Informatics & Computer    
   
Table of contents
(Prev) SohoSoil governance (Next)

Soil

This is a diagram and related photograph of soil layers from bedrock to soil.
A represents soil; B represents laterite, a regolith; C represents saprolite, a less-weathered regolith; the bottom-most layer represents bedrock
Loess field in Germany
Surface-water-gley developed in glacial till, Northern Ireland

Soil is a natural body consisting of layers (soil horizons) that are primarily composed of minerals which differ from their parent materials in their texture, structure, consistency, color, chemical, biological and other characteristics. It is the unconsolidated or loose covering of fine rock particles that covers the surface of the earth.[1] Soil is the end product of the influence of the climate (temperature, precipitation), relief (slope), organisms (flora and fauna), parent materials (original minerals), and time. In engineering terms, soil is referred to as regolith, or loose rock material that lies above the 'solid geology'.[citation needed] In horticulture, the terms 'soil' is defined as the layer that contains organic material that influences and has been influenced by plant roots and may range in depth from centimetres to many metres.

Soil is composed of particles of broken rock (parent materials) which have been altered by physical, chemical and biological processes that include weathering (disintegration) with associated erosion (movement). Soil is altered from its parent material by the interactions between the lithosphere, hydrosphere, atmosphere, and biosphere.[2] It is a mixture of mineral and organic materials in the form of solids, gases and liquids.[3][4] Soil is commonly referred to as "earth" or "dirt"; technically, the term "dirt" should be restricted to displaced soil.[5]

Soil forms a structure filled with pore spaces and can be thought of as a mixture of solids, water, and gases.[6] Accordingly, soils are often treated as a three-state system.[7] Most soils have a density between 1 and 2 g/cm³.[8] Little of the soil of planet Earth is older than the Pleistocene and none is older than the Cenozoic,[9] although fossilised soils are preserved from as far back as the Archean.[10]

Contents

Overview

Darkened topsoil and reddish subsoil layers are typical in some regions.

On a volume basis a good quality soil is one that is 45% minerals (sand, silt, clay), 25% water, 25% air, and 5% organic material, both live and dead. The mineral and organic components are considered a constant while the percentages of water and air are variable where the increase in one is balanced by the reduction in the other.

Given time, the simple mixture of sand, silt, and clay will evolve into a soil profile which consists of two or more layers called horizons that differ in one or more properties such as texture, structure, colour, porosity, consistency, and reaction.[citation needed] The horizons differ greatly in thickness and generally lack sharp boundaries. Mature soil profiles in temperate regions may include three master horizons A, B and C. The A and B horizons are called the solum or “true soil” as most of the chemical and biological activity that has formed soil takes place in those two profiles.[11]

The pore space of soil is shared by gases as well as water.[citation needed] The aeration of the soil influences the health of the soil's flora and fauna and the movement of gases into and out of the soil.

Of all the factors influencing the evolution of soil, water is the most powerful due to its involvement in the solution, erosion, transportation, and deposition of the materials of which a soil is composed. The mixture of water and dissolved and suspended materials is called the soil solution. Water is central to the solution, precipitation and leaching of minerals from the soil profile. Finally, water affects the type of vegetation that grows in a soil, which in turn affects the development of the soil profile.

The most influential factor in stabilizing soil fertility are the soil colloidal particles, clay and humus, which behave as repositories of nutrients and moisture and act to buffer the variations of soil solution ions and moisture. Their contributions to soil nutrition are out of proportion to their part of the soil. Colloids act to store nutrients that might otherwise be leached from the soil or to release those ions in response changes to soil pH.[citation needed]

The greatest influence on plant nutrition is soil pH, which is a measure of the hydrogen ion (acid-forming) soil reactivity, and is in turn a function of the soil materials, precipitation level, and plant root behavior. Soil pH strongly affects the availability of nutrients.

Most nutrients, with the exception of nitrogen, originate from minerals and are stored in organic materials both live and dead and on colloidal particles. Some nitrogen originates from rain, but most of the nitrogen available in soils is the result of nitrogen fixation by bacteria. The action of microbes on organic matter and minerals may be to free nutrients for use, sequester them, or cause their loss from the soil by their volatilisation to gases or their leaching from the soil. The nutrients may be stored on soil colloids, and live or dead organic matter, but may not be accessible to plants due to extremes of pH.

The organic material of the soil has a powerful effect on its development, fertility, and available moisture. Following water and soil colloids, organic material is next in importance to soil's formation and fertility.

History of the study of soil

Studies concerning soil fertility

The history of the study of soil is intimately tied to our urgent need to provide food for ourselves and forage for our animals. Throughout history, civilizations have prospered or declined as a function of the availability and productivity of their soils.

The Greek historian Xenophon (450–355 B.C.) is credited with being the first to expound upon the merits of green-manuring crops: "But then whatever weeds are upon the ground, being turned into earth, enrich the soil as much as dung."[12]

Columella’s "Husbandry," circa 60 A.D., was used by 15 generations (450 years) under the Roman Empire until its collapse. From the fall of Rome to the French Revolution, knowledge of soil and agriculture was passed on from parent to child and as a result, crop yields were low. During the European Dark Ages, Yahya Ibn_al-'Awwam’s handbook guided the people of North Africa, Spain and the Middle East with its emphasis on irrigation; a translation of this work was finally carried to the southwest of the United States.

Experiments into what made plants grow first led to the idea that the ash left behind when plant matter was burned was the essential element but overlooked the role of nitrogen, which is not left on the ground after combustion. In about 1635, the Flemish chemist Jan Baptist van Helmont thought he had proved water to be the essential element from his famous five years' experiment with a willow tree grown with only the addition of rainwater. His conclusion came from the fact that the increase in the plant's weight had been produced only by the addition of water, with no reduction in the soil's weight.[13] John Woodward (d. 1728) experimented with various types of water ranging from clean to muddy and found muddy water the best, and so he concluded that earthy matter was the essential element. Others concluded it was humus in the soil that passed some essence to the growing plant. Others held that the vital growth principal was something passed from dead plants or animals to the new plants. At the start of the 18th century, Jethro Tull demonstrated that it was beneficial to cultivate the soil, but his opinion that the stirring made the fine parts of soil available for plant absorption was erroneous.[13]

As chemistry developed, it was applied to the investigation of soil fertility. The French chemist Antoine Lavoisier showed in about 1778 that plants and animals must “combust” oxygen internally to live and was able to deduce that most of the 165-pound weight of van Helmont’s willow tree derived from air. It was the French agriculturalist Jean-Baptiste Boussingault who by means of experimentation obtained evidence showing that the main sources of carbon, hydrogen and oxygen for plants were the air and water. Justus von Liebig in his book Organic Chemistry in its Applications to Agriculture and Physiology (published 1840), he asserted that the chemicals in plants must have come from the soil and air and that to maintain soil fertility, the used minerals must be replaced. Liebig nevertheless believed the nitrogen was supplied from the air. The enrichment of soil with guano by the Incas was rediscovered in 1802, by Alexander von Humboldt. This led to its mining and that of Chilean nitrate and to its application to soil in the United States and Europe after 1840.

The work of Liebig was a revolution for agriculture, and so other investigators started experimentation based on it. In England John Bennet Lawes and Joseph Henry Gilbert worked in the Rothamsted Experimental Station, founded by the former, and discovered that plants took nitrogen from the soil, and that salts needed to be in an available state to be absorbed by plants. Their investigations also produced the "superphosphate", consisting in the acid treatment of phosphate rock.[13] This led to the invention and use of salts of potassium (K) and nitrogen (N) as fertilizers. Finally, the chemical basis of nutrients delivered to the soil in manure was understood and in the mid-19th century chemical fertilisers were applied. The dynamic interaction of soil and its life forms awaited discovery.

In 1856 J. T. Way discovered that ammonia contained in fertilisers was transformed into nitrates, and twenty years later R. W. Warington proved that this transformation was done by living organisms. In 1890 Sergei Winogradsky announced he had found the bacteria responsible for this transformation.[13]

It was known that certain legumes could take up nitrogen from the air and fix it to the soil. The development of bacteriology towards the end of the 19th century led to an understanding of the role played in nitrogen fixation by bacteria. The symbiosis of bacteria and leguminous roots, and the fixation of nitrogen by the bacteria, were simultaneously discovered by German agronomist Hermann Hellriegel and Dutch microbiologist Martinus Beijerinck.

Crop rotation, mechanisation, chemical and natural fertilisers led to a doubling of wheat yields in Western Europe between 1800 and 1900.[14]

Studies concerning soil formation

The scientists who studied the soil in connection with agricultural practices had considered it mainly as a static substrate. However, soil is the result of evolution from more ancient geological materials. Other scientists later began to study soil genesis and as a result also soil types and classifications.

In 1860, in Mississippi, Eugene W. Hilgard studied the relationship among rock material, climate, and vegetation, and the type of soils that were developed. He realised that the soils were dynamic, and considered soil types classification. Unfortunately his work was not continued. At the same time Vasily Dokuchaev was leading a team of soil scientists in Russia who conducted an extensive survey of soils, finding that similar basic rocks, climate and vegetation types lead to similar soil layering and types, and established the concepts for soil classifications. The work of this team was communicated to Western Europe in 1914 by a publication in German by K. D. Glinka, a member of the Russian team.

Curtis F. Marbut was influenced by the work of the Russian team, translated Glinka's publication into English, and as he was placed in charge of the U. S. National Cooperative Soil Survey, applied it to a national soil classification system.[13]

Influences on soil formation

Soil formation, or pedogenesis, is the combined effect of physical, chemical, biological and anthropogenic processes on soil parent material. Soil is said to be formed when organic matter has accumulated and colloids are washed downward, leaving deposits of clay, humus, iron oxide, carbonate, and gypsum. These constituents are moved (translocated) from one level to another by water and animal activity. As a result, layers (horizons) form in the soil profile. The alteration and movement of materials within a soil causes the formation of distinctive soil horizons.

How soil formation proceeds is influenced by at least five classic factors that are intertwined in the evolution of a soil. They are: parent material, climate, topography (relief), organisms, and time. When reordered to climate, relief, organisms, parent material, and time, they form the acronym CROPT.[15]

An example of the development of a soil would begin with the weathering of lava flow bedrock, which would produce the purely mineral-based parent material from which the soil texture forms. Soil development would proceed most rapidly from bare rock of recent flows in a warm climate, under heavy and frequent rainfall. Under such conditions, plants become established very quickly on basaltic lava, even though there is very little organic material. The plants are supported by the porous rock as it is filled with nutrient-bearing water that carries dissolved minerals from the rocks and guano. Crevasses and pockets, local topography of the rocks, would hold fine materials and harbour plant roots. The developing plant roots are associated with mycorrhizal fungi[16] that assist in breaking up the porous lava, and by these means organic matter and a finer mineral soil accumulate with time.

Parent material

The mineral material from which a soil forms is called parent material. Rock, whether its origin is igneous, sedimentary, or metamorphic, is the source of all soil mineral materials and origin of all plant nutrients with the exceptions of nitrogen, hydrogen and carbon. As the parent material is chemically and physically weathered, transported, deposited and precipitated, it is transformed into a soil.

Typical soil mineral materials are:[17]

  • Quartz: SiO2
  • Calcite: CaCO3
  • Feldspar: KAlSi3O8
  • Mica (biotite): K(Mg,Fe)3AlSi3O10(OH)2

Classification of parent material

Parent materials are classified according to how they came to be deposited. Residual materials are mineral materials that have weathered in place from primary bedrock. Transported materials are those that have been deposited by water, wind, ice or gravity. And cumulose material is organic matter that has grown and accumulates in place.
Residual soils are soils that develop from their underlying parent rocks and have the same general chemistry as those rocks. The soils found on mesas, plateaux, and plains are residual soils. In the United States as little as three percent of the soils are residual.[18]
Most soils derive from transported materials that have been moved many miles by wind, water, ice and gravity.[19]
  • Aeolian processes (movement by wind) are capable of moving silt and fine sand many hundreds of miles, forming loess soils (60–90 percent silt),[20] common in the Midwest of North America and in Central Asia. Clay is seldom moved by wind as it forms stable aggregates.
  • Water-transported materials are classed as either alluvial, lacustrine, or marine. Alluvial materials are those moved and deposited by flowing water. Sedimentary deposits settled in lakes are called lacustrine. Lake Bonneville and many soils around the Great Lakes of the United States are examples. Marine deposits, such as soils along the Atlantic and Gulf Coasts and in the Imperial Valley of California of the United States, are the beds of ancient seas that have been revealed as the land uplifted.
  • Ice moves parent material and makes deposits in the form of terminal and lateral moraines in the case of stationary glaciers. Retreating glaciers leave smoother ground moraines and in all cases, outwash plains are left as alluvial deposits are moved downstream from the glacier.
  • Parent material moved by gravity is obvious at the base of steep slopes as talus cones and is called colluvial material.
Cumulose parent material is not moved but originates from deposited organic material. This includes peat and muck soils and results from preservation of plant residues by the low oxygen content of a high water table. While peat may form sterile soils, muck soils may be very fertile.

Weathering of parent material

The weathering of parent material takes the form of physical disintegrating and chemical decomposition and transformation.
  • Physical disintegration is the first stage in the transformation of parent material into soil. The freezing of absorbed water causes the physical splitting of material along a path toward the center of the rock, while temperature gradients within the rock can cause exfoliation of "shells". Cycles of wetting and drying cause soil particles to be ground down to a finer size, as does the physical rubbing of material caused by wind, water, and gravity. Organisms also reduce parent material in size through the action of plant roots or digging on the part of animals.
  • Chemical decomposition results when minerals are made soluble by water or are changed in structure. The first three of the following list are solubility changes and the last three are structural changes.[21]
  1. The solution of salts in water results from the action of bipolar water on ionic salt compounds.
  2. Hydrolysis is the transformation of minerals into polar molecules by the splitting of the mineral and the intervening water. This results in soluble acid-base pairs. For example, the hydrolysis of orthoclase-feldspar transforms it to acid silicate clay and basic potassium hydroxide, both of which are more soluble.
  3. In carbonation, the reaction of carbon dioxide in solution with water forms carbonic acid. Carbonic acid will transform calcite into more soluble calcium bicarbonate.
  4. Hydration is the inclusion of water in a mineral structure, causing it to swell and leaving it more stressed and easily decomposed.
  5. Oxidation of a mineral compound causes it to swell and increase its oxidation number, leaving it more easily attacked by water or carbonic acid.
  6. Reduction means the oxidation number of some part of the mineral is reduced, which occurs when oxygen is scarce. The reduction of minerals leaves them electrically unstable, more soluble and internally stressed and easily decomposed.
Of the above, hydrolysis and carbonation are the most effective.
Saprolite is a particular example of a residual soil formed from the transformation of granite, metamorphic and other types of bedrock into clay minerals. Often called "weathered granite", saprolite is the result of weathering processes that include: hydrolysis, chelation from organic compounds, hydration[disambiguation needed] (the solution of minerals in water with resulting cation and anion pairs) and physical processes that include freezing and thawing.[22] The mineralogical and chemical composition of the primary bedrock material, its physical features, including grain size and degree of consolidation, and the rate and type of weathering transform the parent material into a different mineral. The texture, pH and mineral constituents of saprolite are inherited from its parent material.

