal absorption of SO2. Assuming these sinks are functioning, then gas absorption is
largely limited by physical processes.

Figure 5.4 The dry deposition process (from Fowler, 1980). Reproduced by permission of
Norwegain SNSF Project

Figure 5.4, from Fowler (1980), describes the dry deposition process as a series of
mechanisms for penetrating two boundary layers and ultimate sorption on the surface. These
mechanisms are somewhat different for gases than for particles, but the principles of
measurement are the same for each, so inertial impaction and molecular diffusion processes
can be treated simultaneously: Inertial impaction is the major depositional mechanisms for
particles > l µm and < 10 µm in diameter; molecular diffusion is the depositional mechanism
for particles < 1 µm in diameter. A convenient means of estimating flux from the atmosphere
to a surface is to convert diffusion rates across boundary layers to resistance as an analog to
the analysis of electrical circuits. In Table 5.3, resistances to gaseous deposition are shown
for a variety of circumstances for vegetated surfaces The resistances representing the
turbulent boundary layer and laminar boundary layer are pooled and termed atmospheric
resistance. Surface sorption characteristics are described as canopy resistance in this table.
Atmospheric resistance is inversely related to canopy roughness, which is crudely
proportional to height. It is also inversely related to wind speed. Hence, other factors being
equal, dry deposition will be least in still air over a smooth surface. It will be greatest in
windy conditions over forest vegetation (Figure 5.5). Looking back at Table 5.3, we see that
for nitrogen and sulphur gases, resistances are least for wet surfaces, low for open stomata,
higher for closed stomata, and highest for senescent surface tissues. Perception of these
factors in terms of deposition rates is made easier by the deposition velocity column in Table
5.3. Deposition velocity can be viewed as a coefficient for capture. Deposition is the product
of deposition velocity times exposure (concentration).

Taken together, estimates in Table 5.3 enable us to assess the kinds of conditions and
surfaces that would lead to more or less dry deposition for a given exposure. Wet climates
will enhance not only wet deposition but also generally dry deposition, as well. Vegetated
surfaces are more efficient collectors than non-vegetated surfaces. This suggests not only
ways of rating ecosystem types for deposition but also ways of rating seasonal effects.
Cropland ought to be superior to urban areas, forests ought to be superior to croplands and
grasslands, and tall, moist forests ought to be the best sinks for dry deposition. Except under
special meteorological conditions, dry deposition of gases, and especially particles, is
generally higher on land surfaces than on lake or ocean surfaces (Sheih et al., 1979). The
potential of this method for estimating sink characteristics on a geographic basis will be
discussed further in a later section.

5.3 INTER-BIOME TRANSPORT VIA THE ATMOSPHERE

The foregoing discussion included processes involved in transport at all scales: at the local
ecosystem level, along soil catenas, within watersheds, over landscapes, in regional land
systems, and on a global scale (Woodmansee and Adamsen, 1981). One of the conclusions of
this survey is that mobilization is critical for transport at any scale. The same processes that
mobilize for small-scale (intra-ecosystem) transport are necessary for long range transport.
On the other hand, high rates of mobilization within an ecosystem are necessary but not
sufficient for long range transport at the regional to global scale. There may be cases in which
the system having the highest mobilization rate may also have the highest deposition rate. For
example, the generation rate of suspended aerosols may be highest in tropical forests, but
deposition velocities for these aerosols may be highest there also. Because mobilization is
high in an ecosystem system, it does not mean that deposition, as well, may not be high there.

Table 5.3 Rates of dry deposition of SO2, NO2, and HNO3 on vegetation (from Fowler,
1980). Reproduced by permission of Norwegian SNSF Project

Atmospheric Canopy resistance ‡ Deposition velocity*


resistance*† rc(s m-1) Vg (mm s-1)
r(a + b)(s m-
1
)

Vegetation SO2, NO2, Surface


height (m) HNO3 condition SO2 HNO3 NO2 SO2 HNO3 NO2

0.1 90 Stomata open 100 100 200 5 5 3


Stomata
300 200300 400600 3 4 2
closed
Surface
500 300500 500+ 2 2 1
senescent
Wet 0 0 ? 11 11 ?

1.0 35 Stomata open 100 100 200 8 8 4


Stomata
300 200300 400600 3 3 2
closed
Surface
500 300500 500+ 2 2 1
senescent
Wet 0 0 ? 28 28 ?

10.0 5 Stomata open 100§ 100 200 10 10 5


Stomata 400-
400500 300500 2 3 2
closed 600
Wet 0 0 ? 200|| 200 ?

*r(a + b) and Vg reference height 1 m above surface at which a wind speed of 2.5 m s-1 is
assumed.
†Small differences in the component rb due to different molecular diffusion coefficients for
SO2, NO2, HNO3 are ignored.
‡rc values for 32SO2 assumed similar or slightly smaller for HNO3 values for NO2.
§Assuming Rc SO2:rc H2O is the same as ratio of molecular diffusion coefficients D SO2:D
H2O (2) and taking a value for rc with stomata open for sitka spruce (Picea stichensis).
||Though very large, this value would rarely be found unless the water on the trees has not
reached equilibrium with ambient SO2 concentration; in practice, this value would only be
applicable for a very short time (minutes) unless there is a mechanism for removing S(IV)
from solution and neutralizing the H+ generated by the solution and oxidation of SO2.

Figure 5.5 Variation of aerodynamic resistance with wind speed and vegetation height (from
Fowler, 1980). Reproduced by permission of Norwegian SNSF Project

A second conclusion is an obvious one: that the most rapid, long distance transfers are via the
very fluid atmosphere. Thus, mobilization processes that lead to production of gases or
aerosols are most likely to lead to rapid, interregional transport.
The third general conclusion is that, whereas the greatest net transport of elements away from
terrestrial systems may be via fluvial agencies, the return of material to the land from aquatic
systems, particularly the sea, can be affected in the short term (<1 year) only through the
intermediary of atmospheric transport. Thus, while rock weathering is a major, long term
input to terrestrial systems, the atmosphere is the principal source of short term inputs for
land as a whole. On a smaller spatial scale, fluvial transport from uplands to lowlands is very
important, but on the larger spatial scale it is not important for transport to terrestrial systems.

The fourth general conclusion is that the influences of one terrestrial system on another as
affected by material transfer must, in the short term, be mainly through atmospheric transport.

Given these generalizations, if we take a primarily terrestrial and short term focus on the role
of transport in biogeochemical cycles, then it is logical to concentrate primarily on
atmospheric transport processes. For the remainder of this section, I will review the general
mobility of the major airborne compounds of C, N, P, and S. This review will set the scale at
which geographic modelling of material transfers via the atmosphere is best attempted. For
compounds that enter large atmospheric reservoirs and are long-lived, global mixing dictates
global-scale modelling; for other compounds, regional modelling is more sensible (e.g.,
Rodhe, 1978). Direction and speed of atmospheric transport by global circulation patterns is a
problem of climatology over long distances and time intervals and of meteorology over
shorter space and time intervals. Deposition rates over these transport tracks depend on
chemical transformation rates, precipitation rates, and the nature of the ground surface for dry
deposition processes. Stochastic variation in all of these factors can, of course, affect
deposition at any single place and time.

5.3.1. Residence Times and Travel Distances

A. Carbon Compounds

Although carbon dioxide emissions vary in time and space, the carbon dioxide pool is so vast
and well mixed that circulation is essentially global. The same is true for methane and
unreactive, volatile hydrocarbons and essentially true for carbon monoxide (Table 5.4). Only
C-containing particulates are deposited in reasonably short distances, so that the vectors of
travel for these might be interesting.

The residence time of carbon dioxide was calculated from the reservoir size and the sum of
inputs in the model developed by Bolin et al. (1979a). Residence time for dust came from
Kempe (1979); all other values came from Freyer (1979).

B. Nitrogen Compounds

Nitrogen gas and nitrous oxide are essentially globally distributed due to the very large
atmospheric turn-over times (i.e. reservoir content compared to inputs and output fluxes).
NH3, NOx and HNO3 all have residence times in the lower atmosphere of the order of only
one day due to efficient transformation and removal processes. Nitrogen compounds in
secondary aerosols (mainly NH4+ and NO3-) have a residence time determined by that of the
aerosol particles, which may vary from a few days to a few weeks depending on climatic
conditions.
Table 5.4 Residence times and relative transport distances for major chemical species of
carbon, nitrogen, phosphorus, and sulphur. Sources for residence times are described in the
text. Relative transport distances are the same as those defined in Table 1

Transport distance
Compound Residence time
(km)

Carbon Compounds
CO2 Years Global
CO Weeks Continental
CH4 Years Global
Volatile Hydrocarbons
Vapour phase Variable Global
Particulate phase Daysweeks Regionalcontinental
Dust and spray particulates
d > 1 µm Days Regional
d < 1 µm Weeks Continental

Nitrogen Compounds
N2 2 x 107 years Global
NH3 Days Regional
N2O < 2 x 102 years Global
NOx (gas) Days Regional
NOx (solid) Days Regional

Organic particulate Daysweeks Regionalcontinental

Phosphorus Compounds
Dust and spray PO43- particulates
d > 1 µm Days Regional
d < 1 µm Daysweeks Regionalcontinental

Sulphur Compounds

SO2 (combined dry and wet) Ca. 1 day Mesoscaleregional


SO42- (combined dry and wet) A few days Regional
H2S (as H2S and converted to SO2,
Days Regional
SO42-
(CH3)2S (converted to SO2, SO42-) Days Regional
COS > 1 year Global
Dust and spray particulates
d > 1 µm Days Regional
d < 1 µm Daysweeks Regionalcontinental

Residence times for all N compounds were taken from Svensson and Söderlund (1976) or
calculated from their data.

C. Phosphorus Compounds

Atmospheric transfer of P is an important part of the global phosphorus budget, although


small quantities of matter are involved. Atmospheric deposition may be a critical input to
very oligotrophic terrestrial and fresh-water ecosystems, where mineral weathering provides
little P. No stable or abundant gases of P are known; Particulate P is the only important form.
A global budget for atmospheric P, prepared by Graham and Duce (1979) and reviewed in
Duce, Chapter 16, this volume, gives valuable perspectives on atmospheric transport of P.
About 4.6 T g/yr are deposited from the atmosphere to earth, 70% falling on land. About 83%
of this deposition is estimated to be of crustal origin (mostly dust from arid regions) and 7%
from sea-salt particles. The remainder is anthropogenic, with the largest single fraction of that
being derived from soil dust raised by man's activities (2850%). Travel distances calculated
in Table 5.4 are divided in terms of particle size following the approach of Kempe (1979).
Travel time and distances for particulates are very sensitive to size. If combustion-derived
phosphate leads to small particles sizes, then such particles are carried long distances and
probably highly diffused over vast areas. More interesting, from a modelling point of view,
are the trajectories of coarser particles derived, mostly, from soil dust.

D. Sulphur Compounds

The behaviour and deposition of S compounds have been well studied due to the large
volume of anthropogenic S emissions and their ecological importance. The regional transport
of S has been estimated with considerable success by the OECD program LRTAP (Doveland
and Semb, 1980) and others in Europe. Other programs for mapping atmospheric transport in
North America are described by Pack et al. (1978), Perhac (1978), Wilson (1978), and
Whelpdale (1978). Sulphur has been, perhaps, an ideal case for mapping trajectories of
emissions due to the relatively short transit distances of some of the compounds but mostly
because of the localized sources. `Hot spots' of intense anthropogenic emissions generate
60% of global S from 1% of the earth's surface (Husar and Husar, 1978). The approaches of
these programs in modelling regional transfers (Figure 5.6) and global transfers (Figure 5.7)
serve as examples for the modelling of other substances.

Except for the case of carbonyl sulphide (COS), residence times, and thus transit times, of S
compounds are relatively short, so that intraregional transport would probably be the most
reasonable scale for modelling. Residence times for sulphur dioxide and sulphate came from
Rodhe (1978) and thus represent conditions in Northern Europe. Residence time for hydrogen
sulphide was approximated from a variety of sources (Granat et al., 1976; Rodhe, 1978;
Georgii, 1978). Residence time for dimethyl sulphide was extrapolated from Granat et al.
(1976) for sulphur dioxide and sulphate together, with the assumptions of a slight dry
deposition of dimethyl sulphide itself and a relatively slow conversion rate to sulphur
dioxide. Data on carbonyl sulphide came from Graedel et al. (1981). Estimates for dust and
spray particulates followed the approach of Kempe (1979).
Figure 5.6 Emissions puff advection and diffusion scheme used in EURMAP (from Johnson
et al., 1978) Reproduced by permission of Pergamon Press Ltd.

5.4 SUMMARIZING CONCLUSIONS

1. Transport, as a part of biogeochemical cycles, can be effectively analysed in terms of


three phases: mobilization, transport, and deposition. Each phase involves separate
processes.
2. Collectively, all the processes involved in mobilization, transport, and deposition
embrace a very broad spectrum of the natural sciences, necessitating a
multidisciplinary approach to analysis. These processes can be organized into classes
more amenable to generalization; but transport, in the broad sense, is not easily
reduced to a few generalizations or principles.
3. Transport occurs on a continuum of scale in terms of space and time.
4. The same mobilization processes may lead to either short or long range transport, but
processes leading to formation of gases (e.g. methanogenesis) are more likely to
ensure long range transport.
5. Chemical interactions in the transport of C, N, P, and S are most pronounced in four
areas:

a. Biological mechanisms of mobilization.


b. Mobilization processes engendered by human activity, especially those
powered by fossil hydrocarbons.
c. Biological mechanisms of deposition.
d. Influences of anthropogenically derived acids.

6. The behaviour of P is remarkably conservative in all aspects of transport compared


wtih C, N, and S. This can give P a special regulatory role in biological processes.
7. Mobilization and deposition rates may both be high for the same substance in the
same system. High rates of mobilization are necessary but not sufficient for net flux
from one system to another.
8. Atmospheric transport is of special importance for

a. Rapid, long distance transport.


b. Deposition to terrestrial systems.
c. Land-to-land transfers and effects.

9. Fluvial transport is of special importance for

a. Massive land-to-sea transport. (Sea-to-land transport occurs only through the


intermediary atmosphere.)
b. Involvement of huge pools.
c. Relatively slow transport rates.
d. Slow time scale and long term regulator role.

10. Biomes or other biologically defined units are not necessarily the best units for
defining source and sink regions for material transfer. Furthermore, air mass
trajectories, drainage systems, and ocean currents do not usually conform to
ecological patterns of gradients or boundaries. Different geographic criteria are
necessary for different substances and modes of transport for resolution below the
land, sea, atmosphere scale of definition.

Figure 5.7 Global-scale mapping of SO42-, the long-lived product of H2S emissions (upper),
and combined H2S and SO2 emissions (lower). The upper panel is a calculated distribution of
SO42- for July resulting from a release of 40 Tg H2S S/year (pptv), more or less representing
background emissions to the atmosphere. In the lower panel 80 Tg anthropogenic SO2 S/year
(pptv) for July has been added to the background H2S. Numbers in the circles give percentage
contributions of man-made sources to local concentrations of SO42- at different heights and
latitudes (from Rodhe and Isaksen, 1980) Reproduced by permission of American
Geophysical Union.

ACKNOWLEDGEMENTS

The author acknowledges with sincere thanks the constructive suggestions and information
provided by Professor Henning Rodhe, Department of Meteorology, University of
Stockholm, especially on aspects of atmospheric transport. Likewise, Professor Robert
Woodmansee, Department of Range Science, Colorado State University, made very
significant improvements on the treatment of entrainment and fluvial processes.

5.5 REFERENCES

Beauford, W., Barber, J., and Barringer, A.R. (1977) Release of particles containing metals
from vegetation into the atmosphere, Science, 195, 571-573.

Birkeland, P. W. (1974) Pedology, Weathering, and Geomorphological Research, New York,


Oxford University Press.

Bolin, B., Degens, E. T., Duvigneaud, P., and Kemp, S. (1979a) The global biogeochemical
carbon cycle, in Bolin, B., Degens, E. T., Kempe, S., and Ketner, P., (eds) The Global
Carbon Cycle, SCOPE Report No. 13, Chichester, Wiley, 1-53.

Bolin, B., Degens, E. T., Kempe, S., and Ketner, P. (eds) (1979b) The Global Carbon Cycle,
SCOPE Report No 13, Chichester, Wiley.

Branson, F. A., Gifford, G. E., Renard, K. G., and Hadley, R. F. (1981) Rangeland
Hydrology, 2nd Edn, Range Science Series No. 1, Dubuque, Iowa, Kendall/ Hunt Publishing
Co.

Broecker, W. S. (1974) Chemical Oceanography. New York, Harcourt Brace Jovanovich,


Inc. 214 pages.

Butcher, S. S., and Charlson, R. J. (1972) Introduction to Air Chemistry, New York,
Academic Press.

Cook, R. B. The impact of acid deposition on the cycles of C, N, and S, Chapter 12, this
volume.
Cronan, C. S., and Schofield, C. L. (1979) Aluminum leaching response to acid precipitation:
Effects on high-elevation watersheds in the northeast, Science, 204, 304-306.

Crozat, G. (1978) On the emission of potassium-rich aerosols from a tropical forest, Tellus,
31, 52-57 (in French).

Crutzen, P. J., Heidt, L. E., Krasnec, J. P., Pollack, W. H., and Seiler, W. (1979) Biomass
burning as a source of atmospheric gases CO, H2, N2O, NO, CH3Cl, and COS, Nature, 282,
253-256.

Crutzen, P. J. Atmospheric interactionshomogeneous gas reactions of C, N, and S containing


compounds, Chapter 3, this volume.

Dana, M. T. (1980) SO2 versus sulfate wet deposition in the eastern United States, J.
Geophys. Res., 5, 4475-4480.

Doveland, H., and Semb, A. (1980) Atmospheric transport of pollutants, in Drabløs D., and
Tollan, A. (eds) Ecological Impact of Acid Precipitation, Proceedings of an International
Conference, Sandefjord, Norway, March 11-14, 1980, Oslo, SNSF, 14-21.

Duce, R. A. Biogeochemical cycles and air/sea exchange of aerosols, Chapter 16, this
volume.

Fiadiero, M. Physical-chemical processes in the open ocean, Chapter 17, this volume.

Fish, B. R. (1972) Electrical generation of natural aerosols from vegetation, Science, 175,
1239-1240.,

Fletcher, J. E., Sorensen, D. L., and Porcella, D. F. (1978) Erosional transfer of nitrogen in
desert ecosystems, in West, N.E., and Skujins, J. (eds) Nitrogen in Desert Ecosystems,
Stroudsburg, Pennsylvania, Dowden, Hutchinson & Ross, Inc., 171-181.

Foster, G. R. (ed.) (1977) Soil Erosion: Prediction and Control, Soil Conservation Society of
America, Spec. Publ. No. 2. Ankeny, Iowa, Soil Conservation Society of America.

Fowler, D. (1980) Removal of sulphur and nitrogen compounds from the atmosphere in rain
and by dry deposition, in Drabløs, D., and Tollan, A. (eds) Ecological Impact of Acid
Precipitation, Proceedings of an International Conference, Sandefjord, Norway, March 11-
14, 1980, Oslo, SNSF, 22-32.

Freney, J., and Rodhe, H. The sulphur cycle, Chapter 2, this volume.

Freyer, H. D. (1979) Atmospheric cycles of trace gases containing carbon, in Bolin, B.,
Degens, E. T., Kempe, S., and Ketner, P. (eds) The Global Carbon Cycle, SCOPE Report No.
13, Chichester, Wiley,101-128.

Galloway, J. N., and Whelpdale, D. M. (1980) An atmospheric sulfur budget for eastern
North America, Atmos. Environ., 14, 409-417.
Garrels, R. M., Mackenzie, F. T., and Hunt, C. (1975) Chemical Cycles and the Global
Environment, Los Altos, California, William Kaufmann, Inc.

Georgii, H.-W. (1978) Large scale spatial and temporal distribution of sulfur compounds,
Atmos. Environ., 12, 681-690.

Gibbs, R. J. (1967) The geochemistry of the Amazon River system. Part I. The factors that
control salinity and the composition and concentration of the suspended solids, Geol. Soc.
Am., 78, 1203-1232.

Gibbs, R. J. (ed.) (1977) Transport Processes in Lakes and Oceans, New York, Plenum
Press.

Gorham, E., and McFee, W. W. (1980) Effects of acid deposition upon outputs from
terrestrial to aquatic ecosystems, in Hutchinson, T. C., and Havas, M. (eds) Effects of Acid
Precipitation on Terrestrial Ecosystems, New York, Plenum Press, 465-480.

Gorham, E., Vitousek, P. M., and Reiners, W. A. (1979) The regulation of chemical budgets
over the course of terrestrial ecosystem succession. Ann. Rev. Ecol. Syst., 10, 53-88

Graedel, T. E., Kammlott, G. W., and Franey, J. P. (1981) Carbonyl sulfide: Potential agent
of atmospheric sulfur corrosion, Science, 212, 663-665.

Graham, W. F., and Duce, R. A. (1979) Atmospheric pathways of the phosphorus cycle,
Geochim. Cosmochim. Acta, 43,1195-1208.

Graham, W. F., Piotrowicz, S. R., and Duce, R. A. (1979) The sea as a source of atmospheric
phosphorus, Mar. Chem., 7, 325-342.

Granat, L., Hallberg, R. O., and Rodhe, H. (1976) The global sulphur cycle, in Svensson, B.
H., and Söderlund, R. (eds) Nitrogen, Phosphorus and SulphurGlobal Cycles, SCOPE Report
No. 7, Ecol. Bull. (Stockholm), 22, 89-134.

Granhall, U. (1981) Biological nitrogen fixation in relation to environmental factors and


functioning of natural ecosystems, in Clark, F. E., and Rosswall, T. (eds) Terrestrial Nitrogen
Cycles, Ecol. Bull. (Stockholm), 33, 131-144.

Grennfelt, P., Bengston, C., and Skarby, L. (1980) Estimation of the atmospheric input of
acidifying substances to a forest ecosystem, in Hutchinson, T. C., and Havas, M. (eds) Effects
of Acid Precipitation on Terrestrial Ecosystems. NATO Conference Series I, Ecology, New
York, Plenum Press, 29-40.

Hogan, A. W., and Mohnen, V. A. (1979) On the global distribution of aerosols, Science,
205, 1373-1375.

Husar, R., and Husar, J. (1978) Forward, Atmos. Environ., 12, 3-5.

Hynes, H. B. N. (1970) The Ecology of Running Waters, Toronto, University of Toronto


Press.
Idso, S. B. (1976) Dust storms, Sci. Am., 235(4),108-114.

Ivanov, M. V. (1981) The global biogeochemical sulphur cycle, in Likens, G. E. (ed.) Some
Perspectives of the Major Biogeochemical Cycles, SCOPE Report No. 17, Chichester, Wiley,
61-78.

