You are on page 1of 50

Soil

Soil, also commonly referred to as earth, is a


mixture of organic matter, minerals, gases, liquids,
and organisms that together support life of plants
and soil organisms. Some scientific definitions
distinguish dirt from soil by restricting the former
term specifically to displaced soil.

Soil consists of a solid phase of minerals and


organic matter (the soil matrix), as well as a
porous phase that holds gases (the soil
atmosphere) and water (the soil solution).[1][2]
Accordingly, soil is a three-state system of solids,
liquids, and gases.[3] Soil is a product of several
factors: the influence of climate, relief (elevation,
orientation, and slope of terrain), organisms, and Surface-water-gley developed in glacial till in Northern
the soil's parent materials (original minerals) Ireland
[4]
interacting over time. It continually undergoes
development by way of numerous physical,
chemical and biological processes, which include weathering with associated erosion.[5] Given its
complexity and strong internal connectedness, soil ecologists regard soil as an ecosystem.[6]

Most soils have a dry bulk density (density of soil taking into account voids when dry) between 1.1 and
1.6 g/cm3 , though the soil particle density is much higher, in the range of 2.6 to 2.7 g/cm3 .[7] Little of the
soil of planet Earth is older than the Pleistocene and none is older than the Cenozoic,[8] although fossilized
soils are preserved from as far back as the Archean.[9]

Collectively the Earth's body of soil is called the pedosphere. The pedosphere interfaces with the
lithosphere, the hydrosphere, the atmosphere, and the biosphere.[10] Soil has four important functions:

as a medium for plant growth


as a means of water storage, supply and purification
as a modifier of Earth's atmosphere
as a habitat for organisms

All of these functions, in their turn, modify the soil and its properties.

Soil science has two basic branches of study: edaphology and pedology. Edaphology studies the influence
of soils on living things.[11] Pedology focuses on the formation, description (morphology), and
classification of soils in their natural environment.[12] In engineering terms, soil is included in the broader
concept of regolith, which also includes other loose material that lies above the bedrock, as can be found on
the Moon and other celestial objects.[13]

Processes
Soil is a major component of the Earth's ecosystem. The world's ecosystems are impacted in far-reaching
ways by the processes carried out in the soil, with effects ranging from ozone depletion and global warming
to rainforest destruction and water pollution. With respect to Earth's carbon cycle, soil acts as an important
carbon reservoir,[14] and it is potentially one of the most reactive to human disturbance[15] and climate
change.[16] As the planet warms, it has been predicted that soils will add carbon dioxide to the atmosphere
due to increased biological activity at higher temperatures, a positive feedback (amplification).[17] This
prediction has, however, been questioned on consideration of more recent knowledge on soil carbon
turnover.[18]

Soil acts as an engineering medium, a habitat for soil organisms, a recycling system for nutrients and
organic wastes, a regulator of water quality, a modifier of atmospheric composition, and a medium for plant
growth, making it a critically important provider of ecosystem services.[19] Since soil has a tremendous
range of available niches and habitats, it contains a prominent part of the Earth's genetic diversity. A gram
of soil can contain billions of organisms, belonging to thousands of species, mostly microbial and largely
still unexplored.[20][21] Soil has a mean prokaryotic density of roughly 108 organisms per gram,[22]
whereas the ocean has no more than 107 prokaryotic organisms per milliliter (gram) of seawater.[23]
Organic carbon held in soil is eventually returned to the atmosphere through the process of respiration
carried out by heterotrophic organisms, but a substantial part is retained in the soil in the form of soil
organic matter; tillage usually increases the rate of soil respiration, leading to the depletion of soil organic
matter.[24] Since plant roots need oxygen, aeration is an important characteristic of soil. This ventilation can
be accomplished via networks of interconnected soil pores, which also absorb and hold rainwater making it
readily available for uptake by plants. Since plants require a nearly continuous supply of water, but most
regions receive sporadic rainfall, the water-holding capacity of soils is vital for plant survival.[25]

Soils can effectively remove impurities,[26] kill disease agents,[27] and degrade contaminants, this latter
property being called natural attenuation.[28] Typically, soils maintain a net absorption of oxygen and
methane and undergo a net release of carbon dioxide and nitrous oxide.[29] Soils offer plants physical
support, air, water, temperature moderation, nutrients, and protection from toxins.[30] Soils provide readily
available nutrients to plants and animals by converting dead organic matter into various nutrient forms.[31]

Composition
A typical soil is about 50% solids (45% mineral and 5% organic matter), and 50% voids (or pores) of which
half is occupied by water and half by gas.[32] The percent soil mineral and organic content can be treated as
a constant (in the short term), while the percent soil water and gas content is considered highly variable
whereby a rise in one is simultaneously balanced by a reduction in the other.[33] The pore space allows for
the infiltration and movement of air and water, both of which are critical for life existing in soil.[34]
Compaction, a common problem with soils, reduces this space, preventing air and water from reaching
plant roots and soil organisms.[35]

Given sufficient time, an undifferentiated soil will evolve a soil profile which consists of two or more
layers, referred to as soil horizons. These differ in one or more properties such as in their texture, structure,
density, porosity, consistency, temperature, color, and reactivity.[8] The horizons differ greatly in thickness
and generally lack sharp boundaries; their development is dependent on the type of parent material, the
processes that modify those parent materials, and the soil-forming factors that influence those processes.
The biological influences on soil properties are strongest near the surface, though the geochemical
influences on soil properties increase with depth. Mature soil profiles typically include three basic master
horizons: A, B, and C. The solum normally includes the A and B horizons. The living component of the
soil is largely confined to the solum, and is generally more prominent in the A horizon.[36] It has been
suggested that the pedon, a column of soil extending vertically from the surface to the underlying parent
material and large enough to show the characteristics of all its
horizons, could be subdivided in the humipedon (the living part,
where most soil organisms are dwelling, corresponding to the
humus form), the copedon (in intermediary position, where most
weathering of minerals takes place) and the lithopedon (in contact
with the subsoil).[37]

The soil texture is determined by the relative proportions of the


individual particles of sand, silt, and clay that make up the soil.

The interaction of the individual mineral particles with organic


matter, water, gases via biotic and abiotic processes causes those
particles to flocculate (stick together) to form aggregates or peds.[38]
Where these aggregates can be identified, a soil can be said to be A, B, and C represent the soil
developed, and can be described further in terms of color, porosity, profile, a notation firstly coined by
consistency, reaction (acidity), etc. Vasily Dokuchaev (1846–1903), the
father of pedology. Here, A is the
Water is a critical agent in soil development due to its involvement topsoil; B is a regolith; C is a
in the dissolution, precipitation, erosion, transport, and deposition of saprolite (a less-weathered regolith);
the materials of which a soil is composed.[39] The mixture of water the bottom-most layer represents the
and dissolved or suspended materials that occupy the soil pore bedrock.
space is called the soil solution. Since soil water is never pure water,
but contains hundreds of dissolved organic and mineral substances,
it may be more accurately called the soil solution. Water is central to
the dissolution, 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, a complex
feedback which is exemplified in the dynamics of banded vegetation
patterns in semi-arid regions.[40]

Soils supply plants with nutrients, most of which are held in place by
particles of clay and organic matter (colloids)[41] The nutrients may be
adsorbed on clay mineral surfaces, bound within clay minerals
(absorbed), or bound within organic compounds as part of the living
organisms or dead soil organic matter. These bound nutrients interact
Components of a silt loam
with soil water to buffer the soil solution composition (attenuate soil by percent volume
changes in the soil solution) as soils wet up or dry out, as plants take
up nutrients, as salts are leached, or as acids or alkalis are added.[42]    Water (25%)
   Gases (25%)
Plant nutrient availability is affected by soil pH, which is a measure of
   Sand (18%)
the hydrogen ion activity in the soil solution. Soil pH is a function of
   Silt (18%)
many soil forming factors, and is generally lower (more acid) where
   Clay (9%)
weathering is more advanced.[43]
   Organic matter (5%)
Most plant nutrients, with the exception of nitrogen, originate from the
minerals that make up the soil parent material. Some nitrogen
originates from rain as dilute nitric acid and ammonia,[44] but most of the nitrogen is available in soils as a
result of nitrogen fixation by bacteria. Once in the soil-plant system, most nutrients are recycled through
living organisms, plant and microbial residues (soil organic matter), mineral-bound forms, and the soil
solution. Both living soil organisms (microbes, animals and plant roots) and soil organic matter are of
critical importance to this recycling, and thereby to soil formation and soil fertility.[45] Microbial soil
enzymes may release nutrients from minerals or organic matter for
use by plants and other microorganisms, sequester (incorporate)
them into living cells, or cause their loss from the soil by
volatilisation (loss to the atmosphere as gases) or leaching.[46]

Formation
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, producing a distinct layer called
the B horizon. This is a somewhat arbitrary definition as mixtures of
sand, silt, clay and humus will support biological and agricultural
activity before that time.[47] These constituents are moved from one
level to another by water and animal activity. As a result, layers A soil texture triangle plot is a visual
(horizons) form in the soil profile. The alteration and movement of representation of the proportions of
materials within a soil causes the formation of distinctive soil sand, silt, and clay in a soil sample.
horizons. However, more recent definitions of soil embrace soils
without any organic matter, such as those regoliths that formed on
Mars[48] and analogous conditions in planet Earth deserts.[49]

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 (in a first stage nitrogen-fixing lichens and
cyanobacteria then epilithic higher plants) become established very quickly on basaltic lava, even though
there is very little organic material.[50] Basaltic minerals commonly weather relatively quickly, according to
the Goldich dissolution series.[51] The plants are supported by the porous rock as it is filled with nutrient-
bearing water that carries minerals dissolved from the rocks. Crevasses and pockets, local topography of the
rocks, would hold fine materials and harbour plant roots. The developing plant roots are associated with
mineral-weathering mycorrhizal fungi[52] that assist in breaking up the porous lava, and by these means
organic matter and a finer mineral soil accumulate with time. Such initial stages of soil development have
been described on volcanoes,[53] inselbergs,[54] and glacial moraines.[55]

How soil formation proceeds is influenced by at least five classic factors that are intertwined in the
evolution of a soil: parent material, climate, topography (relief), organisms, and time.[56] When reordered to
climate, relief, organisms, parent material, and time, they form the acronym CROPT.[57]

Physical properties
The physical properties of soils, in order of decreasing importance for ecosystem services such as crop
production, are texture, structure, bulk density, porosity, consistency, temperature, colour and resistivity.[58]
Soil texture is determined by the relative proportion of the three kinds of soil mineral particles, called soil
separates: sand, silt, and clay. At the next larger scale, soil structures called peds or more commonly soil
aggregates are created from the soil separates when iron oxides, carbonates, clay, silica and humus, coat
particles and cause them to adhere into larger, relatively stable secondary structures.[59] Soil bulk density,
when determined at standardized moisture conditions, is an estimate of soil compaction.[60] Soil porosity
consists of the void part of the soil volume and is occupied by gases or water. Soil consistency is the ability
of soil materials 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
which are buried in soil.[61] These properties vary through the depth of a soil profile, i.e. through soil
horizons. Most of these properties determine the aeration of the soil and the ability of water to infiltrate and
to be held within the soil.[62]

Soil moisture
Soil water content can be measured as volume or weight. Soil moisture levels, in order of decreasing water
content, are saturation, field capacity, wilting point, air dry, and oven dry. Field capacity describes a drained
wet soil at the point water content reaches equilibrium with gravity. Irrigating soil above field capacity risks
percolation losses. Wilting point describes the dry limit for growing plants. During growing season, soil
moisture is unaffected by functional groups or specie richness.[63]

Available water capacity is the amount of water held in a soil profile available to plants. As water content
drops, plants have to work against increasing forces of adhesion and sorptivity to withdraw water. Irrigation
scheduling avoids moisture stress by replenishing depleted water before stress is induced.[64][65]

Capillary action is responsible for moving groundwater from wet regions of the soil to dry areas.
Subirrigation designs (e.g., wicking beds, sub-irrigated planters) rely on capillarity to supply water to plant
roots. Capillary action can result in an evaporative concentration of salts, causing land degradation through
salination.