Climate

Climate is the dominant factor in soil formation, and soils show the distinctive characteristics of the climate zones in which they form.[23] Mineral precipitation and temperature are the primary climatic influences on soil formation.

The direct influences of climate include:[24]

  • A shallow accumulation of lime in low rainfall areas as caliche
  • Formation of acid soils in humid areas
  • Erosion of soils on steep hillsides
  • Deposition of eroded materials downstream
  • Very intense chemical weathering, leaching, and erosion in warm and humid regions where soil does not freeze

Climate directly affects the rate of weathering and leaching. Soil is said to be formed when detectable layers of clays, organic colloids, carbonates, or soluble salts have been moved downward. Wind moves sand and smaller particles, especially in arid regions where there is little plant cover. The type and amount of precipitation influence soil formation by affecting the movement of ions and particles through the soil, and aid in the development of different soil profiles. Soil profiles are more distinct in wet and cool climates, where organic materials may accumulate, than in wet and warm climates, where organic materials are rapidly consumed. The effectiveness of water in weathering parent rock material depends on seasonal and daily temperature fluctuations. Cycles of freezing and thawing constitute an effective mechanism which breaks up rocks and other consolidated materials.

Climate also indirectly influences soil formation through the effects of vegetation cover and biological activity, which modify the rates of chemical reactions in the soil.

Topography

The topography, or relief, characterised by the inclination of the surface, determines the rate of precipitation runoff and rate of formation or erosion of the surface soil profiles. Steep slopes allow rapid runoff and erosion of the top soil profiles and little mineral deposition in lower profiles. Depressions allow the accumulation of water, minerals and organic matter and in the extreme, the resulting soils will be saline marshes or peat bogs. Intermediate topography affords the best conditions for the formation of an agriculturally productive soil.

Organisms

Soil is the most abundant ecosystem on Earth, but the vast majority of organisms in soil are microbes, a great many of which have not been described.[25][26] There may be a population limit of around one billion cells per gram of soil, but estimates of the number of species vary widely.[26][27] One estimate put the number at over a million species per gram of soil, although a later study suggests a maximum of just over 50,000 species per gram of soil.[27][28] The total number of organisms and species can vary widely according to soil type, location, and depth.[26]

Plants, animals, fungi, bacteria and humans affect soil formation (see soil biomantle and stonelayer). Animals, soil mesofauna and micro-organisms mix soils as they form burrows and pores, allowing moisture and gases to move about. In the same way, plant roots open channels in soils. Plants with deep taproots can penetrate many metres through the different soil layers to bring up nutrients from deeper in the profile. Plants with fibrous roots that spread out near the soil surface have roots that are easily decomposed, adding organic matter. Micro-organisms, including fungi and bacteria, effect chemical exchanges between roots and soil and act as a reserve of nutrients. Humans impact soil formation by removing vegetation cover with erosion as the result. Their tillage also mixes the different soil layers, restarting the soil formation process as less weathered material is mixed with the more developed upper layers.

Vegetation impacts soils in numerous ways. It can prevent erosion caused by excessive rain that results in surface runoff. Plants shade soils, keeping them cooler and slowing evaporation of soil moisture, or conversely, by way of transpiration, plants can cause soils to lose moisture. Plants can form new chemicals that can break down minerals and improve soil structure. The type and amount of vegetation depends on climate, topography, soil characteristics, and biological factors. Soil factors such as density, depth, chemistry, pH, temperature and moisture greatly affect the type of plants that can grow in a given location. Dead plants and fallen leaves and stems begin their decomposition on the surface. There, organisms feed on them and mix the organic material with the upper soil layers; these added organic compounds become part of the soil formation process.

Time

Time is a factor in the interactions of all the above. Over time, soils evolve features that are dependent on the interplay of other soil forming factors. Soil is always changing. It takes about 800 to 1000 years for a 2.5 cm (0.98 in) thick layer of fertile soil to be formed in nature.[citation needed] For example, recently deposited material from a flood exhibits no soil development because there has not been enough time for the material to form a structure that further defines soil. The original soil surface is buried, and the formation process must begin anew for this deposit. Over a period of between hundreds and thousands of years, the soil will develop a profile that depends on the intensities of biota and climate. While soil can achieve relative stability of its properties for extended periods, the soil life cycle ultimately ends in soil conditions that leave it vulnerable to erosion. Despite the inevitability of soil retrogression and degradation, most soil cycles are long.[citation needed]

Soil-forming factors continue to affect soils during their existence, even on “stable” landscapes that are long-enduring, some for millions of years. Materials are deposited on top or are blown or washed from the surface. With additions, removals and alterations, soils are always subject to new conditions. Whether these are slow or rapid changes depends on climate, topography and biological activity.

Physical properties of soils

The physical properties of soils, in order of decreasing importance, are texture, structure, density, porosity, consistency, temperature, colour and resistivity. Most of these determine the aeration of the soil and the ability of water to infiltrate and to be held in the soil. Soil texture is determined by the relative proportion of the three kinds of soil particles, called soil "separates": sand, silt, and clay. Larger soil structures called "peds" are created from the separates when iron oxides, carbonates, clay, and silica with the organic constituent humus, coat particles and cause them to adhere into larger, relatively stable secondary structures. Soil density, particularly bulk density, is a measure of soil compaction. Soil porosity consists of the part of the soil volume occupied by air and water. Soil consistency is the ability of soil to stick together. Soil temperature and colour are self-defining. Resistivity refers to the resistance to conduction of electric currents and affects the rate of corrosion of metal and concrete structures. The properties may vary through the depth of a soil profile.

Texture

Soil types by clay, silt and sand composition as used by the USDA
Iron-rich soil near Paint Pots in Kootenay National Park, Canada

The mineral components of soil, sand, silt and clay, determine a soil's texture. In the illustrated USDA textural classification triangle, the only soil that does not exhibit one of these predominately is called "loam". While even pure sand, silt or clay may be considered a soil, from the perspective of food production a loam soil with a small amount of organic material is considered ideal. The mineral constituents of a loam soil might be 40% sand, 40% silt and the balance 20% clay by weight. Soil texture affects soil behaviour, in particular its retention capacity for nutrients and water.[29]

Sand and silt are the products of physical and chemical weathering; clay, on the other hand, is a product of chemical weathering but often forms as a secondary mineral precipitated from dissolved minerals. It is the specific surface area of soil particles and the unbalanced ionic charges within them that determine their role in the cation exchange capacity of soil, and hence its fertility. Sand is least active, followed by silt; clay is the most active. Sand's greatest benefit to soil is that it resists compaction and increases porosity. Silt is mineralogically like sand but with its higher specific surface area it is more chemically active than sand. But it is the clay content, with its very high specific surface area and generally large number of negative charges, that gives a soil its high retention capacity for water and nutrients. Clay soils also resist wind and water erosion better than silty and sandy soils, as the particles are bonded to each other.

Sand is the most stable of the mineral components of soil; it consists of rock fragments, primarily quartz particles, ranging in size from 2.0 to 0.05 mm (0.079 to 0.0020 in) in diameter. Silt ranges in size from 0.05 to 0.002 mm (0.002 to 0.00008 in). Clay cannot be resolved by optical microscopes as its particles are 0.002 mm (7.9×10−5 in) or less in diameter.[30] In medium-textured soils, clay is often washed downward through the soil profile and accumulates in the subsoil.

Soil components larger than 2.0 mm (0.079 in) are classed as rock and gravel and are removed before determining the percentages of the remaining components and the texture class of the soil, but are included in the name. For example, a sandy loam soil with 20% gravel would be called gravelly sandy loam.

When the organic component of a soil is substantial, the soil is called organic soil rather than mineral soil. A soil is called organic if:

  1. Mineral fraction is 0% clay and organic matter is 20% or more
  2. Mineral fraction is 0% to 50% clay and organic matter is between 20% and 30%
  3. Mineral fraction is 50% or more clay and organic matter 30% or more.[31]

Structure

The clumping of the soil textural components of sand, silt and clay forms aggregates and the further association of those aggregates into larger units forms soil structures called peds. The adhesion of the soil textural components by organic substances, iron oxides, carbonates, clays, and silica, and the breakage of those aggregates due to expansion-contraction, freezing-thawing, and wetting-drying cycles, shape soil into distinct geometric forms. These peds evolve into units which may have various shapes, sizes and degrees of development.[32] A soil clod, however, is not a ped but rather a mass of soil that results from mechanical disturbance. The soil structure affects aeration, water movement, conduction of heat, plant root growth and resistance to erosion. Water has the strongest effect on soil structure due to its solution and precipitation of minerals and its effect on plant growth.

Soil structure often gives clues to its texture, organic matter content, biological activity, past soil evolution, human use, and the chemical and mineralogical conditions under which the soil formed. While texture is defined by the mineral component of a soil and is an innate property of the soil that does not change with agricultural activities, soil structure can be improved or destroyed by the choice and timing of farming practices.

Soil Structural Classes:[33]

1. Types: Shape and arrangement of peds
a. Platy: Peds are flattened one atop the other 1-10 mm thick.
Found in the A-horizon of forest soils and lake sedimentation.
b. Prismatic and Columnar: Prismlike peds are long in the
vertical dimension, 10-100 mm wide. Prismatic peds have flat
tops, columnar peds have rounded tops. Tend to form in the B-
horizon in high sodium soil where clay has accumulated.
c. Angular and subangular: Blocky peds are imperfect cubes,
5-50 mm, angular have sharp edges, subangular have rounded
edges. Tend to form in the B-horizon where clay has
accumulated and indicate poor water penetration.
d. Granular and Crumb: Spheroid peds of polyhedrons, 1-10 mm,
often found in the A-horizon in the presence of organic
material. Crumb peds are more porous and are considered ideal.
2.Classes: Size of peds whose ranges depend upon the above type
a. Very fine or very thin: <1 mm platy and spherical; <5 mm
blocky; <10 mm prismlike.
b. Fine or thin: 1-2 mm platy, and spherical; 5-10 mm blocky;
10-20 mm prismlike.
c. Medium: 2-5 mm platy, granular; 10-20 mm blocky; 20-50
prismlike.
d. Coarse or thick: 5-10 mm platy, granular; 20-50 mm blocky;
50-100 mm prismlike.
e. Very coarse or very thick: >10 mm platy, granular; >50 mm
blocky; >100 mm prismlike.
3. Grades: Is a measure of the degree of development or cementation within the
peds that results in their strength and stability.
a. Weak: Weak cementation allows peds to fall apart into the
three textural constituents, sand, silt and clay.
b. Moderate: Peds are not distinct in undisturbed soil but when
removed they break into aggregates, some broken aggregates and
little unaggregated material. This is considered ideal.
c. Strong:Peds are distinct before removed from the profile and
do not break apart easily.
d. Structureless: Soil is entirely cemented together in one
great mass such as slabs of clay or no cementation at all such
as with sand.

At the largest scale, the forces that shape a soil's structure result from swelling and shrinkage that initially tend to act horizontally, causing vertically oriented prismatic peds. Clayey soil, due to its differential drying rate with respect to the surface, will induce horizontal cracks, reducing columns to blocky peds. Roots, rodents, worms, and freezing-thawing cycles further break the peds into a spherical shape.

At a smaller scale, plant roots extend into voids and remove water and cause the open spaces to increase, and further decrease physical aggregation size. At the same time roots, fungal hyphae and earthworms create microscopic tunnels that break up peds.

At an even smaller scale, soil aggregation continues as bacteria and fungi exude sticky polysaccharides which bind soil into small peds. The addition of the raw organic matter that bacteria and fungi feed upon encourages the formation of this desirable soil structure.

At the lowest scale, the soil chemistry affects the aggregation or dispersal of soil particles. The clay particles contain polyvalent cations which give the faces of clay layers a net negative charge. At the same time the edges of the clay plates have a slight positive charge, thereby allowing the edges to adhere to the faces of other clay particles or to flocculate. On the other hand, when monovalent ions such as sodium invade and displace the polyvalent cations, they weaken the positive charges on the edges, while the negative surface charges are relatively strengthened. This leaves a net negative charge on the clay, causing the particles to push apart, and so prevents the flocculation of clay particles into larger assemblages. As a result, the clay disperses and settles into voids between peds, causing them to close. In this way the soil aggregation is destroyed and the soil made impenetrable to air and water. Such sodic soil tends to form columnar structures near the surface.[34]

Density

Density is the weight per unit volume of an object. Particle density is the density of only the mineral particles that make up a soil; i.e., it excludes pore space and organic material. Particle density averages approximately 2.65 g/cc (165 lbm/ft3). Soil bulk density, a dry weight, includes air space and organic materials of the soil volume. A high bulk density indicates either compaction of the soil or high sand content. The bulk density of cultivated loam is about 1.1 to 1.4 g/cc (for comparison water is 1.0 g/cc).[35] A lower bulk density by itself does not indicate suitability for plant growth due to the influence of soil texture and structure.

Representative bulk densities of soils. The percentage pore space was calculated using 2.7 g/cc for particle density except for the peat soil, which is estimated.[36]
Soil treatment and identificationBulk density g/ccPore space %
Tilled surface soil of a cotton field1.351
Trafficked inter-rows where wheels passed surface1.6737
Traffic pan at 25 cm deep1.736
Undisturbed soil below traffic pan, clay loam1.543
Rocky silt loam soil under aspen forest1.6240
Loamy sand surface soil1.543
Decomposed peat0.5565

Porosity

Pore space is that part of the bulk volume that is not occupied by either mineral or organic matter but is open space occupied by either air or water. Ideally, the total pore space should be 50% of the soil volume. The air space is needed to supply oxygen to organisms decomposing organic matter, humus, and plant roots. Pore space also allows the movement and storage of water and dissolved nutrients.