Jenny, H. (1980) The Soil Resource. Origin and Behavior. New York, Springer-Verlag.

Johnson, N. M., Likens, G. E., Bormann, F. H., and Pierce, R. S. (1968) Rate of chemical
weathering of silicate minerals in New Hampshire, Geochim. Cosmochim. Acta, 32, 531-545.

Johnson, N. M., Reynolds, R. C., and Likens, G. E. (1972) Atmospheric sulfur: Its effect on
the chemical weathering of New England, Science, 177, 514-516.

Johnson, N. M., Driscoll, C. T., Eaton, J. S., Likens, G. E., and McDowell, W. H. (1981)
`Acid rain', dissolved aluminum and chemical weathering at the Hubbard Brook
Experimental Forest, New Hampshire, Geochim. Cosmochim. Acta, 45,1421-1437.

Johnson, W. B., Wolf, D. E., and Mancuso, R. L. (1978) Long term regional patterns and
transfrontier exchanges of airborne sulfur pollution in Europe, Atmos. Environ., 12, 511-527.

Jørgensen, B. Processes at the sediment-water interface, Chapter 18, this volume.

Junge, C. E. (1958) The distribution of ammonia and nitrate in rain water over the United
States, Trans. Am. Geophys. Union, 39, 241-248.

Junge, C. E. (1963) Air Chemistry and Radioactivity, New York, Academic Press.

Kempe, S. (1979) Carbon in the rock cycle, in Bolin, B., Degens, E. T., Kempe, S., and
Ketner, P. (eds) The Global Carbon Cycle, SCOPE Report No. 13, Chichester, Wiley, 343-
378.

Kilmer, V. J. (1979) Minerals and agriculture, in Trudinger, P. A., and Swaine, D. J. (eds)
Biogeochemical Cycling of Mineral forming Elements, Amsterdam, Elsevier Scientific
Publishing Co., 515-558.

Lawson, D. R., and Winchester, J. W. (1979) Sulfur, potassium, and phosphorus associations
in aerosols from South American tropical rain forest, J. Geophys. Res., 84, 3723-3727.

Liss, P. S. The exchange of biogeochemically-important gases across the air-sea interface.


Chapter 15, this volume.

Livingstone, D. A. (1963) Chemical composition of rivers and lakes, in Data of


Geochemistry, 6th Edn, U. S. Geological Survey Professional Paper 440G, Washington, D.C.,
U.S. Government Printing Office.

Loughnan, F. C. (1969) Chemical Weathering of the Silicate Minerals, New York, American
Elsevier.

Lowenstam, H. A. (1981) Minerals formed by organisms, Science, 211, 1126-1131.


Marshall, J. K. (1973) Drought, land use and soil erosion, in Lovett, J. V. (ed.) The
Environmental, Economic, and Social Significance of Drought, Sydney, Angus and
Robertson, Publishers, 55-77.

Mooney, H., Bonnicksen, J. M., Christensen, N. L., Lotan, J. E., and Reiners, W. A. (eds)
(1981) Fire Regimes and Ecosystem Properties, USDA Forest Service, General Technical
Report, Washington, D. C., U.S. Government Printing Office (in press).

Morales, C. (ed.) (1979) Saharan Dust. Mobilization, Transport, Deposition, SCOPE Report
No. 14 Chichester, Wiley.

Neumann, G. (1968) Ocean Currents, Amsterdam, Elsevier Publishing Co.

Nye, P. H., and Tinker, P. B. (1977) Solute Movement in the Soilroot System, Berkeley and
Los Angeles, California, University of California Press.

Pack, D. H., Ferber, G. J., Heffter, J. L., Telogadas, K., Angell, J. K., Hoecker, W. H., and
Machta, L. (1978) Meteorology of long-range transport, Atmos. Environ., 12, 425-444.

Perhac, R. M. (1978) Sulfate regional experiment in northeastern United States: The SURE
program, Atmos. Environ., 12, 641-647.

Pierrou, U. (1976) The global phosphorus cycle, in Svensson, B. H., and Söderlund, R. (eds)
Nitrogen, Phosphorus and SulphurGlobal Cycles, SCOPE Report No. 7, Ecol. Bull
(Stockholm), 22, 75-88.

Rahn, K. A., Borys, R. D., Shaw, G. E., Schutz, L., and Jaenicke, R. (1979) Long-range
impact of desert aerosol on atmospheric chemistry: Two examples, in Morales, C. (ed.)
Saharan Dust. Mobilization, Transport, Deposition, SCOPE Report No. 14, Chichester
Wiley, 243-266.

Raison, R. J. (1979) Modification by the soil environment by vegetation fires, with particular
reference to nitrogen transformations: A review, Plant Soil, 51, 72-108.

Redfield, A. C. (1958) The biological control of chemical factors in the environment, Am.
Sci., 46,205-221.

Reiners, W. A. (1981) Nitrogen cycling in relation to ecosystem succession, in Clark, F. E.,


and Rosswall, T. (eds) Terrestrial Nitrogen Cycles. Processes, Ecosystem Strategies and
Management Impacts, Ecol. Bull (Stockholm), 33, 507-528.

Reynolds, R. C., and Johnson, N. M. (1972) Chemical weathering in the temperate glacial
environment of the northern Cascade Mountains, Geochim. Cosmochim. Acta, 36, 537-554.

Richey, J. E. Interactions in freshwater ecosystems, Chapter 13, this volume.

Richey, J. E. The phosphorus cycle, Chapter 2, this volume.

Rodhe, H. (1978) Budgets and turnover times of atmospheric sulfur compounds, Atmos
Environ., 12, 671-680.
Rodhe, H., and Isaksen, I. (1980) Global distribution of sulfur compounds in the troposphere
estimated in a height/latitude transport model, J. Geophys. Res., 85, 7401-7409.

Rosswall, T. (1976) The internal nitrogen cycle between micro-organisms, vegetation and
soil, in Svensson, B. H., and Söderlund, R. (eds) Nitrogen, Phosphorus and Sulphur-Global
Cycles, SCOPE Report No. 7, Ecol. Bull. (Stockholm), 22, 157-167.

Sheih, C. M., Wesely, M. L., and Hicks, B. B. (1979) Estimated dry deposition velocities of
sulfur over the eastern United States and surrounding regions, Atmos. Environ., 13, 1361-
1368.

Slinn, W. G. N. (1976) Dry deposition and resuspension of aerosol particlesa new look at
some old problems, in Englemann, R. J., and Sehmel, G. A. (coordinators)
AtmosphericSurface Exchange of Particulate and Gaseous Pollutants (1974), CONF-740921,
Tech. Information Center, ERDA, Springfield, Virginia, 1-40.

Söderlund, R. (1977) NOx pollutants and ammonia emissionsa mass balance for the
atmosphere over northwestern Europe, Ambio, 6, 118-122.

Söderlund, R., and Svensson, B. H. (1976) The global nitrogen cycle, in Svensson, B. H., and
Söderlund, R. (eds) Nitrogen, Phosphorus and SulphurGlobal Cycles, SCOPE Report No. 7,
Ecol. Bull. (Stockholm), 22, 23-73.

Stern, M. E. (1975) Ocean Circulation Physics, New York, Academic Press.

Strakhov, N. M. (1967) Principles of Lithogenesis, Vol. 1, Edinburgh, Oliver and Boyd,


(English translation).

Statham, I. (1977) Earth Surface Sediment Transport, Oxford, Clarendon Press.

Study of Man's Impact on Climate (SMIC) (1971) Inadvertent Climate Modification,


Cambridge, Mass. M.I.T. Press.

Svensson, B. H., and Söderlund, R. (eds) (1976) Nitrogen, Phosphorus and SulphurGlobal
Cycles, SCOPE Report No. 7, Ecol. Bull. (Stockholm), 22.

Swanson, J. F., Fredricksen, R. L., and McCorison, F. M. (1981) Material transfer in a


western Oregon forested watershed, in Coniferous Forest Biome Synthesis Volume,
Stroudsburg, Pennsylvania, Dowden, Hutchinson and Ross (in press).

Taylor, G. S., Baker, M. B., and Charlson, R. J. Atmospheric interactions of the C, N, and S
cycles: the role of aerosols and clouds, Chapter 4, this volume.

Turk, J. T., and Peters, N. E. (1978) Acid-rain weathering of a metasedimentary rock basin,
Herkimer County, New York, in Izard, H. H., and Jacobson, J. S. (eds) Scientific Papers from
the Public Meeting on Acid Precipitation, May 4-5, 1978, Lake Placid, New York, Science
and Technology Staff, New York State Assembly, Albany, New York, 136-145.

Twomey, S. (1977) Atmospheric Aerosols. Amsterdam, Elsevier Scientific Publishing Co.


Vitousek, P. M., Gosz, J. R., Grier, C. C., Melillo, J. M., Reiners, W. A., and Todd, R. L.
(1979) Nitrate losses from disturbed ecosystems, Science, 204, 469-474.

Whelpdale, D. M. (1978) Large-scale atmospheric studies in Canada, Atmos. Environ., 12,


661-670.

White, E., Hem, J. D., and Waring, G. A. (1963) Chapter F. Chemical composition of
subsurface waters, in Data of Geochemistry, 6th Edn, Geological Survey Professional Paper
440-F, Washington, D.C., U.S. Government Printing Office.

Wiley, M. (1977) Estuarine Proceses, Vol. 2, Circulation, Sediments, and Transfer of


Material in the Estuary, New York, Academic Press.

Wilson, W. E. (1978) Sulfates in the atmosphere: A progress report on Project MISTT,


Atmos. Environ., 12, 537-547.

Wollast, R. Interactions in estuaries and coastal waters, Chapter 14, this volume.

Woodmansee, R. G., and Adamsen, F. J. (1981) Biogeochemical cycles and ecological


hierarchies, in Todd, R. L. (ed.) Nutrient Cycling in Agricultural Ecosystems, Ann Arbor,
Michigan, Ann Arbor Science Publishers, Inc. (in press).

Woodmansee, R. G., Vallis, I., and Mott, J. J. (1981) Grassland Nitrogen, in Clark, E. E., and
Rosswall, T. (eds) Terestrial Nitrogen Cycles, Ecol. Bull (Stockholm), 33, 443-462.

Woodmansee, R. G., and Wallach, L. S. (1981) Effects of fire regimes on biogeochemical


cycles, in Clark, F. E., and Rosswall, T. (eds) Terrestrial Nitrogen Cycles, Ecol. Bull.
(Stockholm), 33, 649-669.

Back to Table of Contents

The electronic version of this publication has been prepared at


the M S Swaminathan Research Foundation, Chennai, India

SCOPE 21 -The Major Biogeochemical Cycles and Their


Interactions
6 Interactions of Biogeochemical Cycles in Forest Ecosystems
J. M. MELILLO AND J. R. GOSZ

Abstract
6.1 Introduction
6.2 Element Ratios in Foliage and Litter
6.2.1 Element Ratios in Foliage
6.2.2 Nutrient Translocation within Trees and Nutrient Use Efficiency
6.2.3 Summary
6.3 Carbon Nutrient Interactions at the Ecosystem Level
6.3.1 Photosynthesis and Net Primary Production
6.3.2 Shoot:Root Ratio and the Response of Shoots and Roots to Fertilization
6.3.3 Carbon Allocation to Perennial Versus Deciduous Plant Parts
6.3.4 Litter Decomposition
6.3.5 Summary
6.4 Element Interactions and the Global Carbon Cycle
6.4.1 The Global Carbon Cycle: An Overview
6.4.2 Element Composition of Forest Ecosystems
6.4.3 Fossil Fuel Combustion and the Acceleration of the N, P, and S Cycles
6.4.4 CarbonNitrogen Interactions
6.4.5 Phosphorus and Sulphur Inputs and Carbon Storage
6.4.6 Summary
6.5 Phosphorus 'The Master Element'
6.5.1 Phosphorus Availability and Carbon, Nitrogen and Sulphur Accumulation in
Soils
6.5.2 Pedogenesis and The Global Carbon Cycle
6.5.3 Concluding Statement
References

ABSTRACT

This paper presents four topics in the study of interactions among carbon, nitrogen,
phosphorus, and sulphur in forest ecosystems. In section 6.2, we consider element ratios in
foliage and litter, and discuss what these ratios imply about the availability of nutrients to
higher plants and the efficiency of nutrient-use by plants. The carbon to nitrogen ratio in litter
varies inversely with the mass of nitrogen in the litter across a broad range of forest types. A
qualitatively similar pattern is described for phosphorus. From extant data we conclude that a
high carbon to nutrient ratio in litter is characteristic of an efficient use of the nutrient under
consideration, with translocation of nutrients out of senescing tissue contributing to this
efficiency. We also conclude that the efficiency of nutrient use is inversely related to the
availability or rate of circulation of that nutrient.

In Section 6.3, we review the ways in which plant nutrient status can influence the amount of
carbon flowing through an ecosystem, the pathways of flow, and the proximate fate of the
carbon. We develop the argument that net carbon storage in forest ecosystems is determined
by the balance among three distinct processes: (1) the net amount of carbon fixed by the
vegetation (i.e. carbon fixed in excess of plant respiratory demand); (2) the relative amounts
of this fixed carbon allocated to vegetation growth increment versus plant litter; and (3) the
decomposition dynamics of carbon compounds entering the soil system. This argument has
several implications for the study of the global carbon budget. First, to determine whether or
not elevated atmospheric levels of CO2 will enhance carbon storage rate in terrestrial
ecosystems, it is insufficient to consider only the relationship between CO2 concentration and
photosynthetic rate. The subsequent fate of the fixed carbon must be considered. Second, the
chronic low level additions of fertilizer to forests, such as is occuring through air pollution
and subsequent rain-out may be: (1) enhancing overall carbon storage in ecosystems; and (2)
shifting a larger fraction of the annual carbon storage to the vegetation component of the
system.

We explore how man's impact on the global cycles of nitrogen, phosphorus, and sulphur may
influence the ability of forest ecosystems to store carbon in Section 6.4. Man may be
inadervertently fertilizing the world's forests with 6 Tg N through the burning of fossil fuels.
We estimate that the maximum amount of additional carbon storage that could be promoted
by a nitrogen fertilization of this magnitude is 300 Tg C/yr.

In section 6.5, we evaluate the possibility that the terrestrial biosphere is on a carbon
accumulation trend that is dependent on the dynamics of phosphorus availability during soil
development and that is independent of the activities of man. We estimate that on an annual
basis only a small amount of carbon, in the range of 79 Tg C, could be stored by this
mechanism.

6.1 INTRODUCTION

The forests and woodlands of the world occupy between 50 x 1012 and 60 x 1012 m2 or
slightly more than one-third of the earth's land area (Whittaker and Likens, 1973), and they
contain major fractions of the terrestrial stocks of C, N, S, and P. During the past two
decades, studies of the fluxes of elements among the structural components of forests have
intensified. Today, carbon and nitrogen budgets are available for a variety of forest
ecosystems, although most are incomplete; sulphur and phosphorus budgets are much less
common and they are fragmentary. However, as the work proceeds, we are beginning to
recognize that an understanding of the mechanisms that control cycling rates of an element
requires consideration of element interactions.

Element interactions fall into two general categories; carbon-nutrient interactions, and
nutrient-nutrient interactions. An example of a carbon-nutrient interaction is the stimulation
of net primary productivity in a forest by added nitrogen (Miller and Miller, 1976). An
example of a nutrient-nutrient interaction is the stimulation of nitrogen fixation by the
addition of phosphorus (Griffith, 1978). Carbon-nutrient interactions are the central concern
of this paper.

This paper presents four topics in the study of interactions among carbon, nitrogen,
phosphorus, and sulphur in forest ecosystems. First, we consider element ratios in foliage and
litter, and discuss what these ratios imply about the availability of nutrients to higher plants
and the efficiency of nutrient use by the plants. Second, we review the ways in which plant
nutrient status can influence the amount of carbon flowing through an ecosystem, the
pathways of flow, and the proximate fate of the carbon. Third, we explore how man's impact
on the global cycles of nitrogen, phosphorus, and sulphur may influence the ability of forest
ecosystems to store carbon. And fourth, we evaluate the possibility that the terrestrial
biosphere is on a carbon accumulation trend that is dependent on the dynamics of phosphorus
availability during soil development and that is independent of the activities of man. J. R.
Gosz is the principle author of section 6.2, and J. M. Melillo is the principle author of
sections 6.36.5.

6.2 ELEMENT RATIOS IN FOLIAGE AND LITTER


Uptake of a nutrient is reflected in the amounts accumulated in plant tissues, and so nutrient
concentrations and ratios in plant tissues have long been of interest to scientists concerned
with mineral nutrition of higher plants. Nutrient concentrations in the green foliage of a tree
species can vary from site to site because of different supplies of individual nutrients at the
sites, and these variations are usually reflected in nutrient concentrations in leaf litter. In this
section of the paper we consider some of the recent analyses relating variations in element
concentrations and ratios in green foliage and leaf litter to nutrient availabilities in forests.
This information relates to the uptake and return portions of the biogeochemical cycles.

6.2.1 Element Ratios in Foliage

The demand for nitrogen is closely related to tree growth, and nitrogen deficiency is, after
water stress, the most frequently reported limitation to growth (Kozlowski, 1971; Kramer and
Kozlowski, 1979). Phosphorus, like nitrogen, is in short supply in forest ecosystems in many
parts of the world and the roles of phosphorus and nitrogen in plant metabolism are clearly
interrelated in a number of ways (Loveless, 1962). Sulphur is primarily a constituent of
amino acids and proteins. The stoichiometry between nitrogen and sulphur is fairly constant,
indicating that on the average there are 36 atoms of nitrogen for each atom of sulphur in
proteins. The ratio of total nitrogen to total sulphur in plants is close to this value under
conditions where there is no luxury consumption of either element (Epstein, 1972).

Although the requirement for these elements in various organic compounds would suggest
rather constant element ratios in green tissues, the ratios may vary considerably because of
different supplies of individual nutrients. Each of the above nutrients has been shown to be
taken up in excess (i.e. luxury consumption) and stored. Slowly growing species that absorb
nutrients in excess of immediate growth requirements during times of high nutrient
availability may use these reserves to support growth after soil reserves are exhausted
(Chapin, 1980). The nutrients available for plant uptake may be present in ratios much
different from those needed by plants. As many studies have shown, nitrogen can be taken up
in excess and stored as amino acids, for example, arginine (Barnes and Bengtson, 1968;
Kramer and Kozlowski, 1979). The absorption of sulphate from excess supply may be faster
than its reduction and assimilation of the sulphur atoms into organic compounds. Thus, an
appreciable fraction of total sulphur in plants may be in the form of sulphate (Epstein, 1972;
Turner et al., 1980). Although phosphorus is absorbed as a complex anion like its nitrogen
and sulphur counterparts (i.e. nitrate, sulphate), the phosphorus atom of phosphate is not
reduced in the cell to a lower oxidation state (Epstein, 1972). In mature plants, phosphorus is
temporarily stored as phosphate while seasonal storage may occur as phospholipids, nucleic
acids, and other complex organic compounds. Luxury consumption and maintenance of
inorganic phosphorus reserves by slowly growing species from infertile habitats are
responsible for the high proportion of inorganic phosphorus and low proportion of
structurally bound phosphorus characteristic of these species (Chapin, 1980).

Although concentration in foliage is commonly used to identify nutrient deficiencies and


imbalances, variations in nutrient ratios have also been found useful. For example, conifers
have total foliar nitrogen very nearly equal to organic foliar nitrogen. However, any sulphur
in excess of that required to balance the nitrogen in protein formation is accumulated as
sulphate-sulphur. Foliar sulphate-sulphur is low when sulphur is deficient and nitrogen is
adequate, and sulphate-sulphur is high when sulphur is abundant and nitrogen is deficient.
Turner et al. (1980) demonstrated that on nitrogen-deficient sites, additions of nitrogen
resulted in the incorporation of sulphate-sulphur into organic forms. Thus, while the cycles of
organic nitrogen and organic sulphur are closely coupled, excess sulphur can cycle as
sulphate and operate somewhat independently of the nitrogen or carbon cycles (Turner et al.,
1980). Some plants accumulate inorganic nitrogen (e.g., nitrate) and phosphate during
conditions of excess supply (Dijkshoorn and Van Wijk, 1967), and therefore, these elements
also can be expected to cycle independently. Relatively large differences in total nitrogen to
total phosphorus ratios in tissue occur for various ecosystems, suggesting different cycling
rates. Cole and Rapp (1980) report nitrogen to phosphorus ratios in uptake ranging from
about 4:1 to 17:1 and 11:1 to 22:1 for 13 temperate conifer and 14 temperate deciduous
forests, respectively. The nitrogen: phosphorus ratios for the requirements of these forests
ranged from 5:1 to 16:1 and 8:1 to 19:1, respectively. The range of phosphorus uptake values
was larger than the range of nitrogen uptake values, indicating that phosphorus availability
may have caused most of the variation in the ratios.

6.2.2 Nutrient Translocation within Trees and Nutrient Use Efficiency

Very low uptake rates of an element in relation to demand by current growth results in
efficient conservation and re-use of that element within the tree. Translocation of nutrients
within the individual is an important mechanism for this efficiency. In much the same way
that translocation of carbohydrates and their partitioning are controlled by the size of the food
supply and relative sizes of various sinks, we suggest the translocation of nutrient elements is
controlled by strength of the sink (i.e. demand) and magnitude of the source (i.e. supply).
During the height of the growing season the relative strengths of the various sinks are: fruits
and seeds > young leaves and stem tips > mature leaves > cambia > roots > storage (Kramer
and Kozlowski, 1979). Relative distance from source to sink is also important, because sinks
are supplied from the nearest source. Due to the asynchrony of growth of different plant
parts, the same nutrient capital can serve several functions during the growing season.
Tissues with an imbalance of nutrients such as an excess of sulphur and deficiency of
nitrogen will have proportionately more of the nitrogen translocated (Turner et al., 1980).
Nutrients in excess can be transported to storage areas. For example, leaves with high
concentrations of nitrogen and phosphorus have a larger percentage of soluble and inorganic
forms and may actually retranslocate a larger total quantity of nitrogen and phosphorus from
leaves than would occur in leaves with nutrient deficiencies (Chapin, 1980). The most
deficient element may possibly be transported to stronger sinks while storage, being a weaker
sink, may draw a larger proportion of elements in excess. The true measure of the efficiency
of this internal cycle may not be the total quantity translocated, but the ability to withdraw
nutrients, leaving very low levels in senescing tissues. Thus, the percentage translocation may
not be as important as the level to which the nutrient concentration can be decreased. A better
measure of efficiency may be the quantity translocated divided by the amount remaining.