Soil moisture measurement — measuring the water content of the soil, as can be expressed in terms of
volume or weight — can be based on in situ probes (e.g., capacitance probes, neutron probes), or remote
sensing methods. Soil moisture measurement is an important factor in determining changes in soil
activity.[63]

Soil gas
The atmosphere of soil, or soil gas, is very different from the atmosphere above. The consumption of
oxygen by microbes and plant roots, and their release of carbon dioxide, decreases oxygen and increases
carbon dioxide concentration. Atmospheric CO2 concentration is 0.04%, but in the soil pore space it may
range from 10 to 100 times that level, thus potentially contributing to the inhibition of root respiration.[66]
Calcareous soils regulate CO2 concentration by carbonate buffering, contrary to acid soils in which all CO2
respired accumulates in the soil pore system.[67] At extreme levels, CO2 is toxic.[68] This suggests a
possible negative feedback control of soil CO2 concentration through its inhibitory effects on root and
microbial respiration (also called soil respiration).[69] In addition, the soil voids are saturated with water
vapour, at least until the point of maximal hygroscopicity, beyond which a vapour-pressure deficit occurs in
the soil pore space.[34] 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, the
diffusion coefficient decreasing with soil compaction.[70] Oxygen from above atmosphere diffuses in the
soil where it is consumed and levels of carbon dioxide in excess of above atmosphere diffuse out with other
gases (including greenhouse gases) as well as water.[71] Soil texture and structure strongly affect soil
porosity and gas diffusion. It is the total pore space (porosity) of soil, not the pore size, and the degree of
pore interconnection (or conversely pore sealing), together with water content, air turbulence and
temperature, that determine the rate of diffusion of gases into and out of soil.[72][71] Platy soil structure and
soil compaction (low porosity) impede gas flow, and a deficiency of oxygen may encourage anaerobic
bacteria to reduce (strip oxygen) from nitrate NO3 to the gases N2 , N2 O, and NO, which are then lost to
the atmosphere, thereby depleting the soil of nitrogen, a detrimental process called denitrification.[73]
Aerated soil is also a net sink of methane (CH4 )[74] but a net producer of methane (a strong heat-absorbing
greenhouse gas) when soils are depleted of oxygen and subject to elevated temperatures.[75]

Soil atmosphere is also the seat of emissions of volatiles other than carbon and nitrogen oxides from various
soil organisms, e.g. roots,[76] bacteria,[77] fungi,[78] animals.[79] These volatiles are used as chemical cues,
making soil atmosphere the seat of interaction networks[80][81] playing a decisive role in the stability,
dynamics and evolution of soil ecosystems.[82] Biogenic soil volatile organic compounds are exchanged
with the aboveground atmosphere, in which they are just 1–2 orders of magnitude lower than those from
aboveground vegetation.[83]

Humans can get some idea of the soil atmosphere through the well-known 'after-the-rain' scent, when
infiltering rainwater flushes out the whole soil atmosphere after a drought period, or when soil is
excavated,[84] a bulk property attributed in a reductionist manner to particular biochemical compounds such
as petrichor or geosmin.

Solid phase (soil matrix)


Soil particles can be classified by their chemical composition (mineralogy) as well as their size. The particle
size distribution of a soil, its texture, determines many of the properties of that soil, in particular hydraulic
conductivity and water potential,[85] but the mineralogy of those particles can strongly modify those
properties. The mineralogy of the finest soil particles, clay, is especially important.[86]

Chemistry
The chemistry of a soil determines its ability to supply available plant nutrients and affects its physical
properties and the health of its living population. In addition, a soil's chemistry also determines its
corrosivity, stability, and ability to absorb pollutants and to filter water. It is the surface chemistry of mineral
and organic colloids that determines soil's chemical properties.[87] A colloid is a small, insoluble particle
ranging in size from 1 nanometer to 1 micrometer, thus small enough to remain suspended by Brownian
motion in a fluid medium without settling.[88] Most soils contain organic colloidal particles called humus as
well as the inorganic colloidal particles of clays. The very high specific surface area of colloids and their net
electrical charges give soil its ability to hold and release ions. Negatively charged sites on colloids attract
and release cations in what is referred to as cation exchange. Cation-exchange capacity is the amount of
exchangeable cations per unit weight of dry soil and is expressed in terms of milliequivalents of positively
charged ions per 100  grams of soil (or centimoles of positive charge per kilogram of soil; cmolc/kg).
Similarly, positively charged sites on colloids can attract and release anions in the soil, giving the soil anion
exchange capacity.

Cation and anion exchange

The cation exchange, that takes place 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 or positive charges on colloid particles make them able to hold cations or anions, respectively,
to their surfaces. The charges result from four sources.[89]

1. Isomorphous substitution occurs in clay during its formation, when lower-valence cations
substitute for higher-valence cations in the crystal structure.[90] Substitutions in the
outermost layers are more effective than for the innermost layers, as the electric charge
strength drops off as the square of the distance. The net result is oxygen atoms with net
negative charge and the ability to attract cations.
2. Edge-of-clay oxygen atoms are not in balance ionically as the tetrahedral and octahedral
structures are incomplete.[91]
3. Hydroxyls may substitute for oxygens of the silica layers, a process called hydroxylation.
When the hydrogens of the clay hydroxyls are ionised into solution, they leave the oxygen
with a negative charge (anionic clays).[92]
4. Hydrogens of humus hydroxyl groups may also be ionised into solution, leaving, similarly to
clay, an oxygen with a negative charge.[93]

Cations held to the negatively charged colloids resist being washed downward by water and are out of
reach of plant roots, thereby preserving the soil fertility in areas of moderate rainfall and low
temperatures.[94][95]

There is a hierarchy in the process of cation exchange on colloids, as cations differ in the strength of
adsorption by the colloid and hence their ability to replace one another (ion exchange). If present in equal
amounts in the soil water solution:

Al3+ replaces H+ replaces Ca2+ replaces Mg2+ replaces K+ same as NH+4 replaces Na+[96]

If one cation is added in large amounts, it may replace the others by the sheer force of its numbers. This is
called law of mass action. This is largely what occurs with the addition of cationic fertilisers (potash,
lime).[97]

As the soil solution becomes more acidic (low pH, meaning an abundance of H+), the other cations more
weakly bound to colloids are pushed into solution as hydrogen ions occupy exchange sites (protonation). A
low pH may cause the hydrogen of hydroxyl groups to be pulled into solution, leaving charged sites on the
colloid available to be occupied by other cations. This ionisation of hydroxy groups on the surface of soil
colloids creates what is described as pH-dependent surface charges.[98] Unlike permanent charges
developed by isomorphous substitution, pH-dependent charges are variable and increase with increasing
pH.[99] Freed cations can be made available to plants but are also prone to be leached from the soil,
possibly making the soil less fertile.[100] Plants are able to excrete H+ into the soil through the synthesis of
organic acids and by that means, change the pH of the soil near the root and push cations off the colloids,
thus making those available to the plant.[101]

Cation exchange capacity (CEC)

Cation exchange capacity is 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.[102] 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) × 1 milliequivalent = 20 milliequivalents of hydrogen ion per
100 grams of dry soil or 20 meq/100 g.[103] 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
(such as tropical rainforests), due to leaching and decomposition, respectively, explains the apparent sterility
of tropical soils.[104] Live plant roots also have some CEC, linked to their specific surface area.[105]
Cation exchange capacity for soils; soil textures; soil colloids[106]
Soil State CEC meq/100 g
Charlotte fine sand Florida 1.0

Ruston fine sandy loam Texas 1.9

Glouchester loam New Jersey 11.9


Grundy silt loam Illinois 26.3

Gleason clay loam California 31.6

Susquehanna clay loam Alabama 34.3


Davie mucky fine sand Florida 100.8

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 is the soil's ability to remove anions (such as nitrate, phosphate) from the soil
water solution and sequester those for later exchange as the plant roots release carbonate anions to the soil
water solution.[107] Those colloids which have low CEC tend to have some AEC. Amorphous and
sesquioxide clays have the highest AEC,[108] followed by the iron oxides.[109] Levels of AEC are much
lower than for CEC, because of the generally higher rate of positively (versus negatively) charged surfaces
on soil colloids, to the exception of variable-charge soils.[110] Phosphates tend to be held at anion exchange
sites.[111]

Iron and aluminum hydroxide clays are able to exchange their hydroxide anions (OH−) for other
anions.[107] The order reflecting the strength of anion adhesion is as follows:

H2PO−4 replaces SO2− −


4 replaces NO3 replaces Cl

The amount of exchangeable anions is of a magnitude of tenths to a few milliequivalents per 100  g dry
soil.[106] 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).[112]

Reactivity (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 hydronium 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.[113]

At 25 °C an aqueous solution that has a pH of 3.5 has 10−3.5 moles H3 O+ (hydronium ions) per litre of
solution (and also 10−10.5 moles per litre OH−). A pH of 7, defined as neutral, has 10−7 moles of
hydronium 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 has 10−9.5 moles hydronium ions per litre of
solution (and also 10−2.5 moles per litre OH−). A pH of 3.5 has one million times more hydronium ions per
litre than a solution with pH of 9.5 (9.5 − 3.5 = 6 or 106 ) and is more acidic.[114]

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.[115] As a result of a trade-off between
toxicity and requirement most nutrients are better available to plants at moderate pH,[116] although 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.[117] Given that at low pH toxic
metals (e.g. cadmium, zinc, lead) are positively charged as cations and organic pollutants are in non-ionic
form, thus both made more available to organisms,[118][119] it has been suggested that plants, animals and
microbes commonly living in acid soils are pre-adapted to every kind of pollution, whether of natural or
human origin.[120]

In high rainfall areas, soils tend to acidify as the basic cations are forced off the soil colloids by the mass
action of hydronium ions from usual or unusual rain acidity against those attached to the colloids. High
rainfall rates can then wash the nutrients out, leaving the soil inhabited only by those organisms which are
particularly efficient to uptake nutrients in very acid conditions, like in tropical rainforests.[121] Once the
colloids are saturated with H3 O+, the addition of any more hydronium ions or aluminum hydroxyl cations
drives the pH even lower (more acidic) as the soil has been left with no buffering capacity.[122] In areas of
extreme rainfall and high temperatures, the clay and humus may be washed out, further reducing the
buffering capacity of the soil.[123] In low rainfall areas, unleached calcium pushes pH to 8.5 and with the
addition of exchangeable sodium, soils may reach pH  10.[124] Beyond a pH of 9, plant growth is
reduced.[125] High pH results in low micro-nutrient mobility, but water-soluble chelates of those nutrients
can correct the deficit.[126] 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.[127][128]

Base saturation percentage

There are acid-forming cations (e.g. hydronium, aluminium, iron) and there are base-forming cations (e.g.
calcium, magnesium, sodium). The fraction of the negatively-charged soil colloid exchange sites (CEC)
that are occupied by base-forming cations is called base saturation. If a soil has a CEC of 20  meq and
5  meq are aluminium and hydronium cations (acid-forming), the remainder of positions on the colloids
(20 − 5 = 15 meq) are assumed occupied by base-forming cations, so that the base saturation is
15 ÷ 20 × 100% = 75% (the compliment 25% is assumed acid-forming cations). Base saturation is almost in
direct proportion to pH (it increases with increasing pH).[129] It is of use in calculating the amount of lime
needed to neutralise an acid soil (lime requirement). The amount of lime needed to neutralize a soil must
take account of the amount of acid forming ions on the colloids (exchangeable acidity), not just those in the
soil water solution (free acidity).[130] The addition of enough lime to neutralize the soil water solution will
be insufficient to change the pH, as the acid forming cations stored on the soil colloids will tend to restore
the original pH condition as they are pushed off those colloids by the calcium of the added lime.[131]

Buffering

The resistance of soil to change in pH, as a result of the addition of acid or basic material, is a measure of
the buffering capacity of a soil and (for a particular soil type) increases as the CEC increases. Hence, pure
sand has almost no buffering ability, though soils high in colloids (whether mineral or organic) have high
buffering capacity.[132] Buffering occurs by cation exchange and neutralisation. However, colloids are not
the only regulators of soil pH. The role of carbonates should be underlined, too.[133] More generally,
according to pH levels, several buffer systems take precedence over each other, from calcium carbonate
buffer range to iron buffer range.[134]

The addition of a small amount of highly basic aqueous ammonia to a soil will cause the ammonium to
displace hydronium ions from the colloids, and the end product is water and colloidally fixed ammonium,
but little permanent change overall in soil pH.