There are four categories of pores:

  1. Very fine pores: < 2 microns
  2. Fine pores: 2-20 microns
  3. Medium pores: 20-200 microns
  4. Coarse pores: 200 microns-0.2 mm

In comparison, root hairs are 8 to 12 microns in diameter. When pore space is less than 30 microns, the forces of attraction that hold water in place are greater than those acting to drain the water. At that point, soil becomes water-logged and it cannot breathe. For a growing plant, pore size is of greater importance than total pore space. A medium-textured loam provides the ideal balance of pore sizes. Having large pore spaces that allow rapid air and water movement is superior to smaller pore space but has a greater percentage pore space.[37] Tillage has the short-term benefit of temporarily increasing the number of pores of largest size, but in the end those will be degraded by the destruction of soil aggregation.[38]

Consistency

Consistency is the ability of soil to stick together and resist fragmentation. It is of rough use in predicting cultivation problems and the engineering of foundations. Consistency is measured at three moisture conditions: air-dry, moist and wet. The measures of consistency border on subjective as they employ the "feel" of the soil in those states. A soil's resistance to fragmentation and crumbling is assessed in the dry state by rubbing the sample. Its resistance to shearing forces is assessed in the moist state by thumb and finger pressure. Finally, a soil's plasticity is measured in the wet state by moulding with the hand.

The terms used to describe a soil in those three moisture states and a last state of no agricultural value are as follows:

  1. Consistency of Dry Soil: loose, soft, hard, extremely hard
  2. Consistency of Moist Soil: loose, friable, firm, extremely firm
  3. Consistency of Wet Soil: non-sticky, sticky or non-plastic, plastic
  4. Consistency of Cemented Soil: weakly cemented, indurated (cemented)

Soil consistency is useful in estimating the ability of soil to support buildings and roads. More precise measures of soil strength are often made prior to construction.

Temperature

Soil temperature regulates seed germination, root growth and the availability of nutrients. Soil temperatures range from permafrost at a few inches below the surface to 38°C (100°F) in Hawaii on a warm day. The colour of the ground cover and its insulating ability have a strong influence on soil temperature. Snow cover will reflect light and heavy mulching will slow the warming of the soil, but at the same time they will reduce the fluctuations in the surface temperature.

Below 50 cm (20 in), soil temperature seldom changes and can be approximated by adding 1.8°C (2°F) to the mean annual air temperature.

Most often, soil temperatures must be accepted and agricultural activities adapted to them to:

  1. maximize germination and growth by timing of planting
  2. optimise use of anhydrous ammonia by applying to soil below 10°C (50°F)
  3. prevent heaving and thawing due to frosts from damaging shallow-rooted crops
  4. prevent damage to desirable soil structure by freezing of saturated soils
  5. improve uptake of phosphorus by plants

Otherwise soil temperatures can be raised by drying soils or the use of clear plastic mulches. Organic mulches slow the warming of the soil.

Colour

Soil colour is often the first impression one has when viewing soil. Striking colours and contrasting patterns are especially noticeable. The Red River (Mississippi watershed) carries sediment eroded from extensive reddish soils like Port Silt Loam in Oklahoma. The Yellow River in China carries yellow sediment from eroding loess soils. Mollisols in the Great Plains of North America are darkened and enriched by organic matter. Podsols in boreal forests have highly contrasting layers due to acidity and leaching.

In general, colour is determined by organic matter content, drainage conditions, and the degree of oxidation. Soil colour, while easily discerned, has little use in predicting soil characteristics.[39] It is of use in distinguishing boundaries within a soil profile, determining the origin of a soil's parent material, as an indication of wetness and waterlogged conditions, and as a qualitative means of measuring organic, salt and carbonate contents of soils. Colour is recorded in the Munsell color system as for instance 10YR3/4.

Soil colour is primarily influenced by soil mineralogy. Many soil colours are due to various iron minerals. The development and distribution of colour in a soil profile result from chemical and biological weathering, especially redox reactions. As the primary minerals in soil parent material weather, the elements combine into new and colourful compounds. Iron forms secondary minerals of a yellow or red colour, organic matter decomposes into black and brown compounds, and manganese, sulfur and nitrogen can form black mineral deposits. These pigments can produce various colour patterns within a soil. Aerobic conditions produce uniform or gradual colour changes, while reducing environments (anaerobic) result in rapid colour flow with complex, mottled patterns and points of colour concentration.[40]

Resistivity

Soil resistivity is a measure of a soil's ability to retard the conduction of an electric current. The electrical resistivity of soil can affect the rate of galvanic corrosion of metallic structures in contact with the soil. Higher moisture content or increased electrolyte concentration can lower resistivity and increase conductivity, thereby increasing the rate of corrosion.[41][42] Soil resistivity values typically range from about 2 to 1000 Ω·m, but more extreme values are not unusual.[43]

Soil water

Water affects soil formation, structure, stability and erosion but is of primary concern with respect to plant growth. Water is essential to plants for four reasons:

  1. It constitutes 85%-95% of the plant's protoplasm.
  2. It is essential for photosynthesis.
  3. It is the solvent in which nutrients are carried to, into and throughout the plant.
  4. It provides the turgidity by which the plant keeps itself in proper position.[44]

In addition, water alters the soil profile by dissolving and redepositing minerals, often at lower levels, and possibly leaving the soil sterile in the case of extreme rainfall and drainage. In a loam soil, solids constitute half the volume, air one-quarter of the volume, and water one-quarter of the volume, of which only half will be available to most plants.

Water retention forces

Water is retained in a soil when the adhesive force of attraction of water for soil particles and the cohesive forces water feels for itself are capable of resisting the force of gravity which tends to drain water from the soil. When a field is flooded, the air space is displaced by water. The field will drain under the force of gravity until it reaches what is called field capacity, at which point the smallest pores are filled with water and the largest with water and air.[45] The total amount of water held when field capacity is reached is a function of the specific surface area of the soil particles. As a result, high clay and high organic soils have higher field capacities. The total force required to pull or push water out of soil is termed suction and usually expressed in units of bars (105 pascal) which is just a little less than one-atmosphere pressure. Alternatively, the terms "tension" or "moisture potential" may be used.[46]

Moisture classification

The forces with which water is held in soils determine its availability to plants. Forces of adhesion hold water strongly to mineral and humus surfaces and less strongly to itself by cohesive forces. A plant's root may penetrate a very small volume of water that is adhering to soil and be initially able to draw in water that is only lightly held by the cohesive forces. But as the droplet is drawn down, the forces of adhesion of the water for the soil particles make reducing the volume of water increasingly difficult until the plant cannot produce sufficient suction to use the remaining water.[47] The remaining water is considered unavailable. The amount of available water depends upon the soil texture and humus amounts and the type of plant attempting to use the water. Cacti, for example, can produce greater suction than can agricultural crop plants.

The following description applies to a loam soil and agricultural crops. When a field is flooded, it is said to be saturated and all available air space is occupied by water. The suction required to draw water into a plant root is zero. As the field drains under the influence of gravity (drained water is called gravitational water or drain-able water), the suction a plant must produce to use such water increases to 1/3 bar. At that point, the soil is said to have reached field capacity, and plants that use the water must produce increasingly higher suction, finally up to 15 bar. At 15 bar suction, the soil water amount is called wilting percent. At that suction the plant cannot sustain its water needs as water is still being lost from the plant by transpiration; the plant's turgidity is lost, and it wilts. The next level, called air-dry, occurs at 1000 bar suction. Finally the oven dry condition is reached at 10,000 bar suction. All water below wilting percentage is called unavailable water.[48]

Soil moisture content

The amount of water remaining in a soil drained to field capacity and the amount that is available are functions of the soil type. Sandy soil will retain very little water, while clay will hold the maximum amount. The time required to drain a field from flooded condition for a clay loam that begins at 43% water by weight to a field capacity of 21.5% is six days, whereas a sand loam that is flooded to its maximum of 22% water will take two days to reach field capacity of 11.3% water. The available water for the clay loam might be 11.3% whereas for the sand loam it might be only 7.9% by weight.[49]

Wilting point, field capacity, and available water capacity of various soil textures[50]
Soil TextureWilting PointField CapacityAvailable water capacity
Water per foot of soil depthWater per foot of soil depthWater per foot of soil depth
 %in. %in. %in.
Medium sand1.70.36.81.25.10.9
Fine sand2.30.48.51.56.21.1
Sandy loam3.40.611.32.07.91.4
Fine sandy loam4.50.814.72.610.21.8
Loam6.81.218.13.211.32.0
Silt loam7.91.419.83.511.92.1
Clay loam10.21.821.53.811.32.0
Clay14.72.622.64.07.91.4

The above are average values for the soil textures as the percentage of sand, silt and clay vary within the listed soil textures.

Water flow in soils

Water moves through soil due to the force of gravity, osmosis and capillarity. At zero to one-third bar suction, water moves through soil due to gravity; this is called saturated flow. At higher suction, water movement is called unsaturated flow.[51]

Water infiltration into soil is controlled by six factors:

  1. Soil texture
  2. Soil structure. Fine-textured soils with granular structure are most favourable to infiltration of water.
  3. The amount of organic matter. Coarse matter is best and if on the surface helps prevent the destruction of soil structure and the creation of crusts.
  4. Depth of soil to impervious layers such as hardpans or bedrock
  5. The amount of water already in the soil
  6. Soil temperature. Warm soils take in water faster while frozen soils may not be able to absorb depending on the type of freezing.[52]

Water infiltration rates range from 0.25 cm (0.098 in) per hour for high clay soils to 2.5 cm (0.98 in) per hour for sand and well stabilised and aggregated soil structures.[53] Water flows through the ground unevenly, called "gravity fingers", because of the surface tension between water particles. [54] Tree roots create paths for rainwater flow through soil by breaking though soil including clay layers: one study showed roots increasing infiltration of water by 153% and another study showed an increase by 27 times. [55] Flooding temporarily increases soil permeability in river beds, helping recharge aquifers.[56]

Saturated flow

Once soil is completely wetted, any more water will move downward, or percolate, carrying with it clay, humus and nutrients, primarily cations, out of the range of plant roots and result in acid soil conditions. In order of decreasing solubility, the leached nutrients are:
  • Calcium
  • Magnesium, Sulfur, Potassium; depending upon soil composition
  • Nitrogen; usually little, unless nitrate fertiliser was applied recently
  • Phosphorus; very little as its forms in soil are of low solubility.[57]
In the United States percolation water due to rainfall ranges from zero inches just east of the Rocky Mountains to twenty or more inches in the Appalachian Mountains and the north coast of the Gulf of Mexico.[50]

Unsaturated flow

At suctions less than one-third bar, water moves in all directions via unsaturated flow at a rate that is dependent on the square of the diameter of the water-filled pores. Water is pushed by pressure gradients from the point of its application where it is saturated locally, and pulled by capillary action due to adhesion force of water to the soil solids, producing a suction gradient from wet towards drier soil. Doubling the diameter of the pores increases the flow rate by a factor of four. Large pores drained by gravity and not filled with water do not greatly increase the flow rate for unsaturated flow. Water flow is primarily from coarse-textured soil into fine-textured soil and is slowest in fine-textured soils such as clay.[58]

Water uptake by plants

Of equal importance to the storage and movement of water in soil is the means by which plants acquire it and their nutrients. Ninety percent of water is taken up by plants as passive absorption caused by the pulling force of water evaporating (transpiring) from the long column of water that leads from the plant's roots to its leaves. In addition, the high concentration of salts within plant roots creates an osmotic pressure gradient that pushes soil water into the roots. Osmotic absorption becomes more important during times of low water transpiration caused by lower temperatures (for example at night) or high humidity. It is the process that causes guttation.[59]

Root extension is vital for plant survival. A study of a single winter rye plant grown for four months in one cubic foot of loam soil showed that the plant developed 13,800,000 roots a total of 385 miles in length and 2,550 square feet in surface area and 14 billion hair roots of 6,600 miles total length and 4,320 square feet total area, for a total surface area of 6,870 square feet (83 ft squared). The total surface area of the loam soil was estimated to be 560,000 square feet.[60] In other words the roots were in contact with only 1.2% of the soil. Roots must seek out water as the unsaturated flow of water in soil can move only at a rate of up to 2.5 cm (0.98 in) per day; as a result they are constantly dying and growing as they seek out high concentrations of soil moisture.