If there are ample soil supplies of a nutrient element, then less demand is put on the internal
cycle and greater quantities may remain to accumulate in tissues as they age; an accumulation
that is eventually reflected in higher nutrient concentrations in litterfall. This has been
demonstrated in a number of fertilization studies (Barnes and Bengston, 1968; Miller et al.,
1979; Turner et al., 1980). Similarly, studies that have decreased availability, and hence
uptake, by practices such as sugarsawdust application, caused increased translocation
efficiency in senescing tissues and decreased nutrient levels in leaching and litterfall (Turner,
1977; Turner et al., 1980). This also has been demonstrated across gradients of nutrient
availability within natural stands. Sites with low nutrient availability have individuals and
species that translocate proportionately more out of leaves before leaf fall than do the species
on sites with abundant nutrients (Lamb, 1975; Stachurski and Zimka, 1975; Zimka and
Stachurski, 1976). The relatively high concentrations of nutrients left in senescing leaves on
nutrient-rich sites speed the decomposition and mineralization processes that enhance the
high availability (Gosz, 1981). Figure 6.1 presents data for six communities showing a very
good relationship between k values for nitrogen in leaf litter, a measure of the decomposition
and release of this element, and percentage retranslocation out of leaves during senescence
(Zimka and Stachurski, 1976).

Figure 6.1 The relationship between percent leaf nitrogen translocated prior to leaf abcission
and the decay rate (k) for nitrogen in the litter for six forest ecosystems along a nutrient
availability gradient in Poland (constructed from data of Zimka and Stachurski, 1976, and
Stachurski and Zimka, 1975)

While many studies have demonstrated a reduction in nutrient concentrations between green
and senescent tissues, Chapin (1980) reports that the limited evidence available does not
confirm that species adapted to infertile soils are particularly effective in retranslocating
nutrients prior to leaf abscission. Much more work is necessary in this important area.

In an attempt to determine whether nutrient cycling and nutrient use efficiency vary with
forest type, we analysed 102 data sets taken from boreal to tropical sites in the northern and
southern hemispheres for nutrients in litterfall. The quantity of a nutrient in litterfall is a
measure of nutrient circulation (Gosz, 1981; Vitousek et al., 1982), particularly for nutrients
such as nitrogen and phosphorus that are lost from the plant primarily through litterfall (Cole
and Rapp, 1980). Vitousek (1982) suggested two extreme hypotheses to aid the interpretation
of patterns of nutrient use efficiency in litterfall. One is that the efficiency of nutrient use, as
measured by the amount of organic matter discarded in litter per unit of N, is constant for any
level of nutrient supply and circulation; the concentration of a nutrient in litterfall would then
be constant. The variable nutrient concentrations in litter prove this hypothesis false.
Alternatively, it could be hypothesized that nutrient circulation is unrelated to litterfall mass.
Litterfall mass would vary independently or randomly from nutrient circulation. If this were
true then a plot of the carbon:nutrient ratio of litterfall versus the nutrient content of litterfall
would result in a random scatter of points with an upper limit of the form Y = 1/X (Figure
6.2). The upper limit is a result of autocorrelation; the X-axis values are used in the
calculation of Y-axis values (Vitousek, 1982).

Figure 6.2 The relationship between the amount of nitrogen in litterfall and the
carbon:nitrogen ratio of that litterfall. Data from 102 forest sites world-wide (T-tropical, D-
deciduous, C-conifer, E-Eucalyptus). The dashed line is the upper limit for the relationship
for randomly generated data. See text for explanation

In Figure 6.2 we have plotted carbon:nitrogen ratio versus nitrogen content of litterfall for
tropical hardwood, temperate conifer, temperate deciduous, Eucalyptus (Australia), and
Nothofagus (New Zealand) forests. The overall plot reveals a strong inverse relationship
between carbon:nitrogen ratio and the mass of nitrogen in litterfall. This inverse relationship
differs from one generated with random data by being more confined (less variation) and
having a steeper logarithmic function.
An analysis by vegetation types shows that the relationship between carbon:nitrogen ratio and
nitrogen in litterfall varies markedly. Regression coefficients for these communities were
significantly different (P < 0.05) from coefficients of randomly generated data. Tropical
vegetation has a very high litter nitrogen content and, although the slope of the regression
lines is significantly different from 0 (P < 0.05), it is the closest to a horizontal line (i.e. the
hypothesis stating that the efficiency of nitrogen use is unchanged at different rates of
nitrogen circulation). The higher nitrogen levels in litter suggest nitrogen circulation is
always high, however. One interpretation is that nitrogen is in excess in these forests with
respect to other elements and may be cycling somewhat independently.

The regression for temperate deciduous forests has a slope somewhat greater than that for the
tropics and differs primarily in the lower nitrogen content of the litter. These two vegetation
types seem to comprise a continuum across a large portion of the range of nitrogen contents
of litterfall.

Conifer forests show a marked increase in slope, and have some of the lowest litter nitrogen
contents. The evergreen Nothofagus forests of New Zealand show similar plots. The conifer
vegetation type shows a curvilinear relationship (Figure 6.2) and may be divided into two
parts; more than or less than 3 g m-2 yr-1 nitrogen in litterfall. Forests with nitrogen masses of
less than 3 g m-2 yr-1 in litter have an almost vertical slope for their plots. For forests with
nitrogen values greater than 3 g m-2 yr-1, the plots are similar to deciduous forest plots.
Interestingly, these sites are 2033 year-old plantations of various conifer species on former
hardwood sites (Gloaguen and Touffet, 1976). Vitousek (1982) also reports a high nitrogen
content in the litter of a balsam fir forest in New Hampshire. However, this native conifer
forest is nitrogen-rich perhaps because of high nitrogen inputs in acid precipitation. The
Eucalyptus forests of Australia also plot along a steep slope (P < 0.05, Figure 6.2) showing a
strong relationship between nitrogen circulation and litterfall carbon to nitrogen ratio.

The data of Figure 6.2 are based on annual litterfall measurements. A comparison of tropical
forests with temperate forests on an annual basis may not be appropriate because
decomposition rates are so rapid in the tropics. With litter decomposing in a matter of
months, a pool of high available ntirogen may be circulated several times in the course of a
year. Assuming that the nitrogen pool was used three times per year (i.e. nitrogen content of
litter/3) would increase the carbon:nitrogen ratio of annual litterfall by a factor of three. A
plot of such data would result in a steep slope more similar to those of conifer and Eucalyptus
communities. These data suggest a much more efficient use of nitrogen at the ecosystem level
in the tropics than annual litterfall nitrogen values indicate. The resolution of this difference
is an important research objective.

Figure 6.3 The relationship between the amount of phosphorus in litterfall and the carbon:
phosphorus ratio of that litterfall. Data from 102 forest sites world-wide (T-tropical, D-
deciduous, C-conifer, E-Eucalyptus)

Although the data of Figure 6.2 indicate that tropical forests have high nitrogen levels in
litterfall, this is not the case for all tropical forests. Many areas in the tropics (e.g., certain
parts of Venezuela and Brazil) are nutrient-poor and have lower nitrogen levels in litterfall
(R. Herrera, personal communication). These sites plot with the conifer forests in Figure 6.2.
Similarly, conifer and deciduous forests can have high nitrogen levels in litterfall and plot
more like tropical sites.

The results for phosphorus (Figure 6.3) differ somewhat from those for nitrogen in that there
are steeper slopes for tropical and Eucalyptus forests, while slopes for conifer and deciduous
forests are similar. Also, the range of phosphorus in tropical litterfall is very large. The data
for Eucalyptus forests give a plot that is nearly a vertical line; a situation where phosphorus
circulation is low and almost constant. This is of interest because of the marked phosphorus
deficiency cited for most of Australia (Loveless, 1962). The results for conifer, Nothofagus,
and deciduous forests are less obvious and although they have similar and significant
regression lines (P < 0.05), the scatter of points is appreciable.

Figure 6.4 The relationship between the amount of phosphorus in litterfall and the nitrogen:
phosphorus ratio of that litterfall. Data from 102 forest sites world-wide

The relationship between nitrogen and phosphorus in litterfall appears different for different
forests. In Figure 6.4 we have plotted nitrogen:phosphorus ratios versus the phosphorus
content of litterfall. A highly significant relationship exists for the tropical forests (P <
0.001). The tropical forests are normally described as nitrogen-rich forests, as can be seen
from Figure 6.2. The major factor in the relationship seems to be the quantity of phosphorus
in litter, with the range of phosphorus values almost four times the range of nitrogen values
(Figure 6.5). High phosphorus in litter reduces the nitrogen to phosphorus ratio in litter and
low phosphorus increases the ratio. This suggests that nitrogen acts like an element in excess
(i.e. the efficiency of nutrient use is unchanged at various levels of nutrient circulation) while
phosphorus use efficiency is more strongly related to phosphorus circulation. The relationship
for conifers also is statistically significant (P < 0.05), although the scatter of points is
appreciable. The other vegetation types did not have significant relationships.

Figure 6.5 The relationship between the phosphorus and nitrogen concentrations in litterfall.
Data from 102 forest sites world-wide

The relatively strong relationships between carbon and nutrients (Figures 6.2, 6.3) and the
lack of a relationship between nutrients (Figures 6.4, 6.5) for most forests again indicate the
nutrient use efficiency of one nutrient is somewhat independent of that for another nutrient.
Unfortunately, little data are available for sulphur, preventing us from testing this result with
other nutrient combinations.

If correct, these data support our previous discussions that the efficiency of nutrient use is
inversely related to the availability or rate of circulation of that nutrient. At high levels of
availability, the mass of litterfall (a measure of productivity) is not related to levels of the
nutrient in litterfall. At lower nutrient levels, the carbon to nutrient ratio in litterfall is
markedly influenced and the efficiency of use of the nutrient increases. This also agrees with
previous discussions that the requirement for this nutrient causes strong sinks and
translocation out of senescent tissues, resulting in high carbon to nutrient ratios in litterfall.
At very low nutrient levels (nutrient deficiency) litterfall mass again appears unrelated to the
nutrient content of litterfall. It is as if there is a minimum level to which the nutrient content
of litterfall can be reduced despite a strong sink effect and effective translocation.

Figure 6.6 The relationship between phosphorus content of litterfall and the fibre:protein
ratio of that litterfall (data of Loveless, 1962)

A high carbon to nutrient ratio or high litterfall mass with low nutrient content can be
described as an efficient use of the nutrient (i.e. high litter production per unit of nutrient
(Vitousek, 1982)). Translocation of nutrients out of senescing tissue would contribute to this
efficiency, although the extremely high carbon to nutrient ratios in litter of some conifer and
Eucalyptus forests suggest other physiological processes are involved. Loveless (1962)
proposed that sclerophylly was related to nutrient availability. Sclerophyllous leaves have a
high fibre:protein ratio caused by reductions in protein content and concomitant increases in
fibre (cellulose, lignin). A plot of fibre to protein versus phosphorus concentration is
suggestive of a limiting factor curve (Loveless, 1962); that is, the fibre:protein ratio decreases
with increased phosphorus content up to a certain level (i.e. 0.3%) above which increased
phosphorus content does not result in a further proportional decrease in the fibre to protein
ratio (Figure 6.6). The roles of phosphorus and nitrogen in plant metabolism are interrelated
in many ways and both are essential for protein synthesis. Thus, lowered protein levels could
be a result of either low nitrogen or low phosphorus content. Furthermore, it is reasonable to
expect that the intermediate products of metabolism that otherwise might have formed protein
should, in the absence of either adequate phosphorus or nitrogen, be diverted along
alternative metabolic pathways to form other end-products, including fibre (Loveless, 1962;
Neish, 1964; Gosz, 1981).

Gnanam et al. (1980) proposed a mechanism of action for ammonium in regulating the
photosynthetic carbon flow. In this scheme ammonium ions seem to regulate the
photosynthetic carbon flow by abolishing the light activation of the enzymes that would
normally favour the flow of carbon toward sugar biosynthesis, thereby facilitating the
increased synthesis of amino acids.

These results suggest that in addition to translocation of scarce nutrients from senescing
tissues to other sinks within the plant, tissue chemistry in nutrient deficient sites also is
different. For nutrient deficient sites, the diversion of proportionately more carbon into
fibrous material along with a more complete removal of nutrients by translocation may
account for the very high carbon to nutrient ratios in litterfall.

6.2.3 Summary

Several conclusions can be drawn from an analysis of element ratios in foliage and litter:

1. Element ratios in green foliage and leaf litter vary among sites because of different
supplies of individual nutrients at the sites.
2. The withdrawal or translocation of nutrients from leaves prior to abscission is an
important nutrient conservation mechanism in forest ecosystems.
3. A high carbon:nutrient ratio or a high litterfall mass with low nutrient content can be
described as an efficient use of the nutrient under consideration. Translocation of
nutrients out of senescing tissue contributes to this efficiency.
4. Nutrient use efficiency of one nutrient is often independent of nutrient use efficiencies
of other nutrients; for example, efficient use of nitrogen in a forest ecosystem does not
necessarily imply efficient use of phosphorus in that system.
5. The efficiency of nutrient use is strongly related to the quantity of the nutrient cycling
within the forest ecosystem. The quantity of a nutrient in litterfall is a measure of the
nutrient cycling rate.

6.3 CARBONNUTRIENT INTERACTIONS AT THE ECOSYSTEM LEVEL

Net carbon storage in forest ecosystems is determined by the balance among three distinct
processes: (1) the net amount of carbon fixed by the vegetation (i.e. carbon fixed in excess of
plant respiratory demand); (2) the relative amounts of this fixed carbon allocated to
vegetation growth increment versus plant litter; and (3) the decomposition of carbon
compounds entering the soil system. All three of these processes are regulated by
temperature, moisture and the availability of key nutrients such as nitrogen, phosphorus, and
sulphur. Light and carbon dioxide concentration of the atmosphere serve as additional
regulating factors on the process of carbon fixation by vegetation.

The primary objective of this section is to examine how the cycles of nutrients are linked to
both the paths of carbon transfer and the amounts of carbon storage in forest ecosystems. We
present a simple model (summarized in Figure 6.7) that we use to identify the linkages
between carbon and nutrient dynamics in forests. Our analysis suggests that the nutrient
dynamics of a system can control the amount of carbon that moves through a system, the
pathways of its movement, the amount of carbon accumulated in the system, and the
distribution of the accumulated carbon in the system. We envision a series of switches in the
carbon flow pathway, which are in part under the control of the nutrient status of the plants in
a system.

6.3.1 Photosynthesis and Net Primary Production

Photosynthesis and net primary production (net dry matter production by green plants) are
distinct components of the carbon budget of a forest ecosystem. Carbon flux through a forest
begins with photosynthesis, or the conversion of CO2 to organic carbon compounds. Some
fraction of the total photosynthate is consumed in plant respiration, with the carbon returned
to the atmosphere as C02. The remainder of the photosynthate is used to build plant tissues
(e.g. leaves, stems, roots) and soluble organic compounds (e.g. root exudates). The tissues
and soluble compounds produced during the course of a year comprise the system's annual
net primary production.

A. Photosynthesis

Plant nutrient status along with light, temperature, and water availability are important
controlling parameters for photosynthetic rate. Several comprehensive reviews document the
relationship between leaf nutrient status and photosynthetic rate (e.g. Keller, 1967; Natr 1972,
1975). According to these reviews, nitrogen occupies a special place among nutrients
involved in photosynthesis. One generalization that emerges is that a quantitative relationship
exists between rate of photosynthesis and nitrogen content of leaves. Photosynthetic capacity
often increases linearly with increases in leaf nitrogen concentration if light, water and other
nutrients are not limiting. This relationship is well documented for crop plants and non-
woody wild plants (Natr, 1975), and for both deciduous (Keller, 1960) and coniferous
(Keller, 1971) tree species.

The nutrient status of plant parts other than leaves also influences photosynthetic rate through
complex source-sink control mechanisms (Wareing and Patrick, 1975). Under conditions
where the potential rate of carbon fixation in the leaves exceeds the rate of carbon
consumption throughout the rest of the plant, there is a feedback mechanism whereby the rate
of assimilation is regulated to meet demand. The demand for carbon by various plant parts is,
in turn, related to their nutrient status. For example, nitrogen deficiency limits the formation
of new tissue and thus the demand for photosynthate (Kramer, 1981).

By influencing a plant's demand for photosynthate, plant nutrient status exerts an influence
on the ability of a plant to respond to elevated levels of CO2. Working with tobacco plants,
Raper and Peedin (1978) reported a close relationship between nitrogen supply and the ability
of the plants to respond to enhanced CO2 levels. During a 35 day study, tobacco plants
growing at `low' levels of nitrogen supply in CO2 concentrations of both 400 ppm and 1000
ppm had per plant photosynthetic rates that were only 60 percent of the photosynthetic rates
exhibited by tobacco plants growing at `high' levels of nitrogen supply in the same two CO2
concentrations. Other nutrients besides nitrogen can also influence photosynthetic rates.
Sulphur deficiency prevented photosynthesis of sugar beet from responding to an increase in
CO2 concentration (Thomas and Hill, 1949). Unfortunately, no studies have been conducted
to examine the relationship between the nutrient status of forest trees and their ability to
respond to elevated levels of CO2. Such studies are important for our understanding of the
carbon flux through forest ecosystems and the role of forest vegetation in the global carbon
budget.

B. Partitioning of Total Photosynthate between Respiration and Net Primary Production

All biomass production ultimately depends on photosynthesis. This fact does not imply that
the rate or extent of net primary production bears a close relationship to the photosynthetic
rate, or is determined by it. The processes that follow photosynthesis, such as respiration, can
be major determinants of productivity. The loss of photosynthate by dark respiration can be
substantial, particularly in communities having a large biomass, and growing under high
temperatures. In the massive forests of southern Thailand, for example, Kira (1975) estimates
that about three-quarters of the carbon assimilated in photosynthesis is lost by dark
respiration.

The partitioning of the total photosynthate between respiration and net primary productivity
is controlled by a variety of factors. In non-woody plants there is some evidence that plant
nutrient status is in part responsible for determining the fate of photosynthate. Based on work
with the grassland species Plantago lanceolate, Lambers et al. (1981) have suggested the
existence of an overflow metabolism or SHAM (salicylhydroxamic acid) pathway in roots of
plants growing in environments with fluctuating nutrient availabilities. In such environments,
plants may maintain sink strength in leaves that allows high rates of photosynthesis. To
maintain photosynthate demand in leaves, which have limited carbon storage capacity, the
fixed carbon is transported to other plant parts such as the roots. At times of high nutrient
availability, and thus high nutrient status throughout the plant, there will be high demand for
the photosynthate and it will be incorporated into plant tissues and become part of the net
primary production of the system. At times of low nutrient availability and thus low nutrient
status throughout the plant, there will be low demand for the photosynthate and it will be
respired via the SHAM pathway.

The importance of the SHAM pathway and other biochemical pathways for wasteful
oxidation (Solomos, 1977) has not been established for forest tree species. Both woody and
non-woody plants certainly have other strategies for photosynthate management in
environments of fluctuating nutrient availability. As mentioned in the previous section, the
capacity of a plant to engage in luxury consumption of nutrients may be an important
mechanism for dealing with the asynchronies that often exist between nutrient availability
and carbon fixation.

6.3.2 Shoot:Root Ratio and the Response of Shoots and Roots to Fertilization

The relative distribution of plant carbon between above-ground parts (shoots) and below-
ground parts (roots) is closely related to plant nutrient status, which in turn reflects soil
nutrient availability (cf. Figure 6.7). Shoot:root ratios of both deciduous and coniferous tree
species tend to be lower in infertile habitats than in fertile habitats (Yen et al., 1978; Grier et
al., 1980; Keyes and Grier, 1981).

The initial response of forest trees to improved soil nutrient availability, and thus plant
nutrient status, seems to be an increase in the relative amount of annual net primary
production allocated to roots (Safford, 1974; Miller and Miller, 1976; Brix and Mitchell,
1980). The enhanced root growth presumably increases uptake of both water and nutrients
and permits increased above-ground growth.

Above-ground growth response to fertilization appears to be closely related to increases in


leaf area. Brix and Ebell (1969) reported that the only above-ground plant factor associated
with increased diameter growth of a 20-year-old Douglas fir stand fertilized with nitrogen
was increased leaf area. In a subsequent experiment on Douglas fir, Brix (1971) found that,
although rates of photosynthesis and dark respiration increased after nitrogen fertilization,
most of the increase in diameter growth was caused by the increase in leaf area. Fertilization
was most effective in open stands where the leaf area was below the optimum and water
supply was not limiting. Tamm (1979) also reported that the close correlation between leaf
area and stem growth observed in young conifer stands decreased as the stands closed and
shading effects increased in importance.

Figure 6.7 Model describing the interaction between the nutrient and carbon budgets of
forest ecosystems.  Symbols indicate switches in the carbon budget that can be influenced
by the nutrient status of the plants or nutrient availability in the system

In closed canopy stands, fertilization may affect stand composition more than it affects
overall stand growth. The larger trees in the stand will often accumulate nutrients more
rapidly than smaller trees, with the result being more rapid growth of the larger trees. This
causes the loss of lower crown classes from the stand, but the rapid growth of the dominant
trees maintains the stand's leaf area and sapwood cross-sectional area (Jarvis, 1975).
However, increases in sapwood cross-sectional area are not always related in a simple way to
biomass increment. For example, Brix and Ebell (1969) reported that the effect of nitrogen
fertilization on dry matter production in Douglas fir is lower than it is on volume increment,
since the specific gravity of treated trees consistently decreased after fertilization and was 12
percent lower than that of the controls.

6.3.3 Carbon Allocation to Perennial Versus Deciduous Plant Parts

The carbon component of net primary production in forest ecosystems can have three fates:
(1) it can be allocated to perennial parts and become incorporated into the vegetation's growth
increment (the vegetation increment of net ecosystem production); (2) it can be transferred
from the vegetation to the soil in plant parts as litter; and (3) it can be transferred from the
vegetation to the soil as soluble organic compounds that are leached from above-ground plant
parts or exuded from roots.

The evidence to date suggests that soluble carbon transfer to the soil in leachates and
exudates is a small part of the carbon component of net ecosystem production in forests (e.g.
Gosz et al., 1976). And at this time it is not clear how plant nutrient status affects the relative
magnitude of this transfer.