The addition of a small amount of lime, Ca(OH)2 , will displace hydronium ions from the soil colloids,
causing the fixation of calcium to colloids and the evolution of CO2 and water, with little permanent change
in soil pH.

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 often related to the CEC of the soil; the greater the CEC, the greater the
buffering capacity of the soil.[135]

Redox

Soil chemical reactions involve some combination of proton and electron transfer. Oxidation occurs if there
is a loss of electrons in the transfer process while reduction occurs if there is a gain of electrons. Reduction
potential is measured in volts or millivolts. Soil microbial communities develop along electron transport
chains, forming electrically conductive biofilms, and developing networks of bacterial nanowires.

Redox factors in soil development, where formation of redoximorphic color features provides critical
information for soil interpretation. Understanding the redox gradient is important to managing carbon
sequestration, bioremediation, wetland delineation, and soil-based microbial fuel cells.

Nutrients
Seventeen elements or nutrients are essential for plant growth and reproduction. They are carbon (C),
hydrogen (H), oxygen (O), nitrogen (N), phosphorus (P), potassium (K), sulfur (S), calcium (Ca),
magnesium (Mg), iron (Fe), boron (B), manganese (Mn), copper (Cu), zinc (Zn), molybdenum (Mo),
nickel (Ni) and chlorine (Cl).[137][138][139] Nutrients required for plants to complete their life cycle are
considered essential nutrients. Nutrients that enhance the growth of plants but are not necessary to complete
the plant's life cycle are considered non-essential. With the exception of carbon, hydrogen and oxygen,
which are supplied by carbon dioxide and water, and nitrogen, provided through nitrogen fixation,[139] the
nutrients derive originally from the mineral component of the soil. The Law of the Minimum expresses that
when the available form of a nutrient is not in enough proportion in the soil solution, then other nutrients
cannot be taken up at an optimum rate by a Plant nutrients, their chemical symbols, and the ionic
plant.[140] A particular nutrient ratio of the soil forms common in soils and available for plant uptake[136]
solution is thus mandatory for optimizing plant Element Symbol Ion or molecule
growth, a value which might differ from nutrient
CO2 (mostly through leaves)
ratios calculated from plant composition.[141] Carbon C

Hydrogen H H+, H2O (water)


Plant uptake of nutrients can only proceed when
they are present in a plant-available form. In most Oxygen O O2−, OH−, CO2− 2−
3 , SO4 , CO2
situations, nutrients are absorbed in an ionic form
Phosphorus P − 2−
from (or together with) soil water. Although H2PO4, HPO4 (phosphates)
minerals are the origin of most nutrients, and the Potassium K K+
bulk of most nutrient elements in the soil is held in
Nitrogen N + −
crystalline form within primary and secondary NH4, NO3 (ammonium, nitrate)
minerals, they weather too slowly to support rapid Sulfur S 2−
SO4
plant growth. For example, the application of
finely ground minerals, feldspar and apatite, to soil Calcium Ca Ca2+
seldom provides the necessary amounts of
Iron Fe Fe2+, Fe3+ (ferrous, ferric)
potassium and phosphorus at a rate sufficient for
good plant growth, as most of the nutrients remain Magnesium Mg Mg2+
bound in the crystals of those minerals.[142]
Boron B H3BO3, H2BO−3, B(OH)−4
The nutrients adsorbed onto the surfaces of clay Manganese Mn Mn2+
colloids and soil organic matter provide a more
accessible reservoir of many plant nutrients (e.g. Copper Cu Cu2+
K, Ca, Mg, P, Zn). As plants absorb the nutrients Zinc Zn Zn2+
from the soil water, the soluble pool is replenished
2−
from the surface-bound pool. The decomposition Molybdenum Mo MoO4 (molybdate)
of soil organic matter by microorganisms is Chlorine Cl Cl− (chloride)
another mechanism whereby the soluble pool of
nutrients is replenished – this is important for the
supply of plant-available N, S, P, and B from soil.[143]

Gram for gram, the capacity of humus to hold nutrients and water is far greater than that of clay minerals,
most of the soil cation exchange capacity arising from charged carboxylic groups on organic matter.[144]
However, despite the great capacity of humus to retain water once water-soaked, its high hydrophobicity
decreases its wettability.[145] All in all, small amounts of humus may remarkably increase the soil's capacity
to promote plant growth.[146][143]

Soil organic matter


The organic material in soil is made up of organic compounds and includes plant, animal and microbial
material, both living and dead. A typical soil has a biomass composition of 70% microorganisms, 22%
macrofauna, and 8% roots. 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.[147]

A few percent of the soil organic matter, with small residence time, consists of the microbial biomass and
metabolites of bacteria, molds, and actinomycetes that work to break down the dead organic
matter.[148][149] Were it not for the action of these micro-organisms, the entire carbon dioxide part of the
atmosphere would be sequestered as organic matter in the soil. However, in the same time soil microbes
contribute to carbon sequestration in the topsoil through the formation of stable humus.[150] In the aim to
sequester more carbon in the soil for alleviating the greenhouse effect it would be more efficient in the long-
term to stimulate humification than to decrease litter decomposition.[151]

The main part of soil organic matter is a complex assemblage of small organic molecules, collectively called
humus or humic substances. The use of these terms, which do not rely on a clear chemical classification,
has been considered as obsolete.[152] Other studies showed that the classical notion of molecule is not
convenient for humus, which escaped most attempts done over two centuries to resolve it in unit
components, but still is chemically distinct from polysaccharides, lignins and proteins.[153]

Most living things in soils, including plants, animals, bacteria, and fungi, are dependent on organic matter
for nutrients and/or energy. Soils have organic compounds in varying degrees of decomposition which rate
is dependent on temperature, soil moisture, and aeration. Bacteria and fungi feed on the raw organic matter,
which are fed upon by protozoa, which in turn are fed upon by nematodes, annelids and arthropods,
themselves able to consume and transform raw or humified organic matter. This has been called the soil
food web, through which all organic matter is processed as in a digestive system.[154] 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.[155] In its earliest stage of decomposition, the original
organic material is often called raw organic matter. The final stage of decomposition is called humus.

In grassland, much of the organic matter added to the soil is from the deep, fibrous, grass root systems. By
contrast, tree leaves falling on the forest floor are the principal source of soil organic matter in the forest.
Another difference is the frequent occurrence in the grasslands of fires that destroy large amounts of
aboveground material but stimulate even greater contributions from roots. Also, the much greater acidity
under any forests inhibits the action of certain soil organisms that otherwise would mix much of the surface
litter into the mineral soil. As a result, the soils under grasslands generally develop a thicker A horizon with
a deeper distribution of organic matter than in comparable soils under forests, which characteristically store
most of their organic matter in the forest floor (O horizon) and thin A horizon.[156]

Humus

Humus refers to organic matter that has been decomposed by soil microflora and fauna to the 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.[157] Humus also feeds arthropods, termites and earthworms which further improve the soil.[158]
The end product, humus, is suspended in colloidal form in the soil solution and forms a weak acid that can
attack silicate minerals by chelating their iron and aluminum atoms.[159] Humus has a high cation and anion
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.[160]

Humic acids and fulvic acids, which begin as raw organic matter, are important constituents of humus.
After the death of plants, animals, and microbes, microbes begin to feed on the residues through their
production of extra-cellular soil enzymes, resulting finally in the formation of humus.[161] As the residues
break down, only molecules made of aliphatic and aromatic hydrocarbons, assembled and stabilized by
oxygen and hydrogen bonds, remain in the form of complex molecular assemblages collectively called
humus.[153] Humus is never pure in the soil, because it reacts with metals and clays to form complexes
which further contribute to its stability and to soil structure.[160] Although the structure of humus has in
itself few nutrients (with the exception of constitutive metals such as calcium, iron and aluminum) it is able
to attract and link, by weak bonds, cation and anion nutrients that can further be released into the soil
solution in response to selective root uptake and changes in soil pH, a process of paramount importance for
the maintenance of fertility in tropical soils.[162]

Lignin is resistant to breakdown and accumulates within the soil. It also reacts with proteins,[163] which
further increases its resistance to decomposition, including enzymatic decomposition by microbes.[164] Fats
and waxes from plant matter have still more resistance to decomposition and persist in soils for thousand
years, hence their use as tracers of past vegetation in buried soil layers.[165] 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.[166] Proteins normally decompose readily, to the exception of scleroproteins, but
when bound to clay particles they become more resistant to decomposition.[167] As for other proteins clay
particles absorb the enzymes exuded by microbes, decreasing enzyme activity while protecting extracellular
enzymes from degradation.[168] 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.[169] A study showed increased
soil fertility following the addition of mature compost to a clay soil.[170] High soil tannin content can cause
nitrogen to be sequestered as resistant tannin-protein complexes.[171][172]

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.[156] 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.[155] Humus also absorbs water, and expands and
shrinks between dry and wet states to a higher extent than clay, increasing soil porosity.[173] Humus is less
stable than the soil's mineral constituents, as it is reduced by microbial decomposition, and over time its
concentration diminishes without the addition of new organic matter. However, humus in its most stable
forms may persist over centuries if not millennia.[174] Charcoal is a source of highly stable humus, called
black carbon,[175] which had been used traditionally to improve the fertility of nutrient-poor tropical soils.
This very ancient practice, as ascertained in the genesis of Amazonian dark earths, has been renewed and
became popular under the name of biochar. It has been suggested that biochar could be used to sequester
more carbon in the fight against the greenhouse effect.[176]

Climatological influence

The production, accumulation and degradation of organic matter are greatly dependent on climate. For
example, when a thawing event occurs, the flux of soil gases with atmospheric gases is significantly
influenced.[177] 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[178] or excess moisture which results in anaerobic
conditions.[179] Conversely, excessive rain and high temperatures of tropical climates enables rapid
decomposition of organic matter and leaching of plant nutrients. Forest ecosystems on these soils rely on
efficient recycling of nutrients and plant matter by the living plant and microbial biomass to maintain their
productivity, a process which is disturbed by human activities.[180] Excessive slope, in particular in the
presence of cultivation for the sake of agriculture, may encourage the erosion of the top layer of soil which
holds most of the raw organic material that would otherwise eventually become humus.[181]

Plant residue

Cellulose and hemicellulose undergo fast decomposition by fungi and bacteria, with a half-life of 12–
18  days in a temperate climate.[182] Brown rot fungi can decompose the cellulose and hemicellulose,
leaving the lignin and phenolic compounds behind. Starch, which is an energy storage system for plants,
undergoes fast decomposition by bacteria and fungi. Lignin consists of
polymers composed of 500 to 600 units with a highly branched,
amorphous structure, linked to cellulose, hemicellulose and pectin in
plant cell walls. Lignin undergoes very slow decomposition, mainly by
white rot fungi and actinomycetes; its half-life under temperate
conditions is about six months.[182]

Horizons
A horizontal layer of the soil, whose physical features, composition
and age are distinct from those above and beneath, is referred to as a
soil horizon. The naming of a horizon is based on the type of material Typical types and
of which it is composed. Those materials reflect the duration of percentages of plant
specific processes of soil formation. They are labelled using a residue components
shorthand notation of letters and numbers which describe the horizon
in terms of its colour, size, texture, structure, consistency, root quantity,    Cellulose (45%)
pH, voids, boundary characteristics and presence of nodules or    Lignin (20%)
concretions.[183] No soil profile has all the major horizons. Some,    Hemicellulose (18%)
called entisols, may have only one horizon or are currently considered    Protein (8%)
as having no horizon, in particular incipient soils from unreclaimed    Sugars and starches (5%)
mining waste deposits,[184] moraines,[185] volcanic cones[186] sand
   Fats and waxes (2%)
dunes or alluvial terraces.[187] Upper soil horizons may be lacking in
truncated soils following wind or water ablation, with concomitant
downslope burying of soil horizons, a natural process aggravated by agricultural practices such as
tillage.[188] The growth of trees is another source of disturbance, creating a micro-scale heterogeneity
which is still visible in soil horizons once trees have died.[189] By passing from a horizon to another, from
the top to the bottom of the soil profile, one goes back in time, with past events registered in soil horizons
like in sediment layers. Sampling pollen, testate amoebae and plant remains in soil horizons may help to
reveal environmental changes (e.g. climate change, land use change) which occurred in the course of soil
formation.[190] Soil horizons can be dated by several methods such as radiocarbon, using pieces of charcoal
provided they are of enough size to escape pedoturbation by earthworm activity and other mechanical
disturbances.[191] Fossil soil horizons from paleosols can be found within sedimentary rock sequences,
allowing the study of past environments.[192]

The exposure of parent material to favourable conditions produces mineral soils that are marginally suitable
for plant growth, as is the case in eroded soils.[193] The growth of vegetation results in the production of
organic residues which fall on the ground as litter for plant aerial parts (leaf litter) or are directly produced
belowground for subterranean plant organs (root litter), and then release dissolved organic matter.[194] The
remaining surficial 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.[195] After sufficient time, humus moves downward
and is deposited in a distinctive organic-mineral surface layer called the A horizon, in which organic matter
is mixed with mineral matter through the activity of burrowing animals, a process called pedoturbation.
This natural process does not go to completion in the presence of conditions detrimental to soil life such as
strong acidity, cold climate or pollution, stemming in the accumulation of undecomposed organic matter
within a single organic horizon overlying the mineral soil[196] and in the juxtaposition of humified organic
matter and mineral particles, without intimate mixing, in the underlying mineral horizons.[197]

Classification
One of the first soil classification systems was developed by Russian scientist Vasily Dokuchaev around
1880.[198] It was modified a number of times by American and European researchers and was 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[199] aims to establish an international reference base for soil classification.