Insufficient soil moisture to the point of wilting will cause permanent damage and crop yields will suffer. When grain sorghum was exposed to soil suction as low as 13.0 bar during the seed head emergence through bloom and seed set stages of growth, its production was reduced by 34%.[61]

Consumptive use and water efficiency

Only a small fraction (0.1% to 1%) of the water used by a plant is held within the plant. The majority is ultimately lost via transpiration, while evaporation from the soil surface is also substantial. Transpiration plus evaporative soil moisture loss is called evapotranspiration. Evapotranspiration plus water held in the plant totals consumptive use, which is nearly identical to evapotranspiration.[61]

The total water used in an agricultural field includes runoff, drainage and consumptive use. The use of loose mulches will reduce evaporative losses for a period after a field is irrigated, but in the end the total evaporative loss will approach that of an uncovered soil. The benefit from mulch is to keep the moisture available during the seedling stage. Water use efficiency is measured by transpiration ratio, which is the ratio of the total water transpired by a plant to the dry weight of the harvested plant. Transpiration ratios for crops range from 300 to 700. For example alfalfa may have a transpiration ratio of 500 and as a result 500 kilograms of water will produce one kilogram of dry alfalfa. [62]

Soil atmosphere

The atmosphere of soil is radically different from the atmosphere above. The consumption of oxygen, by microbes and plant roots and their release of carbon dioxide, decrease oxygen and increase carbon dioxide concentration. Atmospheric CO2 concentration is 0.03%, but in the soil pore space it may range from 10 to 100 times that level. At extreme levels CO2 is toxic. In addition, the soil voids are saturated with water vapour. Adequate porosity is necessary not just to allow the penetration of water but also to allow gases to diffuse in and out. Movement of gases is by diffusion from high concentrations to lower. Oxygen diffuses in and is consumed and excess levels of carbon dioxide, diffuse out with other gases as well as water. Soil texture and structure strongly affect soil porosity and gas diffusion.[63] Platy and compacted soils impede gas flow, and a deficiency of oxygen may encourage anaerobic bacteria to reduce nitrate to the gases N2, N2O, and NO, which are then lost to the atmosphere. Aerated soil is also a net sink of methane CH4 but a net producer of greenhouse gases when soils are depleted of oxygen and subject to elevated temperatures.[64]

Chemical and colloidal properties

The chemistry of soil determines the availability of nutrients, the health of microbial populations, and its physical properties. In addition, soil chemistry also determines its corrosivity, stability, and ability to absorb pollutants and to filter water. It is the surface chemistry of clays and humus colloids that determines soil's chemical properties. The very high specific surface area of colloids and their net negative charges, gives soil its great ability to hold and release cations in what is referred to as cation exchange. Cation-exchange capacity (CEC) is the amount of exchangeable cations per unit weight of dry soil and is expressed in terms of milliequivalents of hydrogen ion per 100 grams of soil. “A colloid is a small, insoluble, nondiffusible particle larger than a molecule but small enough to remain suspended in a fluid medium without settling. Most soils contain organic colloidal particles as well as the inorganic colloidal particles of clays.”[65]

Soil clays

Due to its high specific surface area and its unbalanced negative charges, clay is the most active mineral component of soil. It is a colloidal and most often a crystalline material. In soils, clay is defined in a physical sense as any mineral particle less than two microns (2x10−6 inches) in effective diameter. Chemically, clay is a range of minerals with certain reactive properties. Clay is also a soil textural class. Many soil minerals, such as gypsum, carbonates, or quartz, are small enough to be classified physically as clay but chemically do not afford the same utility as do clay minerals.[66]

Clay was once thought to be very small particles of quartz, feldspar, mica, hornblende or augite, but it is now known to be (with the exception of mica-based clays) a precipitate with a mineralogical composition that is dependent on but different from its parent materials and is classed as a secondary mineral. The type of clay that is formed is a function of the parent material and the composition of the minerals in solution. Mica-based clays result from a modification of the primary mica mineral in such a way that it behaves and is classed as a clay. Most clays are crystalline, but some are amorphous. The clays of a soil are a mixture of the various types of clay, but one type predominates.[67]

Most clays are crystalline and most are made up of three or four planes of oxygen held together by planes of aluminium and silicon by way of ionic bonds that together form a single layer of clay. The spatial arrangement of the oxygen atoms determines clay's structure. Half of the weight of clay is oxygen, but on a volume basis oxygen is ninety percent.[68] The layers of clay are sometimes held together through hydrogen bonds or potassium bridges and as a result swell less in the presence of water. Other clays, such as montmorillonite, have layers that are loosely attached and will swell greatly when water intervenes.[69]

There are three groups of clays:

  1. Crystalline alumino-silica clays: montmorillonite, illite, vermiculite, chlorite, kaolinite.
  2. Amorphous clays: young mixtures of silica (SiO2-OH) and alumina (Al(OH)3) which have not had time to form regular crystals.
  3. Sesquioxide clays: old, highly leached clays which result in oxides of iron, aluminium and titanium.[70]

Silica clays

Alumino-silica clays are characterised by their regular crystalline structure. Oxygen in ionic bonds with silicon forms a tetrahedral coordination which in turn forms sheets of silica. Two sheets of silica are bonded together by a plane of aluminium which forms an octahedral coordination, called alumina, with the oxygens of the silica sheet above and that below it. Hydroxyl ions (OH-) sometimes substitute for oxygen. As much as one fourth of the aluminium Al3+ may be substituted by Zn2+, Mg2+ or Fe2+, and Si4+ may be substituted by Al3+. The substitution of lower-valence cations for higher-valence cations (isomorphic substitution) gives clay a net negative charge that attracts and holds soil solution cations, some of which are of value for plant growth. Isomorphic substitution occurs during the clay's formation and does not change with time.[71]
  • Montmorillonite clay is made of four planes of oxygen with two silicon and one central aluminium plane intervening. The alumino-silicate montmorillonite clay is said to have a 2:1 ratio of silicon to aluminium. The seven planes together form a single layer of montmorillonite. The layers are weakly held together and water may intervene, causing the clay to swell up to ten times its dry volume. It occurs in soils which have had little leaching, hence it is found in arid regions. The entire surface is exposed and available for surface reactions and it has a high cation exchange capacity (CEC).[72]
  • Illite is a 2:1 clay similar in structure to montmorillonite but has potassium bridges between the clay layers and the degree of swelling depends on the degree of weathering of the potassium. The active surface area is reduced due to the potassium bonds. Illite originates from the modification of mica, a primary mineral. It is often found together with montmorillonite and its primary minerals. It has moderate CEC.[73]
  • Vermiculite is a mica-based clay similar to illite, but the layers of clay are held together more loosely by hydrated magnesium and it will swell, but not as much as does montmorillonite. It has very high CEC.[74]
  • Chlorite is similar to vermiculite, but the loose bonding by occasional hydrated magnesium is replaced by a hydrated magnesium sheet, firmly bonding the planes above and below it. It has two planes of silicon, one of aluminium and one of magnesium; hence it is a 2:2 clay. Chlorite does not swell and it has low CEC.[74]
  • Kaolinite is very common, more common than montmorillonite in acid soils. It has one silica and one alumina sheet per layer; hence it is a 1:1 type clay. One layer of oxygen is replaced with hydroxyls, which produces strong hydrogen bonds to the oxygen in the next layer of clay. As a result kaolinite does not swell in water and has a low specific surface area, and as almost no isomorphic substitution has occurred it has a low CEC. Where rainfall is high, acid soils selectively leach more silica than alumina from the original clays, leaving kaolinite. Even heavier weathering results in sesquioxide clays.[75]
silica-sesquioxide

Amorphous clays

Amorphous clays are young, and commonly found in volcanic ash. They are mixtures of alumina and silica which have not formed the ordered crystal shape of alumino-silica clays which time would provide. The majority of their negative charges originates from hydroxyl ions, which can gain or lose a hydrogen ion (H+) in response to soil pH, and hence buffer the soil pH. They may have either a negative charge provided by the attached hydroxyl ion (OH-), which can attract a cation, or lose the hydrogen of the hydroxyl to solution and display a positive charge which can attract anions. As a result they may display either high CEC, in an acid soil solution, or high anion exchange capacity, in a basic soil solution.[75]

Sesquioxide clays

Sesquioxide clays are a product of heavy rainfall that has leached most of the silica and alumina from alumino-silica clay, leaving the less soluble oxides of iron Fe2O3 and iron hydroxide (Fe(OH)3) and aluminium hydroxides (Al(OH)3). It takes hundreds of thousands of years of leaching to create sesquioxide clays. Sesqui is Latin for "one and one-half": there are three parts oxygen to two parts iron or aluminium; hence the ratio is one and one-half. They are hydrated and act as either amorphous or crystalline. They are not sticky and do not swell, and soils high in them behave much like sand and can rapidly pass water. They are able to hold large quantities of phosphates. Sesquioxides have low CEC. Such soils range from yellow to red in color. Such clays tend to hold phosphorus tightly rendering them unavailable for absorption by plants.[76][77]

Organic colloids

Humus is the penultimate state of decomposition of organic matter; while it may linger for a thousand years, on the larger scale of the age of the other soil components, it is temporary. It is composed of the very stable lignins (30%) and complex sugars (polyuronides, 30%). Its chemical assay is 60% carbon, 5% nitrogen, some oxygen and the remainder hydrogen, sulfur, and phosphorus. On a dry weight basis, the CEC of humus is many times greater than that of clay. Plant roots also have cation exchange sites.[78]

Carbon and terra preta

In the extreme environment of heavy rain and high temperatures of tropical rain forests, the clay and organic colloids are largely destroyed. The heavy rain washes the alumino-silicate clays from the soil leaving only sesquioxide clays of low CEC. The high temperatures and humidity allow bacteria and fungi to virtually dissolve any organic matter on the rain-forest floor overnight and much of the nutrients are volatilized or leached from the soil and lost. Carbon, however, is far more stable than soil colloids and is capable of performing many of the functions of the soil colloids of sub-tropical soils. Research into terra-preta is still young but is promising. Fallow periods "on the Amazonian Dark Earths can be as short as 6 months, whereas fallow periods on Oxisols are usually 8 to 10 years long"[79]

Cation and anion exchange

Cation exchange, between colloids and soil water, buffers (moderates) soil pH, alters soil structure, and purifies percolating water by adsorbing cations of all types, both useful and harmful.

The negative charges on a colloid particle make it able to hold cations to its surface. The charges result from four sources.[80]

  1. Isomorphous substitution occurs in clay when lower-valence cations substitute for higher-valence cations in the crystal structure. Substitutions in the outermost layers are more effective than for the innermost layers, as the charge strength drops off as the square of the distance. The net result is a negative charge.
  2. Edge-of-clay oxygen atoms are not in balance ionically as the tetrahedral and octahedral structures are incomplete at the edges of clay.
  3. Hydrogens of the clay hydroxyls may be ionised into solution, leaving an oxygen with a negative charge.
  4. Hydrogens of humus hydroxyl groups may be ionised into solution, leaving an oxygen with a negative charge.

Cations held to the negatively charged colloids resist being washed downward by water and out of reach of plants' roots, thereby preserving the fertility of soils in areas of moderate rainfall and low temperatures.

There is a hierarchy in the process of cation exchange on colloids, as they differ in the strength of adsorption and their ability to replace one another. If present in equal amounts:

Al3+ replaces H+ replaces Ca2+ replaces Mg2+ replaces K+ same as NH4+ replaces Na+[65]

If one cation is added in large amounts, it may replace the others by the sheer force of its numbers (mass action). This is largely what occurs with the addition of fertiliser.

As the soil solution becomes more acidic (an abundance of H+), the other cations bound to colloids are pushed into solution. This is caused by the ionisation of hydroxyl groups on the surface of soil colloids in what is described as pH-dependent charges. Unlike permanent charges developed by isomorphous substitution, pH-dependent charges are variable and increase with increasing pH.[81] As a result, those cations can be made available to plants but are also able to be leached from the soil, possibly making the soil less fertile. Plants will excrete H+ into the soil and by that means, push cations off the colloids, thus making those cations absorbable by the plant.

Cation exchange capacity (CEC)

Cation exchange capacity should be thought of as the soil's ability to remove cations from the soil water solution and sequester those to be exchanged later as the plant roots release hydrogen ions to the solution. CEC is the amount of exchangeable hydrogen cation (H+) that will combine with 100 grams dry weight of soil and whose measure is one milliequivalents per 100 grams of soil (1 meq/100 g). Hydrogen ions have a single charge and one-thousandth of a gram of hydrogen ions per 100 grams dry soil gives a measure of one milliequivalent of hydrogen ion. Calcium, with an atomic weight 40 times that of hydrogen and with a valence of two, converts to (40/2) x 1 milliequivalent = 20 milliequivalents of hydrogen ion per 100 grams of dry soil or 20 meq/100 g.[82] The modern measure of CEC is expressed as centimoles of positive charge per kilogram (cmol/kg) of oven-dry soil.
Most of the soil's CEC occurs on clay and humus colloids, and the lack of those in hot, humid, wet climates, due to leaching and decomposition respectively, explains the relative sterility of tropical soils.
Cation exchange capacity for soils; soil textures; soil colloids[83]
SoilStateCEC meq/100 g
Charlotte fine sandFlorida1.0
Ruston fine sandy loamTexas1.9
Glouchester loamNew Jersey11.9
Grundy silt loamIllinois26.3
Gleason clay loamCalifornia31.6
Susquehanna clay loamAlabama34.3
Davie mucky fine sandFlorida100.8
Sands------1 - 5
Fine sandy loams------5-10
Loams and silt loams-----5-15
Clay loams-----15-30
Clays-----over 30
Sesquioxides-----0-3
Kaolinite-----3-15
Illite-----25-40
Montmorillonite-----60-100
Vermiculite (similar to illite)-----80-150
Humus-----100-300

Anion exchange capacity (AEC)

Anion exchange capacity should be thought of as the soil's ability to remove anions from the soil water solution and sequester those for later exchange as the plant roots release carbonate anions to the solution. Those colloids which have low CEC tend to have some AEC. Amorphous and sesquioxide clays have the highest AEC, followed by the iron oxides. Levels of AEC are much lower than for CEC. Phosphates tend to be held at anion exchange sites.
Iron and aluminum hydroxide clays are able to exchange their hydroxide anions (OH-) for other anions. The order reflecting the strength of anion adhesion is as follows:
H2PO4- replaces SO42- replaces NO3- replaces Cl-
The amount of exchangeable anions is of a magnitude of tenths to a few milliequivalents per 100 g dry soil.[83] As pH rises, there are relatively more hydroxyls, which will displace anions from the colloids and force them into solution and out of storage; hence AEC decreases with increasing pH (alkalinity).

Soil reaction (pH)

Soil reactivity is expressed in terms of pH and is a measure of the acidity or alkalinity of the soil. More precisely, it is a measure of hydrogen ion concentration in an aqueous solution and ranges in values from 0 to 14 (acidic to basic) but practically speaking for soils, pH ranges from 3.5 to 9.5, as pH values beyond those extremes are toxic to life forms.