As an aside, it is interesting to note that in plants with very limited storage capacity, the
management of photosynthate during periods of low nutrient availability involves the
exudation of photosynthate. This is evident, for example, in some communities of algae
growing in oligotrophic waters (Fogg, 1975). For many of these algae, growth is limited to
such an extent by nutrient supply that they cannot use most of their photosynthate and may
excrete up to 90% of it. At first glance, this means of managing excess photosynthate would
seem to further lower nutrient availability by promoting blooms of micro-organisms that have
nutrient demands. If nitrogen is the limiting nutrient, as it may be in some river waters, the
excretion of simple carbon compounds by algae may promote nitrogen fixation by micro-
organisms, thus aleviating some of the nutrient stress. An analogue of this scenario may be
occurring in forest ecosystems. Trees growing in nitrogen-poor soils may promote nitrogen
fixation by micro-organisms in the root zone by allocating a larger percentage of their
photosynthate to root exudation than trees growing on nitrogen-rich soils. There is some
evidence that this may be occurring in dwarf willows in Alaska, with mycorrhiza acting as
intermediates in the process (Linkins, personal communication). Mycorrhizal fungi
associated with the dwarf willows can convert the sucrose they receive from the willows to
mannitol, some fraction of which is then exuded from the mycorrhizae into the soil. Since
mannitol is a preferred carbon form for free living nitrogen-fixers, the allocation of
photosynthate to the root-mycorrhizal complex of plants growing in nitrogen-poor sites can
promote fixation.

Turning now to a consideration of the major fates of the carbon component of net primary
production, we ask the following question: Does the nutrient status of the trees in a forest
influence the relative distribution of net primary production between litter and the vegetation
growth increment? Studies relevant to this question are few. One of the most useful is the
work of Turner (1977) on the effects of nitrogen availability on nitrogen cycling in a Douglas
fir stand. Turner created a gradient of nitrogen availabilities in a 50-year old Douglas fir stand
in Washington State, U.S.A. The annual net primary production was highest in fertilized sites
and lowest in the site where nitrogen availability was lowered relative to the control by the
additions of carbohydrate to the soil, which promoted microbial immobilization of nitrogen.
In the first year of the study, trees growing in high nitrogen availability sites allocated a
smaller percentage of their net primary production to litter than to growth increment
compared to trees growing in the control and nutrient stressed sites (Table 6.1). Part of this
response is associated with the fact that fertilization increases leaf longevity of conifers in the
first year of application (Miller et al. 1976, Chapin 1980). The relative distribution of the net
primary production among stem increment, branch increment, and new foliage production
was similar in the control and fertilized trees. In the stressed trees, the relative distribution of
the net primary production was similar to that of the others for branch increment and new
foliage production, but substantially less than the others for stem increment (Table 6.1).
These data suggest that the relative amount of net primary production allocated to stem
growth is variable under nutrient stress conditions in the Douglas fir.

From a study of the effect of nitrogen supply on net primary production in Corsican pine,
Miller and Miller (1976) reported that repeated fertilization over a 3-year period: (1)
increased the amount of net primary production substantially; and (2) reduced the relative
amount of net primary production allocated to litter (Table 6.2). In fertilized trees the relative
allocation of net primary production to new foliage production, stem increment, and root
growth generally increased compared to non-fertilized trees.

6.3.4 Litter Decomposition

In section 6.2, we suggested that litter quality is related to the nutrient status of the plant. We
argued that nutrient stressed plants retranslocate large amounts of nutrients into woody parts
before leaf fall, thus lowering the nutrient content of the litter. Also, plants growing on
nutrient-poor sites contain a larger proportion of complex carbon compounds (e.g., lignin and
polyphenols) in their litter than plants growing on nutrient-rich sites (Loveless, 1962; Davies
et al., 1964 a, b; Lamb, 1976; Gosz, 1981). We will now present evidence that litter quality
can affect both the decomposition rate of fresh litter and the relative amount of that litter that
is ultimately transformed to meta-stable `humus'.

Table 6.1 Net primary production rates and allocation of net primary production in treated
and control Douglas-fir forests (calculated from Turner, 1977)

22 g N m-2
Treatment 88 g N m-2 yr-1 Control Carbohydrate
hr-1
Net Primary
512 512 492 405
Production*
(g C m-2 hr-1)
Allocation of
NPP
(percentage)
Litterfall 12.6 16.5 19.0 29.5
Total Growth
87.4 83.5 81.0 70.5
Increment
Stem Increment 57.2 54.1 53.7 41.9
Branch
5.2 4.9 4.8 6.4
Increment
New Foliage
25.0 24.5 22.5 22.5
Production
*Above-ground components only. Carbon calculated as 50%
of dry matter.
Table 6.2 Net primary production rates and allocation of net primary production rates in
fertilized and control Corsican pine forests (Miller and Miller, 1976)

Treatment 0 8.4 16.8 33.6 50.4


(g N m-2 yr-1
Net Primary
560 845 930 935 950
Production*
(g C m-2 yr-1)

Allocation of NPP
(percentage)
Foliage 8 11 14 16 16
Live Branches and
12 11 9 9 6
twigs
Dead Branches 4 2 2 1 4
Stem Wood 32 29 37 40 30
Stem Bark 7 6 6 5 6
Root System 13 26 17 15 23
Litterfall 24 15 15 15 15
*Carbon calculated as 50% of dry matter.

A. Litter Quality and Decomposition Rate

Little quality, as defined by chemical composition of the material, has long been considered a
critical factor in determining rate of decay (Waksman and Tenney, 1927). Chemical indices
of litter quality include element concentrations and concentrations of various classes of
organic compounds.

Cromack (1973), Cromack and Monk (1975), and Fogel and Cromack (1977) have reported
that the initial lignin concentration in the litter is an excellent index to the rate of weight loss
of litter samples. Initial lignin concentration also gave a high correlation with decomposition
rate when the data of Lockett (1937) were re-analysed by Cromack (1973).

A high correlation between initial nitrogen concentration and decomposition rate has been
demonstrated by a number of researchers, including Bal (1922), Hill (1926), Waksman and
Tenny (1928), Waksman and Gerretsen (1931), Monnier and Jeanson (1964), Cowling and
Merrill (1966), Satchell and Lowe (1966), Witkamp (1966), and Zimka and Stachurski
(1976). But the high correlation between initial nitrogen concentration and decomposition is
not universal. Melin (1930), and Daubenmire and Prusso (1963) have found poor correlation
between decomposition rates and the initial nitrogen percentage in litter.

Figure 6.8 The relationship between the percent biomass remaining in decomposing leaf
litter at the end of 12 months of decomposition in the field and the initial lignin:nitrogen ratio
of the various litter materials. At the New Hampshire site the leaf litters are as follows: Be,
American beech; SM, sugar maple; PB, paper birch; RM, red maple; PC, pin cherry; and A,
ash. At the North Carolina site the leaf litters are as follows: WP, white pine; CO, chestnut
oak; WO, white oak; RM red maple; and FD, flowering dogwood. From Melillo et al. (1982)
Melillo et al. (1982) have found that the ratio of initial lignin concentration to initial nitrogen
concentration is a better predictor of decomposition rate than is either initial lignin
concentration or initial nitrogen concentration (Figure 6.8). Since lignin is among the most
difficult organic compounds to decompose, and since nitrogen is, for many ecosystem
processes, the most limiting nutrient, it is reasonable that the lignin: nitrogen ratio of litter
would be a good predictor of decomposition rate.

Slow rates of litter decomposition can result in the accumulation of large unavailable nutrient
stocks in a forest soil's surface horizons, and nutrient limitations for primary producers
(Siren, 1955; Weetman, 1962; Heilman and Gessel, 1963; Florence, 1965; Watt and
Heinselman, 1965; Heilman, 1966; Miller, 1969; Adams et al., 1970; Lamb, 1971). It is not
difficult to envision a positive feedback loop that would result in the perpetuation of an
unproductive forest stand: plant nutrient stress promotes the production of low-quality litter
that decomposes slowly, releasing nutrients slowly, and thereby perpetuating plant nutrient
stress.

B. Litter Quality and the Formation of Humus

Formation of humus, the meta-stable organic fraction of soils, is still a poorly understood
process. Evidence suggests that lignin and polyphenols contribute to humus formation
(Allison, 1973). DeHaan (1977) found a high correlation between humus formation and the
amount of lignin introduced into soils over a 10 year period. Given that lignin concentrations
are generally higher in litter of nutrient-poor sites, it may be that in relative terms, more of
the litter entering the soil in a nutrient-poor site would be transformed to humus than would
be the case in a nutrient-rich site. If this speculation is true, then the nutrient status of plants
would affect carbon flow in soils. In absolute terms, the higher litter inputs associated with
nutrient-rich sites may ultimately lead to more humus formation per unit time, despite the fact
that a smaller percentage of the litter will be transformed into humus. This may be a critical
component of calculations on the effects of forest fertilization on carbon storage.

C. Root Litter and Soil Carbon Stocks

To this point, we have only considered above-ground litter input. Root litter input to soils,
especially fine root litter input, can be very large. For example, Edwards and Harris (1977)
estimate a fine root litter input of 900 g C m-2 yr-1 in a temperate zone hardwood forest.

The fraction of root litter input transformed to refractory humus in forests is not known.
Relatively labile carbon compounds can also have a prolonged residence time in soils if they
are physically protected from decomposition. Allison (1973) suggested that this process may
be very important in determining the fate of root litter input. He notes that roots have some
advantages over top residues as carbon sources to soils. Roots are intimately mixed with the
soil at all times and, as they decompose, produce a gum-like material, that is well distributed.
These polysaccharide `gums' are in a position to act as cements between soil particles as they
are being formed into aggregates by various forces. When fixed in aggregates there is much
evidence that the polysaccharide `gums' are protected for a time against oxidation by micro-
organisms.

Table 6.3 Total net primary production in g m-2 yr-1 for two 40-year-old Douglas fir stands
(Keyes and Grier, 1981)
Component Poor site % of total Good site % of total

Stem wood 420 27 .3 820 46 .0


Stem bark 90 5.8 170 9.5
Living branch 20 1 .3 60 3 .4
Foliage 200 13 .0 320 18 .0
Large root (> 5 mm) 110 7 .1 160 9 .0
Small root (25 mm) 140 9 .1 110 6 .2
Fine root (< 2 mm) 560 36 .4 140 7 .9
Total 1540 100 1780 100

Polysaccharides are also produced during decomposition of surface litter, but the
polysaccharides are usually not in intimate contact with the soil. Furthermore, the
polysaccharides produced in the surface litter are decomposed so rapidly that there is only
limited movement from the immediate areas where they are formed.

As mentioned earlier, both relative and absolute amounts of net primary production allocated
to roots may be larger in nutrient-stressed stands than in nutrient-rich stands. This is clearly
seen in the data of Keyes and Grier (1981; Table 6.3). On the nutrient-rich site the net
primary production allocated to fine roots in a 40-year old Douglas fir stand was 140 g C m-2
yr-1, which amounted to 7.9 percent of the net primary production. In the nutrient-poor site,
the primary production allocated to fine roots in a 40-year old Douglas fir stand was 560 g C
m-2 yr-1, which amounted to 36.4 percent of the net primary production. It is currently
hypothesized that the annual fine root production dies off and becomes root litter. Combining
the concept of physical protection from decomposition of relatively labile compounds with
the fact that both the relative and absolute amounts of carbon entering the soil in root litter
can be larger in nutrient-poor sites than in nutrient-rich sites, it is logical to suspect that plant
nutrient status acting on the shootroot ratio switch may affect the respirationnet soil
increment switch in the soils (cf Figure 6.7). By increasing the amount of carbon input in
litter from above-ground entry to below-ground entry, the rate of carbon storage at depth in
soils may increase.

6.3.5 Summary

In this section we have discussed the ways in which plant nutrient status can influence the
amounts of carbon flowing through an ecosystem, the pathways of flow, and the proximate
fate of the carbon. We envision nutrients acting on a series of switches in the carbon flow
pathway. It is not surprising that our arguments suggest that the net primary production per
unit area in forests growing on nutrient-rich sites is greater than that on nutrient-poor sites.
More interestingly, on nutrient-rich sites as compared with nutrient-poor sites, a greater
percentage of the net primary production is allocated to plant growth increment. Carbon
accumulations in soils of nutrient-poor sites may be larger per unit of carbon in net primary
production because of the increased fraction of net primary production that is allocated to fine
roots and eventually fine root litter.
These considerations raise a series of interesting questions at the global scale. Will chronic
low level additions of fertilizer to forests, such as may be occurring through air pollution and
subsequent rain-out, shift the pathways of carbon flux in these ecosystems? Is the fertilization
causing not only an increase in net primary production, but an increase in the amount of net
primary production that is stored in the vegetation? If we define the efficiency of carbon
storage in vegetation as the annual plant growth increment divided by annual net primary
production, the fertilizer studies just reviewed would suggest that an improvement in plant
nutrient status results in increased carbon storage efficiency. This result is consistent with the
sourcesink concept of photosynthate use discussed earlier.

6.4 ELEMENT INTERACTIONS AND THE GLOBAL CARBON CYCLE

The global carbon cycle is linked to other element cycles in many complex ways. For
example, the burning of fossil fuels not only releases large amounts of carbon to the
atmosphere, but it also increases the input of nitrogen, phosphorus, and sulphur to the
atmosphere. Some of this nitrogen, phosphorus, and sulphur may enter terrestrial ecosystems
in bulk precipitation, resulting in an increase in nutrient availability in these systems. This
rise in available nutrients may in turn stimulate both carbon fixation and storage in terrestrial
ecosystems, as has been suggested by Deevey (1970) and Simpson et al. (1977), thereby
altering the dynamics of the global carbon cycle. In this section of the paper we consider how
man's alteration of the global cycles of nitrogen, sulphur, and phosphorus could potentially
increase the carbon storage capacity of the world's forests and thereby alter the global carbon
balance.

6.4.1 The Global Carbon Cycle: An Overview

The carbon dioxide concentration of the atmosphere has risen at least 40 ppm since the
beginning of the industrial revolution. The present annual rate of increase is between 1.0 and
1.5 ppm, translating to an increase in the atmospheric carbon load of about 2700 Tg of carbon
per year.

The annual carbon budget for the atmosphere can be formally stated as a mass balance
equation (Equation 1) with two source terms and two sink terms. The two source terms are:
(1) the release of carbon dioxide to the atmosphere from the combustion of fossil fuels (FFR
in Equation 1); and (2) the net release of carbon dioxide to the atmosphere resulting from
land use changes (e.g., forest to cultivated field, pasture to forest) and the oxidation of
harvested renewable resources (TR in Equation 1). The two sink terms are: (1) the net uptake
of carbon from the atmosphere by a variety of processes in the world's oceans (OU in
Equation 1); and (2) the net uptake of carbon from the atmosphere by terrestrial ecosystems
due to `fertilization' of these systems with carbon dioxide or nutrients, or both (TU in
Equation 1). In the mass balance equation, AI is the annual increase in the carbon stock of the
atmosphere.

AI = FFR + TR OU TU (1)

Table 6.4 contains the current best estimates of the variables in Equation 1. The magnitude of
the release of carbon dioxide to the atmosphere from the combustion of fossil fuels is well
known; This process is now estimated to release about 5200 Tg of carbon annually (Rotty,
1981). Estimates have ranged over an order of magnitude for the net release of carbon
dioxide to the atmosphere from terrestrial vegetation and soils as a consequence of land use
changes and the oxidation of harvested renewable resources (Table 6.5). To date, the most
systematic analysis of the terrestrial source term has been conducted by Houghton et al
(1983). According to their `population' based estimate, the magnitude of this source term is
currently 2600 Tg of carbon annually. The net annual uptake of carbon dioxide from the
atmosphere by the world's oceans is now estimated to be 2100 Tg of carbon (40% of the
amount of carbon released from the combustion of fossil fuel).

With four of the five terms in the mass balance equation defined, we can calculate TU, the
net annual storage of carbon by terrestrial ecosystems due to `fertilization' of these systems
with carbon dioxide or nutrients, or both. By difference, the `fertilization' factor required to
balance the global carbon budget is 3000 Tg of carbon.

A central question in the global carbon cycle is: Will carbon storage on land increase in
response to man's acceleration of the global cycles of N, S, and P? We will assume that N, S,
and P added to forests in bulk precipitation act as fertilizer. In this section of the paper we
attempt to evaluate the consequences of such a fertilization.

We will use a simple element-matching approach to evaluate the expected increase in carbon
storage in a terrestrial ecosystem following the addition of a given amount of a plant nutrient.
Two assumptions are made in an element-matching analysis: (1) carbon will be stored in
some constant proportion to nitrogen, sulphur, and phosphorus; and (2) the concept of a
single limiting nutrient for plant growth and organic matter decomposition is valid.

Table 6.4 Components of the atmospheric carbon budget for 1980

Amount
(Tg
Symbol Commments
C/yr)
Source of Estimate Rotty (1981). The
Inputs to atmosphere FFR 5200
error
associated with this estimate is
Fossil fuel CO2
probably
12-15% (Keeling, 1973).
Net CO2 flux due to forest Population-based estimate from
TR 2600
cutback, etc. Houghton et
al. (1983). The range for net release
from the
biosphere reported by Houghton et al.
(1982)
is 1,8004,700 T g C/yr in 1980.
Outputs from atmosphere
Estimate calculated as 40% of the
CO2 uptake by oceans OU 2100
amount
released from fossil fuels (Broecker et
al.
1979). The error associated with this
estimate
is probably 2025% (Broecker et al.,
1979).
CO2 taken up and stored in
plants and
organic residues TU 3000 Calculated by difference
Estimated from measurements made
Accumulation in atmosphere 2700
by
Bacastow and Keeling (1981). The
error
associated with this estimate is
probably less
than 23% (Bacastow and Keeling,
1981).
Table 6.5 Estimates of annual net carbon exchange between terrestrial ecosystems and the
atmosphere in or about 1980. Positive values indicate net terrestrial releases to the
atmosphere

Author Amount (Tg C yr-1)

Adams et al. (1978) 400 to 4,000


Bolin (1977) 400 to 1,600
Wong (1978) 1,900
Hampicke (1979) 1,500 to 4,500
Woodwell et al. (1978) 2,000 to 18,000
Moore et al. (1981) 2,200 to 4,700
Houghton et al. (1982) 2,600 population-based estimate
(range 1,8004,700)

6.4.2 Element Composition of Forest Ecosystems

The mean element composition (by weight) of the three major components of forest
ecosystemsvegetation, surface litter (i.e., forest floor) and mineral soilare listed in Table 6.6.
In computing the element composition of the forest components, only the organic fractions of
the various elements were used. Table 6.6 also contains the carbon to element ratios, which
are greatest in the vegetation and smallest in the soils.

Table 6.6 Element ratios (by weight) in the three major components of forest ecosystems.
Values for vegetation are for woody tissue only. They were derived from a broad survey of
the literature, with heavy reliance on Rodin and Bazilevish (1967)

Element Ratio
System
component C N S P C:N C:So C:Po

Vegetation 1500 10 1 1 150 1500 1500


Litter 500 10 1 1 50 500 500
Soil 120 9 .4 1 1 13 120 120

These ratios are the heart of the element-matching analysis. For example, if one unit of
nitrogen is added to a forest ecosystem and it is stored in the vegetation component which has
a C:N ratio of 150:1, we would predict the storage of an additional 150 units of carbon in the
vegetation component of the forest ecosystem.

Although element matching is a simple form of element interactions analysis, it can lead to
significant insights (e.g., Redfield, 1958). With respect to the analysis that follows, the power
of the element matching technique is in its implication of constraints on the increase in the
ability of forest ecosystems to act as sites of carbon storage, as a consequence of inadvertant
eutrophication.

6.4.3 Fossil Fuel Combustion and the Acceleration of the N, P, and S Cycles

The activities of industrial man are accelerating the global cycles of nitrogen, sulphur, and
phosphorus. Through the burning of fossil fuels, the industrial fixation of nitrogen, and the
mining of phosphate, man is increasing the inputs of nitrogen, sulphur, and phosphorus to the
land. Table 6.7 gives estimates of the amounts of C, N, S, and P introduced to the biosphere
annually through man's activities. A major portion of the industrially fixed nitrogen and
mined phosphorus is applied directly to agricultural fields. Very little N and P fertilizer is
added directly to forests. Fossil fuel emissions of nitrogen, phosphorus and sulphur are
dispersed much more widely over the biosphere, and some fraction of the fossil fuel N, S, and
P finds its way into forest ecosystems.