Uses
Soil is used in agriculture, where it serves as the anchor and primary nutrient base for plants. The types of
soil and available moisture determine the species of plants that can be cultivated. Agricultural soil science
was the primeval domain of soil knowledge, long time before the advent of pedology in the 19th century.
However, as demonstrated by aeroponics, aquaponics and hydroponics, soil material is not an absolute
essential for agriculture, and soilless cropping systems have been claimed as the future of agriculture for an
endless growing mankind.[200]

Soil material is also a critical component in mining, construction and landscape development industries.[201]
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. Many building materials are soil based. Loss
of soil through urbanization is growing at a high rate in many areas and can be critical for the maintenance
of subsistence agriculture.[202]

Soil resources are critical to the environment, as well as to food and fibre production, producing 98.8% of
food consumed by humans.[203] Soil provides minerals and water to plants according to several processes
involved in plant nutrition. Soil absorbs rainwater and releases it later, thus preventing floods and drought,
flood regulation being one of the major ecosystem services provided by soil.[204] Soil cleans water as it
percolates through it.[205] Soil is the habitat for many organisms: the major part of known and unknown
biodiversity is in the soil, in the form of 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.[206]
Above-ground and below-ground biodiversities are tightly interconnected,[156][207] 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 in deserts, cyanobacteria, lichens and mosses form biological soil crusts which
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 the effect of increases in greenhouse gas emissions and slow global warming, while
improving crop yields and reducing water needs.[208][209][210]

Waste management often has a soil component. Septic drain fields treat septic tank effluent using aerobic
soil processes. Land application of waste water relies on soil biology to aerobically treat BOD.
Alternatively, landfills use soil for daily cover, isolating waste deposits from the atmosphere and preventing
unpleasant smells. Composting is now widely used to treat aerobically solid domestic waste and dried
effluents of settling basins. Although compost is not soil, biological processes taking place during
composting are similar to those occurring during decomposition and humification of soil organic
matter.[211]
Organic soils, especially peat, serve as a significant fuel and horticultural resource. Peat soils are also
commonly used for the sake of agriculture in Nordic countries, because peatland sites, when drained,
provide fertile soils for food production.[212] However, wide areas of peat production, such as rain-fed
sphagnum bogs, also called blanket bogs or raised bogs, are now protected because of their patrimonial
interest. As an example, Flow Country, covering 4,000 square kilometres of rolling expanse of blanket bogs
in Scotland, is now candidate for being included in the World Heritage List. Under present-day global
warming peat soils are thought to be involved in a self-reinforcing (positive feedback) process of increased
emission of greenhouse gases (methane and carbon dioxide) and increased temperature,[213] a contention
which is still under debate when replaced at field scale and including stimulated plant growth.[214]

Geophagy is the practice of eating soil-like substances. Both animals and humans occasionally consume
soil for medicinal, recreational, or religious purposes.[215] It has been shown that some monkeys consume
soil, together with their preferred food (tree foliage and fruits), in order to alleviate tannin toxicity.[216]

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.[217] Soil organisms metabolise them or immobilise them in their
biomass and necromass,[218] thereby incorporating them into stable humus.[219] The physical integrity of
soil is also a prerequisite for avoiding landslides in rugged landscapes.[220]

Degradation
Land degradation is a human-induced or natural process which impairs the capacity of land to
function.[221] Soil degradation involves acidification, contamination, desertification, erosion or
salination.[222]

Acidification

Soil acidification is beneficial in the case of alkaline soils, but it degrades land when it lowers crop
productivity, soil biological activity and increases soil vulnerability to contamination and erosion. Soils are
initially acid and remain such when their parent materials are low in basic cations (calcium, magnesium,
potassium and sodium). On parent materials richer in weatherable minerals acidification occurs when basic
cations are leached from the soil profile by rainfall or exported by 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. Deforestation is another cause of soil acidification, mediated by increased leaching of
soil nutrients in the absence of tree canopies.[223]

Contamination

Soil contamination at low levels is often within a soil's capacity to treat and assimilate waste material. Soil
biota can treat waste by transforming it, mainly through microbial enzymatic activity.[224] Soil organic
matter and soil minerals can adsorb the waste material and decrease its toxicity,[225] although when in
colloidal form they may transport the adsorbed contaminants to subsurface environments.[226] Many waste
treatment processes rely on this natural bioremediation 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,
soil conditioners, phytoremediation, bioremediation and Monitored Natural Attenuation. An example of
diffuse pollution with contaminants is copper accumulation in vineyards and orchards to which fungicides
are repeatedly applied, even in organic farming.[227]

Microfibres from synthetic textiles are another type of plastic soil contamination, 100% of agricultural soil
samples from southwestern China contained plastic particles, 92% of which were microfibres. Sources of
microfibres likely included string or twine, as well as irrigation water in which clothes had been
washed.[228]

The application of biosolids from sewage sludge and compost can introduce microplastics to soils. This
adds to the burden of microplastics from other sources (e.g. the atmosphere). Approximately half the
sewage sludge in Europe and North America is applied to agricultural land. In Europe it has been estimated
that for every million inhabitants 113 to 770 tonnes of microplastics are added to agricultural soils each
year.[228]

Desertification

Desertification, an environmental process of ecosystem degradation


in arid and semi-arid regions, is often caused by badly adapted
human activities such as overgrazing or excess harvesting of
firewood. It is a common misconception that drought causes
desertification.[229] 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.[230]
These practices help to control erosion and maintain productivity
during periods when moisture is available. Continued land abuse Desertification
during droughts, however, increases land degradation. Increased
population and livestock pressure on marginal lands accelerates
desertification.[231] It is now questioned whether present-day climate warming will favour or disfavour
desertification, with contradictory reports about predicted rainfall trends associated with increased
temperature, and strong discrepancies among regions, even in the same country.[232]

Erosion

Erosion of soil is caused by water, wind, ice, and movement in response to


gravity. More than one kind of erosion can occur simultaneously. Erosion is
distinguished from weathering, since erosion also transports eroded soil
away from its place of origin (soil in transit may be described as sediment).
Erosion is an intrinsic natural process, but in many places it is greatly
increased by human activity, especially unsuitable land use practices.[233]
These include agricultural activities which leave the soil bare during times
of heavy rain or strong winds, overgrazing, deforestation, and improper
construction activity. Improved management can limit erosion. Soil
conservation techniques which are employed include changes of land use
(such as replacing erosion-prone crops with grass or other soil-binding
plants), changes to the timing or type of agricultural operations, terrace
building, use of erosion-suppressing cover materials (including cover crops Erosion control
and other plants), limiting disturbance during construction, and avoiding
construction during erosion-prone periods and in erosion-prone places such as steep slopes.[234]
Historically, one of the best examples of large-scale soil erosion due to unsuitable land-use practices is wind
erosion (the so-called dust bowl) which ruined American and Canadian prairies during the 1930s, when
immigrant farmers, encouraged by the federal government of both countries, settled and converted the
original shortgrass prairie to agricultural crops and cattle ranching.

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.[235]

Soil piping is a particular form of soil erosion that occurs below the soil surface.[236] It causes levee and
dam failure, as well as sink hole formation. Turbulent flow removes soil starting at the mouth of the seep
flow and the subsoil erosion advances up-gradient.[237] The term sand boil is used to describe the
appearance of the discharging end of an active soil pipe.[238]

Salination

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.[239] 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.[240][241]

Reclamation
Soils which contain high levels of particular clays with high swelling properties, such as smectites, are often
very fertile. For example, the smectite-rich paddy soils of Thailand's Central Plains are among the most
productive in the world. However, the overuse of mineral nitrogen fertilizers and pesticides in irrigated
intensive rice production has endangered these soils, forcing farmers to implement integrated practices
based on Cost Reduction Operating Principles.[242]

Many farmers in tropical areas, however, struggle to retain organic matter and clay in the soils they work. In
recent years, for example, productivity has declined and soil erosion has increased in the low-clay soils of
northern Thailand, following the abandonment of shifting cultivation for a more permanent land use.[243]
Farmers initially responded by adding organic matter and clay from termite mound material, but this was
unsustainable in the long-term because of rarefaction of termite mounds. 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 (IWMI) 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 kilograms per rai (1,300 kg/ha; 1,100 lb/acre) of bentonite resulted in an
average yield increase of 73%.[244] Other studies showed that applying bentonite to degraded sandy soils
reduced the risk of crop failure during drought years.[245]
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.[246]

If the soil is too high in clay or salts (e.g. saline sodic soil), adding gypsum, washed river sand and organic
matter (e.g.municipal solid waste) will balance the composition.[247]

Adding organic matter, like ramial chipped wood or compost, to soil which is depleted in nutrients and too
high in sand will boost its quality and improve production.[248][249]

Special mention must be made of the use of charcoal, and more generally biochar to improve nutrient-poor
tropical soils, a process based on the higher fertility of anthropogenic pre-Columbian Amazonian Dark
Earths, also called Terra Preta de Índio, due to interesting physical and chemical properties of soil black
carbon as a source of stable humus.[250] However, the uncontrolled application of charred waste products
of all kinds may endanger soil life and human health.[251]

History of studies and research


The history of the study of soil is intimately tied to humans' urgent need to provide food for themselves and
forage for their animals. Throughout history, civilizations have prospered or declined as a function of the
availability and productivity of their soils.[252]

Studies of soil fertility

The Greek historian Xenophon (450–355 BCE) 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.'[253]

Columella's Of husbandry, circa 60  CE, advocated the use of lime and that clover and alfalfa (green
manure) should be turned under,[254] and was used by 15 generations (450 years) under the Roman Empire
until its collapse.[253][255] 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
Middle Ages, Yahya Ibn al-'Awwam's handbook,[256] with its emphasis on irrigation, guided the people of
North Africa, Spain and the Middle East; a translation of this work was finally carried to the southwest of
the United States when under Spanish influence.[257] Olivier de Serres, considered the father of French
agronomy, was the first to suggest the abandonment of fallowing and its replacement by hay meadows
within crop rotations. He also highlighted the importance of soil (the French terroir) in the management of
vineyards. His famous book Le Théâtre d'Agriculture et mesnage des champs[258] contributed to the rise of
modern, sustainable agriculture and to the collapse of old agricultural practices such as soil amendment for
crops by the lifting of forest litter and assarting, which ruined the soils of western Europe during the Middle
Ages and even later on according to regions.[259]

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, a belief which prevailed until the 19th century.[260] 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 apparently been produced only by the addition of water, with no
reduction in the soil's weight.[261][262][263] 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. Still 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 (stir) the soil, but his opinion that the stirring made the fine parts of soil available for plant
absorption was erroneous.[262][264]

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. He was
able to deduce that most of the 165-pound (75  kg) weight of van Helmont's willow tree derived from
air.[265] 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 air and
water, while nitrogen was taken from soil.[266] Justus von Liebig in his book Organic chemistry in its
applications to agriculture and physiology (published 1840), 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.[267] 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.[268]

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 (re)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.[269] This led to the invention and use of
salts of potassium (K) and nitrogen (N) as fertilizers. Ammonia generated by the production of coke was
recovered and used as fertiliser.[270] Finally, the chemical basis of nutrients delivered to the soil in manure
was understood and in the mid-19th century chemical fertilisers were applied. However, the dynamic
interaction of soil and its life forms was still not understood.