Soil pH

At 25°C an aqueous solution that has a pH of 3.5 has 10-3.5 moles hydrogen ions per litre of solution (and also 10-10.5 mole/litre OH-) . A pH of 7, defined as neutral, has 10−7 moles hydrogen ions per litre of solution and also 10−7 moles of OH- per litre; since the two concentrations are equal, they are said to neutralise each other. A pH of 9.5 is 10-9.5 moles hydrogen ions per litre of solution (and also 10-2.5 mole per litre OH-) . A pH of 3.5 has one million times more hydrogen ions per litre than a solution with pH of 9.5 (9.5 - 3.5 = 6 or 106) and is more acidic.[84]
The effect of pH on a soil is to remove from the soil or to make available certain ions. Soils with high acidity tend to have toxic amounts of aluminium and manganese. Plants which need calcium need moderate alkalinity, but most minerals are more soluble in acid soils. Soil organisms are hindered by high acidity, and most agricultural crops do best with mineral soils of pH 6.5 and organic soils of pH 5.5.[85]
In high rainfall areas, soils tend to acidity as the basic cations are forced off the soil colloids by the mass action of hydrogen ions from the rain as those attach to the colloids. High rainfall rates can then wash the nutrients out, leaving the soil sterile. Once the colloids are saturated with H+, the addition of any more hydrogen ions or aluminum hydroxyl cations drives the pH even lower (more acidic) as the soil is left with no buffering capacity. In extreme rainfall areas and high temperatures, the clay and humus may be washed out, further reducing the buffering capacity of the soil. In low rainfall areas, unleached calcium pushes pH to 8.5 and with the addition of exchangeable sodium, soils may reach pH 10. Beyond a pH of 9, plant growth is reduced. High pH results in low micro-nutrient mobility, but water-soluble chelates of those nutrients can supply the deficit. Sodium can be reduced by the addition of gypsum (calcium sulphate) as calcium adheres to clay more tightly than does sodium causing sodium to be pushed into the soil water solution where it can be washed out by an abundance of water.[86]

Base saturation percentage

There are acid-forming cations (hydrogen and aluminium) and there are base-forming cations. The fraction of the base-forming cations that occupy positions on the soil colloids is called the base saturation percentage. If a soil has a CEC of 20 meq and 5 meq are aluminium and hydrogen cations (acid-forming), the remainder of positions on the colloids (20-5 = 15 meq) are assumed occupied by base-forming cations, so that the percentage base saturation is 15/20 x 100% = 75% (the compliment 25% is assumed acid-forming cations). When the soil pH is 7 (neutral), base saturation is 100 percent and there are no hydrogen ions stored on the colloids. Base saturation is almost in direct proportion to pH and except for its use in calculating the amount of lime needed to neutralise an acid soil, it is of little use.[87]

Buffering of soils

The resistance of soil to changes in pH and available cations from the addition of acid or basic material is a measure of the buffering capacity of a soil and increases as the CEC increases. Hence, pure sand has almost no buffering ability, while soils high in colloids have high buffering capacity. Buffering occurs by cation exchange and neutralisation.

The addition of highly basic aqueous ammonia to a soil will cause the ammonium to displace hydrogen ions from the colloids, and the end product is water and colloidally fixed ammonium, but no permanent change overall in soil pH.

The addition of lime, CaCO3, will displace hydrogen ions from the soil colloids, causing the fixation of calcium and the evolution of CO2 and water, with no permanent change in soil pH.

The addition of carbonic acid (the solution of CO2 in water) will displace calcium from colloids, as hydrogen ions are fixed to the colloids, evolving water and slightly alkaline (temporary increase in pH) highly soluble calcium bicarbonate, which will then precipitate as lime (CaCO3) and water at a lower level in the soil profile, with the result of no permanent change in soil pH.

All of the above are examples of the buffering of soil pH. The general principal is that an increase in a particular cation in the soil water solution will cause that cation to be fixed to colloids (buffered) and a decrease in solution of that cation will cause it to be withdrawn from the colloid and moved into solution (buffered). The degree of buffering is limited by the CEC of the soil; the greater the CEC, the greater the buffering capacity of the soil.[88]

Nutrients

Sixteen nutrients are essential for plant growth and reproduction. They are carbon, hydrogen, oxygen, nitrogen, phosphorus, potassium, sulfur, calcium, magnesium, iron, boron, manganese, copper, zinc, molybdenum, and chlorine. Nearly all plant nutrients are taken up in ionic forms from the soil solution as cations or as anions. Plants release bicarbonate and hydroxyl (OH-) anions or hydrogen cations from their roots in an effort to cause nutrient ions to be freed from sequestration on colloids and so forced into the soil solution where they can be picked up. Nitrogen is available in soil organic material but is unusable by plants until it is made available by that material's decomposition by micro-organisms into cation or anion forms.[89]

Plant nutrients, their chemical symbols, and the ionic forms common in soils and available for plant uptake[90]
ElementSymbolIon or molecule
CarbonCCO2 (mostly through leaves)
HydrogenHH+, HOH (water)
OxygenOO2-, OH -, CO32-, SO42-, CO2
PhosphorusPH2PO4 -, HPO42- (phosphates)
PotassiumKK+
NitrogenNNH4+, NO3 - (ammonium, nitrate)
SulfurSSO42-
CalciumCaCa2+
IronFeFe2+, Fe3+ (ferrous, ferric)
MagnesiumMgMg2+
BoronBH3BO3, H2BO3 -, B(OH)4 -
ManganeseMnMn2+
CopperCuCu2+
ZincZnZn2+
MolybdenumMoMoO42- (molybdate)
ChlorineClCl - (chloride)

Mechanism of nutrient uptake

All the nutrients with the exception of carbon are taken up by the plant through its roots. All those brought through the roots, with the exception of hydrogen, which is derived from water, are taken up in the form of ions. Carbon, in the form of carbon dioxide, enters primarily through the leaf stomata. All the hydrogen utilised by the plant originates from soil water and participates with the carbon dioxide in the photosynthetic production of sugars and release of oxygen as a byproduct. Plants may have their nutrient needs supplemented by spraying a water solution of nutrients on their leaves, but nutrients are typically received through the roots by:

  1. Mass flow
  2. Diffusion
  3. Root interception

The nutrient needs of a plant may be carried to the plant by the movement of the soil water solution in what is called mass flow. The absorption of nutrients from the soil solution with which the roots are in contact causes the concentration of nutrients in that area to be reduced. Diffusion of nutrients from areas with high concentration to those of lower concentration moves nutrients near the roots as they take up those nutrients. Plants constantly send out roots to seek new sources of nutrients in a process called root interception. Meanwhile older, less effective roots die back. Water is lifted to the leaves, where it is lost by transpiration and in the process it brings soil nutrients with it. A maize plant, for example, will use one quart of water per day at the height of its growing season.[91]

Estimated relative importance of mass flow, diffusion and root interception as mechanisms in supplying plant nutrients to corn plant roots in soils[92]
NutrientApproximate percentage supplied by:
Mass flowRoot interceptionDiffusion
Nitrogen98.81.20
Phosphorus6.32.890.9
Potassium20.02.377.7
Calcium71.428.60
Sulfur95.05.00
Molybdenum95.24.80

In the above table, phosphorus and potassium nutrients move more by diffusion than they do by mass flow in water solution, as they are rapidly taken up by the roots creating a concentration near zero near the roots. The very steep concentration gradient is of greater influence in the movement of those ions than is the movement of those by mass flow.[93] The movement by mass flow requires the transpiration of water from the plant causing water and solution ions to also move toward the roots. Movement by root interception is slowest as the plants must extend their roots. Plants move ions out of their roots in proportion to the amount of nutrients they move in. Hydrogen H+ is exchanged for cations, and carbonate (HCO3-) and hydroxide (OH-) anions are exchanged for nutrient anions. Plants derive most of their anion nutrients from decomposing organic matter, which holds 95 percent of the nitrogen, 5 to 60 percent of the phosphorus and 80 percent of the sulfur. As plant roots remove nutrients from the soil water solution, nutrients are added to the soil water as other ions move off of clay and humus, are added from the decomposition of soil minerals, and are released by the decomposition of organic matter. Where crops are produced, the replenishment of nutrients in the soil must be augmented by the addition of fertiliser or organic matter.[92]

Carbon

Plants obtain their carbon from atmospheric carbon dioxide. A plant's weight is forty-five percent carbon. Elementally, carbon is 50% of plant material. Plant residues have a carbon to nitrogen ratio (C/N) of 50:1. As the soil organic material is digested by arthropods and micro-organisms, the C/N decreases as the carbonaceous material is metabolised and carbon dioxide (CO2) is released as a byproduct and finds its way out of the soil and into the atmosphere. The nitrogen, however, is sequestered in the bodies of the live matter and so it builds up in the soil. Normal CO2 concentration in the atmosphere is 0.03%, which is probably the factor limiting plant growth. In a field of maize on a still day during high light conditions in the growing season, the CO2 concentration drops very low, but under such conditions the crop could use up to 20 times the normal concentration. The respiration of CO2 by soil micro-organisms decomposing soil organic matter contributes an important amount of CO2 to the photosynthesising plants. Within the soil, CO2 concentration is 10 to 100 times atmospheric but may rise to toxic levels if the soil porosity is low or if diffusion is impeded by flooding.[94]

Nitrogen

Nitrogen is the most critical element obtained by plants from the soil and is a bottleneck in plant growth.[95] Plants can use the nitrogen as either the ammonium cation ammonium (NH4+) or the anion nitrate (NO3-). Nitrogen is seldom missing in the soil, but is often in the form of raw organic material which cannot be used directly.

Carbon/Nitrogen Ratio of Various Organic Materials[96]Sortable table
Organic MaterialC:N Ratio
Alfalfa13
Bacteria4
Clover, green sweet16
Clover, mature sweet23
Fungi9
Forest litter30
Humus in warm cultivated soils11
Legume-grass hay25
Legumes (alfalfa or clover), mature20
Oat straw80
Straw, cornstalks90
Sawdust250

Some micro-organisms are able to metabolise the organic matter and release ammonium in a process called mineralisation. Others take free ammonium and oxidise it to nitrate. Particular bacteria are capable of metabolising N2 into the form of nitrate in a process called nitrogen fixation. Both ammonium and nitrate can be lost from the soil by incorporation into the microbes' living cells, where it is temporarily immobilised or sequestered. Nitrate may also be lost from the soil when bacteria metabolise it to the gases N2 and N2O. In that gaseous form, nitrogen escapes to the atmosphere in a process called denitrification. Nitrogen may also be leached from the soil if it is in the form of nitrate or lost to the atmosphere as ammonia due to a chemical reaction of ammonium with alkaline soil by way of a process called volatilisation. Ammonium may also be sequestered in clay by fixation. Nitrogen is added to soil by rainfall. [97][98]

Nitrogen gains

In a process called mineralisation, certain bacteria feed on organic matter, releasing ammonia (NH3) (which may be reduced to ammonium NH4+) and other nutrients. As long as the carbon to nitrogen ratio (C/N) in the soil is above 30:1, nitrogen will be in short supply and other bacteria will feed on the ammonium and incorporate its nitrogen into their cells. In that form the nitrogen is said to be immobilised. Later, when such bacteria die, they too are mineralised and some of the nitrogen is released as ammonium and nitrate. If the C/N is less than 15, ammonia is freed to the soil, where it may be used by bacteria which oxidise it to nitrate in a process called nitrification. Bacteria may on average add 25 pounds nitrogen per acre, and in an unfertilised field, this is the most important source of usable nitrogen. In a soil with 5 percent organic matter perhaps 2 to 5 percent of that is released to the soil by such decomposition. It occurs fastest in warm, moist, well aerated soil. The mineralisation of 3 percent of a soil that is 4 percent organic matter would release 120 pounds of nitrogen as ammonium per acre.[99]
In symbiotic fixation, Rhizobium bacteria convert N2 to nitrate by way of nitrogen fixation. They have a symbiotic relationship with host plants, wherein they supply the host with nitrogen and the host provides the bacteria with nutrients and a safe environment. It is estimated that such symbiotic bacteria in the root nodules of legumes add 45 to 250 pounds of nitrogen per acre per year, which may be sufficient for the crop. Other, free-living nitrogen-fixing bacteria and blue-green algae live independently in the soil and release nitrate when their dead bodies are converted by way of mineralisation.[100]
Some amount of usable nitrogen is fixed by lightning as nitric acid (HNO3). Ammonia, NH3, previously released from the soil or from combustion, may fall with precipitation as nitric acid at a rate of about five pounds nitrogen per acre per year.[101]

Nitrogen sequestration

When bacteria feed on soluble forms of nitrogen (ammonium and nitrite), they temporarily sequester that nitrogen in their bodies in a process called immobilisation. At a later time when those bacteria die, their nitrogen may be released as ammonium by the processes of mineralisation.
Protein material is easily broken down, but the rate of its decomposition is slowed by its attachment to the crystalline structure of clay and trapped between the clay layers. The layers are small enough that bacteria cannot enter. Some organisms can exude extracellular enzymes that can act on the sequestered proteins. However, those enzymes too may be trapped on the clay crystals.
Ammonium fixation occurs when ammonium replaces the potassium ions that normally exist between the layers of clay such as illite or montmorillonite. Only a small fraction of nitrogen is held this way.[102]

Nitrogen losses

Usable nitrogen may be lost from soils when it is in the form of nitrate, as it is easily leached. Further losses of nitrogen occur by denitrification, the process whereby soil bacteria convert nitrate (NO3-) to nitrogen gas, N2 or N2O. This occurs when poor soil aeration limits free oxygen, forcing bacteria to use the oxygen in nitrate for their respiratory process. Denitrification increases when oxidisable organic material is available and when soils are warm and slightly acidic. Denitrification may vary throughout a soil as the aeration varies from place to place. The conversion of nitrate to gases causes nitrogen to be lost from the soil to the atmosphere. Denitrification may cause the loss of 10 to 20 percent of the available nitrates within a day and when conditions are favourable to that process, losses of up to 60 percent of nitrate applied as fertiliser may occur.[103]
Ammonium volatilisation occurs when ammonium reacts chemically with an alkaline soil, converting NH4+ to NH3. The application of ammonium fertiliser to such a field can result in volatilisation losses of as much as 30 percent.[104]

Phosphorus

Phosphorus is the second most critical plant nutrient. The soil mineral apatite is the most common mineral source of phosphorus. While there is on average 1000 lb of phosphorus per acre in the soil, it is generally in unavailable forms. The available portion of phosphorus is low as it is in the form of phosphates of low solubility. Total phosphorus is about 0.1 percent by weight of the soil, but only one percent of that is available. Of the part available, more than half comes from the mineralisation of organic matter. Agricultural fields may need to be fertilised to make up for the phosphorus that has been removed in the crop.[105]

When phosphorus does form solubilised ions of H2PO4-, they rapidly form insoluble phosphates of calcium or hydrous oxides of iron and aluminum. Phosphorus is largely immobile in the soil and is not leached but actually builds up in the surface layer if not cropped. The application of soluble fertilisers to soils may result in zinc deficiencies as zinc phosphates form. Conversely, the application of zinc to soils may immobilise phosphorus as zinc phosphate. Lack of phosphorus may interfere with the normal opening of the plant leaf stomata, resulting in plant temperatures 10 percent higher than normal. Phosphorus is most available when soil pH is 6.5 in mineral soils and 5.5 in organic soils.[104]

Potassium

The amount of potassium in a soil may be as much as 80,000 lb per acre, of which only 150 lb or 2 percent is available for plant growth. When solubilised, half will be held as exchangeable cations on clay while the other half is in the soil water solution. Potassium fixation occurs when soils dry and the potassium is bonded between layers of clay. Under certain conditions, dependent on the soil texture, intensity of drying, and initial amount of exchangeable potassium, the fixed percentage may be as much as 90 percent within ten minutes. Potassium may be leached from soils low in clay.[106]

Calcium

Calcium is 1 percent by weight of soils and is generally available but may be low as it is soluble and can be leached. It is thus low in sandy and heavily leached soil or strongly acidic mineral soil. Calcium is supplied to the plant in the form of exchangeable ions and moderately soluble minerals. Calcium is more available on the soil colloids than is potassium because the common mineral calcite, CaCO3, is more soluble than potassium-bearing minerals.[107]

Magnesium

Magnesium is central to chlorophyll and aids in the uptake of phosphorus. The minimum amount of magnesium required for plant health is not sufficient for the health of forage animals. Magnesium is generally available, but is missing from some soils along the Gulf and Atlantic coasts of the United States due to leaching by heavy precipitation.[108]

Sulfur

Most sulfur is made available to plants, like phosphorus, by its release from decomposing organic matter.[108] Deficiencies may exist in some soils and if cropped, sulfur needs to be added. A 15-ton crop of onions uses up to 19 lb of sulfur and 4 tons of alfalfa uses 15 lb per acre. Sulfur abundance varies with depth. In a sample of soils in Ohio, United States, the sulfur abundance varied with depths, 0-6 inches, 6-12 inches, 12-18 inches, 18-24 inches in the amounts: 1056, 830, 686, 528 lb per acre respectively.