Table 6.7 Element loading of the biosphere by industrial activity (in Tg) (modified from
Peterson, 1981)

C N S P

Fossil fuel emissions 6000 24 65 1*


Industrial nitrogen
50
fixation
Phosphate mining 10

*All values rounded to nearest whole number.,

The forests of the world cover a land area of approximately 57 x 1012 m2. If, each year, all of
the fossil fuel N, S, and P were to be evenly distributed over the world's forests, the input
rates would be 0.4 g N m-2 yr-1, 1.1 g S m-2,and0.02 g P m-2 yr-1.These values are very similar
to the N, S, and P bulk precipitation input values reported for a variety of forests around the
world (see Table 6.8).The mean annual bulk precipitation input values reported in Table 6.8
for N, S and P are 0.6 g m-2, 1.1 g m-2,and 0.02 g m-2 respectively. Certainly, fossil fuel
burning is not the sole source of the N, S, and P found in bulk precipitation; moreover, fossil
fuel N, S, and P are not uniformly distributed. For example, nitrogen input at the Solling
beech forest (2.26 g N m-2 yr-1), which is in the heart of one of the industrial regions of West
Germany, is approximately eleven times higher than nitrogen input at the H. J. Andrews
Douglas fir forest (0.2 g N m-2 yr-1), which is in a rural, heavily forested region of the United
States.
Table 6.8 Nutrient budgets for various terrestrial ecosystems of the world (g m-2 yr-1)
(Modified from Likens et al.,1977)

NITROGEN
Stream-Net
Location Precipitation Net gain
water gain or
input output loss Input
Temperate: mostly
angiosperm
Coshocton, OH,
and deciduous forest 2.00 0.25 +1.75
U.S.A.
Hubbard Brook,
0.65 0.40 +0.25
U.S.A.
S.E., U.S.A. 0.20 0.10 +0.10
Silverstream, New
0.22 0.18 +0.04
Zealand
Taughannock Creek,
0.97 0.56 +0.41
NY, U.S.A.
Walker Branch, TE,
0.87 0.18 +0.69
U.S.A.
Soiling, W.
2.26 0.60 +1.66
Germany
Mean
0.55
Temperate: mostly
coniferous
Birkenes Watershed,
and evergreen forest 1.45 0.22 +1.23
Norway
Carnation Creek,
Vancouver
Island, Canada 0.27 0.11 +0.16
Cedar River, WA,
0.11 0.06 +0.05
U.S.A.
ELA, Ontario,
0.64 0.09 +0.55
Canada
Finland 0.60 0.20 +0.40
Storsj6n, Sweden 1.00 0.23 +0.77
Velen, Sweden 0.59 0.04 +0.55
Western Cascades
Range,
OR, U.S.A. 0.25 0.12 +0.13
H. J. Andrews 0.20 0.17 +0.03
Coweeta 0.55 0.02 +0.53
Mean
0.71
Tropical: angiosperm
0.16
mostly
evergreen forest Rio Negro, Brazil 0.56 0.47 +0.09
Mean 0.63
PHOSPHORUS
Temperate: mostly
angiosperm
Coshocton, OH,
and deciduous forest 0.018 0.005 +0.013
U.S.A.
Hubbard Brook,
0.0036 0.0019 +0.0017
U.S.A.
Pago Catchment,
0.033 0.026 +0.007
Australia
Silverstream, New
0.020 0.003 +0.017
Zealand
Taughannock Creek,
0.007 0.020 -0.013
NY, U.S.A.
Walker Branch, TE,
0.054 0.002 +0.052
U.S.A.
Temperate: mostly
coniferous
Blue Range
and evergreen forest 0.039 0.042 -0.003
Catchment, Australia
Boundary Waters
Canoe Area,
MN, U.S.A. (24) 0.014 0.0015 +0.013
Carnation Creek,
Vancouver
Island, Canada 0.011 0.005 +0.006
Cedar River, WA,
0.002
U.S.A.
Clear Lake, Ontario,
0.035 0.009 +0.026
Canada
ELA, Ontario,
0.032 0.005 +0.027
Canada
Finland 0.001 0.003 -0.002
Storsjön, Sweden 0.014 0.002 +0.012
Western Cascades
Range,
OR, U.S.A. 0.029 0.051 -0.022
Tropical: angiosperm
mostly
evergreen forest Rio Negro, Brazil 0.020 0.010 +0.010
Mean 0.023
SULPHATE-SULPHUR
Temperate: mostly
angiosperm
Hubbard Brook,
and deciduous forest 1.88 1.76 +0.12
U.S.A.
S.E., U.S.A. 0.80 0.70 +0.10
Silverstream, New
0.70 1.30 -0.60
Zealand
Taughannock
2.10 3.80 -1.70
Creek, NY, U.S.A.
Walker Branch, TE,
1.88 1.13 +0.75
U.S.A.
Temperate: mostly
coniferous
Birkenes Watershed,
and evergreen forest 1.56 2.69 -1.13
Norway
Carnation Creek,
Vancouver
Island, Canada 0.87 2.80 -1.93
ELA, Ontario,
0.30 0.32 -0.02
Canada
Finland 0.14 0.47 -0.33
Storsjön, Sweden 1.15 2.53 -1.38
Velen, Sweden 1.03 0.94 +0.09
Mean 1.13

At present we lack a detailed understanding of what fraction of the bulk precipitation inputs
of N, S, and P to various forests result from the burning of fossil fuels. For illustrative
purposes, we will begin our element-matching analysis by assuming that 25% of the fossil
fuel N, S, and P is evenly distributed over the world's forests. This assumption results in
annual nutrient inputs of the following magnitudes to the world's forests: 6.00 Tg N, 16.25 Tg
S and 0.25 Tg P.

6.4.4 Carbon-Nitrogen Interactions

Nitrogen is frequently a limiting factor for forest growth. The application of nitrogen
fertilizers to forest stands often results in dramatic increases in carbon storage in forests (e.g.,
Miller et al., 1976). The amount of carbon stored per unit of nitrogen applied depends
ultimately on how much of the nitrogen remains in the forest system and how the nitrogen is
distributed among the forest's three components: vegetation, litter, and soil (section 6.3).

The maximum carbon storage in forests would occur if all added nitrogen remained in the
system and if all of the added nitrogen were stored in the vegetation, since it has the largest
carbon to nitrogen ratio of the three forest components. Under these conditions, addition of
the 6 Tg N fixed each year through fossil fuel combustion would result in the storage of 900
Tg C in trees.

But not all of the nitrogen entering a forest in bulk precipitation will be retained; some of it
will be leached from the system and some, generally a much smaller amount, will be lost
through denitrification. To estimate the efficiency with which forests retain nitrogen entering
them in bulk precipitation, let us turn again to Table 6.8. A word of caution is warranted at
this point. The measurement of nitrogen entering a forest ecosystem is relatively
straightforward. Measurements of gaseous losses of nitrogen are very few and difficult to
interpret (Melillo et al., 1983) and measurements of solution losses of nitrogen from forests
are difficult to make. Data on leaching losses are often scanty and derived from systems with
poorly defined hydrology. Therefore, the efficiency calculations we derive must be
considered as suggestive, not definitive.

Nitrogen retention efficiencies for forest systems considered in Table 6.8 were calculated by
dividing the net gain of nitrogen by the input of nitrogen. The mean retention efficiency value
for all forest systems described in Table 6.8 is about 60%. Using the 60% retention
efficiency, we can recalculate the estimated carbon storage per unit nitrogen due to forest
fertilization. We estimate a nitrogen-fertilizer induced carbon storage in the world's forests of
540 Tg C/yr given the following assumptions: 1) of the 6 Tg N entering the world's forests,
60% or 3.6 Tg N is retained; 2) all of the added nitrogen is stored in the vegetation; and 3) the
storage of one unit of nitrogen results in the storage of 150 units of carbon.

When a forest is fertilized, the total amount of fertilizer retained is more than actually ends up
in the vegetation. Some will end up in the litter (surface organic horizon), and some will end
up in the soil (Mead and Pritchett, 1975a, b). We therefore needed a set of fractionation
factors that will allow us to distribute the fertilizer among forest components. We chose to
use the relative distribution of carbon within the system, reasoning that this distribution
integrates a variety of plant and soil processes responsible for element distribution. We have
not included the more subtle components of fractionation analysis discussed in section 6.3.
Since the major forest types of the world have distinctly different carbon distributions, we
considered it necessary to regionalize our analysis.

Table 6.9 gives mean values for the absolute and relative distributions of carbon in the
world's four major woody vegetation types, and the total area of each type. These relative
distributions are then used in conjunction with the carbon to nitrogen ratios of the three
components and the nitrogen input per unit area to estimate carbon storage per unit nitrogen
due to forest `fertilization'. Assuming that 6 Tg N is uniformly distributed across the world's
forests with 60% retained in the forests in some organic form (an effective nitrogen fertilizer
addition to the world's forests of 3.6 Tg N or 0.06 g N m-2) the net rate of carbon storage in
the world's forests due to nitrogen fertilization is 320 Tg C.

It is clear that the inadvertent nitrogen enrichment of forests is patchy; that is, some systems
receive high nitrogen inputs while others receive low inputs. In an attempt to improve our
estimate of increased carbon storage resulting from inadvertent nitrogen fertilization, we
weighted the nitrogen inputs to the world's four major forest types. We again used a nitrogen
input to the world's forests of 6 Tg. Given that much of the fossil fuel use occurs in close
proximity to temperate forests, they were allotted an annual input of 4 Tg N or 0.3 g N m-2.
The annual nitrogen input for the boreal forest was set at 1 Tg or 0.08 g N m-2 and that for the
tropical forest was set at 0.5 Tg N each year or 0.02 g N m-2. Like the tropical forest, the
annual nitrogen input for the woodlands and shrublands was set at 0.5 Tg, resulting in an
enrichment rate of 0.06 g N m-2 yr-1. And once again we used a 60% retention efficiency for
all systems and the fractionation factors developed in Table 6.8 to allocate nitrogen and new
carbon among the vegetation, litter, and soil of each of the four forest types. Based on these
assumptions, the annual rate of fertilizer-induced carbon storage in the world's forests was
about 310 Tg (Table 6.10).

In this analysis, the temperate forests accumulate the largest amount of carbon and the
tropical forests accumulate the least amount of carbon as a consequence of inadvertant
nitrogen fertilization. The rates of annual carbon increment are small relative to the standing
stocks of carbon in forests. This is most obvious for tropical forests, where an annual carbon
increment of 1.3 g m-2 represents a change in carbon stocks of about 0.0041% or 1 part in
24,000. Such a small change in a large store of carbon will be undetectable in field
measurements for many decades.
Table 6.9 Absolute and relative distributions of carbon in the world's forests, and estimates
of forest area

Woodland and
Forest Tropics Temperate Boreal shrubland
components
Absolute* RelativeAbsolute*RelativeAbsolute*RelativeAbsolute*Relative
a
Vegetation 20,969 0.665 16,042 0.552 10,000 0.372 6,000 0.455
Littera 147 0.005 1,208 0.042 2.000 0.074 282 0.021
Soilb 10,400 0.330 11,800 0.406 14,500 0.554 6.900 0,523
Total 31,516 1.000 29,050 1.000 26,500 1.000 13,182 1.000
Forest
24.5 12 12 8.5
Areat†
*g C m-2
†1012m2.
a
Whittaker and Likens (1973).
b
Schlesinger (1977).

Table 6.10 Potential carbon storage in the world's forests as a consequence of nitrogen
loading from fossil fuel combustion (assuming all nitrogen entering forests is stored in
organic form see text for details of calculations)

Ratio of
Carbon storage Ecosystem ecosystem
mean C C density
2
C/region C/m density to annual
Component (Tg C) (g C m ) (g C m-2) C storage
-2

Tropics
Vegetation 29.9 1.2
Litter 0.1 Trace
Soil 1.3 0.1
Total 31.3 1.3 31,500 24,200
Temperate
Vegetation 198.7 16.6
Litter 0.5 Trace
Soil 12.6 1.1
Total 211.8 17.7 29,100 1,600
Boreal
Vegetation 33.5 2.8
Litter 2.2 0.2
Soil 4.3 0.4
Total 40.0 3.4 26,900 7,900
Woodland and Shrubland
Vegetation 20.5 2.4
Litter 0.3 Trace
Soil 2.0 0.2
Total 22.8 2.6 13,200 5,100
Grand Total 305.9

Annual net carbon accumulation per unit area is equivalent to the ecologist's net ecosystem
production term. Several measurements of net ecosystem production (NEP) have been made
in the temperate forest regions and they are given in Table 6.11. These values represent the
total amount of net carbon storage in these ecosystems in a single year. The mean NEP for
these four sites is 200 g C m-2 yr-1. If we assume that these forests have been subjected to
nitrogen fertilization from the combustion of fossil fuels, we would have to argue that
approximately 8.8% of the NEP was due to fertilization (17.7 g C m-2 divided by 200 g C m-
2
).

Table 6.11 Net ecosystem production estimates for four temperate zone forests

Site g C m-2 yr-1 Reference

Hubbard Brook 218 Whittaker et al. (1974)


Walker Branch 160 Reichle et al. (1973)
Whittaker and Woodwell
Brookhaven 270
(1969)
H. J. Andrews 155 Grier and Logan (1977)

Mean 200

Our analysis suggests that man can have an impact on the global carbon balance by changing
the flows of other elements such as nitrogen. 'Fertilization' of forests with nitrogen from fossil
fuel sources will at most result in the storage of about 300 Tg C in these systems annually.
This amount of carbon storage, an upper estimate, is approximately 10% of the amount
needed to balance the global carbon budget as we currently understand it.

6.4.5 Phosphorus and Sulphur Inputs and Carbon Storage

Phosphorus is the one major element that must be supplied almost entirely by the parent
material of unfertilized soils because of the low atmospheric inputs. In tropical forests with
highly weathered soils, phosphorus may be a limiting nutrient for plant growth. In northern
forests, massive surface organic matter accumulations may isolate plants from unweathered
mineral soils, thereby depriving them access to phosphorus. Carbon accumulation in plants
and soils of both regions may be controlled by phosphorus availability.
Table 6.12 gives the regional summary of amounts of carbon storage, assuming that 0.25 Tg
P is uniformly distributed across the world's forests and retained in the forests in organic form
in proportion to mean carbon distribution by forest type. The net carbon storage in forests due
to P fertilization is about 220 Tg C. There is probably relatively little chance that this
potential will be realized, since most of the phosphorus that enters a forest will probably be
bound chemically in soils as compounds of Al, Fe, and Ca.

Table 6.12 Potential carbon storage in the world's forests as a consequence of phosphorus
loading from fossil fuel combustion (assumptions stated in the text)

Ratio of
Ecosystem ecosystem
Carbonstorage mean C C density to
annual C
C/region C/m2 density
storage
(g C m-
Component (Tg C) 2 (g C m-2) mass
)
Tropics
Vegetation 107.2 4.3
Litter 0.3 Trace
Soil 4.3 0.2
Total 111.8 4.5 31,500 7,000
Temperate
Vegetation 43.6 3.6
Litter 1.1 0.1
Soil 2.6 0.2
Total 47.3 3.9 29,100 7,500
Boreal
Vegetation 29.4 2.4
Litter 1.9 0.2
Soil 3.5 0.3
Total 34.8 2.9 26,900 9,300
Woodland and
Shrubland
Vegetation 25.5 3.0
Litter 0.4 Trace
Soil 2.3 0.3
Total 28.2 3.3 13,200 4,000
Grand Total 222

Sulphur has been identified as a limiting nutrient in only a few forest types. Thus while
sulphur loading of the atmosphere is great (Table 6.7), there is little chance that it will
stimulate forest growth on the global scale. And, in fact, this high sulphur loading may be
limiting forests' ability to respond to either nitrogen or phosphorus fertilization due to the
acid precipitation problem associated with sulphur loading of the atmosphere.

6.4.6 Summary
In this section we have considered the possibility that carbon storage in forest ecosystems is
stimulated by the increased inputs of N, S, and P that result from the combustion of fossil
fuels. Our most extensive analysis related elevated N inputs with increased carbon storage in
forests. From this analysis we have concluded that the maximum amount of additional carbon
storage that could be promoted by this form of N fertilization is about 300 Tg C per year. As
we currently understand the global carbon budget, the fate of 3000 Tg C that fluxes into the
atmosphere each year cannot be accounted for. If 300 Tg C per year were stored in the
world's forests as a result of inadvertent N fertilization, the amount of carbon unaccounted for
would be reduced by 10%.

Finally, we must stress that the inadvertent nitrogen fertilization of the world's forests may
not promote an annual increment of carbon storage as large as 300 Tg/yr for a variety of
reasons, several of which are listed below:

1. Other limiting factorsOur analysis has not fully considered the possibility that water
availability and the availability of nutrients such as phosphorus may limit the
effectiveness of N fertilization.
2. Toxic substancesThe N fertilization of forests that is caused by fossil fuel combustion
is accompanied by hydrogen ion loading (i.e. acid precipitation) and heavy metals
input. These factors may reduce the ability of forest ecosystems to respond to N
fertilization.

While the inadvertent N fertilization of the world's forests will probably not promote a carbon
storage as large as 300 Tg per year, an N-fertilization effect may be of sufficient magnitude
to be important in the global carbon budget. Further analysis of this problem is necessary.

6.5 PHOSPHORUS'THE MASTER ELEMENT'

As soils develop through geological time, there is an extended period during the early and
middle stages of pedogenesis when carbon accumulates. In this section of the paper, we
discuss how phosphorus availability in developing soils is thought to control this carbon
accumulation. We also explore, in a very preliminary way, how important this mechanism of
carbon accumulation is in the context of the global carbon budget.

6.5.1 Phosphorus Availability and Carbon, Nitrogen and Sulphur Accumulation in Soils

A. C. Redfield (1958), in his classic paper `The biological control of chemical factors in the
environment', hypothesized that in ocean systems, phosphorus is the master-element.
Redfield argued that through its influence on the oceanic carbon, nitrogen and sulphur cycles,
phosphorus controls the quantity of nitrate in the sea, and the partial pressure of oxygen in the
atmosphere.

T. W. Walker and his co-workers have hypothesized that in terrestrial systems, phosphorus is
also the master element, regulating the accumulation of carbon, nitrogen and organic sulphur
in soils (Walker, 1964; Williams and Walker, 1969a; Syers et al., 1970; Walker and Syers,
1976). They have shown that the total amount and chemical forms of phosphorus in an
ecosystem change predictably during soil development (Figure 6.9). The total amount of
phosphorus declines with time as leaching losses exceed inputs. In their scheme the soil's
total phosphorus pool is divided into four components: (1) easily weathered (primary)
minerals such as apatite, Ca-P; (2) available P or non-occluded P; (3) difficult-to-weather
(secondary) minerals or occluded P; and (4) organically bound P. At the initiation of soil
development, phosphorus is present in weatherable minerals. In the early and middle stages
of soil development, phosphorus is present in all four forms, with available P making up a
sizeable fraction of the total phosphorus pool. Late in soil development phosphorus is present
mainly as difficult-to-weather mineral forms or as organically bound phosphorus.

Figure 6.9 Changes in total phosphorus and phosphorus fractions during soil development.
Non-occluded P is basically plant-available P. Occluded P is highly recalcitrant secondary
mineral P and inaccessible primary mineral P. Modified from Walker and Syers (1976)

Phosphorus exerts control over nitrogen accumulation by influencing nitrogen fixation. An


ample supply of phosphorus and a scarcity of nitrogen are necessary for high rates of nitrogen
fixation (Griffith, 1978). In the very early stages of soil formation on parent material devoid
of nitrogen, nitrogen fixation is an important process. With high phosphorus availability and
increasing nitrogen availability, plant production increases and organic matter begins to
accumulate in the soil. In the middle stages of soil formation, losses of P are equalled by P
input from weathering, the N:P ratio increases and non-N-fixers compete successfully for the
N and P being mineralized from soil organic matter. The amounts of available N and P are in
the optimum range for maximum plant production and organic matter continues to
accumulate in the soil. Late in soil development, soil organic matter begins to decline. This
occurs when apatite has disappeared or fallen to such a low value that the rate of release of P
by dissolution of apatite is less than the loss of P from the system either by leaching or
conversion to unavailable forms or both (Williams and Walker, 1969b, Syers et al., 1970).
Further development of the ecosystem in climates where leaching occurs causes additional
loss of P and leads to declining levels of unavailable inorganic P and organically bound P as
well as organically bound C, N, and S (Walker and Syers, 1976).

6.5.2 Pedogenesis and The Global Carbon Cycle

Given the control exerted on C, N, and S soil stocks by P on a geological time scale, the
question arises as to the mean position of the world's soils on the Walker curve. Is the mean
pedogenic position of the world's soils in an early enough stage of development to be
accumulating carbon even though external P inputs are insignificant? And if soil C
accumulation is regulated by the rate of supply of available P through weathering, what is the
magnitude of the resulting C storage on an annual basis? Is the C storage rate large enough to
be significant in short term global carbon balance considerations?

Let us address these questions by attempting to establish the order of magnitude of carbon
accumulation that could possibly result annually if the world's soils were considered to be in
the middle stages of pedogenesis; that is, if phosphorus were not limiting the rate of C
accumulation. One such estimate can be derived from Syers et al. (1970). They studied
changes in C and organic stocks of N, S, and P along a 10,000 yr chronosequence of soils
developed on windblown sands in New Zealand. During the middle stages of pedogenesis
(years 3,00010,000) they calculated a mean carbon accumulation of 1.4 g C m-2 yr-1. If the 57
x 1012 m2 of the world's forests were accumulating C at this rate, a total of 79 Tg C would be
accumulating each year. This amount of carbon storage in soils is small in the context of the
overall global carbon budget. A storage of 79 Tg C per year in the soils of the world's forests
identifies the fate of only about 2 percent of the 3000 Tg C that is unaccounted for in the
global carbon budget as we currently understand it.

6.5.3 Concluding Statement

We consider this section of the paper important not for the specific questions it has raised, but
rather because of the general class of questions it has raised. Is it possible that the terrestrial
biosphere is on a carbon accumulation trend that is independent of the activities of man? If
yes, what are the mechanisms and the magnitude of the accumulation? Do element
interactions play an important role in controlling the rate of the accumulation? These are
questions we must address to make complete an analysis of the role of terrestrial biosphere in
the global carbon cycle.

6.6 REFERENCES

Adams, S. N., Jack, W., H., and Dickson, D. A. (1970) The growth of Sitka spruce on poorly
drained sites in northern Ireland, Forestry, 43, 125-133.

Adams, J. A. S., Mantovani, M. S. M., and Lundell, L. L. (1977) Wood versus fossil fuel as a
source of excess carbon dioxide in.the atmosphere: A preliminary report, Science, 196, 54-56.

Allison, F. E. (1973) Soil Organic Matter and Its Role in Crop Production, New York,
Elsevier Scientific.

Bacastow, R., and Keeling, C. D. (1981) Atmospheric carbon dioxide concentration and the
observed airborne fraction, in Bolin, B. (ed.) Carbon Cycle Modelling, SCOPE Report No.
16, New York, Wiley, 103-112.

Bal, D. V. (1922) Studies on the decomposition of some common green manuring plants at
different stages of growth in the black cotton soil of Central Provinces, Agron, J. India, 17,
133-155.

Barnes, R. L., and Bengtson, G. W. (1968) Some aspects of nitrogen nutrition and
metabolism in relation to fertilizer responses in southern pines, in Forest Fertilization,
Theory and Practice, Tennessee Valley Authority, National Fertilizer Development Center,
Muscle Shoals, Alabama, 58-63.

Bolin, B. (1977) Changes of land biota and their importance for the carbon cycle, Science,
196, 613-615.

Brix, H. (1971) Effects of nitrogen fertilization on photosynthesis and respiration of Douglas-


fir, For. Sci., 17, 407-414.

Brix, H., and Ebell, L. F. (1969) Effects of nitrogen fertilization on growth, leaf area, and
photosynthesis rate in Douglas-fir, For. Sci., 15, 189-196.

Brix, H., and Mitchell, A. K. (1980) Effects of thinning and nitrogen fertilization on xylem
development in Douglas-fir, Can. J. For. Res., 10,121-128.
Broecker, W. S., Takahashi, T., Simpson, H. J., and Peng, T.-H. (1979) Fate of fossil fuel
carbon dioxide and the global carbon budget, Science, 206, 409-418.

Chapin, F. S. III (1980) The mineral nutrition of wild plants, Ann. Rev. Ecol. Syst., 11, 233-
260.

Cole, D. W., and Rapp, M. (1980) Element cycling in forest ecosystems, in D. E. Reichle
(ed.) Dynamics Properties of Forest Ecosystems, International Biological Programme 23,
Cambridge, Cambridge University Press, 341-408.

Cowling, E. B., and Merrill, W. (1966) Nitrogen in wood and its role in wood deterioration,
Can. J. Bot., 44, 1539-1554.

Cromack, K. Jr. (1973) Litter production and litter decomposition in a mixed hardwood
watershed and in a white pine watershed at Coweeta Hydrologic Station, North Carolina.
Ph.D. thesis, University of Georgia, Athens, Georgia. 180 pages.

Cromack, K., and Monk, D. (1975) Litter production, decomposition, and nutrient cycling in
a mixed hardwood watershed and a white pine watershed, in Howell, F. G., Gentry, J. B., and
Smith, M. H. (eds) Mineral Cycling in Southeastern Ecosystems, U. S. Energy Research and
Development Admin. Symposium Series, CONF-740513. Washington, D.C., ERDA, 609-624.