In 1856, J. Thomas Way discovered that ammonia contained in fertilisers was transformed into nitrates,[271]
and twenty years later Robert Warington proved that this transformation was done by living organisms.[272]
In 1890 Sergei Winogradsky announced he had found the bacteria responsible for this transformation.[273]

It was known that certain legumes could take up nitrogen from the air and fix it to the soil but it took the
development of bacteriology towards the end of the 19th century to lead 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 the German agronomist Hermann Hellriegel
and the Dutch microbiologist Martinus Beijerinck.[269]

Crop rotation, mechanisation, chemical and natural fertilisers led to a doubling of wheat yields in western
Europe between 1800 and 1900.[274]

Studies of 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, under the
action of biotic and abiotic processes. After studies of the improvement of the soil commenced, other
researchers began to study soil genesis and as a result also soil types and classifications.
In 1860, while in Mississippi, Eugene W. Hilgard (1833–1916) studied the relationship between rock
material, climate, vegetation, and the type of soils that were developed. He realised that the soils were
dynamic, and considered the classification of soil types.[275] His work was not continued. At about the
same time, Friedrich Albert Fallou was describing soil profiles and relating soil characteristics to their
formation as part of his professional work evaluating forest and farm land for the principality of Saxony.
His 1857 book, Anfangsgründe der Bodenkunde (First principles of soil science) established modern soil
science.[276] Contemporary with Fallou's work, and driven by the same need to accurately assess land for
equitable taxation, Vasily Dokuchaev led a team of soil scientists in Russia who conducted an extensive
survey of soils, observing that similar basic rocks, climate and vegetation types lead to similar soil layering
and types, and established the concepts for soil classifications. Due to language barriers, the work of this
team was not communicated to western Europe until 1914 through a publication in German by Konstantin
Glinka, a member of the Russian team.[277]

Curtis F. Marbut, influenced by the work of the Russian team, translated Glinka's publication into
English,[278] and, as he was placed in charge of the U.S. National Cooperative Soil Survey, applied it to a
national soil classification system.[262]

See also
Environment
portal
Geology portal

Acid sulfate soil


Agrophysics
Crust
Agricultural science
Factors affecting permeability of soils
Index of soil-related articles
Mycorrhizal fungi and soil carbon storage
Shrink–swell capacity
Soil biodiversity
Soil liquefaction
Soil moisture velocity equation
Soil zoology
Tillage erosion
World Soil Museum
Red soil