Micronutrients

Micronutrients include iron, manganese, zinc, copper, boron, chlorine and molybdenum. The term refers to plants' needs, not to their abundance in soil. They are required in very small amounts but are essential to plant health in that most are required parts of some enzyme system which speeds up plants' metabolisms. They are generally available in the mineral component of the soil, but the heavy application of phosphates can cause a deficiency in zinc and iron by the formation of insoluble phosphates. Iron deficiency may also result from excessive amounts of heavy metals or calcium minerals (lime) in the soil. Excess amounts of soluble boron, molybdenum and chloride are toxic.[109]

Organic matter

The organic soil matter includes all the dead plant material and all creatures, live and dead. The living component of an acre of soil may include 900 lb of earthworms, 2400 lb of fungi, 1500 lb of bacteria, 133 lb of protozoa and 890 lb of arthropods and algae.[110]

Most living things in soils, including plants, insects, bacteria and fungi, are dependent on organic matter for nutrients and energy. Soils have varying organic compounds in varying degrees of decomposition. Organic matter holds soils open, allowing the infiltration of air and water, and may hold as much as twice its weight in water. Many soils, including desert and rocky-gravel soils, have little or no organic matter. Soils that are all organic matter, such as peat (histosols), are infertile.[111] In its earliest stage of decomposition, the original organic material is often called raw organic matter. The final stage of decomposition is called humus.

Humus

Humus refers to organic matter that has been decomposed by bacteria, fungi, and protozoa to the final point where it is resistant to further breakdown. Humus usually constitutes only five percent of the soil or less by volume, but it is an essential source of nutrients and adds important textural qualities crucial to soil health and plant growth. Humus also hold bits of undecomposed organic matter which feed arthropods and worms which further improve the soil. Humus has a high cation exchange capacity that on a dry weight basis is many times greater than that of clay colloids. It also acts as a buffer, like clay, against changes in pH and soil moisture.

Humic acids and fulvic acids, which begin as raw organic matter, are important constituents of humus. After the death of plants and animals, microbes begin to feed on the residues, resulting finally in the formation of humus. With decomposition, there is a reduction of water-soluble constituents including cellulose and hemicellulose and nutrients such as nitrogen, phosphorus, and sulfur. As the residues break down, only complex molecules made of aromatic carbon rings, oxygen and hydrogen remain in the form of humin, lignin and lignin complexes as humus. While the structure of humus has few nutrients, it is able to attract and hold cation and anion nutrients by weak bonds that can be released in response to changes in soil pH.

Lignin is resistant to breakdown and accumulates within the soil. It also reacts with amino acids, which further increases its resistance to decomposition, including enzymatic decomposition by microbes. Fats and waxes from plant matter have some resistance to decomposition and persist in soils for a while. Clay soils often have higher organic contents that persist longer than soils without clay as the organic molecules adhere to and are stabilised by the clay. Proteins normally decompose readily, but when bound to clay particles, they become more resistant to decomposition. Clay particles also absorb the enzymes exuded by microbes which would normally break down proteins. The addition of organic matter to clay soils can render that organic matter and any added nutrients inaccessible to plants and microbes for many years. High soil tannin (polyphenol) content can cause nitrogen to be sequestered in proteins or cause nitrogen immobilisation.[112][113]

Humus formation is a process dependent on the amount of plant material added each year and the type of base soil. Both are affected by climate and the type of organisms present. Soils with humus can vary in nitrogen content but typically have 3 to 6 percent nitrogen. Raw organic matter, as a reserve of nitrogen and phosphorus, is a vital component affecting soil fertility.[111] Humus also absorbs water, and expands and shrinks between dry and wet states, increasing soil porosity. Humus is less stable than the soil's mineral constituents, as it is reduced by microbial decomposition, and over time its concentration diminshes without the addition of new organic matter. However, humus may persist over centuries if not millennia.

Climate and organics

The production, accumulation and degradation of organic matter are greatly dependent on climate. Temperature, soil moisture and topography are the major factors affecting the accumulation of organic matter in soils. Organic matter tends to accumulate under wet or cold conditions where decomposer activity is impeded by low temperature[114] or excess moisture which results in anaerobic conditions.[115] Excessive slope may encourage the erosion of the top layer of soil which holds most of the raw organic material that will eventually become humus.

Soil horizons

A horizontal layer of the soil, whose physical features, composition and age are distinct from those above and beneath, are referred to as a soil horizon. The naming of a horizon is based on the type of material of which it is composed. Those materials reflect the duration of specific processes of soil formation. They are labelled using a shorthand notation of letters and numbers[116] which describe the horizon in terms of its colour, size, texture, structure, consistency, root quantity, pH, voids, boundary characteristics and presence of nodules or concretions.[117] Few soil profiles have all the major horizons. Some may have only one horizon.

The exposure of parent material to favourable conditions produces mineral soils that are marginally suitable for plant growth. That growth often results in the accumulation of organic residues. The accumulated organic layer called the O horizon produces a more active soil due to the effect of the organisms that live within it. Organisms colonise and break down organic materials, making available nutrients upon which other plants and animals can live. After sufficient time, humus moves downward and is deposited in a distinctive organic surface layer called the A horizon.

Classification

Soil is classified into categories in order to understand relationships between different soils and to determine the suitability of a soil for a particular use. One of the first classification systems was developed by the Russian scientist Dokuchaev around 1880. It was modified a number of times by American and European researchers, and developed into the system commonly used until the 1960s. It was based on the idea that soils have a particular morphology based on the materials and factors that form them. In the 1960s, a different classification system began to emerge which focused on soil morphology instead of parental materials and soil-forming factors. Since then it has undergone further modifications. The World Reference Base for Soil Resources (WRB)[118] aims to establish an international reference base for soil classification.

USDA soil taxonomy

A taxonomy is an arrangement in a systematic manner. Soil taxonomy has six categories. They are, from most general to specific: order, suborder, great group, subgroup, family and series. The soil properties that can be measured quantitatively are used to classify soils. A partial list is: depth, moisture, temperature, texture, structure, cation exchange capacity, base saturation, clay mineralogy, organic matter content and salt content.

In the United States, soil orders are the top hierarchical level of soil classification in the USDA soil taxonomy[119] . The names of the orders end with the suffix -sol. There are 12 soil orders in Soil Taxonomy:[120][121] The criteria for the order divisions include properties that reflect major differences in the genesis of soils.

  • Alfisol - soils with aluminium and iron. They have horizons of clay accumulation, and form where there is enough moisture and warmth for at least three months of plant growth. They constitute 10.1% of soils worldwide.
  • Andisols - volcanic ash soils. They are young and very fertile. They cover 1% of the world's ice-free surface.
  • Aridisol - dry soils forming under desert conditions which have fewer than 90 consecutive days of moisture during the growing season. They include nearly 12% of soils on Earth. Soil formation is slow, and accumulated organic matter is scarce. They may have subsurface zones of caliche or duripan. Many aridisols have well-developed Bt horizons showing clay movement from past periods of greater moisture.
  • Entisol - recently formed soils that lack well-developed horizons. Commonly found on unconsolidated river and beach sediments of sand and clay or volcanic ash, some have an A horizon on top of bedrock. They are 18% of soils worldwide.
  • Gelisols - permafrost soils with permafrost within two metres of the surface or gelic materials and permafrost within one metre. They constitute 9.1% of soils worldwide.
  • Histosol - organic soils, formerly called bog soils, are 1.2% of soils worldwide.
  • Inceptisol - young soils. They have subsurface horizon formation but show little eluviation and illuviation. They constitute 15% of soils worldwide.
  • Mollisols - soft, deep, dark fertile soil formed in grasslands and some hardwood forests with very thick A horizons. They are 7% of soils worldwide.
  • Oxisol - are the most weathered, are rich in iron and aluminum oxides (sesquioxides) and kayolin but low in silica. They have only trace nutrients due to heavy tropical rainfall and high temperatures. They are 7.5% of soils worldwide.
  • Spodosol - acid soils with organic colloid layer complexed with iron and aluminium leached from a layer above. They are typical soils of coniferous and deciduous forests in cooler climates. They constitute 4% of soils worldwide.
  • Ultisol - acid soils in humid climates, tropical to subtropical temperatures, which are heavily leached of Ca, Mg, and K nutrients. They are not quite Oxisols. They are 8.1% of the soil worldwide.
  • Vertisol - inverted soils. They are clay-rich and tend to swell when wet and shrink upon drying, often forming deep cracks into which surface layers can fall. They are difficult to farm and on which to construct roads and buildings due to their high expansion rate. They constitute 2.4% of soils worldwide.

The percentages listed above[122] are for land area free of ice. "Soils of Mountains", which constitute the balance (11.6%), have a mixture of those listed above, or are classified as "Rugged Mountains" which have no soil.

The above soil orders in sequence of increasing degree of development are Entisols, Inceptisols, Aridisols, Mollisols, Alfisols, Spodosols, Ultisols, and Oxisols. Histosols and Vertisols may appear in any of the above at any time during their development.

The soil suborders within an order are differentiated on the basis of soil properties and horizons which depend on soil moisture and temperature. Forty-seven suborders are recognized in the United States.[123]

The soil great group category is a subdivision of a suborder distinguishing one soil from another by the kind and sequence of soil horizons. About 185 great groups are recognized in the United States. Horizons marked by clay, iron, humus and hard pans and soil features such as the expansion-contraction of clays that produce self-mixing provided by clay, temperature, and marked quantities of various salts are used as distinguishing features.[123]

The great group categories are divided into three kinds of soil subgroups: typic, intergrade and extragrade. A typic subgroup represents the basic or 'typical' concept of the great group to which the described subgroup belongs. An intergrade subgroup describes the properties that suggest how it grades towards (is similar to) soils of other soil great groups, suborders or orders. These properties are not developed or expressed well enough to cause the soil to be included within the great group towards which they grade, but suggest similarities. Extragrade features are aberrant properties which prevent that soil from being included in another soil classification. About 1,000 soil subgroups are defined in the United States.[123]

A soil family category is a group of soils within a subgroup and describes the physical and chemical properties which affect the response of soil to agricultural management and engineering applications. The principal characteristics used to differentiate soil families include texture, mineralogy, pH, permeability, structure, consistency, the locale's precipitation pattern, and soil temperature. For some soils the criteria also specify the percentage of silt, sand and coarse fragments such as gravel, cobbles and rocks. About 4,500 soil families are recognised in the United States.[124]

A family may contain several soil series which describe the physical location using the name of a prominent physical feature such as a river or town near where the soil sample was taken. An example would be Merrimac for the Merrimack River in New Hampshire, USA. More than 14,000 soil series are recognised in the United States. This permits very specific descriptions of soils.[125]

A soil phase of series, originally called 'soil type' describes the soil surface texture, slope, stoniness, saltiness, erosion, and other conditions.[125]

Australian soil taxonomy

There are fourteen soil orders at the top level of the Australian Soil Classification. They are: Anthroposols, Organosols, Podosols, Vertosols, Hydrosols, Kurosols, Sodosols, Chromosols, Calcarosols, Ferrosols, Dermosols, Kandosols, Rudosols and Tenosols.

Uses

Soil is used in agriculture, where it serves as the anchor and primary nutrient base for plants; however, as demonstrated by hydroponics, it is not essential to plant growth if the soil-contained nutrients can be dissolved in a solution. The types of soil and available moisture determine the species of plants that can be cultivated.

Soil material is also a critical component in the mining and construction industries. Soil serves as a foundation for most construction projects. The movement of massive volumes of soil can be involved in surface mining, road building and dam construction. Earth sheltering is the architectural practice of using soil for external thermal mass against building walls.

Soil resources are critical to the environment, as well as to food and fibre production. Soil provides minerals and water to plants. Soil absorbs rainwater and releases it later, thus preventing floods and drought. Soil cleans water as it percolates through it. Soil is the habitat for many organisms: the major part of known and unknown biodiversity is in the soil, in the form of invertebrates (earthworms, woodlice, millipedes, centipedes, snails, slugs, mites, springtails, enchytraeids, nematodes, protists), bacteria, archaea, fungi and algae; and most organisms living above ground have part of them (plants) or spend part of their life cycle (insects) below-ground. Above-ground and below-ground biodiversities are tightly interconnected,[126][127] making soil protection of paramount importance for any restoration or conservation plan.

The biological component of soil is an extremely important carbon sink since about 57% of the biotic content is carbon. Even on desert crusts, cyanobacteria lichens and mosses capture and sequester a significant amount of carbon by photosynthesis. Poor farming and grazing methods have degraded soils and released much of this sequestered carbon to the atmosphere. Restoring the world's soils could offset some of the huge increase in greenhouse gases causing global warming, while improving crop yields and reducing water needs.[128][129][130]

Waste management often has a soil component. Septic drain fields treat septic tank effluent using aerobic soil processes. Landfills use soil for daily cover. Land application of waste water relies on soil biology to aerobically treat BOD.

Organic soils, especially peat, serve as a significant fuel resource; but wide areas of peat production, such as sphagnum bogs, are now protected because of patrimonial interest.