Daubenmire, R., and Prusso, D. C. (1963) Studies on the decomposition rates of tree litter,
Ecology, 44, 589-592.

Davies, R. I., Coulson, C. B., and Lewis, D. A. (1964a) Polyphenols in plant, humus, and
soil. III. Stabilization of gelatin by polyphenol tanning, J. Soil Sci., 15, 299-309.

Davies, R. I., Coulson, C. B., and Lewis, D. A. (1964b) Polyphenols inplant, humus, and soil.
IV. Factors leading to increase in synthesis of polyphenol in leaves and their relationship to
mull and mor formation, J. Soil Sci., 15, 310-318.

Deevey, E. S. (1970) Mineral cycles, Sci. American, 223, 148-158.

DeHaan, S. (1977). Humus, its formation, its relation with the mineral part of the soil, and its
significance for soil productivity, in Soil Organic Matter Studies I, Vienna, International
Atomic Energy Agency, 21-30.

Dijkshoorn, W., and Van Wijk, A. L. (1967) The sulphur requirements of plants as evidence
by the sulphur-nitrogen ratio in organic matter. A review of published data, Plant Soil, 26,
129-157.

Edwards, N. T., and Harris, W. F. (1977) Carbon cycling in a mixed deciduous forest floor,
Ecology, 58, 431-437.

Epstein, E. (1972) Mineral Nutrition of Plants: Principles and Perspectives, New York,
Wiley.

Florence, R. G. (1965) Decline of old-growth redwood forests in relation to some soil


microbiological processes, Ecology, 46, 52-64.
Fogel, R., and Cromack, K. Jr (1977) Effects of habitat and substrate quality on Douglas fir
litter decomposition in western Oregon, Can. J. Bot., 55, 1632-1640.

Fogg, G. E. (1975) Biochemical pathways in unicellular plants, in Cooper, J. P. (ed.)


Photosynthesis and Productivity in Different Environments, Cambridge, Cambridge
University Press, 437-458.

Gloaguen, J. C., and Touffet, J. (1976) Production de litiere et apport au sol d'elements
minerauk dans quelques peuplements resineaux de BRetagne, Ann. Sci. For. (Paris), 33, 87-
107.

Gnanam, A., Habib Mohamed, A., and Seetha, R. (1980) Comparative studies on the effect of
ammonia and blue light on the regulation of photosynthetic carbon metabolism in higher
plants, in Senger, H. (ed.) The Blue Light Syndrome, New York, Springer-Verlag, 435-443.

Gosz, J. R. (1981) Nitrogen cycling in coniferous ecosystems, in Clark, F. E., and Rosswall,
T. (eds) Terrestrial Nitrogen Cycles, Ecol. Bull. (Stockholm), 33, 405-426.

Gosz, J. R., Likens, G. E., and Bormann, F. H. (1976) Organic matter and nutrient dynamics
of the forest and forest floor in the Hubbard Brook Forest, Oecologia, 22, 305-320.

Grier, C. C., and Logan, R. S. (1977) Old-growth Pseudotsuga menziesii communities of a


western Oregon watershed: Biomass distribution and production budgets, Ecol. Monogr., 47,
373-400.

Grier, C. C., Vogt, K. A., Keyes, M. R., and Edmonds, R. L. (1980) Biomass distribution and
above- and below-ground production in young and mature Abies amabilis zone ecosystems of
the Washington Cascades, Can J. For. Res., 10, 118-130.

Griffith, W. F. (1978) Effects of phosphorus and potassium on nitrogen fixation, in


Phosphorus for Agriculture: A Situation Analysis, Potash/Phosphate Institute, 80-94.

Hampicke, U. (1979) Net transfer of carbon between the land biota and the atmosphere
induced by man, in Bolin, B., Degens, E. J., Kempe, S., and Ketner, P., (eds) The Global
Carbon Cycle, SCOPE Report No. 13, New York, Wiley, 219-236.

Heilman, P. E. (1966) Change in distribution and availability of nitrogen with forest


succession on north slopes in interior Alaska, Ecology, 49, 331-339.

Heilman, P. E., and Gessel, S. P. (1963) Nitrogen requirements and the biological cycling of
nitrogen in Douglas fir in relationship to the effects of nitrogen fertilization, Plant Soil, 18,
386-402.

Hill, H. H. (1926) Decomposition of organic matter in soil, J. Agr. Res., 333, 77-99.

Houghton, R. A., Hobbie, J. E., Melillo, J. M., Moore, B., Peterson, B. J., Shaver, G. R., and
Woodwell, G. M. (1983) Changes in the carbon content of terrestrial biota and soils between
1860 and 1980: A net release of CO2 to the atmosphere, Ecol. Monogr. (in press).
Jarvis, P. G. (1975) Water transfer in plants, in de Vries, D. A., and van Alfen, N. K. (eds)
Heat and Mass Transfer in the Environment of Vegetation. Washington, D.C., Scripta Book
Co., 369-394.

Keeling, C. D. (1973) Industrial production of carbon dioxide from fossil fuels and limestone,
Tellus, 25, 174-198.

Keller, T. (1967) The influence of fertiliation on gaseous exchange of forest tree species, in
Proceedings of the Colloquium on Forest Fertilization, Jyväskylä/Finland, 65-79.

Keller, T. (1970) Gaseous exchangea good indicator of nutritional status and fertilizer
response of forest trees, Proceedings of the 6th International Colloquium on Plant Analysis
and Fertilizer Problems (ISHS), Tel Aviv, Israel, 669-678.

Keller, T. (1971) Der Einfluss der Stickstoffernährung auf den Gaswechsel der Fichte, Allg.
Forst Jagdztg., 142, 89-93.

Keyes, M. R., and Grier, C. C. (1981) Above- and below-ground net production in 40-year-
old Douglas-fir stands on high and low productivity sites, Can. J. For. Res., 11, 599-605.

Kira, T. (1975) Primary production of forests, in Cooper, J. P. (ed.) Photosynthesis and


Productivity in Different Environments. Cambridge, Cambridge University Press, 5-40.

Kozlowski, T. T. (1971) Growth and Development of Trees, Vol. I, Seed Germination,


Ontogeny, and Shoot Growth, New York, Academic Press.

Kramer, P. J. (1981) Carbon dioxide concentration, photosynthesis, and dry matter


production, BioScience, 31, 29-33.

Kramer, P. J., and Kozlowski, T. T. (1979) Physiology of Woody Plants, New York,
Academic Press.

Lamb, D. (1971) Litter decomposition and nutrient release in Pinus radiata plantations. Ph.D.
thesis, Australian National University, Canberra, Australia, v + 213 pages.

Lamb, D. (1975) Patterns of nitrogen mineralization in the forest floor of stands of Pinus
radiata on different soils, J. Ecol., 63, 615-625.

Lamb, D. (1976) Decomposition of organic matter on the forest floor of Pinus radiata
plantations. J. Soil Sci., 27, 206-217.

Lambers, H., Posthumus, F., Stulen, I., Lanting, L., van de Dijk, S. J., and Hofstra, R. (1981)
Energy metabolism of Plantago lanceolata as dependent on the supply of mineral nutrients,
Physiol. Plant., 51, 85-92.

Likens, G. E., Bormann, F. H., Pierce, R. S., Eaton, J. S., and Johnson, N. M. (1977)
Biogeochemistry of a Forested Ecosystem, New York, Springer-Verlag.

Lockett, J. L. (1937) Microbial aspects of decomposition of clover and rye plants at different
growth stages, Soil Sci., 44, 425-435.
Loveless, A. R. (1962) Further evidence to support a nutritional interpretation of
sclerophylly, Ann Bot., 26, 551-561.

Mead, D. J., and Pritchett, W. L. (1975a) Fertilizer movement in a slash pine ecosystem. I.
Uptake of N and P and N movement in the soil, Plant Soil, 43, 451-465.

Mead, D. J., and Pritchett, W. L. (1975b) Fertilizer movement in a slash pine ecosystem. II. N
distribution after two growing seasons, Plant Soil, 43, 467-478.

Melillo, J. M., Aber, J. D., and Muratore, J. F. (1982) Nitrogen and lignin control of
hardwood leaf litter decomposition dynamics, Ecology, 63, 621-626.

Melillo, J. M., Aber, J. D., Steudler, P. A., and Schimel, J. (1983) Denitrification potentials in
a successional sequence of northern hardwood forest stands, in Hallberg, R. (ed.)
Environment Biogeochemistry, Ecol. Bull. (Stockholm), 35.

Melin, E. (1930) Biological decomposition of some types of litter from North American
forests, Ecology, 11, 72-101.

Miller, H. G. (1969) Nitrogen nutrition on the sands of Culbin Forest, Morayshire, J. Sci.
Food Agric., 20, 417-419.

Miller, H. G., and Miller, J. D. (1976) Effect of nitrogen supply on net primary production in
Corsican pine, J. Appl. Ecol., 13, 249-256.

Miller, H. G., Miller, J. D., and Pauline, O. L. (1976) Effect of nitrogen supply on nutrient
uptake in Corsican Pine, J. Appl. Ecol., 13, 955-966.

Miller, J. G., Cooper, J. M., Miller, J. D., and Pauline, O. J. L. (1979) Nutrient cycles in pine
and their adaptation to poor soils, Can. J. For. Res., 9,19-26.

Monnier, G., and Jeanson, C. (1964) Studies on the stability of soil structure: influence of
moulds and soil fauna, in Hallsworth, H. G. and Crawford, D. V. (eds) Experimental
Pedology, London, Butterworths, 244-254.

Moore, B., Boone, R. D., Hobbie, J. E., Houghton, R. A., Melillo, J. M., Peterson, B. J.,
Shaver, G. R., Vörösmarty, C. J., and Woodwell, G. M. (1981) A simple model for analysis
of the role of terrestrial ecosystems in the global carbon budget, in Bolin, B. (ed.) Carbon
Cycle Modelling, SCOPE Report No. 16, New York, Wiley, 365-386.

Natr, L. (1972) Influence of mineral nutrients on photosynthesis of higher plants,


Photosynthetica, 6, 80-99.

Natr, L. (1975) Influence of mineral nutrition on photosynthesis and the use of assimilates, in
Cooper, J. P. (ed.) Photosynthesis and Productivity in Different Environments, Cambridge,
Cambridge University Press, 537-556.

Neish, A. C. (1964) Major pathways of biosynthesis of phenols, in Harbourne, J. D. (ed.)


Biochemistry of the Phenolic Compounds, London, Academic Press, 293-359.
Peterson, B. J. (1981) Perspectives on the importance of the oceanic particulate flux in the
global carbon cycle, Ocean Sci. Engn., 6, 71-108.

Raper, C. D. Jr., and Peedin, G. F. (1978) Photosyntfietic rate during steady-state growth as
influenced by carbon-dioxide concentration, Bot. Gaz., 139, 147-149.

Redfield, A. C. (1958) The biological control of chemical factors in the environment, Am.
Sci., 46, 205-221.

Reichle, D. E., Dinger, B. E., Edwards, N. T., Harris, W. F., and Sollins, P. (1973) Carbon
flow and storage in a forest ecosystem, in Woodwell, G. M., and Pecan, E. V. (ed.) Carbon in
the Biosphere. AEC Symposium Series No. 30. Springfield, Virginia, National Technical
Information Service, 345-365.

Rodin, L. E., and Bazilevish, N.I. (1967) Production and Mineral Cycling in Terrestrial
Vegetation, Edinburgh and London, Oliver and Boyd.

Rotty, R. (1981) Distribution and changes in industrial carbon dioxide production, in Papers
presented at the WMO/ICSU/UNEP Scientific Conference on Analysis and Interpretation of
Atmospheric CO2 Data, World Climate Programme 14, Geneva, Switzerland.

Safford, L. O. (1974) Effect of fertilization on biomass and nutrient content of fine roots in a
beechbirchmaple stand, Plant Soil, 40, 349-363.

Satchell, J. E., and Lowe, D. G. (1966) Selection of leaf litter by Lumbricus terrestris, in
Graff, G., and Satchell, J. E. (eds) Progress in Soil Biology, Amsterdam, North Holland, 180-
196.

Schlesinger, W. H. (1977) Carbon balance in terrestrial detritus, Ann. Rev. Ecol. Syst., 8, 51-
81.

Simpson, H. J. et al. (1977) Man and the global nitrogen cycle group report, in Stumm, W.
(ed.) Global Chemical Cycles and their Alterations by Man, Berlin, Dahlem Konferenzen,
253-274.

Siren, G. (1955) The development of spruce forest on raw humus sites in northern Finland
and its ecology, Acta For. Fenn., 62, 1-405.

Solomos, T. (1977) Cyanide resistant respiration in higher plants, Ann. Rev. Plant Physiol.,
28, 279-297.

Stachurski, A., and Zimka, J. R. (1975) Methods of studying forest ecosystems: Leaf area,
leaf production and withdrawal of nutrients from leaves of trees, Ekol. Pol., 23, 637-316.

Syers, J. K., Adams, J. A., and Walker, T. W. (1970) Accumulation of organic matter in a
chronosequence of soils developed on wind-blown sand in New Zealand, J. Soil Sci., 21, 146-
153.
Tamm, C. O. (1979) Nutrient cycling and productivity of forest ecosystems, in Leaf, A. L.
(ed.) Impact of Intensive Harvesting on Foresting Nutrient Cycling, State University of New
York, College of Environmental Science and Forestry, Syracuse, 2-21.

Thomas, J. D., and Hill, G. R. (1949) Photosynthesis under field conditions, in Franck, J., and
Loomis, W. E. (eds) Photosynthesis in Plants, Ames, Iowa, Iowa State College Press, 19-52.

Turner, J. (1977) Effects of nitrogen availability on nitrogen cycling in a Douglas-fir stand,


For. Sci., 23, 307-316.

Turner, J., Johnson, D. W., and Labert, M. (1980) Sulphur cycling in a Douglas-fir forest and
its modification by nitrogen applications, Acta Oecol., 1, 27-35.

Vitousek, P. (1982) Nutrient cycling and nutrient use efficiency. Am. Nat., 119, 553-572.

Vitousek, P., Gosz, J. R., Grier, C. C., Melillo, J. M., and Reiners, W. A. (1982) A
comparative analysis of nitrification and nitrate mobility in forest ecosystems, Ecol. Monogr.,
52, 155-177.

Waksman, S. A., and Gerretsen, F. C. (1931) Influences of temperature and moisture upon
the nature and extent of decomposition of plant residues by micro-organisms, Ecology, 12,
33-60.

Waksman, S. A., and Tenney, F. G. (1927) The composition of natural organic materials and
their decomposition in soil. II. Influence of age of plant upon the rapidity and nature of its
decomposition-rye plants, Soil Sci., 24, 317-334.

Waksman, S. A., and Tenney, F. G. (1928) Composition of natural organic materials and
their decomposition in the soil. III. The influence of nature of plant upon the rapidity of its
decomposition, Soil Sci., 26, 155-171.

Walker, T. W. (1964) The significance of phosphorus in pedogenesis, in Hallsworth, E. G.,


and Crawford, D. V. (eds) Experimental Pedology, Butterworths, London, 295-315.

Walker, T. W., and Syers, J. K. (1976) The fate of phosphorus during pedogenesis,
Geoderma, 15, 1-19.

Wareing, P. F., and Patrick, J. (1975) Sourcesink relations and the partitioning of assimilates
in the plant, in Cooper, J. P. (ed.) Photosynthesis and Productivity in Different Environments,
Cambridge, Cambridge University Press, 481-500.

Watt, R. F., and Heinselman, M. L. (1965) Foliar nitrogen and phosphorus level related to
site quality in a northern Minnesota spruce bog, Ecology, 46, 357-361.

Weetman, G. F. (1962) Nitrogen relations in a black spruce (Picea mariana mill.) stand
subject to various fertilizer and soil treatments, Pulp and Paper Res. Inst. Can., Woodl. Res.
Index 129.
Whittaker, R. H., and Likens, G. E. (1973) Carbon in the biota, in Woodwell, G. M., and
Pecan, E. V. (eds) Carbon and the Biosphere, AEC Symposium Series No. 30, Springfield,
Virginia, National Technical Information Service, 281-302.

Whittaker, R. H., and Woodwell, G. M. (1969) Structure, production, and diversity of the
oak-pine forest at Brookhaven, New York, J. Ecol., 57, 155-174.

Whittaker, R. H., Bormann, F. H., Likens, G. E., and Siccama, T. G. (1974) The Hubbard
Brook ecosystem study: Forest biomass and production, Ecol. Monogr., 44, 233-252.

Williams, J. D. H., and Walker, T. W. (1969a) Fractionation of phosphate in a maturity


sequence of New Zealand basaltic soil profiles. II, Soil Sci., 107, 213-219.

Williams, J. D. H., and Walker, T. W. (1969b) Fractionation of phosphate in a maturity


sequence of New Zealand basaltic soil profiles. I, Soil Sci., 107, 22-30.

Witkamp, M. (1966) Decomposition of leaf litter in relation to environment, microflora and


microbial respiration, Ecology, 47, 194-201.

Wong, C. S. (1978) Atmospheric input of carbon dioxide from burning wood, Science, 200,
197-200.

Woodwell, G. M., Whittaker, R. H., Reiners, W. A., Likens, G. E., Delwiche, C. C., and
Botkin, D. B. (1978) The biota and the world carbon budget, Science, 199, 141-146.

Yen, C. P., Pham, C. H., Cox, G. S., and Garrett, H. E. (1978) Soil depth and root
development patterns of Missouri black walnut and certain Taiwan hardwoods, in Eerdenand,
E. V., and Kinghorn, J. M. (eds) Root Form of Planted Trees, Can. For. Serv. Joint Report
No. 8, British Columbia Ministry of Forests, 18 pages.

Zimka, J. R., and Stachurski, A. (1976) Regulation of C and N transfer to the soil of forest
ecosystems and the rate of litter decomposition, Bull. Acad. Pol. Sci., 24, 127-132.

Back to Table of Contents

The electronic version of this publication has been prepared at


the M S Swaminathan Research Foundation, Chennai, India

SCOPE 21 -The Major Biogeochemical Cycles and Their


Interactions
7 The Effects of Deforestation on Air, Soil, and Water
P. M. VITOUSEK

Abstract
7.1 Introduction
7.2 What is Deforestation?
7.3 The Effects of Deforestation
7.3.1 Effects on Temperature, Water and Erosion
7.3.2 Effects on Carbon
7.3.3 Effects on Nitrogen
7.3.4 Effects on Phosphorus
7.3.5 Effects on Sulphur
7.3.6 Reforestation
7.4 Rates of Response to Deforestation
7.5 Interactions of C, N, P, and S
7.5.1 Element Supply
7.5.2 Ratios of Element Availability
7.5.3 Specific Effects
7.5.4 Interactions Off of the Site
7.6 Conclusions
Acknowledgements
References

ABSTRACT

Deforestation causes increased losses of carbon, nitrogen, phosphorus, and sulphur from
terrestrial ecosystems. Where deforestation is followed by conversion to other than forest
land uses, the effects of deforestation are magnified. The major causes of organic carbon
losses are harvest, burning of forest residue, accelerated decomposition, decreased production
of wood and roots, and erosion. Nitrogen and sulphur are lost by the same pathways, and
additionally by leaching of nitrate and sulphate to stream-water and ground-water and by the
anaerobic production of N- and S-containing gases. Phosphorus is lost primarily through
harvest and erosion. More than half of the C and N and somewhat less P and S can be lost in
sites where forest land is converted to other uses.

Losses of these elements following deforestation are most rapid in sites with high
decomposition rates, especially in the tropics and on fertile soils. The interactions of the C, N,
P, and S cycles affect losses of any element through nutrient limitations to biological
transformations, ratios of element availability, which cause either biological mobilization or
immobilization, and anion/anion interactions in the soil solution.

7.1 INTRODUCTION

Deforestation has provided a major focus for process-level (cf. Hesselman, 1917), budgetary
(Cole and Gessel, 1965; Likens et al., 1970), and modelling (Aber et al., 1979) studies of
ecosystem-level nutrient cycling and flux. The cycles of carbon, nitrogen, phosphorus, and
sulphur have received particular attention, in part because bicarbonate, organic anions,
sulphate, and (in disturbed forests or agricultural sites) nitrate are the most important anions
in the soil solution (Johnson and Cole, 1980). As such, their concentrations and mobilities
control the losses of cations as well as anions to stream-water and ground-water.
Additionally, the erosion of organic matter and phosphorus to streams and lakes contributes
to aquatic production and eutrophication. More recently, the need to evaluate forest
ecosystems as net sources or sinks for atmospheric CO2 (Woodwell et al., 1978; Broecker et
al., 1979), oxides of nitrogen (Crutzen and Ehhalt, 1977), and sulphur gases (Eaton et al.,
1978; Rice et al., 1981) has become apparent.

In this paper, I examine the major effects of deforestation on water flux through ecosystems
and on carbon, nitrogen, phosphorus, and sulphur transformations and losses. This
examination will necessarily be a rather general survey. I evaluate some of the processes
causing major differences between forest types in the pattern of their responses to
deforestation. Finally, I examine how interactions among the carbon, nitrogen, phosphorus,
and sulphur cycles control large scale responses to deforestation.

7.2 WHAT IS DEFORESTATION?

A wide range of forest land use practices can be termed `deforestation.' Two important
practices that strongly differ in intensity are forest clear-cutting and forest land conversion. In
clear-cutting all tree stems over some minimum diameter are cut and stem wood is removed,
and then the site is either replanted with tree seedlings, or natural revegetation is allowed to
occur. The major variants of this practice include whole-tree harvest and complete forest
removal (in which other parts of the trees in addition to just stem wood are removed) and
slash burning (in which the debris remaining on site after logging is burned).

Forest land conversion involves the removal of trees (as above) followed by the conversion of
the land to agriculture, pasture, development, or some other non-forest use. The intensive
harvesting of forest lands for fuel-wood in much of the world fits into this category. The
effects of forest land conversion (especially to agricultural use) are generally more severe
than those of clear-cutting. The shifting cultivation system which is widely practiced in the
tropics fits between these extremes. Essentially, it involves the temporary conversion of
forest to agriculture, followed by the natural re-establishment of forest cover.

7.3 THE EFFECTS OF DEFORESTATION

7.3.1 Effects on Temperature, Water and Erosion

The conversion of forest land to other uses decreases above- and below-ground biomass on a
site. Shading of the soil surface is thus decreased, and soil temperature increases (Stone,
1973; Harcombe, 1977). Additionally, plant uptake and transpiration of soil water and
mineral nutrient uptake are usually decreased for at least 2-3 years even in sites that rapidly
regrow to forests (Marks and Bormann, 1972; Gholz, 1980; Boring et al., 1981). With
reduced evapotranspiration, water flux through the soil is increased (Figure 7.1), and so
losses of nutrients through leaching to ground-water and stream-water can be increased.