References
1. Voroney, R. Paul; Heck, Richard J. (2007). "The soil habitat" (https://fr.art1lib.org/book/34240
339/73458e). In Paul, Eldor A. (ed.). Soil microbiology, ecology and biochemistry (3rd ed.).
Amsterdam, the Netherlands: Elsevier. pp. 25–49. doi:10.1016/B978-0-08-047514-1.50006-
8 (https://doi.org/10.1016%2FB978-0-08-047514-1.50006-8). ISBN 978-0-12-546807-7.
Archived (https://web.archive.org/web/20180710102532/http://csmi.issas.ac.cn/uploadfiles/S
oil%20Microbiology%2C%20Ecology%20%26%20Biochemistry.pdf) (PDF) from the original
on 10 July 2018. Retrieved 27 March 2022.
2. Taylor, Sterling A.; Ashcroft, Gaylen L. (1972). Physical edaphology: the physics of irrigated
and nonirrigated soils (https://archive.org/details/physicaledapholo0000tayl). San Francisco,
California: W.H. Freeman. ISBN 978-0-7167-0818-6.
3. McCarthy, David F. (2014). Essentials of soil mechanics and foundations: basic geotechnics
(https://fr.book4you.org/book/3555343/0f8f97) (7th ed.). London, United Kingdom: Pearson.
ISBN 9781292039398. Retrieved 27 March 2022.
4. Gilluly, James; Waters, Aaron Clement; Woodford, Alfred Oswald (1975). Principles of
geology (https://archive.org/details/principlesofgeol0000gill) (4th ed.). San Francisco,
California: W.H. Freeman. ISBN 978-0-7167-0269-6.
5. Huggett, Richard John (2011). "What is geomorphology?". Fundamentals of geomorphology
(https://cc1lib.vip/book/1220195/5e924c). Routledge Fundamentals of Physical Geography
Series (3rd ed.). London, United Kingdom: Routledge. pp. 148–150. ISBN 978-0-203-86008-
3. Retrieved 16 October 2022.
6. Ponge, Jean-François (2015). "The soil as an ecosystem" (https://www.researchgate.net/pub
lication/276090499). Biology and Fertility of Soils. 51 (6): 645–648. doi:10.1007/s00374-
015-1016-1 (https://doi.org/10.1007%2Fs00374-015-1016-1). S2CID 18251180 (https://api.s
emanticscholar.org/CorpusID:18251180). Retrieved 3 April 2022.
7. Yu, Charley; Kamboj, Sunita; Wang, Cheng; Cheng, Jing-Jy (2015). "Data collection
handbook to support modeling impacts of radioactive material in soil and building structures"
(https://resrad.evs.anl.gov/docs/data_collection.pdf) (PDF). Argonne National Laboratory.
pp. 13–21. Archived (https://web.archive.org/web/20180804105951/http://resrad.evs.anl.gov/
docs/data_collection.pdf) (PDF) from the original on 4 August 2018. Retrieved 3 April 2022.
8. Buol, Stanley W.; Southard, Randal J.; Graham, Robert C.; McDaniel, Paul A. (2011). Soil
genesis and classification (https://fr1lib.org/book/2156097/707d35) (6th ed.). Ames, Iowa:
Wiley-Blackwell. ISBN 978-0-470-96060-8. Retrieved 3 April 2022.
9. Retallack, Gregory J.; Krinsley, David H.; Fischer, Robert; Razink, Joshua J.; Langworthy,
Kurt A. (2016). "Archean coastal-plain paleosols and life on land" (https://cpb-us-e1.wpmucd
n.com/blogs.uoregon.edu/dist/d/3735/files/2013/07/Retallack-et-al.-2016-Farrel-1gt7uft.pdf)
(PDF). Gondwana Research. 40: 1–20. Bibcode:2016GondR..40....1R (https://ui.adsabs.har
vard.edu/abs/2016GondR..40....1R). doi:10.1016/j.gr.2016.08.003 (https://doi.org/10.1016%2
Fj.gr.2016.08.003). Archived (https://web.archive.org/web/20181113075710/https://cpb-us-e
1.wpmucdn.com/blogs.uoregon.edu/dist/d/3735/files/2013/07/Retallack-et-al.-2016-Farrel-1
gt7uft.pdf) (PDF) from the original on 13 November 2018. Retrieved 3 April 2022.
10. Chesworth, Ward, ed. (2008). Encyclopedia of soil science (https://fr1lib.org/book/563235/8e
916e) (1st ed.). Dordrecht, The Netherlands: Springer. ISBN 978-1-4020-3994-2. Archived (h
ttps://web.archive.org/web/20180905002957/http://www.encyclopedias.biz/dw/Encyclopedi
a%20of%20Soil%20Science.pdf) (PDF) from the original on 5 September 2018. Retrieved
27 March 2022.
11. "Glossary of terms in soil science" (https://sis.agr.gc.ca/cansis/glossary/e/index.html).
Agriculture and Agri-Food Canada. 13 December 2013. Archived (https://web.archive.org/we
b/20181027045042/http://sis.agr.gc.ca/cansis/glossary/e/index.html) from the original on 27
October 2018. Retrieved 3 April 2022.
12. Amundson, Ronald. "Soil preservation and the future of pedology" (https://web.archive.org/w
eb/20180612140029/http://natres.psu.ac.th/Link/SoilCongress/bdd/symp45/75-t.pdf) (PDF).
CiteSeerX 10.1.1.552.237 (https://citeseerx.ist.psu.edu/viewdoc/summary?doi=10.1.1.552.2
37). Archived from the original (http://natres.psu.ac.th/Link/SoilCongress/bdd/symp45/75-t.pd
f) (PDF) on 12 June 2018.
13. Küppers, Michael; Vincent, Jean-Baptiste. "Impacts and formation of regolith" (https://www.m
ps.mpg.de/phd/planetary-interiors-and-surfaces-2011-part-05). Max Planck Institute for Solar
System Research. Archived (https://web.archive.org/web/20180804200824/https://www.mps.
mpg.de/phd/planetary-interiors-and-surfaces-2011-part-05) from the original on 4 August
2018. Retrieved 3 April 2022.
14. Amelung, Wulf; Bossio, Deborah; De Vries, Wim; Kögel-Knabner, Ingrid; Lehmann,
Johannes; Amundson, Ronald; Bol, Roland; Collins, Chris; Lal, Rattan; Leifeld, Jens;
Minasny, Buniman; Pan, Gen-Xing; Paustian, Keith; Rumpel, Cornelia; Sanderman,
Jonathan; Van Groeningen, Jan Willem; Mooney, Siân; Van Wesemael, Bas; Wander,
Michelle; Chabbi, Abad (27 October 2020). "Towards a global-scale soil climate mitigation
strategy" (https://www.nature.com/articles/s41467-020-18887-7.pdf) (PDF). Nature
Communications. 11 (1): 5427. Bibcode:2020NatCo..11.5427A (https://ui.adsabs.harvard.ed
u/abs/2020NatCo..11.5427A). doi:10.1038/s41467-020-18887-7 (https://doi.org/10.1038%2F
s41467-020-18887-7). ISSN 2041-1723 (https://www.worldcat.org/issn/2041-1723).
PMC 7591914 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7591914). PMID 33110065
(https://pubmed.ncbi.nlm.nih.gov/33110065). Retrieved 3 April 2022.
15. Pouyat, Richard; Groffman, Peter; Yesilonis, Ian; Hernandez, Luis (2002). "Soil carbon pools
and fluxes in urban ecosystems" (https://www.researchgate.net/publication/11526697).
Environmental Pollution. 116 (Supplement 1): S107–S118. doi:10.1016/S0269-
7491(01)00263-9 (https://doi.org/10.1016%2FS0269-7491%2801%2900263-9).
PMID 11833898 (https://pubmed.ncbi.nlm.nih.gov/11833898). Retrieved 3 April 2022. "Our
analysis of pedon data from several disturbed soil profiles suggests that physical
disturbances and anthropogenic inputs of various materials (direct effects) can greatly alter
the amount of C stored in these human "made" soils."
16. Davidson, Eric A.; Janssens, Ivan A. (2006). "Temperature sensitivity of soil carbon
decomposition and feedbacks to climate change" (https://www.nature.com/articles/nature045
14.pdf) (PDF). Nature. 440 (9 March 2006): 165‒73. Bibcode:2006Natur.440..165D (https://u
i.adsabs.harvard.edu/abs/2006Natur.440..165D). doi:10.1038/nature04514 (https://doi.org/1
0.1038%2Fnature04514). PMID 16525463 (https://pubmed.ncbi.nlm.nih.gov/16525463).
S2CID 4404915 (https://api.semanticscholar.org/CorpusID:4404915). Retrieved 3 April
2022.
17. Powlson, David (2005). "Will soil amplify climate change?" (https://fr.art1lib.org/book/105433
01/528a68). Nature. 433 (20 January 2005): 204‒05. Bibcode:2005Natur.433..204P (https://
ui.adsabs.harvard.edu/abs/2005Natur.433..204P). doi:10.1038/433204a (https://doi.org/10.1
038%2F433204a). PMID 15662396 (https://pubmed.ncbi.nlm.nih.gov/15662396).
S2CID 35007042 (https://api.semanticscholar.org/CorpusID:35007042). Retrieved 3 April
2022.
18. Bradford, Mark A.; Wieder, William R.; Bonan, Gordon B.; Fierer, Noah; Raymond, Peter A.;
Crowther, Thomas W. (2016). "Managing uncertainty in soil carbon feedbacks to climate
change" (http://fiererlab.org/wp-content/uploads/2014/09/Bradford_etal_2016_NCC.pdf)
(PDF). Nature Climate Change. 6 (27 July 2016): 751–758. Bibcode:2016NatCC...6..751B
(https://ui.adsabs.harvard.edu/abs/2016NatCC...6..751B). doi:10.1038/nclimate3071 (https://
doi.org/10.1038%2Fnclimate3071). hdl:20.500.11755/c1792dbf-ce96-4dc7-8851-
1ca50a35e5e0 (https://hdl.handle.net/20.500.11755%2Fc1792dbf-ce96-4dc7-8851-1ca50a
35e5e0). S2CID 43955196 (https://api.semanticscholar.org/CorpusID:43955196). Retrieved
3 April 2022.
19. Dominati, Estelle; Patterson, Murray; Mackay, Alec (2010). "A framework for classifying and
quantifying the natural capital and ecosystem services of soils" (https://www.researchgate.ne
t/publication/223852147). Ecological Economics. 69 (9): 1858‒68.
doi:10.1016/j.ecolecon.2010.05.002 (https://doi.org/10.1016%2Fj.ecolecon.2010.05.002).
Archived (https://web.archive.org/web/20170808082847/http://esanalysis.colmex.mx/Sorte
d%20Papers/2010/2010%20NZL%20-3F%20Phys.pdf) (PDF) from the original on 8 August
2017. Retrieved 10 April 2022.
20. Dykhuizen, Daniel E. (1998). "Santa Rosalia revisited: why are there so many species of
bacteria?" (https://www.researchgate.net/publication/13682480). Antonie van Leeuwenhoek.
73 (1): 25‒33. doi:10.1023/A:1000665216662 (https://doi.org/10.1023%2FA%3A100066521
6662). PMID 9602276 (https://pubmed.ncbi.nlm.nih.gov/9602276). S2CID 17779069 (https://
api.semanticscholar.org/CorpusID:17779069). Retrieved 10 April 2022.
21. Torsvik, Vigdis; Øvreås, Lise (2002). "Microbial diversity and function in soil: from genes to
ecosystems" (https://www.academia.edu/13038690). Current Opinion in Microbiology. 5 (3):
240‒45. doi:10.1016/S1369-5274(02)00324-7 (https://doi.org/10.1016%2FS1369-5274%28
02%2900324-7). PMID 12057676 (https://pubmed.ncbi.nlm.nih.gov/12057676). Retrieved
10 April 2022.
22. Raynaud, Xavier; Nunan, Naoise (2014). "Spatial ecology of bacteria at the microscale in
soil" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3905020). PLOS ONE. 9 (1): e87217.
Bibcode:2014PLoSO...987217R (https://ui.adsabs.harvard.edu/abs/2014PLoSO...987217R).
doi:10.1371/journal.pone.0087217 (https://doi.org/10.1371%2Fjournal.pone.0087217).
PMC 3905020 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3905020). PMID 24489873
(https://pubmed.ncbi.nlm.nih.gov/24489873).
23. Whitman, William B.; Coleman, David C.; Wiebe, William J. (1998). "Prokaryotes: the
unseen majority" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC33863). Proceedings of the
National Academy of Sciences of the USA. 95 (12): 6578‒83.
Bibcode:1998PNAS...95.6578W (https://ui.adsabs.harvard.edu/abs/1998PNAS...95.6578W).
doi:10.1073/pnas.95.12.6578 (https://doi.org/10.1073%2Fpnas.95.12.6578). PMC 33863 (htt
ps://www.ncbi.nlm.nih.gov/pmc/articles/PMC33863). PMID 9618454 (https://pubmed.ncbi.nl
m.nih.gov/9618454).
24. Schlesinger, William H.; Andrews, Jeffrey A. (2000). "Soil respiration and the global carbon
cycle" (https://www.researchgate.net/publication/51997678). Biogeochemistry. 48 (1): 7‒20.
doi:10.1023/A:1006247623877 (https://doi.org/10.1023%2FA%3A1006247623877).
S2CID 94252768 (https://api.semanticscholar.org/CorpusID:94252768). Retrieved 10 April
2022.
25. Denmead, Owen Thomas; Shaw, Robert Harold (1962). "Availability of soil water to plants
as affected by soil moisture content and meteorological conditions" (https://www.researchgat
e.net/publication/250098028). Agronomy Journal. 54 (5): 385‒90.
doi:10.2134/agronj1962.00021962005400050005x (https://doi.org/10.2134%2Fagronj1962.
00021962005400050005x). Retrieved 10 April 2022.
26. House, Christopher H.; Bergmann, Ben A.; Stomp, Anne-Marie; Frederick, Douglas J.
(1999). "Combining constructed wetlands and aquatic and soil filters for reclamation and
reuse of water" (https://www.researchgate.net/publication/222464331). Ecological
Engineering. 12 (1–2): 27–38. doi:10.1016/S0925-8574(98)00052-4 (https://doi.org/10.101
6%2FS0925-8574%2898%2900052-4). Retrieved 10 April 2022.
27. Van Bruggen, Ariena H.C.; Semenov, Alexander M. (2000). "In search of biological indicators
for soil health and disease suppression" (https://www.researchgate.net/publication/2225209
30). Applied Soil Ecology. 15 (1): 13–24. doi:10.1016/S0929-1393(00)00068-8 (https://doi.or
g/10.1016%2FS0929-1393%2800%2900068-8). Retrieved 10 April 2022.
28. "Community guide to monitored natural attenuation" (https://semspub.epa.gov/work/HQ/401
611.pdf) (PDF). Retrieved 10 April 2022.
29. Linn, Daniel Myron; Doran, John W. (1984). "Effect of water-filled pore space on carbon
dioxide and nitrous oxide production in tilled and nontilled soils" (https://fr.art1lib.org/book/23
108771/821c3f). Soil Science Society of America Journal. 48 (6): 1267–1272.
Bibcode:1984SSASJ..48.1267L (https://ui.adsabs.harvard.edu/abs/1984SSASJ..48.1267L).