Both animals and humans in many cultures occasionally consume soil. It has been shown that some monkeys consume soil, together with their preferred food (tree foliage and fruits), in order to alleviate tannin toxicity.[131]

Soils filter and purify water and affect its chemistry. Rain water and pooled water from ponds, lakes and rivers percolate through the soil horizons and the upper rock strata, thus becoming groundwater. Pests (viruses) and pollutants, such as persistent organic pollutants (chlorinated pesticides, polychlorinated biphenyls), oils (hydrocarbons), heavy metals (lead, zinc, cadmium), and excess nutrients (nitrates, sulfates, phosphates) are filtered out by the soil.[132] Soil organisms metabolise them or immobilise them in their biomass and necromass,[133] thereby incorporating them into stable humus.[134] The physical integrity of soil is also a prerequisite for avoiding landslides in rugged landscapes.[135]

Degradation

Here, land degradation[136] refers to a human-induced or natural process which impairs the capacity of land to function. Soils are the critical component in land degradation when it involves acidification, contamination, desertification, erosion or salination.

While soil acidification is beneficial in the case of alkaline soils, it degrades land when it lowers crop productivity and increases soil vulnerability to contamination and erosion. Soils are often initially acid because their parent materials were acid and initially low in the basic cations (calcium, magnesium, potassium and sodium). Acidification occurs when these elements are removed from the soil profile by normal rainfall or the harvesting of forest or agricultural crops. Soil acidification is accelerated by the use of acid-forming nitrogenous fertilizers and by the effects of acid precipitation.

Soil contamination at low levels is often within soil's capacity to treat and assimilate. Many waste treatment processes rely on this treatment capacity. Exceeding treatment capacity can damage soil biota and limit soil function. Derelict soils occur where industrial contamination or other development activity damages the soil to such a degree that the land cannot be used safely or productively. Remediation of derelict soil uses principles of geology, physics, chemistry and biology to degrade, attenuate, isolate or remove soil contaminants to restore soil functions and values. Techniques include leaching, air sparging, chemical amendments, phytoremediation, bioremediation and natural attenuation.

Desertification is an environmental process of ecosystem degradation in arid and semi-arid regions, often caused by human activity. It is a common misconception that droughts cause desertification. Droughts are common in arid and semiarid lands. Well-managed lands can recover from drought when the rains return. Soil management tools include maintaining soil nutrient and organic matter levels, reduced tillage and increased cover. These practices help to control erosion and maintain productivity during periods when moisture is available. Continued land abuse during droughts, however, increases land degradation. Increased population and livestock pressure on marginal lands accelerates desertification.

Erosion of soil is caused by wind, water, ice and movement in response to gravity. Although the processes may be simultaneous, erosion is distinguished from weathering. Erosion is an intrinsic natural process, but in many places it is increased by human land use. Poor land use practices include deforestation, overgrazing and improper construction activity. Improved management can limit erosion by using techniques like limiting disturbance during construction, avoiding construction during erosion-prone periods, intercepting runoff, terrace-building, use of erosion-suppressing cover materials, and planting trees or other soil-binding plants.

A serious and long-running water erosion problem occurs in China, on the middle reaches of the Yellow River and the upper reaches of the Yangtze River. From the Yellow River, over 1.6 billion tons of sediment flow each year into the ocean. The sediment originates primarily from water erosion (gully erosion) in the Loess Plateau region of northwest China.

Soil piping is a particular form of soil erosion that occurs below the soil surface. It is associated with levee and dam failure, as well as sink hole formation. Turbulent flow removes soil starting from the mouth of the seep flow and subsoil erosion advances upgradient.[137] The term sand boil is used to describe the appearance of the discharging end of an active soil pipe.[138]

Soil salination is the accumulation of free salts to such an extent that it leads to degradation of the agricultural value of soils and vegetation. Consequences include corrosion damage, reduced plant growth, erosion due to loss of plant cover and soil structure, and water quality problems due to sedimentation. Salination occurs due to a combination of natural and human-caused processes. Arid conditions favour salt accumulation. This is especially apparent when soil parent material is saline. Irrigation of arid lands is especially problematic.[139] All irrigation water has some level of salinity. Irrigation, especially when it involves leakage from canals and overirrigation in the field, often raises the underlying water table. Rapid salination occurs when the land surface is within the capillary fringe of saline groundwater. Soil salinity control involves watertable control and flushing with higher levels of applied water in combination with tile drainage or another form of subsurface drainage.[140][141]

Soil salinity models like SWAP,[142] DrainMod-S,[143] UnSatChem,[144] SaltMod[145][146] and SahysMod[147] are used to assess the cause of soil salination and to optimise the reclamation of irrigated saline soils.

Reclamation

Soils which contain high levels of particular clays, such as smectites, are often very fertile. For example, the smectite-rich clays of Thailand's Central Plains are among the most productive in the world.

Many farmers in tropical areas, however, struggle to retain organic matter in the soils they work. In recent years, for example, productivity has declined in the low-clay soils of northern Thailand. Farmers initially responded by adding organic matter from termite mounds, but this was unsustainable in the long-term. Scientists experimented with adding bentonite, one of the smectite family of clays, to the soil. In field trials, conducted by scientists from the International Water Management Institute in cooperation with Khon Kaen University and local farmers, this had the effect of helping retain water and nutrients. Supplementing the farmer's usual practice with a single application of 200 kg bentonite per rai (6.26 rai = 1 hectare) resulted in an average yield increase of 73%. More work showed that applying bentonite to degraded sandy soils reduced the risk of crop failure during drought years.

In 2008, three years after the initial trials, IWMI scientists conducted a survey among 250 farmers in northeast Thailand, half of whom had applied bentonite to their fields. The average improvement for those using the clay addition was 18% higher than for non-clay users. Using the clay had enabled some farmers to switch to growing vegetables, which need more fertile soil. This helped to increase their income. The researchers estimated that 200 farmers in northeast Thailand and 400 in Cambodia had adopted the use of clays, and that a further 20,000 farmers were introduced to the new technique.[148]

If the soil is too high in clay, adding gypsum, washed river sand and organic matter will balance the composition. Adding organic matter to soil which is depleted in nutrients and too high in sand will boost its quality.[149]

See also

  • Acid sulfate soil
  • Agrophysics
  • Alkaline soil
  • Bulk soil
  • Geoponic
  • Hydroponics
  • Index of soil-related articles
  • Manure
  • Mineral matter in plants
  • Nitrogen cycle
  • Red Mediterranean soil
  • Saline soil
  • Shrink-swell capacity
  • Soil management
  • Soil salinity control
  • Soil zoology
  • Terra preta
  • Topsoil