Figure 7.1 The water cycle in an undisturbed forest (above) and a deforested site (below).
The width of the arrows is proportional to the amount of water following each path; the
system represented is a relatively wet forest

While the soil of a deforested site is thus on the average warmer and wetter than a forest soil,
the extremes in temperature and moisture levels are also increased. When subjected to direct
solar radiation, the upper few cm of forest floor or mineral soil can dry to moisture contents
well below those in undisturbed forest (Likens et al., 1978). Similarly, re-radiation from bare
surface soil on clear nights can cause ground-level frosts in midsummer in boreal forests (C.
O. Tamm, personal communication). Surface soils are thus subjected to extremes of heating
and cooling and wetting and drying in deforested sites.

A consequence of these changes in temperature and moisture is an increase in rates of


decomposition and nutrient mineralization in deforested sites (Dominski, 1971; Stone, 1973;
Stone et al., 1979). The forest floor decomposes rapidly (Covington, 1976; Bormann and
Likens, 1979), and without forest regeneration will eventually disappear. The combination of
increased decomposition (which consumes oxygen) and wetter soils (which slow oxygen
diffusion) may also increase the occurrence of anaerobic microsites within the soil.

In sites with even a slight slope, another consequence of deforestation is an increase in


erosion and particulate transport. The delivery of soil to stream courses is increased because:
(i) the wetter soil after deforestation is both heavier and less cohesive, and thus more subject
to both soil creep and more rapid slope failure; (ii) the decay of tree roots reduces the
cohesiveness of the soil and increases both soil creep and the probability of debris avalanches
(Swanson et al., 1981) ; and (iii) the decrease and eventual disappearance of the forest floor
alters the infiltration rate of the soil, allows raindrop impact on the mineral soil, and can thus
increase surface run-off. Once material reaches streams, the increased stream flows in
deforested sites are able to transport more and larger particulates downstream (Bormann et
al., 1974). The relationship between stream flow and particulate transport often has an
increasing exponential form, so the capacity to transport particles increases more rapidly than
increases in peak stream flows. Where deforestation leads to agricultural land use, higher
rates of erosion will be maintained indefinitely (Ritchie et al., 1974; Rapp, 1975).

7.3.2 Effects on Carbon

The carbon cycle in a natural forest and a deforested site are contrasted in Figure 7.2. The
most important consequence of deforestation is a substantial decrease (well over 50%) in
total organic carbon (above and below ground, living and dead) in a deforested site. This
decrease has several important causes:

1. The removal (by harvest) of organic carbon for wood or paper products. Most of our
knowledge about this flux is based on national statistics on the rate of carbon removal
in merchantable stems, and the information appears to be relatively good in many of
the developed countries (Armentano and Ralston, 1980). Less information is available
on the longer term fate of the organic carbon harvested. The amount of carbon in each
class of forest products (i.e., firewood, building material, paper products) and the
mean residence time before oxidation to CO2 of the material in each class is essential
to an evaluation of the importance of forest harvesting in the global CO2 budget
(Armentano and Hett, 1979). It is likely that, on the average, harvested material has a
shorter turn-over time than it would have had it not been harvested.
2. The combustion of residue left after deforestation. Where fire is used for land clearing
and conversion or as a silvicultural practice, a large amount of organic carbon is
rapidly released as C02. Substantial CO emissions also occur during such fires
(Crutzen et al., 1979; Crutzen, Chapter 3 this volume), and an unknown but probably
substantial amount of recalcitrant elemental carbon (charcoal) is produced (Seiler and
Crutzen, 1980).
3. Accelerated decomposition. After deforestation, the warmer, wetter soil conditions
accelerate the decomposition of residues left from land clearing, the forest floor
(Covington, 1976; Bormann and Likens, 1979), and soil organic carbon. Even with
immediate revegetation, considerable losses of forest floor organic carbon (up to
60%) can occur early in succession (Covington, 1976). Most of this carbon is
probably lost as CO2, although some may be incorporated into the mineral soil (at
least temporarily).
4. Lack of replacement of organic carbon. The amount of organic carbon in the soil
declines under continuous cultivation (Haas et al., 1957). This decline is partially due
to accelerated decomposition of the more labile fraction of native soil organic matter,
but another important cause is the smaller amount and greater lability of organic
matter added to the soil by crops as opposed to forests.
5. Erosion of organic carbon (often complexed with clay particles). The organic carbon
removed by erosion may be redistributed to lower-lying areas within the terrestrial
system (McCallan et al., 1980), in which case its turnover is probably little affected.
Alternatively, it may be transferred to lacustrine or marine sediments, where its turn-
over time is probably increased.

Figure 7.2 The carbon cycle in an undisturbed forest and in a deforested site shortly (23
years) after deforestation

Other losses of carbon from deforested systems include CH4 flux (probably somewhat
increased in the warmer, wetter soils of deforested sites), the leaching of dissolved organic
carbon to stream-water and ground-water and the leaching of carbonate species. Bicarbonate
is the most abundant anion in river-water (Garrels and MacKenzie, 1971) and in many soils
(Johnson et al., 1977), and bicarbonate concentrations and losses can be increased by
deforestation (Cole et al., 1975; Snyder et al., 1975). While this increase could significantly
increase cation mobility through soils (Johnson and Cole, 1980), it represents a relatively
minor flux of carbon. Losses of dissolved organic carbon can also be increased by
deforestation (Sollins and McCorison, 1981), but this is also likely to be a minor flux.

Overall, the major effect of deforestation on carbon cycling is the large decrease in the
organic carbon pool on the deforested site. Much of this decrease makes a net contribution
(from that site) to the atmosphere and subsequent pools.

This conclusion appears straightforward, and it has been clearly stated previously (Delcourt
and Harris, 1980). It is important enough, however, and it has been obscured enough (cf.
Brown et al., 1980) to warrant a more extended development. Old-growth forest ecosystems,
when averaged over sufficient space or time to include natural disturbances (White, 1979),
make little or no net contribution to atmospheric CO2. They take up in photosynthesis about
as much as they lose in plant and decomposer respiration, although any small patch of forest
is likely to be either a net source or a sink. The same is true (without the spatial
heterogeneity) of agricultural fields or pastures on land that has been in agriculture for a long
time, as long as the plant or animal harvest is consumed reasonably rapidly. The transition
between the two, though, involves the loss of most of the organic carbon in living biomass,
most of the forest floor, and much of the organic carbon in mineral soil. Some of this organic
carbon may go into wood used in construction, some may go into lacustrine or marine
sediments, and an unknown amount of partially oxidized organic carbon is retained with the
soil. The reminder is released as CO2 and makes a net contribution to atmospheric CO2.

The same argument applies to shifting cultivation systems and even to commercial forestry. If
an area of primary rain-forest is converted to a shifting cultivation mosaic, the total amount of
organic carbon in the area at equilibrium is decreased. At this equilibrium the landscape as a
whole may be neither a source nor a sink, and any fallow patch (of regrowing forest) may be
much more of a sink than it was when it was in primary forest (Lugo, 1980). The difference
in equilibrium (or mean) organic carbon between the primary forest and the shifting
cultivation system represents a loss of carbon to the landscape, however. Unless this amount
is tied up in building materials, charcoal, or aquatic sediments, it represents a loss of CO2 to
the atmosphere. Similarly, where forest harvests are more frequent than natural disturbances,
the mean organic matter content of forest land is decreased in the long run by forest
harvesting.

Viewed in this way, the tropical forests must now be a source for CO2 under almost any set of
assumptions. Temperate forests may presently be a sink for atmospheric CO2 (their total
organic carbon is apparently increasing (Armentano and Ralston, 1980). Viewed over the
longer term, however, the North American forests have lost organic carbon over the time
since European settlement.

7.3.3 Effects on Nitrogen

The nitrogen cycle in an undisturbed forest and a deforested site are contrasted in Figure 7.3.
Nitrogen cycling in undisturbed forests is relatively closed; the internal soilplantmicro-
organism cycle in most forests involves 1030 times more nitrogen than annual nitrogen inputs
or outputs (Rosswall, 1976), and annual inputs in aggrading forests generally exceed annual
outputs (Likens et al., 1977). When this soilplantmicro-organism cycle is disrupted by
deforestation, however, a large fraction of the organic nitrogen present in an ecosystem can
be lost. Deforestation interrupts the cycle both by preventing plant uptake of nitrogen and by
increasing the rate of nitrogen mineralization (Vitousek and Melillo, 1979).

The five major causes of organic carbon losses are also important in causing nitrogen losses
from deforested sites. Additionally, the leaching of nitrogen in the form of nitrate to stream-
water and ground-water can be a significant pathway of nitrogen loss. Nitrate is produced in
large amounts in some disturbed forests (Likens et al., 1970; Vitousek and Melillo, 1979),
and it is often the most important anion in agricultural soils (Nye and Tinker, 1977). The
nitrate anion is relatively mobile in the soil solution, and nitrate and associated cations are
thus easily leached through most soils. Some tropical soils could retain large amounts of
nitrate by anion adsorption, however (Kinjo and Pratt, 1971; Bartholomew, 1977).

Figure 7.3 The nitrogen cycle in an undisturbed forest and a deforested site. The system
represented in a relatively fertile site before and 23 years after deforestation. Dashed arrows
represent possible alternative pathways for losses of nitrogen from the site

There is considerable uncertainty over the most important pathways of nitrogen losses in
deforested ecosystems, especially in some tropical sites for which losses of as much as 1400
kg N/ha from the upper 30 cm of soil in the 2 years following clearing and burning have been
estimated (Nye and Greenland, 1964; Bartholomew, 1977). The nitrogen mineralized
following deforestation may: (i) be held within the site (temporarily or permanently) by
microbial immobilization, clay fixation, and other processes (Vitousek et al., 1979); (ii) be
lost to the atmosphere through ammonia volatilization, N2O production during nitrification
(Bremner and Blackmer, 1978), or denitrification to N2O or N2 (Firestone et al., 1980); or
(iii) be leached from the site as dissolved organic nitrogen or nitrate. N2O production during
nitrification is probably a small proportion of nitrification (Bremner and Blackmer, 1978), but
the relative magnitude of losses by denitrification and nitrate leaching in a range of
ecosystems is largely unknown. The presence of warmer, wetter soils in deforested sites
should increase rates of both nitrification and denitrification.

In general, deforestation causes a decrease in the amount of organic nitrogen on a site. Losses
by harvest, erosion, volatilization, and leaching are all important in at least some sites. As
discussed below, nitrate production and loss (to streams and to the atmosphere) is most rapid
in relatively fertile sites. If forest regeneration is prevented, however, nitrate production will
eventually occur on almost any site (Wiklander, 1981; Vitousek et al., 1982). The major
environmental concerns are: (i) that plant growth on deforested sites will be slowed because
of the nitrogen lost following deforestation; (ii) that the NOx released during burning or the
N2O produced during nitrification or denitrification will adversely affect atmospheric
chemistry (Crutzen, chapter 3, this volume); and (iii) that nitrate leached to stream-water and
ground-water could affect downstream ecosystems and even human health (Magee, 1977).

7.3.4 Effects on Phosphorus

The effects of deforestation on phosphorus are summarized in Figure 7.4. The overall effects
(Haas et al., 1961) are quite different from those on carbon and nitrogen for several reasons.
Phosphorus does not ordinarily undergo oxidationreduction reactions in terrestrial
ecosystems; it is present in both plants and soil as phosphate. The phosphate ion is relatively
immobile in soilsit is immobilized by micro-organisms, precipitated with calcium in
circumneutral and basic soils, and precipitated with iron and aluminium in acid soils.
Accordingly, phosphorus losses upon deforestation occur primarily in forest product removal
and erosion, although substantial leaching losses have been observed after deforestation in
sites on quartz sand parent material (Herrera, personal communication).

Phosphorus cycling in deforested sites may also be different from carbon and nitrogen in that
decomposition and gross nitrogen mineralization are rather closely linked, but phosphorus
mineralization is carried out in large part by extracellular enzymes produced by phosphorous-
requiring organisms (McGill and Cole, 1981).

Figure 7.4 The phosphorus cycle in an undisturbed forest and a deforested site

Most importantly, phosphorus has no significant gaseous form. Once it is either depleted or
converted to biologically unavailable forms on a site, phosphorus inputs from atmospheric
deposition are generally too low to maintain high levels of available phosphorus. Phosphorus
is added to the biological cycle through rock weathering, but very old, deeply leached soils
become depleted in weatherable phosphorus (Walker and Syers, 1976).
The most significant impact of deforestation on phosphorus cycling is an increase in
particulate phosphorus transport to streams and lakes, where it can drive increased production
and eutrophication (Schindler, 1977). The reduction of soil phosphorus content that
accompanies deforestation may be important in reducing the fertility of old soils, particularly
in the tropics.

Figure 7.5 The sulphur cycle in an undisturbed forest and a deforested site. The fluxes to the
atmosphere represented here are largely speculative

7.3.5 Effects on Sulphur

Some of the possible effects of deforestation on sulphur cycling and losses are summarized in
Figure 7.5. Perhaps because sulphur cycling is rarely limiting to forest growth under natural
conditions, sulphur cycling has received relatively less study than the cycles of carbon,
nigrogen, and phosphorus.

The same processes cause losses of sulphur in a deforested site as cause losses of nitrogen
(Bettany et al., 1980). Harvest and fire remove sulphur rapidly, and the erosion of organic
sulphur-containing compounds can be significant. Sulphate leaching is perhaps the most
important form of sulphur loss, but soil and stream-water sulphate concentrations decline
precipitously, in at least some deforested sites. Several mechanisms could cause this decrease
(Bormann and Likens, 1979).

The sulphur cycle in a deforested site is different from that of nitrogen in other ways as well.
Sulphate is more tightly held by anion adsorption, particularly by acid sesquioxides in older
soils (Couto et al., 1979; Johnson et al., 1980). Additionally, only reduced organic sulphur
(carbon-bonded sulphur) is mineralized with organic carbon and nitrogen at accelerated rates
following deforestation. The mineralization of ester-sulphates, which are the major form of
organic sulphur in many sites, may be uncoupled from decomposition in the same way as
phosphate mineralization is (McGill and Cole, 1981).

The effects of deforestation on the net exchange of sulphur with the atmosphere is difficult to
quantity. A large fraction of atmospheric inputs of sulphur to terrestrial ecosystems come in
the form of SO2 absorbed by foliage and to a lesser extent sulphate-containing aerosols
impacted on vegetation (Galloway and Whelpdale, 1980). Deforestation would reduce these
inputs. At the same time, more rapid decomposition rates and warmer, wetter soils could
increase the frequency of anaerobic conditions and thus increase sulphate reduction and H2S
volatilization, and possibly the production of volatile organic sulphur compounds (Trudinger,
1979). If deforestation caused the eutrophication of downstream aquatic systems, sulphate
reduction could also be enhanced in the sediments of those systems.

The overall effects of deforestation on sulphur cycling thus include a decrease in the sulphur
pool size in soils and vegetation, usually without serious consequences for the fertility of the
site. The delivery of sulphate and other soluble sulphur compounds to aquatic systems is
relatively little affected by deforestation. However, deforestation does cause a decrease in
sulphur uptake from the atmosphere and probably an increase in gaseous losses of sulphur to
the atmosphere.
7.3.6 Reforestation

Reforestation of long term agricultural and pasture land essentially reverses the patterns
discussed above for all of the elements. The total pool of each element in organic matter in
the site increases, and losses to downstream ecosystems and to the atmosphere decrease
(Vitousek and Reiners 1975; Bormann and Likens, 1979). The rate of reforestation varies
widely in different sites (Tamm et al., 1974), and the overall rate of accumulation of all four
elements can be controlled by limited supplies of any one of the elements (as discussed
below).

The source of carbon for reforestation is atmospheric CO2. Nitrogen comes from fixed
nitrogen in the atmosphere and from biological nitrogen fixation. Nitrogen fixation often
peaks in early or mid-succession, especially in primary succession or on drastically disturbed
sites (Gorham et al., 1979). Sulphur can come from precipitation, absorption or interception
of atmospheric sulphur by regrowing vegetation, and sulphur minerals on the site.
Phosphorus comes almost entirely from minerals on the site.

7.4 RATES OF RESPONSE TO DEFORESTATION

Given enough time without forest regeneration, most ecosystems follow the patterns outlined
above. They do so, however, at vastly different rates in different sites. The patterns discussed
above thus provide useful statements of direction, but they are virtually useless for evaluating
the timing of response to deforestation in any real system.

Studies of nitrate losses from forest ecosystems following clear-cutting and other
disturbances provide a clear illustration of differences in timing of response. Some sites have
large, rapid losses of nitrate to stream-water and ground-water following cutting; others
respond little if at all to an essentially similar treatment (Vitousek and Melillo, 1979). In
experimental studies with trenched plots in which plant regrowth was prevented, elevated
nitrate losses below the rooting zone were eventually observed in 18 out of 19 sites (Vitousek
et al., 1982). Some responded very rapidly, however, while others had delays of up to 3 years
before elevated nitrate losses to below the rooting zone occurred. Sites which never show
elevated nitrate losses may yet lose nitrate; delays of more than 10 years have been observed
in commercial clear-cuts in Sweden (Wiklander, 1981). Similarly, large amounts of carbon
and nitrogen disappear from surface soils immediately after land clearing and cultivation in
West Africa (Nye and Greenland, 1964), but it takes decades for changes of similar
magnitude to occur following cultivation in the mid-western United States (Haas et al.,
1957).

What are the major causes of differences among sites in the rate of response to deforestation?
Any answer to this question is necessarily partly speculativesome of the important differences
have not been studied directly. The differences that are likely to prove important, however,
include the following.

1. The rate of decomposition. The rate of leaf litter decomposition in undisturbed forests
can be predicted reasonably well from calculated actual evapotranspiration
(Meentemeyer, 1978). Decomposition rates are high (the exponential decay constant
K 4) in tropical forests (Nye, 1961; BernhardReversat, 1977) and much lower (K 0.25)
in some boreal forests (Berg and Staff, 1980), and the breakdown of labile soil
organic matter should also be more rapid in tropical sites. Litter inputs to the forest
floor vary less among forests from different latitudes than do decomposition rates, so
more dead organic matter is accumulated in boreal sites than in tropical sites (Jordan,
1971; Schlesinger, 1977). The much more rapid decomposition rates and relatively
small organic matter pools in tropical forests cause much more rapid losses of a
substantial portion of a site's organic carbon after deforestation (Nye and Greenland,
1964). The loss of elements associated with organic carbon (nitrogen and carbon-
bonded sulphur) may lag behind carbon losses, but they too should be lost more
rapidly where decomposition rates are higher.
2. Site fertility. Both the maximum rate of nitrate losses from experimental trenched
plots and length of delays in those losses are strongly correlated with site fertility in a
range of temperate coniferous and deciduous forests (Vitousek et al., 1979). Litter
with a much wider carbon: nitrogen ratio (and thus with a much greater capacity for
microbial immobilization of nitrogen) is produced in infertile sites. While litter in
tropical sites appears to be generally nitrogen-rich and coniferous forests generally
nitrogen-poor (Figure 7.6), all of the biomes include sites representative of the `fertile'
and `infertile' patterns of nutrient circulation (Herrera and Jordan, 1981).
Additionally, the net mineralizability of the organic nitrogen produced is less in
infertile sites than in fertile sites (Vitousek et al., 1982). The loss of carbon may not
be too dissimilar between a nitrogen-rich site and a nitrogen-poor site following
deforestation, but nitrogen losses are lower and long-delayed in the nitrogen-poor site.
3. Climate and soils. The rate and pathway of element losses varies with the amount of
water entering and percolating through the soil. Soil texture also exerts an important
effect on soil water retention, aeration, and erosion, thus affecting the pathways of
element loss in deforested systems.

Soil age and minerology affect element losses somewhat more subtly. The old, deeply
leached soils of the tropics and sub-tropics have substantial anion adsorption
capacities, and can retain phosphate and sulphate quite effectively. In some sites,
anion adsorption is sufficient to hold even the relatively weakly adsorbed anions
nitrate and chloride against leaching (Singh and Kanehiro, 1969; Kinjo and Pratt,
1971). The retention of sulphate and especially nitrate in the subsoil would reduce the
impacts of deforestation on downstream ecosystems.

4. Slope. The greatest impact on downstream ecosystems caused by deforestation


probably comes about through erosion. Although erosion can be important even on
relatively gentle slopes (Pimentel et al., 1976; Juo and Lal, 1979), its greatest impact
is on steep, unconsolidated slopes.

7.5 INTERACTIONS OF C, N, P, AND S

Deforestation affects the interactions of the carbon, nitrogen, phosphorus, and sulphur cycles
on three levelselement supply, ratios of element availability, and specific interactions.

7.5.1 Element Supply

The most important interaction in the cycling and loss of carbon, nitrogen, phosphorus, and
sulphur in deforested ecosystems is simply that many of the major transformations of these
elements are biological transformations. As such, they are carried out by organisms which
require relatively large amounts of each element to grow and carry out any transformation,
and a low supply of any of these elements can inhibit transformations for all of them.
Figure 7.6 Nitrogen circulation in litterfall and the dry weight:N ratio of that litterfall in a
range of forest ecosystems. `T' represents tropical forests, `D' deciduous forests, `N'
temperate symbiotic N-fixing forests, `C' coniferous forests, and `M' Mediterranean-type
ecosystems. Redrawn from Vitousek (1982)

The best-known role of element supply in affecting biological transformations is the control
of rates of carbon fixation and organic matter accumulation by the availability of nitrogen
and/or phosphorus. Numerous fertilization studies demonstrate that rates of vegetation
regrowth and reforestation are often limited by these nutrients. The supply of phosphorus can
similarly affect rates of nitrogen fixation (Gorham et al., 1979).

Other relatively well-documented examples where a transformation of one element is


inhibited by the lack of another include the suppression of decomposition by low nitrogen
supply (Gadgil and Gadgil, 1978), and the suppression of nitrogen mineralization (Jones and
Richards, 1977) and nitrification (Purchase, 1974) by low levels of available phosphorus.