doi:10.2136/sssaj1984.03615995004800060013x (https://doi.org/10.2136%2Fsssaj1984.03
615995004800060013x). Retrieved 10 April 2022.
30. Gregory, Peter J.; Nortcliff, Stephen (2013). Soil conditions and plant growth (https://fr.book4
you.org/book/2156095/fd863f). Hoboken, New Jersey: Wiley-Blackwell.
ISBN 9781405197700. Retrieved 10 April 2022.
31. Bot, Alexandra; Benites, José (2005). The importance of soil organic matter: key to drought-
resistant soil and sustained food and production (http://www.fao.org/3/a-a0100e.pdf) (PDF).
Rome: Food and Agriculture Organization of the United Nations. ISBN 978-92-5-105366-9.
Retrieved 10 April 2022.
32. McClellan, Tai. "Soil composition" (https://www.ctahr.hawaii.edu/mauisoil/a_comp.aspx).
University of Hawaiʻi at Mānoa, College of Tropical Agriculture and Human Resources.
Retrieved 18 April 2022.
33. "Arizona Master Gardener Manual" (https://web.archive.org/web/20160529015259/http://ag.
arizona.edu/pubs/garden/mg/soils/soils.html). Cooperative Extension, College of Agriculture,
University of Arizona. 9 November 2017. Archived from the original (http://ag.arizona.edu/pu
bs/garden/mg/soils/soils.html) on 29 May 2016. Retrieved 17 December 2017.
34. Vannier, Guy (1987). "The porosphere as an ecological medium emphasized in Professor
Ghilarov's work on soil animal adaptations" (https://link.springer.com/content/pdf/10.1007/BF
00260577.pdf) (PDF). Biology and Fertility of Soils. 3 (1): 39–44. doi:10.1007/BF00260577
(https://doi.org/10.1007%2FBF00260577). S2CID 297400 (https://api.semanticscholar.org/C
orpusID:297400). Retrieved 18 April 2022.
35. Torbert, H. Allen; Wood, Wes (1992). "Effect of soil compaction and water-filled pore space
on soil microbial activity and N losses" (https://www.researchgate.net/publication/24054613
2). Communications in Soil Science and Plant Analysis. 23 (11): 1321‒31.
doi:10.1080/00103629209368668 (https://doi.org/10.1080%2F00103629209368668).
Retrieved 18 April 2022.
36. Simonson 1957, p. 17.
37. Zanella, Augusto; Katzensteiner, Klaus; Ponge, Jean-François; Jabiol, Bernard; Sartori,
Giacomo; Kolb, Eckart; Le Bayon, Renée-Claire; Aubert, Michaël; Ascher-Jenull, Judith;
Englisch, Michael; Hager, Herbert (June 2019). "TerrHum: an iOS App for classifying
terrestrial humipedons and some considerations about soil classification" (https://www.resea
rchgate.net/publication/332080061). Soil Science Society of America Journal. 83 (S1): S42–
S48. doi:10.2136/sssaj2018.07.0279 (https://doi.org/10.2136%2Fsssaj2018.07.0279).
hdl:11577/3315165 (https://hdl.handle.net/11577%2F3315165). S2CID 197555747 (https://a
pi.semanticscholar.org/CorpusID:197555747). Retrieved 18 April 2022.
38. Bronick, Carol J.; Lal, Ratan (January 2005). "Soil structure and management: a review" (htt
p://tinread.usarb.md:8888/tinread/fulltext/lal/soil_structure.pdf) (PDF). Geoderma. 124 (1–2):
3–22. Bibcode:2005Geode.124....3B (https://ui.adsabs.harvard.edu/abs/2005Geode.124....3
B). doi:10.1016/j.geoderma.2004.03.005 (https://doi.org/10.1016%2Fj.geoderma.2004.03.00
5). Retrieved 18 April 2022.
39. "Soil and water" (https://www.fao.org/3/r4082e/r4082e03.htm). Food and Agriculture
Organization of the United Nations. Retrieved 18 April 2022.
40. Valentin, Christian; d'Herbès, Jean-Marc; Poesen, Jean (1999). "Soil and water components
of banded vegetation patterns" (https://www.academia.edu/35300713). Catena. 37 (1): 1‒24.
Bibcode:1999Caten..37....1V (https://ui.adsabs.harvard.edu/abs/1999Caten..37....1V).
doi:10.1016/S0341-8162(99)00053-3 (https://doi.org/10.1016%2FS0341-8162%2899%2900
053-3). Retrieved 18 April 2022.
41. Brady, Nyle C.; Weil, Ray R. (2007). "The colloidal fraction: seat of soil chemical and
physical activity" (https://www.researchgate.net/publication/309630422). In Brady, Nyle C.;
Weil, Ray R. (eds.). The nature and properties of soils (14th ed.). London, United Kingdom:
Pearson. pp. 310–357. ISBN 978-0132279383. Retrieved 18 April 2022.
42. "Soil colloids: properties, nature, types and significance" (http://eagri.org/eagri50/SSAC121/l
ec14.pdf) (PDF). Tamil Nadu Agricultural University. Retrieved 18 April 2022.
43. Miller, Jarrod O. "Soil pH affects nutrient availability" (https://www.researchgate.net/publicati
on/305775103). Retrieved 18 April 2022.
44. Goulding, Keith W.T.; Bailey, Neal J.; Bradbury, Nicola J.; Hargreaves, Patrick; Howe, M.T.;
Murphy, Daniel V.; Poulton, Paul R.; Willison, Toby W. (1998). "Nitrogen deposition and its
contribution to nitrogen cycling and associated soil processes" (https://doi.org/10.1046%2Fj.
1469-8137.1998.00182.x). New Phytologist. 139 (1): 49‒58. doi:10.1046/j.1469-
8137.1998.00182.x (https://doi.org/10.1046%2Fj.1469-8137.1998.00182.x).
45. Kononova, M.M. (2013). Soil organic matter: its nature, its role in soil formation and in soil
fertility (https://fr1lib.org/book/2275488/ea4395) (2nd ed.). Amsterdam, The Netherlands:
Elsevier. ISBN 978-1-4831-8568-2. Retrieved 24 April 2022.
46. Burns, Richards G.; DeForest, Jared L.; Marxsen, Jürgen; Sinsabaugh, Robert L.;
Stromberger, Mary E.; Wallenstein, Matthew D.; Weintraub, Michael N.; Zoppini, Annamaria
(2013). "Soil enzymes in a changing environment: current knowledge and future directions"
(https://www.academia.edu/25235991). Soil Biology and Biochemistry. 58: 216‒34.
doi:10.1016/j.soilbio.2012.11.009 (https://doi.org/10.1016%2Fj.soilbio.2012.11.009).
Retrieved 24 April 2022.
47. Sengupta, Aditi; Kushwaha, Priyanka; Jim, Antonia; Troch, Peter A.; Maier, Raina (2020).
"New soil, old plants, and ubiquitous microbes: evaluating the potential of incipient basaltic
soil to support native plant growth and influence belowground soil microbial community
composition" (https://doi.org/10.3390%2Fsu12104209). Sustainability. 12 (10): 4209.
doi:10.3390/su12104209 (https://doi.org/10.3390%2Fsu12104209).
48. Bishop, Janice L.; Murchie, Scott L.; Pieters, Carlé L.; Zent, Aaron P. (2002). "A model for
formation of dust, soil, and rock coatings on Mars: physical and chemical processes on the
Martian surface" (https://doi.org/10.1029%2F2001JE001581). Journal of Geophysical
Research. 107 (E11): 7-1–7-17. Bibcode:2002JGRE..107.5097B (https://ui.adsabs.harvard.e
du/abs/2002JGRE..107.5097B). doi:10.1029/2001JE001581 (https://doi.org/10.1029%2F20
01JE001581).
49. Navarro-González, Rafael; Rainey, Fred A.; Molina, Paola; Bagaley, Danielle R.; Hollen,
Becky J.; de la Rosa, José; Small, Alanna M.; Quinn, Richard C.; Grunthaner, Frank J.;
Cáceres, Luis; Gomez-Silva, Benito; McKay, Christopher P. (2003). "Mars-like soils in the
Atacama desert, Chile, and the dry limit of microbial life" (https://www.researchgate.net/publi
cation/9020258). Science. 302 (5647): 1018–1021. Bibcode:2003Sci...302.1018N (https://ui.
adsabs.harvard.edu/abs/2003Sci...302.1018N). doi:10.1126/science.1089143 (https://doi.or
g/10.1126%2Fscience.1089143). PMID 14605363 (https://pubmed.ncbi.nlm.nih.gov/146053
63). S2CID 18220447 (https://api.semanticscholar.org/CorpusID:18220447). Retrieved
24 April 2022.
50. Guo, Yong; Fujimura, Reiko; Sato, Yoshinori; Suda, Wataru; Kim, Seok-won; Oshima,
Kenshiro; Hattori, Masahira; Kamijo, Takashi; Narisawa, Kazuhiko; Ohta, Hiroyuki (2014).
"Characterization of early microbial communities on volcanic deposits along a vegetation
gradient on the island of Miyake, Japan" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC404
1228). Microbes and Environments. 29 (1): 38–49. doi:10.1264/jsme2.ME13142 (https://doi.o
rg/10.1264%2Fjsme2.ME13142). PMC 4041228 (https://www.ncbi.nlm.nih.gov/pmc/articles/
PMC4041228). PMID 24463576 (https://pubmed.ncbi.nlm.nih.gov/24463576).
51. Goldich, Samuel S. (1938). "A study in rock-weathering" (https://fr.art1lib.org/book/60175497/
a54b2b). The Journal of Geology. 46 (1): 17–58. Bibcode:1938JG.....46...17G (https://ui.adsa
bs.harvard.edu/abs/1938JG.....46...17G). doi:10.1086/624619 (https://doi.org/10.1086%2F62
4619). ISSN 0022-1376 (https://www.worldcat.org/issn/0022-1376). S2CID 128498195 (http
s://api.semanticscholar.org/CorpusID:128498195). Retrieved 24 April 2022.
52. Van Schöll, Laura; Smits, Mark M.; Hoffland, Ellis (2006). "Ectomycorrhizal weathering of the
soil minerals muscovite and hornblende" (https://doi.org/10.1111%2Fj.1469-8137.2006.017
90.x). New Phytologist. 171 (4): 805–814. doi:10.1111/j.1469-8137.2006.01790.x (https://doi.
org/10.1111%2Fj.1469-8137.2006.01790.x). PMID 16918551 (https://pubmed.ncbi.nlm.nih.g
ov/16918551).
53. Stretch, Rachelle C.; Viles, Heather A. (2002). "The nature and rate of weathering by lichens
on lava flows on Lanzarote" (https://fr.art1lib.org/book/17831662/8253cd). Geomorphology.
47 (1): 87–94. Bibcode:2002Geomo..47...87S (https://ui.adsabs.harvard.edu/abs/2002Geom
o..47...87S). doi:10.1016/S0169-555X(02)00143-5 (https://doi.org/10.1016%2FS0169-555
X%2802%2900143-5). Retrieved 24 April 2022.
54. Dojani, Stephanie; Lakatos, Michael; Rascher, Uwe; Waneck, Wolfgang; Luettge, Ulrich;
Büdel, Burkhard (2007). "Nitrogen input by cyanobacterial biofilms of an inselberg into a
tropical rainforest in French Guiana" (https://www.researchgate.net/publication/224026482).
Flora. 202 (7): 521–529. doi:10.1016/j.flora.2006.12.001 (https://doi.org/10.1016%2Fj.flora.2
006.12.001). Retrieved 21 March 2021.
55. Kabala, Cesary; Kubicz, Justyna (2012). "Initial soil development and carbon accumulation
on moraines of the rapidly retreating Werenskiold Glacier, SW Spitsbergen, Svalbard
archipelago" (https://www.academia.edu/31221217). Geoderma. 175–176: 9–20.
Bibcode:2012Geode.175....9K (https://ui.adsabs.harvard.edu/abs/2012Geode.175....9K).
doi:10.1016/j.geoderma.2012.01.025 (https://doi.org/10.1016%2Fj.geoderma.2012.01.025).
Retrieved 24 April 2022.
56. Jenny, Hans (1941). Factors of soil formation: a system of qunatitative pedology (http://neted
u.xauat.edu.cn/sykc/hjx/content/ckzl/6/2.pdf) (PDF). New York: McGraw-Hill. Archived (http
s://web.archive.org/web/20170808104008/http://netedu.xauat.edu.cn/sykc/hjx/content/ckzl/6/
2.pdf) (PDF) from the original on 8 August 2017. Retrieved 24 April 2022.
57. Ritter, Michael E. "The physical environment: an introduction to physical geography" (http://w
ww.tsu-excel4ed.org/reviews/Geography%20Template_The%20Physical%20Environment_
Cunha.pdf) (PDF). Retrieved 24 April 2022.
58. Gardner, Catriona M.K.; Laryea, Kofi Buna; Unger, Paul W. (1999). Soil physical constraints
to plant growth and crop production (https://web.archive.org/web/20170808175354/http://ww
w.plantstress.com/Files/Soil_Physical_Constraints.pdf) (PDF) (first ed.). Rome, Italy: Food
and Agriculture Organization of the United Nations. Archived from the original (http://www.pla
ntstress.com/Files/Soil_Physical_Constraints.pdf) (PDF) on 8 August 2017.
59. Six, Johan; Paustian, Keith; Elliott, Edward T.; Combrink, Clay (2000). "Soil structure and
organic matter. I. Distribution of aggregate-size classes and aggregate-associated carbon"
(https://www.researchgate.net/publication/280798601). Soil Science Society of America
Journal. 64 (2): 681–689. Bibcode:2000SSASJ..64..681S (https://ui.adsabs.harvard.edu/abs/
2000SSASJ..64..681S). doi:10.2136/sssaj2000.642681x (https://doi.org/10.2136%2Fsssaj2
000.642681x). Retrieved 7 August 2022.
60. Håkansson, Inge; Lipiec, Jerzy (2000). "A review of the usefulness of relative bulk density
values in studies of soil structure and compaction" (http://directory.umm.ac.id/Data%20Elmu/
jurnal/S/Soil%20&%20Tillage%20Research/Vol53.Issue2.Jan2000/1452.pdf) (PDF). Soil
and Tillage Research. 53 (2): 71–85. doi:10.1016/S0167-1987(99)00095-1 (https://doi.org/1
0.1016%2FS0167-1987%2899%2900095-1). S2CID 30045538 (https://api.semanticscholar.
org/CorpusID:30045538). Retrieved 7 August 2022.
61. Schwerdtfeger, William J. (1965). "Soil resistivity as related to underground corrosion and
cathodic protection" (https://nvlpubs.nist.gov/nistpubs/jres/69C/jresv69Cn1p71_A1b.pdf)
(PDF). Journal of Research of the National Bureau of Standards. 69C (1): 71–77.
doi:10.6028/jres.069c.012 (https://doi.org/10.6028%2Fjres.069c.012). Retrieved 7 August
2022.
62. Tamboli, Prabhakar Mahadeo (1961). The influence of bulk density and aggregate size on
soil moisture retention (https://dr.lib.iastate.edu/bitstreams/85621186-4b03-4140-ad1c-b18c3
ab3b4a8/download). Ames, Iowa: Iowa State University. Retrieved 7 August 2022.
63. Spehn, Eva M.; Joshi, Jasmin; Schmid, Bernhard; Alphei, Jörn; Körner, Christian (2000).