References

Citations
  1. ^ Birkeland, Peter W. Soils and Geomorphology. 3rd edition. New York: Oxford University Press, 1999.
  2. ^ Chesworth, Edited by Ward (2008), Encyclopedia of soil science, Dordrecht, Netherlands: Springer, xxiv, ISBN 1-4020-3994-8 
  3. ^ Voroney, R. P., 2006. "The Soil Habitat". In Soil Microbiology, Ecology and Biochemistry, Eldor A. Paul ed. ISBN 0-12-546807-5
  4. ^ James A. Danoff-Burg, Columbia University. The Terrestrial Influence: Geology and Soils
  5. ^ Janet Raloff. Dirt Is Not Soil. ScienceNews. 17 July 2008)
  6. ^ Taylor, S. A., and G. L. Ashcroft. 1972. Physical Edaphology
  7. ^ McCarthy, David F. (1982). Essentials of Soil Mechanics and Foundations: Basic Geotechnics (2nd ed.). Reston, Virginia: Reston Publishing. ISBN 9780835917810. 
  8. ^ Pedosphere.com
  9. ^ Buol, S. W.; Hole, F. D. and McCracken, R. J. (1973), Soil Genesis and Classification (1st ed.), Ames, Iowa: Iowa State University Press, ISBN 0-8138-1460-X 
  10. ^ Retallack, G.J. (2008). Soils of the Past: An Introduction to Paleopedology (2nd ed.). John Wiley & Sons. p. 207. ISBN 9780470698167. Retrieved 18 August 2012. 
  11. ^ Kellogg 1957, p. 17.
  12. ^ Donahue, Miller & Shickluna 1977, p. 4.
  13. ^ a b c d e Brady, Nyle (1984). The Nature and Properties of Soils (9th ed.). USA: Macmillan Publishing Co. pp. 4–7. ISBN 0023133406. 
  14. ^ Kellogg 1957, pp. 1–4.
  15. ^ Michael E. Ritter. Factors Affecting Soil Development, Soil Systems, The Physical Environment: an Introduction to Physical Geography, University of Wisconsin, Stevens Point, 1 October 2009, retrieved 3 January 2012.
  16. ^ Van Schöll, Laura; Smits, Mark M. and Hoffland, Ellis (2006), "Ectomycorrhizal weathering of the soil minerals muscovite and hornblende", New Phytologist 171 (4): 805–814, doi:10.1111/j.1469-8137.2006.01790.x, PMID 16918551 
  17. ^ Donahue, Miller & Shickluna 1977, pp. 20–21.
  18. ^ Donahue, Miller & Shickluna 1977, p. 21.
  19. ^ NASA (broken link)
  20. ^ Donahue, Miller & Shickluna 1977, p. 24.
  21. ^ Donahue, Miller & Shickluna 1977, pp. 31–33.
  22. ^ H.B. Milford, A.J.E. McGaw and K.J. Nixon, Soil Data Entry Handbook for the NSW Soil and Land Information System (SALIS), 3rd ed., New South Wales Department of Land and Water Conservation Resource Information Systems Group, Parramatta, 2001, pdf pp. 30–32.
  23. ^ Gove Hambidge, "Climate and Man—A Summary", in Erwin Raisz, U.S. Department of Agriculture, Climate And Man, Part One, Yearbook of Agriculture 1941, Repr. Honolulu: University Press of the Pacific, 2004, ISBN 978-1-4102-1538-3, pp. 1–66, p. 27.
  24. ^ Donahue, Miller & Shickluna 1977, p. 35.
  25. ^ Copley, Jon (25 August 2005). "Millions of bacterial species revealed underfoot". New Scientist. Retrieved 19 April 2010. 
  26. ^ a b c Amber Dance (2008). "Soil ecology: What lies beneath.". Nature 455 (7214): 724–5. doi:10.1038/455724a. PMID 18843336. 
  27. ^ a b Roesch LF, Fulthorpe RR, Riva A, Casella G, Hadwin AK, Kent AD, Daroub SH, Camargo FA, Farmerie WG, Triplett EW (2007). "Pyrosequencing enumerates and contrasts soil microbial diversity". The ISME Journal 1 (4): 283–90. doi:10.1038/ismej.2007.53. PMC 2970868. PMID 18043639. 
  28. ^ Gans J, Wolinsky M, Dunbar J. (2005). "Computational improvements reveal great bacterial diversity and high metal toxicity in soil". Science 309 (5739): 1387–90. doi:10.1126/science.1112665. PMID 16123304. 
  29. ^ R. B. Brown (September 2003). "Soil Texture". Fact Sheet SL-29. University of Florida, Institute of Food and Agricultural Sciences. Retrieved 8 July 2008. 
  30. ^ Kellogg 1957, pp. 32–33.
  31. ^ Donahue, Miller & Shickluna 1977, p. 53.
  32. ^ Soil Survey Division Staff (1993). "Soil Structure". Handbook 18. Soil survey manual. Soil Conservation Service. U.S. Department of Agriculture. Archived from the original on 16 March 2008. Retrieved 8 July 2008. 
  33. ^ Donahue, Miller & Shickluna 1977, pp. 55–56.
  34. ^ "Soil Structure - Physical Properties". The Cooperative Soil Survey. Retrieved 19 October 2012. 
  35. ^ Donahue, Miller & Shickluna 1977, pp. 59–61.
  36. ^ Donahue, Miller & Shickluna 1977, p. 60.
  37. ^ Donahue, Miller & Shickluna 1977, pp. 62–71.
  38. ^ "Soils - Part 2: Physical Properties of Soil and Soil Water". Retrieved 19 October 2012. 
  39. ^ Donahue, Miller & Shickluna 1977, p. 71.
  40. ^ "The Color of Soil". United States Department of Agriculture - Natural Resources Conservation Service. Archived from the original on 16 March 2008. Retrieved 8 July 2008. 
  41. ^ "Electrical Design, Cathodic Protection". United States Army Corps of Engineers. 22 April 1985. Archived from the original on 12 June 2008. Retrieved 2 July 2008. 
  42. ^ "The why and how to testing the Electrical Conductivity of Soils | Resources". Retrieved 19 December 2010. 
  43. ^ R. J. Edwards (15 February 1998). "Typical Soil Characteristics of Various Terrains". Retrieved 2 July 2008. 
  44. ^ Donahue, Miller & Shickluna 1977, p. 72.
  45. ^ "CHAPTER 2 - SOIL AND WATER". Fao.org. Retrieved 2012-11-07. 
  46. ^ Donahue, Miller & Shickluna 1977, pp. 72–75.
  47. ^ Kellogg 1957, p. 47.
  48. ^ Donahue, Miller & Shickluna 1977, pp. 75–76.
  49. ^ Donahue, Miller & Shickluna 1977, pp. 76–77.
  50. ^ a b Donahue, Miller & Shickluna 1977, p. 80.
  51. ^ Donahue, Miller & Shickluna 1977, p. 85.
  52. ^ Donahue, Miller & Shickluna 1977, p. 86.
  53. ^ Donahue, Miller & Shickluna 1977, p. 88.
  54. ^ Brehm, Denise (December 11, 2008). "CEE researchers explain mystery of gravity fingers". MIT Department of Civil & Environmental Engineering. MIT. Retrieved October 31, 2012. 
  55. ^ "Urban Trees Enhance Water Infiltration". Fisher, Madeline. The American Society of Agronomy. November 17, 2008. Retrieved October 31, 2012. 
  56. ^ "Major floods recharge aquifers". University of New South Wales Science. January, 24, 2011. Retrieved October 31, 2012. 
  57. ^ Donahue, Miller & Shickluna 1977, p. 90.
  58. ^ Donahue, Miller & Shickluna 1977, p. 91.
  59. ^ Donahue, Miller & Shickluna 1977, p. 92.
  60. ^ Kellogg 1957, p. 46.
  61. ^ a b Donahue, Miller & Shickluna 1977, p. 94.
  62. ^ Donahue, Miller & Shickluna 1977, pp. 97–99.
  63. ^ Kellogg 1957, pp. 35–36.
  64. ^ http://atmo.tamu.edu/class/geos489/le cture6/soiltracegasrev.pdf
  65. ^ a b "Hort 403 - Reading - Soil". Hort.purdue.edu. Retrieved 2012-11-07. 
  66. ^ Donahue, Miller & Shickluna 1977, pp. 101–102.
  67. ^ Donahue, Miller & Shickluna 1977, p. 102.
  68. ^ Kellogg 1957, p. 33.
  69. ^ Donahue, Miller & Shickluna 1977, pp. 102–107.
  70. ^ Donahue, Miller & Shickluna 1977, pp. 101–107.
  71. ^ Donahue, Miller & Shickluna 1977, p. 107.
  72. ^ Donahue, Miller & Shickluna 1977, p. 108.
  73. ^ Donahue, Miller & Shickluna 1977, pp. 108–110.
  74. ^ a b Donahue, Miller & Shickluna 1977, p. 110.
  75. ^ a b Donahue, Miller & Shickluna 1977, p. 111.
  76. ^ Donahue, Miller & Shickluna 1977, pp. 103–112.
  77. ^ Kellogg 1957, pp. 33–34.
  78. ^ Donahue, Miller & Shickluna 1977, p. 112.
  79. ^ Lehmann, J. "Terra Preta de Indio". University of Cornell, Dept. of Crop and Soil Sciences. Retrieved 30 March 2013. 
  80. ^ Donahue, Miller & Shickluna 1977, p. 103–106.
  81. ^ "Sources. Negative Charge:". Jan.ucc.nau.edu. Retrieved 2012-11-07. 
  82. ^ Donahue, Miller & Shickluna 1977, p. 114.
  83. ^ a b Donahue, Miller & Shickluna 1977, pp. 115–116.
  84. ^ Chang, Raymond (1984). Chemistry. Random House, Inc. p. 424. ISBN 0-394-32983-X. 
  85. ^ Donahue, Miller & Shickluna 1977, pp. 116–117.
  86. ^ Donahue, Miller & Shickluna 1977, pp. 116–119.
  87. ^ Donahue, Miller & Shickluna 1977, pp. 119–120.
  88. ^ Donahue, Miller & Shickluna 1977, pp. 120–121.
  89. ^ Donahue, Miller & Shickluna 1977, pp. 123–131.
  90. ^ Donahue, Miller & Shickluna 1977, p. 125.
  91. ^ Donahue, Miller & Shickluna 1977, pp. 123–128.
  92. ^ a b Donahue, Miller & Shickluna 1977, p. 126.
  93. ^ "Lecture 22". Northern Arizona University. Retrieved 2013-03-22. 
  94. ^ Kellogg 1957, pp. 41, 80, 153.
  95. ^ Donahue, Miller & Shickluna 1977, p. 128.
  96. ^ Donahue, Miller & Shickluna 1977, p. 145.
  97. ^ Kellogg 1957, pp. 85–94, 152–155.
  98. ^ Donahue, Miller & Shickluna 1977, pp. 128–131.
  99. ^ Donahue, Miller & Shickluna 1977, pp. 129–130.
  100. ^ Donahue, Miller & Shickluna 1977, pp. 128–129.
  101. ^ Kellogg 1957, p. 87.
  102. ^ Kellogg 1957, p. 90.
  103. ^ Donahue, Miller & Shickluna 1977, p. 130.
  104. ^ a b Donahue, Miller & Shickluna 1977, p. 131.
  105. ^ Kellogg 1957, pp. 65–95.
  106. ^ Donahue, Miller & Shickluna 1977, pp. 134–135.
  107. ^ Donahue, Miller & Shickluna 1977, pp. 135–136.
  108. ^ a b Donahue, Miller & Shickluna 1977, p. 136.
  109. ^ Donahue, Miller & Shickluna 1977, pp. 136–137.
  110. ^ Pimentel, D. et al. (1995), "Environmental and economic costs of soil erosion and conservation benefits", Science 267 (24): 1117–22 
  111. ^ a b Foth, Henry D. (1984), Fundamentals of soil science, New York: Wiley, p. 151, ISBN 0-471-88926-1 
  112. ^ Verkaik, Eric; Jongkind, Anne G.; Berendse, Frank (2006), "Short-term and long-term effects of tannins on nitrogen mineralization and litter decomposition in kauri (Agathis australis (D. Don) Lindl.) forests", Plant and Soil 287: 337, doi:10.1007/s11104-006-9081-8 
  113. ^ Fierer, N. (2001), "Influence of balsam poplar tannin fractions on carbon and nitrogen dynamics in Alaskan taiga floodplain soils", Soil Biology and Biochemistry 33 (12–13): 1827, doi:10.1016/S0038-0717(01)00111-0 
  114. ^ Wagai, Rota; Mayer Lawrence M., Kitayama Kanehiro & Knicker Heike (2008), "Climate and parent material controls on organic matter storage in surface soils: A three-pool, density-separation approach", Geoderma 147: 23–33, doi:10.1016/j.geoderma.2008.07.010 
  115. ^ Minayeva, T. Yu.; Trofimov S. Ya., Chichagova O.A., Dorofeyeva E.I., Sirin A.A., Glushkov I.V., Mikhailov N.D. & Kromer B. (2008), "Carbon accumulation in soils of forest and bog ecosystems of southern Valdai in the Holocene", Biology Bulletin 35 (5): 524–532, doi:10.1134/S1062359008050142 
  116. ^ Retallack, G. J. (1990), Soils of the past : an introduction to paleopedology, Boston: Unwin Hyman, p. 32, ISBN 978-0-04-445757-2 
  117. ^ Buol, S.W. (1990), Soil genesis and classification, Ames, Iowa: Iowa State University Press, p. 36, doi:10.1081/E-ESS, ISBN 0-8138-2873-2 
  118. ^ IUSS Working Group WRB (2007). "World Reference Base for soil resources - A framework for international classification, correlation and communication". FAO. 
  119. ^ Soil Survey Staff (1999). Soil taxonomy: A Basic System of Soil Classification for Making and Interpreting Soil Surveys. 2nd edition. Natural Resources Conservation Service. U.S. Department of Agriculture Handbook 436.. USDA: United States Dept. of Agriculture, Naturel Resources Conservation Service. Retrieved 10 March 2013. 
  120. ^ The Soil Orders, Department of Environmental Sciences, University of Virginia, retrieved 23 October 2012.
  121. ^ Donahue, Miller & Shickluna 1977, pp. 411–432.
  122. ^ The Twelve Soil Orders: Soil Taxonomy, Soil & Land Resources Division, College of Agricultural and Life Sciences, University of Idaho
  123. ^ a b c Donahue, Miller & Shickluna 1977, pp. 409.
  124. ^ Donahue, Miller & Shickluna 1977, pp. 409-410.
  125. ^ a b Donahue, Miller & Shickluna 1977, p. 410.
  126. ^ Ponge, Jean-François (2003), "Humus forms in terrestrial ecosystems: a framework to biodiversity", Soil Biology and Biochemistry 35 (7): 935–945, doi:10.1016/S0038-0717(03)00149-4 
  127. ^ De Deyn, Gerlinde B.; Van der Putten Wim H. (2005), "Linking aboveground and belowground diversity", Trends in Ecology & Evolution 20 (11): 625–633, doi:10.1016/j.tree.2005.08.009, PMID 16701446 
  128. ^ Hansen, J., et al. (2008), "Target atmospheric CO2: Where should humanity aim?", Open Atmospheric Science Journal 2: 217–31 
  129. ^ Lal, R. (11 June 2004). "Soil Carbon Sequestration Impacts on Global Climate Change and Food Security". Science 304 (5677): 1623–1627. doi:10.1126/science.1097396. PMID 15192216. 
  130. ^ Blakeslee, Thomas R. (24 February 2010). "Greening Deserts for Carbon Credits". Renewable Energy World.com. Retrieved 23 October 2012. 
  131. ^ Setz, EZF; Enzweiler J, Solferini VN, Amendola MP, Berton RS (1999), "Geophagy in the golden-faced saki monkey (Pithecia pithecia chrysocephala) in the Central Amazon", Journal of Zoology 247 (1): 91–103, doi:10.1111/j.1469-7998.1999.tb00196.x  (subscription required)
  132. ^ Kohne, John Maximilian; Koehne Sigrid, Simunek Jirka (2009), "A review of model applications for structured soils: a) Water flow and tracer transport", Journal of Contaminant Hydrology 104 (1–4): 4–35, Bibcode:2009JCHyd.104....4K, doi:10.1016/j.jconhyd.2008.10.002, PMID 19012994 
  133. ^ Diplock, EE; Mardlin DP, Killham KS, Paton GI (2009), "Predicting bioremediation of hydrocarbons: laboratory to field scale", Environmental Pollution 157 (6): 1831–1840, doi:10.1016/j.envpol.2009.01.022, PMID 19232804 
  134. ^ Moeckel, Claudia; Nizzetto Luca, Di Guardo Antonio, Steinnes Eiliv, Freppaz Michele, Filippa Gianluca, Camporini Paolo, Benner Jessica, Jones Kevin C. (2008), "Persistent organic pollutants in boreal and montane soil profiles: distribution, evidence of processes and implications for global cycling", Environmental Science and Technology 42 (22): 8374–8380, doi:10.1021/es801703k, PMID 19068820 
  135. ^ Rezaei, Khalil; Guest Bernard, Friedrich Anke, Fayazi Farajollah, Nakhaei Mohamad, Aghda Seyed Mahmoud Fatemi, Beitollahi Ali (2009), "Soil and sediment quality and composition as factors in the distribution of damage at the December 26, 2003, Bam area earthquake in SE Iran (M (s)=6.6)", Journal of Soils and Sediments 9: 23–32, doi:10.1007/s11368-008-0046-9 
  136. ^ Johnson, D.L.; Ambrose, S.H.; Bassett, T.J.; Bowen, M.L.; Crummey, D.E.; Isaacson, J.S.; Johnson, D.N.; Lamb, P. et al. (1997). "Meanings of environmental terms". Journal of Environmental Quality 26: 581–589. doi:10.2134/jeq1997.00472425002600030002x. Citation uses old-style implicit et al. for authors
  137. ^ Jones, j. a. a. (1976), "Soil piping and stream channel initiation", Water Resources Research 7 (3): 602–610, Bibcode:1971WRR.....7..602J, doi:10.1029/WR007i003p00602. 
  138. ^ Dooley, Alan (June 2006). "Sandboils 101: Corps has experience dealing with common flood danger". Engineer Update. US Army Corps of Engineers. Archived from the original on 18 April 2008. Retrieved 14 May 2008. 
  139. ^ ILRI (1989), Effectiveness and Social/Environmental Impacts of Irrigation Projects: a Review, In: Annual Report 1988 of the International Institute for Land Reclamation and Improvement (ILRI), Wageningen, The Netherlands, pp. 18–34 
  140. ^ Drainage Manual: A Guide to Integrating Plant, Soil, and Water Relationships for Drainage of Irrigated Lands, Interior Dept., Bureau of Reclamation, 1993, ISBN 0-16-061623-9 
  141. ^ "Free articles and software on drainage of waterlogged land and soil salinity control". Retrieved 28 July 2010. 
  142. ^ SWAP model
  143. ^ DrainMod-S model
  144. ^ UnSatChem model
  145. ^ ILRI (1997), SaltMod: a tool for interweaving of irrigation and drainage for salinity control, In: W.B.Snellen (ed.), Towards integration of irrigation, and drainage management. Special report of the International Institute for Land Reclamation and Improvement (ILRI), Wageningen, The Netherlands, pp. 41–43 
  146. ^ SaltMod, an agro-hydro-soil salinity model 
  147. ^ SahysMod, a spatial agro-hydro-soil salinity cum groundwater model 
  148. ^ Improving soils and boosting yields in Thailand Success stories, Issue 2, 2010, IWMI
  149. ^ "Provide for your garden's basic needs ... and the plants will take it from there". USA Weekend. 10 March 2011. 
Sources
  • Donahue, Roy Luther; Miller, Raymond W.; Shickluna, John C. (1977). Soils: An Introduction to Soils and Plant Growth. Prentice-Hall. ISBN 0-13-821918-4. 
  • Kellogg, Charles E. (1957). In Stefferud, Alfred. Soil: The Yearbook of Agriculture 1957. United States Department of Agriculture. OCLC 036943367. 

Further reading

  • Soil-Net.com A free schools-age educational site teaching about soil and its importance.
  • Adams, J.A. 1986. Dirt. College Station, Texas : Texas A&M University Press ISBN 0-89096-301-0
  • Certini, G., Scalenghe, R., 2006. Soils: Basic concepts and future challenges. Cambridge Univ Press, Cambridge UK.
  • David R. Montgomery, Dirt: The Erosion of Civilizations, ISBN 978-0-520-25806-8
  • Faulkner, Edward H. Plowman's Folly. New York, Grosset & Dunlap. 1943. ISBN 0-933280-51-3
  • LandIS Free Soilscapes Viewer Free interactive viewer for the Soils of England and Wales
  • Geo-technological Research Paper, IIT Kanpur, Dr P P Vitkar - Strip footing on weak clay stabilized with a granular pile National Research Council Canada: From Discovery to Innovation / Conseil national de recherches Canada : de la découverte à l'innovation (English), (French)
  • Jenny, Hans, Factors of Soil Formation: A System of Quantitative Pedology 1941 http://www.soilandhealth.org/01aglibr ary/010159.Jenny.pdf
  • Logan, W. B., Dirt: The ecstatic skin of the earth. 1995 ISBN 1-57322-004-3
  • Mann, Charles C.: " Our good earth" National Geographic Magazine September 2008
  • "97 Flood". USGS. Retrieved 8 July 2008.  Photographs of sand boils.
  • Soil Survey Division Staff. (1999) Soil survey manual. Soil Conservation Service. U.S. Department of Agriculture Handbook 18.
  • Soil Survey Staff. (1975) Soil Taxonomy: A basic system of soil classification for making and interpreting soil surveys. USDA-SCS Agric. Handb. 436. United States Government Printing Office, Washington, DC.
  • Soils, Oregon State University
  • Why Study Soils?
  • Soil notes
  • LandIS Soils Data for England and Wales a pay source for GIS data on the soils of England and Wales and soils data source; they charge a handling fee to researchers.

External links





(Prev) SohoSoil governance (Next)