7.5.2 Ratios of Element Availability

The best-studied example of the importance of the relative availability of different elements
in controlling the rate of a transformation in terrestrial ecosystems is the effect of the
carbon:nitrogen ratio on net nitrogen mineralization (Black, 1968). Where material with a
wide carbon:nitrogen ratio (such as that from the nitrogen-poor sites in Figure 7.6) is
decomposed, decomposers have an abundance of organic carbon (energy) relative to nitrogen
(protein). Accordingly, they release carbon as CO2 but retain nitrogen within their biomass,
and they generally even remove and incorporate available nitrogen from the soil (Aber and
Melillo, 1980; Berg and Staff, 1981). When the carbon:nitrogen ratio is lower, both nitrogen
and carbon are released by decomposers. The addition of easily oxidized organic carbon to
microscosms with a high rate of nitrogen release stops net nitrogen release (Johnson and
Edwards, 1979; Robertson, 1980). Thus, while gross nitrogen mineralization (the breakdown
of organic nitrogen compounds) may be closely associated with the rate of decomposition
(McGill and Cole, 1981), net nitrogen mineralization (the release of ammonium) is controlled
by the relative availability of organic carbon and nitrogen. The effect of deforestation on net
nitrogen mineralization will thus depend on the relative availability of carbon and nitrogen in
the deforested site.

A similar kind of control over rates of phosphorus and sulphur mineralization has been
proposed (Gosz et al., 1973). Alternatively, if McGill and Cole (1981) are correct that
phosphorus and ester-sulphate mineralization occurs primarily by extracellular enzymes
produced by P- or S-deficient organisms, then both gross mineralization rates and microbial
immobilization of P and S should be greatly increased in sites with excess available organic
carbon.

Another important element/element interaction could affect the total anion concentration (and
thus anion and cation leaching) in the soil solution of a deforested site. Nitrification results in
the formation of a mobile anion and hydrogen ions (Figure 7.3). The hydrogen ions produced
can replace cations on exchange sites, leading to nitrate and cation leaching (Nye and
Greenland, 1960; Likens et al., 1969). Alternatively, some of the hydrogen ions produced can
combine with bicarbonate, producing carbonic acid and CO2. In this case the total anion
concentration of the soil solution is buffered, as some of the increase in nitrate concentration
is offset by a decrease in bircarbonate concentration.

7.5.3 Specific Effects

More specific effects of one element on the cycle of another are difficult to document. It has
been suggested that elevated nitrate concentrations can inhibit sulphur-oxidizing bacteria
(Likens et al., 1970), thus decreasing sulphate losses from deforested systems. Alternative
explanations for the decline in sulphate losses observed in deforested sites are possible,
however (Bormann and Likens, 1979).

7.5.4 Interactions Off of the Site

Not all of the important element/element interactions resulting from deforestation would be
expected to occur on the deforested site. One important consequence of deforestation is its
impacts on downstream ecosystems.

Deforestation causes an increase on the erosional loss of phosphorus and organic carbon. The
phosphorus delivered to aquatic systems can drive the eutrophication of lakes, with attendant
hypolinmetic oxygen deficits. The organic carbon added could also contribute to the
development of anaerobic conditions in lake sediments. The productivity of coastal marine
waters could be increased from nitrate lost from deforested sites (Wollast, chapter 14, this
volume).

The increase in productivity and anaerobic conditions in lake sediments may cause a
somewhat increased preservation of organic carbon, representing a small net sink for
atmospheric CO2 (Broecker et al., 1979). More importantly, the increased area of
anaerobiosis will increase the rate of sulphate reduction and H2S volatilization (Trudinger,
1979), even though sulphate delivery to aquatic systems may be little affected by
deforestation.

7.6 CONCLUSIONS

Deforestation causes a substantial decrease in the total carbon and nitrogen pools in terrestrial
ecosystems. Losses of phosphorus and sulphur occur to a lesser extent. Much of the carbon
lost from forest ecosystems eventually enters the atmosphere as CO2. Deforestation is a net
source for CO2 wherever: (i) the mean organic carbon pool over the landscape for the land
use which follows deforestation is less than that in the undisturbed forest; and (ii) the sum of
the organic carbon contained in wood used in building materials, charcoal, and organic matter
in aquatic sediments is less than this decline in the terrestrial carbon pool. These conditions
are probably usually met. There is also a loss of soil structure and fertility resulting from the
loss of soil organic carbon in deforested sites.

The nitrogen lost from deforested sites is lost a little more slowly than the carbon, except in
relatively nitrogen-rich forests. Much of the loss follows oxidation to nitrate; nitrate can be
lost to the atmosphere as N2 or N2O (with an environmental impact if it is N2O) or to stream-
water or ground-water as nitrate (again with an environmental impact). Nitrogen availability
to plants is generally increased shortly after deforestation, but the continued loss of nitrogen
in deforested sites may eventually reduce the availability of nitrogen to levels below those in
the forest (Smith and Young, 1974), necessitating substantial nitrogen fertilization.
Phosphorus is lost primarily through harvest and erosion, and the P eroded from deforested
sites can adversely affect downstream ecosystems. The resulting phosphorus depletion in the
soil is likely to prove most important in old, deeply leached soils like those in much of the
tropics. Losses of sulphur to the atmosphere may be important within deforested sites or
downstream from them.

Many of the biological transformations that lead to losses of these four elements are
biological transformations, and any of these biological processes can be inhibited by an
inadequate supply of any one of these elements. Additionally, the relative availability of these
elements can control their net release or immobilization.

ACKNOWLEDGMENTS

I thank R. Herrera, W. A. Reiners and an anonymous reviewer for their helpful criticisms of
this manuscript.

7.7 REFERENCES

Aber, J. D., Botkin, D. B., and Melillo, J. M. (1979) Predicting the effects of different harvest
regimes on productivity and yield in northern hardwoods, Can. J. Forest Res., 9, 10-18.

Aber, J. D., and Melillo, J. M. (1980) Measuring the relative contributions of organic matter
and nitrogen to forest soils, Can. J. Bot., 58, 416-121.

Armentano, T. V., and Hett, J. (eds) (1979) The Role of Temperate Zone Forests in the World
Carbon CycleProblem Definition and Research Needs, U.S. D.O.E. Symposium CONF-
7903105, Springfield, Virginia, National Technical Information Service.

Armentano, T. V., and Ralston, C. W. (1980) The role of temperate zone forests in the global
carbon cycle, Can. J. Forest Res., 10, 53-60.

Bartholomew, W. V. (1977) Soil nitrogen changes in farming systems in the humid tropics, in
Ayanaba, A., and Dart, P. J. (eds) Biological Nitrogen Fixation in Farming Systems of the
Tropics, New York, Wiley, 27-42.

Berg, B., and Staff, H. (1980) Decomposition and chemical changes in Scots Pine needle
litter. II. Influence of chemical composition, in Persson, T. (ed.) Structure and Function of
Northern Coniferous Forests: An Ecosystem Study, Ecol. Bull. (Stockholm), 32, 375-390.

Berg, B., and Staff, H. (1981) Leaching, accumulation, and release of nitrogen in
decomposing forest litter, in Clark, F. E., and Rosswall, T. (eds) Nitrogen Cycling in
Terrestrial Ecosystems: Processes, Ecosystem Strategies, and Management Implications,
Ecol. Bull. (Stockholm), 33, 163-178.

Bernhard-Reversat, F. (1977) Recherches sur les variations stationelles des cycles


biogéochemiques en fórêt ombrophile de Côte d'Ivoire, Cah. ORSTOM, ser Pédol, 15, 75-
189.
Bettany, J. R., Saggar, S., and Stewart, J. W. B. (1980) Comparison of the amounts and forms
of sulphur in soil organic matter fractions after 65 years of cultivation, Soil Sci. Soc. Amer. J.,
44, 70-75.

Black, C. A. (1968) Soil-Plant Relationships, New York, Wiley.

Boring, L. R., Monk, C. D., and Swank, W. T. (1981) The role of sucessional species in
nutrient conservation on a clearcut Appalachian watershed, Ecology 62, 1244-1253..

Bormann, F. H., and Likens, G. E. (1979) Pattern and Process in a Forested Ecosystem, New
York, Springer-Verlag.

Bormann, F. H., Likens, G. E., Siccama, T. G., Pierce, R. S., and Eaton, J. S. (1974) The
export of nutrients and recovery of stable conditions following deforestation at Hubbard
Brook, Ecol. Monogr., 44, 255-277.

Bremner, J. M., and Blackmer, A. M. (1978) Nitrous oxide: emission from soils during
nitrification of fertilizer nitrogen, Science, 199, 295-296.

Broecker, W. S., Takahashi, T., Simpson, H. J., and Peng, T. H. (1979) Fate of fossil fuel
carbon dioxide and the global carbon budget, Science, 206, 409-418.

Brown, S., Lugo, A. E., and Liegel, B. (eds) (1980) The Role of Tropical Forests on the
World Carbon Cycle, U.S. D.O.E. Symposium CONF-800350, Springfield, Virginia, National
Technical Information Service.

Cole, D. W., and Gessel, S. P. (1965) Movements of elements through forest soil as
influenced by tree removal and fertilizer additions, in Youngberg, C. T. (ed.) ForestSoil
Relationships in North America, Corvallis, Oregon State University Press, 95-104.

Cole, D. W., Crane, W. J. B., and Grier, C. C. (1975) The effect of forest management
practices on water chemistry in a second-growth Douglas-fir ecosystem, in Bernier, B., and
Winget, C. F. (eds) Forest Soils and Land Management, Quebec, Les Presses de L'Université
Laval, 195-208.

Couto, W., Lathwell, D. J., and Bouldin, D. R. (1979) Sulphate sorption by two oxisols and
an alfisol of the tropics, Soil Sci., 127, 108-116.

Covington, W. W. (1976) Forest floor organic matter and nutrient content and leaf fall during
secondary succession in northern hardwoods, Doctoral Thesis, Yale University, New Haven,
Connecticut.

Crutzen, P. J. Atmospheric interactionshomogeneous gas reactions of C, N, and S containing


compounds, Chapter 3, this volume.

Crutzen, P. J., and Ehhalt, D. H. (1977) Effects of nitrogen fertilizer and combustion on the
stratospheric ozone layer, Ambio, 6, 112-117.
Crutzen, P J., Heidt, L. E., Krasnec, J. P., Pollock, W. H., and Seiler, W. (1979) Biomass
burning as a source of atmospheric gases CO, H2, N2O, NO, CH3Cl, and COS, Nature, 253-
256.

Delcourt, H. R., and Harris, W. F. (1980) Carbon budget of the south-eastern U.S. biota:
Analysis of historical change in trend from source to sink, Science, 210, 321-322.

Dominski, A. S. (1971) Nitrogen transformations in a northern-hardwood podzol on cutover


and forested sites, Doctoral Thesis, Yale University, New Haven, Connecticut.

Eaton, J. S., Likens, G. E., and Bormann, F. H. (1978) The input of gaseous and particulate
sulphur to a forest ecosystem, Tellus, 30, 546-551.

Firestone, M. K., Firestone, R. B., and Tiedje, J. M. (1980) Nitrous oxide from soil
denitrification; factors controlling its biological production, Science, 208, 749-751.

Gadgil, R. L., and Gadgil, P. D. (1978) Influence of clearfelling on decomposition of Pinus


radiata litter, N. Z. J. For. Sci., 8, 213-224.

Galloway, J. N., and Whelpdale, D. M. (1980) An atmospheric sulphur budget for eastern
North America, Atmos. Environ., 14, 409-417.

Garrels, R. M., and MacKenzie, F. T. (1971) Evolution of Sedimentary Rocks, New York,
Norton.

Gholz, H. L. (1980) Production and the role of vegetation in element cycles for the first three
years on an unburned clearcut watershed in western Oregon, Ecol. Soc. Amer. Bull., 61,149
(Abstract).

Gorham, E., Vitousek, P. M., and Reiners, W. A. (1979) The regulation of element budgets in
the course of terrestrial ecosystem succession, Rev. Ecol. Syst., 10, 53-84.

Gosz, J. R., Likens, G. E., and Bormann, F. H. (1973) Nutrient release from decomposing
leaf and branch litter in the Hubbard Brook Forest, New Hampshire, Ecol. Monogr., 43, 173-
191.

Haas, H. J., Evans, C. E., and Miles, E. F. (1957) Nitrogen and carbon changes on Great
Plains soils as influenced by cropping and soil treatments, U.S.D.A. Technical Bulletin 1164.

Haas, H. J., Grunes, D. L., and Reichman, G. A. (1961) Phosphorus changes in Great Plains
soils as influenced by cropping and manual applications, Soil Sci. Soc. Amer. Proc., 25, 214-
218.

Harcombe, P. A. (1977) Nutrient accumulation by vegetation during the first year of recovery
of a tropical forest ecosystem, in Cairns, J., Dickison, K. L., and Herricks, E. E. (eds)
Recovery and Restoration of Damaged Ecosystems, Charlottesville, Virginia, University of
Virginia Press, 347-378.

Herrera, R., and Jordan, C. F. (1981) Nitrogen cycle in a tropical Amazonian rain forest: the
caatinga of low mineral nutrient status, in Clark, F. E., and Rosswall, T. (eds) Nitrogen
Cycling in Terrestrial Ecosystems: Processes, Ecosystem Strategies, and Management
Implications, Ecol. Bull. (Stockholm), 33, 493-505.

Hesselman, H. (1917) Om vissa skogsföryngringsåtgarders inverkan på saltpeterbildningen i


marken och dess betydelse für barrskogen föryngring, Medd. Stat. skogsförkningsanst, 13-14,
923.

Johnson, D. W., and Cole, D. W. (1980) Anion mobility in soils: relevance to nutrient
transport from forest ecosystems, Env. Internat., 3, 79-90.

Johnson, D. W., Cole, D. W., Gessel, S. P., Singer, M. J., and Minden, R. B. (1977) Carbonic
acid leaching in a tropical, temperate, subalpine, and northern forest soil, Arc. Alpine Res., 9,
329-343.

Johnson, D. W., and Edwards, N. T. (1979) The effects of stem girdling on biogeochemical
cycles within a mixed deciduous forest in eastern Tennessee. II. Soil nitrogen mineralization
and nitrification rates, Oecologia, 40, 259-271.

Johnson, D. W., Hornbeck, J. W., Kelly, J. M., Swank, W. T., and Todd, D. E. (1980)
Regional patterns of soil sulphate accumulation: relevance to ecosystem sulphur budgets, in
Shriner, D. S., Richmond, C. R., and Lindberg, S. E. (eds) Atmospheric Sulphur Deposition:
Environmental Impact and Health Effects, Michigan, Ann Arbor Science, 507-520.

Jones, J. M., and Richards, B. N. (1977) Effect of reforestation on turnover of 15 N-labelled


nitrate and ammonia in relation to changes in soil microflora, Soil Biol. Biochem., 9, 383-392.

Jordan, C. F. (1971) A world pattern in plant energetics, Am. Sci., 59, 425-433.

Juo, A. S. R., and Lal, R. (1979) Nutrient profile in a tropical alfisol under conventional and
no-till systems, Soil Sci., 127, 168-173.

Kinjo, T., and Pratt, P. F. (1971) Nitrate adsorption. I. In some acid soils of Mexico and
South America, Soil Sci. Soc. Amer. Proc., 31, 722-725.

Likens, G. E., Bormann, F. H., and Johnson, N. M. (1969) Nitrification: importance to


nutrient losses from a cutover forested ecosystem, Science, 163, 1205-1206.

Likens, G. E., Bormann, F. H., Johnson, N. M., Fisher, D. W., and Pierce, R. S. (1970)
Effects of forest cutting and herbicide treatment on nutrient budgets in the Hubbard Brook
ecosystem in New Hampshire, Ecol. Monogr., 40, 23-27.

Likens, G. E., Bormann, F. H., Pierce, R. S., Eaton, J. S., and Johnson, N. M. (1977)
Biogeochemistry of a Forested Ecosystem, New York, Springer-Verlag.

Likens, G. E., Bormann, F. H., Pierce, R. S., and Reiners, W. A. (1978) Recovery of a
deforested ecosystem, Science, 199, 492-496.

Lugo, A. E. (1980) Are tropical forests sources or sinks of carbon? in Brown, S., Lugo, A. E.,
and Liegel, B. (eds) The Role of Tropical Forests on the World Carbon Cycle, U.S. D.O.E.
Symposium CONF-800350. Springfield, Virginia, National Technical Information Service, 1-
18.

McCallan, M. E., O'Learly, B. M., and Rose, C. W. (1980) Redistribution of Caesium-137 by


erosion and deposition on an Australian soil, Aus. J. Soil Res., 18, 119-128.

McGill, W. B., and Cole, C. V. (1981) Comparative aspects of organic C, N, S and P cycling
through soil organic matter during pedogenesis, Geoderma, 26, 267-286.

Magee, P. N. (1977) Nitrogen as a health hazard, Ambio, 6,123-125.

Marks, P. L., and Bormann, F. H. (1972) Revegetation following forest cutting: mechanisms
for return to steady state nutrient cycling, Science, 176, 914-915.

Meentemeyer, V. (1978) Macroclimate and lignin control of litter decomposition rates,


Ecology, 59, 465-472.

Nye, P. H. (1961) Organic matter and nutrient cycles under moist tropical forest, Plant Soil,
13, 333-345.

Nye, P. H., and Greenland D. J. (1960) The soil under shifting cultivation, Commonwealth
Bureau of Soils, Harpenden, England, Technical Bulletin No. 51.

Nye, P. H., and Greenland, D. J. (1964) Changes in the soil after clearing tropical forest,
Plant Soil, 21, 101-112.

Nye, P. H., and Tinker, P. B. (1977) Solute Movement in the SoilRoot System, Oxford,
Blackwell.

Pimentel, D., Terhune, E. C., Dyson-Hudson, R., Rochereau, S., Samis, R., Smith, E. A.,
Denman, D., Reifschneider, D., and Shepard, M. (1976) Land degradation: effects on food
and energy resources, Science, 194,149-155.

Purchase, B. S. (1974) The influence of phosphate deficiency on nitrification, Plant Soil, 41,
541-547.

Rapp, A. (1975) Soil erosion and sedimentation in Tanzania and Lesotho, Ambio, 4, 154-163.

Rice, H., Nochumson, D. H., and Hidy, G. M. (1981) Contributions of anthropogenic and
natural sources to atmospheric sulphur in parts of the United States, Atmos. Environ., 15, 1-9.

Ritchie, J. C., Spraberry, J. A., and McHenry, J. R. (1974) Estimating soil erosion from the
redistribution of fallout 137Cs, Soil Sci. Soc. Amer. Proc., 38, 137-139.

Robertson, G. P. (1980) The characterization and control of nitrification in primary and


secondary succession, Doctoral Thesis, Indiana University. Bloomington, Indiana.

Rosswall, T. (1976) The internal cycle between vegetation, micro-organisms, and soils, in
Svennson, B. H., and Söderlund, R. (eds) Nitrogen, Phosphorus, and SulphurGlobal Cycles,
SCOPE Report No. 7, Ecol. Bull. (Stockholm), 22,157-167.
Schindler, D. W. (1977) Evolution of phosphorus limitation in lakes, Science, 195, 260-262.

Schlesinger, W. H. (1977) Carbon balance in terrestrial detritus, Ann. Rev. Ecol. Syst., 8, 51-
81.

Seiler, W., and Crutzen, P. J. (1980) Estimates of gross and net fluxes of carbon between the
biosphere and the atmosphere from biomass burning, Climatic Change, 2, 207-247.

Singh, B. R., and Kanehiro, Y. (1969) Adsorption of nitrate in amorphous and kaolinitic
Hawaiian soils, Soil Sci. Soc. Amer. Proc., 29, 681-683.

Smith, S. J., and Young, L. B. (1974) Distribution of nitrogen forms in virgin and cultivated
soils, Soil Sci., 120, 354-360.

Snyder, G. G., Haupt, H. F., and Belt, G. H., Jr. (1975) Clearcutting and burning alter quality
of streamwater in northern Idaho. U.S. D.A. Forest Service Research Paper INT-168, Intermt.
Forest and Range Experimental Station.

Sollins, P., and McCorison, F. M. (1981) Changes in solution chemistry after clearcutting in
an old-growth Douglas-fir watershed, Water Resources Research, 17, 1409-1418.

Stone, E. (1973) The impact of timber harvest on soil and water, in President's Advisory
Panel on Timber and Environment Report, Washington, D.C., U.S. Government Printing
Office, 427-467.

Stone, E. L., Swank, W. T., and Hornbeck, J. W. (1979) Impacts of timber harvest and
regeneration on stream flow and soils in the eastern deciduous region, in Youngberg, C. T.
(ed.) Forest Soils and Land Use, Fort Collins, Colorado State University Press, 516-535.

Swanson, F. J., Fredriksen, R. L., and McCorison, F. M. (1981) Material transfer in a western
Oregon forested watershed, in Edmonds, R. L. (ed.) Analysis of Coniferous Forest
Ecosystems in the Western United States, Stroudsberg, Pennsylvania, Dowden, Hutcheson,
and Ross (in press).

Tamm, C. O., Holmen, H., Popovic, B., and Wiklander, G. (1974) Leaching of plant nutrients
from soils as a consequence of forestry operations, Ambio, 3, 211-221.

Trudinger, P. A. (1979) The biological sulphur cycle, in Trudinger, P. A., and Swaine, D. J.
(eds) Biogeochemical Cycling of Mineral-Forming Elements, Amsterdam, Elsevier
Publishing Company, 293-313.

Vitousek, P. M. (1982) Nutrient cycling and nutrient use efficiency, Am. Nat. 119, 553-572.

Vitousek, P. M., and Melillo, J. M. (1979) Nitrate losses from disturbed forests: patterns and
mechanisms, Forest Sci., 25, 605-619.

Vitousek, P. M., and Reiners, W. A. (1975) Ecosystem succession and nutrient retention: a
hypothesis, BioScience, 25, 376-381.
Vitousek, P. M., Gosz, J. R., Grier, C. C., Melillo, J. M., Reiners, W. A., and Todd, R. L.
(1979) Nitrate losses from disturbed ecosystems, Science, 204, 469-474.

Vitousek, P. M., Gosz, J. R., Grier, C. C., Melillo, J. M., and Reiners, W. A. (1982) A
comparative analysis of potential nitrification and nitrate mobilization in forest ecosystems,
Ecol. Monogr., 52, 155-177.

Walker, T. W., and Syers, J. K. (1976) The fate of phosphorus during pedogenesis,
Geoderma, 15, 1-19.

White, P. (1979) Pattern, process, and natural disturbance in vegetation, Bot. Rev., 45, 229-
299.

Wiklander, G. (1981) Rapporteur's comment on clearcutting, in Clark, F. E., and Rosswall, T.


(eds) Nitrogen Cycling in Terrestrial Ecosystems: Processes, Ecosystem Strategies, and
Management Implications, Ecol. Bull. (Stockholm), 33, 642-647.

Woodwell, G. M., Whittaker, R. H., Reiners, W. A., Likens, G. E., Delwiche, C. C., and
Botkin, D. B. (1978) The biota and the world carbon budget, Science, 199, 141-145.

Back to Table of Contents

The electronic version of this publication has been prepared at


the M S Swaminathan Research Foundation, Chennai, India

Das könnte Ihnen auch gefallen