"Plant diversity effects on soil heterotrophic activity in experimental grassland ecosystems"
(http://link.springer.com/10.1023/A:1004891807664). Plant and Soil. 224 (2): 217–230.
doi:10.1023/A:1004891807664 (https://doi.org/10.1023%2FA%3A1004891807664).
S2CID 25639544 (https://api.semanticscholar.org/CorpusID:25639544).
64. "Water holding capacity" (https://forages.oregonstate.edu/ssis/soils/characteristics/water-hol
ding-capacity). Oregon State University. 24 June 2016. Retrieved 9 October 2022. "Irrigators
must have knowledge of the readily available moisture capacity so that water can be applied
before plants have to expend excessive energy to extract moisture"
65. "Basics of irrigation scheduling" (https://extension.umn.edu/irrigation/basics-irrigation-sched
uling). University of Minnesota Extension. Retrieved 9 October 2022. "Only a portion of the
available water holding capacity is easily used by the crop before crop water stress develop"
66. Qi, Jingen; Marshall, John D.; Mattson, Kim G. (1994). "High soil carbon dioxide
concentrations inhibit root respiration of Douglas fir" (https://doi.org/10.1111%2Fj.1469-813
7.1994.tb02989.x). New Phytologist. 128 (3): 435–442. doi:10.1111/j.1469-
8137.1994.tb02989.x (https://doi.org/10.1111%2Fj.1469-8137.1994.tb02989.x).
PMID 33874575 (https://pubmed.ncbi.nlm.nih.gov/33874575).
67. Karberg, Noah J.; Pregitzer, Kurt S.; King, John S.; Friend, Aaron L.; Wood, James R. (2005).
"Soil carbon dioxide partial pressure and dissolved inorganic carbonate chemistry under
elevated carbon dioxide and ozone" (https://www.researchgate.net/publication/8337234).
Oecologia. 142 (2): 296–306. Bibcode:2005Oecol.142..296K (https://ui.adsabs.harvard.edu/
abs/2005Oecol.142..296K). doi:10.1007/s00442-004-1665-5 (https://doi.org/10.1007%2Fs00
442-004-1665-5). PMID 15378342 (https://pubmed.ncbi.nlm.nih.gov/15378342).
S2CID 6161016 (https://api.semanticscholar.org/CorpusID:6161016). Retrieved
13 November 2022.
68. Chang, H.T.; Loomis, Walter E. (1945). "Effect of carbon dioxide on absorption of water and
nutrients by roots" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC437214). Plant
Physiology. 20 (2): 221–232. doi:10.1104/pp.20.2.221 (https://doi.org/10.1104%2Fpp.20.2.2
21). PMC 437214 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC437214). PMID 16653979
(https://pubmed.ncbi.nlm.nih.gov/16653979).
69. McDowell, Nate J.; Marshall, John D.; Qi, Jingen; Mattson, Kim (1999). "Direct inhibition of
maintenance respiration in western hemlock roots exposed to ambient soil carbon dioxide
concentrations" (https://doi.org/10.1093%2Ftreephys%2F19.9.599). Tree Physiology. 19 (9):
599–605. doi:10.1093/treephys/19.9.599 (https://doi.org/10.1093%2Ftreephys%2F19.9.599).
PMID 12651534 (https://pubmed.ncbi.nlm.nih.gov/12651534).
70. Xu, Xia; Nieber, John L.; Gupta, Satish C. (1992). "Compaction effect on the gas diffusion
coefficient in soils" (https://www.academia.edu/6547475). Soil Science Society of America
Journal. 56 (6): 1743–1750. Bibcode:1992SSASJ..56.1743X (https://ui.adsabs.harvard.edu/
abs/1992SSASJ..56.1743X). doi:10.2136/sssaj1992.03615995005600060014x (https://doi.o
rg/10.2136%2Fsssaj1992.03615995005600060014x). Retrieved 13 November 2022.
71. Smith, Keith A.; Ball, Tom; Conen, Franz; Dobbie, Karen E.; Massheder, Jonathan; Rey, Ana
(2003). "Exchange of greenhouse gases between soil and atmosphere: interactions of soil
physical factors and biological processes" (https://www.academia.edu/14433607). European
Journal of Soil Science. 54 (4): 779–791. doi:10.1046/j.1351-0754.2003.0567.x (https://doi.o
rg/10.1046%2Fj.1351-0754.2003.0567.x). S2CID 18442559 (https://api.semanticscholar.org/
CorpusID:18442559). Retrieved 13 November 2022.
72. Russell 1957, pp. 35–36.
73. Ruser, Reiner; Flessa, Heiner; Russow, Rolf; Schmidt, G.; Buegger, Franz; Munch, J.C.
(2006). "Emission of N2O, N2 and CO2 from soil fertilized with nitrate: effect of compaction,
soil moisture and rewetting" (https://www.sciencedirect.com/science/article/abs/pii/S003807
1705001975). Soil Biology and Biochemistry. 38 (2): 263–274.
doi:10.1016/j.soilbio.2005.05.005 (https://doi.org/10.1016%2Fj.soilbio.2005.05.005).
74. Hartmann, Adrian A.; Buchmann, Nina; Niklaus, Pascal A. (2011). "A study of soil methane
sink regulation in two grasslands exposed to drought and N fertilization" (https://www.resear
ch-collection.ethz.ch/bitstream/handle/20.500.11850/34759/2/11104_2010_Article_690.pdf)
(PDF). Plant and Soil. 342 (1–2): 265–275. doi:10.1007/s11104-010-0690-x (https://doi.org/1
0.1007%2Fs11104-010-0690-x). hdl:20.500.11850/34759 (https://hdl.handle.net/20.500.118
50%2F34759). S2CID 25691034 (https://api.semanticscholar.org/CorpusID:25691034).
Retrieved 13 November 2022.
75. Moore, Tim R.; Dalva, Moshe (1993). "The influence of temperature and water table position
on carbon dioxide and methane emissions from laboratory columns of peatland soils" (http
s://www.researchgate.net/publication/229878721). Journal of Soil Science. 44 (4): 651–664.
doi:10.1111/j.1365-2389.1993.tb02330.x (https://doi.org/10.1111%2Fj.1365-2389.1993.tb02
330.x). Retrieved 13 November 2022.
76. Hiltpold, Ivan; Toepfer, Stefan; Kuhlmann, Ulrich; Turlings, Ted C.J. (2010). "How maize root
volatiles affect the efficacy of entomopathogenic nematodes in controlling the western corn
rootworm?" (https://www.researchgate.net/publication/215470509). Chemoecology. 20 (2):
155–162. doi:10.1007/s00049-009-0034-6 (https://doi.org/10.1007%2Fs00049-009-0034-6).
S2CID 30214059 (https://api.semanticscholar.org/CorpusID:30214059). Retrieved
13 November 2022.
77. Ryu, Choong-Min; Farag, Mohamed A.; Hu, Chia-Hui; Reddy, Munagala S.; Wei, Han-Xun;
Paré, Paul W.; Kloepper, Joseph W. (2003). "Bacterial volatiles promote growth in
Arabidopsis" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC153657). Proceedings of the
National Academy of Sciences of the United States of America. 100 (8): 4927–4932.
Bibcode:2003PNAS..100.4927R (https://ui.adsabs.harvard.edu/abs/2003PNAS..100.4927
R). doi:10.1073/pnas.0730845100 (https://doi.org/10.1073%2Fpnas.0730845100).
PMC 153657 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC153657). PMID 12684534 (http
s://pubmed.ncbi.nlm.nih.gov/12684534).
78. Hung, Richard; Lee, Samantha; Bennett, Joan W. (2015). "Fungal volatile organic
compounds and their role in ecosystems" (https://www.researchgate.net/publication/273638
498). Applied Microbiology and Biotechnology. 99 (8): 3395–3405. doi:10.1007/s00253-015-
6494-4 (https://doi.org/10.1007%2Fs00253-015-6494-4). PMID 25773975 (https://pubmed.n
cbi.nlm.nih.gov/25773975). S2CID 14509047 (https://api.semanticscholar.org/CorpusID:145
09047). Retrieved 13 November 2022.
79. Purrington, Foster Forbes; Kendall, Paricia A.; Bater, John E.; Stinner, Benjamin R. (1991).
"Alarm pheromone in a gregarious poduromorph collembolan (Collembola:
Hypogastruridae)" (https://scholar.valpo.edu/cgi/viewcontent.cgi?article=1732&context=tgle).
Great Lakes Entomologist. 24 (2): 75–78. Retrieved 13 November 2022.
80. Badri, Dayakar V.; Weir, Tiffany L.; Van der Lelie, Daniel; Vivanco, Jorge M (2009).
"Rhizosphere chemical dialogues: plant–microbe interactions" (http://www.bicga.org.uk/doc
s/Rhizosphere_chemical_dialogues_plant.pdf) (PDF). Current Opinion in Biotechnology. 20
(6): 642–650. doi:10.1016/j.copbio.2009.09.014 (https://doi.org/10.1016%2Fj.copbio.2009.0
9.014). PMID 19875278 (https://pubmed.ncbi.nlm.nih.gov/19875278). Retrieved
13 November 2022.
81. Salmon, Sandrine; Ponge, Jean-François (2001). "Earthworm excreta attract soil springtails:
laboratory experiments on Heteromurus nitidus (Collembola: Entomobryidae)" (https://www.
academia.edu/20508985). Soil Biology and Biochemistry. 33 (14): 1959–1969.
doi:10.1016/S0038-0717(01)00129-8 (https://doi.org/10.1016%2FS0038-0717%2801%2900
129-8). S2CID 26647480 (https://api.semanticscholar.org/CorpusID:26647480). Retrieved
13 November 2022.
82. Lambers, Hans; Mougel, Christophe; Jaillard, Benoît; Hinsinger, Philipe (2009). "Plant-
microbe-soil interactions in the rhizosphere: an evolutionary perspective" (https://www.acade
mia.edu/25517742). Plant and Soil. 321 (1–2): 83–115. doi:10.1007/s11104-009-0042-x (htt
ps://doi.org/10.1007%2Fs11104-009-0042-x). S2CID 6840457 (https://api.semanticscholar.o
rg/CorpusID:6840457). Retrieved 13 November 2022.
83. Peñuelas, Josep; Asensio, Dolores; Tholl, Dorothea; Wenke, Katrin; Rosenkranz, Maaria;
Piechulla, Birgit; Schnitzler, Jörg-Petter (2014). "Biogenic volatile emissions from the soil" (h
ttps://doi.org/10.1111%2Fpce.12340). Plant, Cell and Environment. 37 (8): 1866–1891.
doi:10.1111/pce.12340 (https://doi.org/10.1111%2Fpce.12340). PMID 24689847 (https://pub
med.ncbi.nlm.nih.gov/24689847).
84. Buzuleciu, Samuel A.; Crane, Derek P.; Parker, Scott L. (2016). "Scent of disinterred soil as
an olfactory cue used by raccoons to locate nests of diamond-backed terrapins (Malaclemys
terrapin)" (http://www.herpconbio.org/Volume_11/Issue_3/Buzuleciu_etal_2016.pdf) (PDF).
Herpetological Conservation and Biology. 11 (3): 539–551. Retrieved 27 November 2022.
85. Saxton, Keith E.; Rawls, Walter J. (2006). "Soil water characteristic estimates by texture and
organic matter for hydrologic solutions" (https://hrsl.ba.ars.usda.gov/SPAW/Soil%20Water%2
0Characteristics-Paper.pdf) (PDF). Soil Science Society of America Journal. 70 (5): 1569–
1578. Bibcode:2006SSASJ..70.1569S (https://ui.adsabs.harvard.edu/abs/2006SSASJ..70.1
569S). doi:10.2136/sssaj2005.0117 (https://doi.org/10.2136%2Fsssaj2005.0117).
S2CID 16826314 (https://api.semanticscholar.org/CorpusID:16826314). Archived (https://we
b.archive.org/web/20180902183902/https://pdfs.semanticscholar.org/5e63/c886c4f68af5e5c
242c006d2d882f0a65bfe.pdf) (PDF) from the original on 2 September 2018. Retrieved
15 January 2023.
86. College of Tropical Agriculture and Human Resources. "Soil mineralogy" (https://www.ctahr.
hawaii.edu/mauisoil/a_factor_mineralogy.aspx). University of Hawaiʻi at Mānoa. Retrieved
15 January 2023.
87. Sposito, Garrison (1984). The surface chemistry of soils (https://epdf.pub/the-surface-chemist
ry-of-soils.html). New York: Oxford University Press. Retrieved 15 January 2023.
88. Wynot, Christopher. "Theory of diffusion in colloidal suspensions" (http://www.owlnet.rice.ed
u/~ceng402/proj02/cwynot/402project.htm). Retrieved 15 January 2023.
89. Donahue, Miller & Shickluna 1977, p. 103–106.
90. Sposito, Garrison; Skipper, Neal T.; Sutton, Rebecca; Park, Sung-Ho; Soper, Alan K.;
Greathouse, Jeffery A. (1999). "Surface geochemistry of the clay minerals" (https://www.ncbi.
nlm.nih.gov/pmc/articles/PMC34275). Proceedings of the National Academy of Sciences of
the United States of America. 96 (7): 3358–3364. Bibcode:1999PNAS...96.3358S (https://ui.
adsabs.harvard.edu/abs/1999PNAS...96.3358S). doi:10.1073/pnas.96.7.3358 (https://doi.or
g/10.1073%2Fpnas.96.7.3358). PMC 34275 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC
34275). PMID 10097044 (https://pubmed.ncbi.nlm.nih.gov/10097044).
91. Bickmore, Barry R.; Rosso, Kevin M.; Nagy, Kathryn L.; Cygan, Randall T.; Tadanier,
Christopher J. (2003). "Ab initio determination of edge surface structures for dioctahedral 2:1
phyllosilicates: implications for acid-base reactivity" (http://randallcygan.com/wp-content/upl
oads/2017/06/Bickmore2003CCM.pdf) (PDF). Clays and Clay Minerals. 51 (4): 359–371.
Bibcode:2003CCM....51..359B (https://ui.adsabs.harvard.edu/abs/2003CCM....51..359B).
doi:10.1346/CCMN.2003.0510401 (https://doi.org/10.1346%2FCCMN.2003.0510401).
S2CID 97428106 (https://api.semanticscholar.org/CorpusID:97428106). Retrieved
15 January 2023.
92. Rajamathi, Michael; Thomas, Grace S.; Kamath, P. Vishnu (2001). "The many ways of
making anionic clays" (https://www.academia.edu/56207482). Journal of Chemical
Sciences. 113 (5–6): 671–680. doi:10.1007/BF02708799 (https://doi.org/10.1007%2FBF027
08799). S2CID 97507578 (https://api.semanticscholar.org/CorpusID:97507578). Retrieved
15 January 2023.
93. Moayedi, Hossein; Kazemian, Sina (2012). "Zeta potentials of suspended humus in
multivalent cationic saline solution and its effect on electro-osomosis behavior" (https://www.
academia.edu/10587240). Journal of Dispersion Science and Technology. 34 (2): 283–294.
doi:10.1080/01932691.2011.646601 (https://doi.org/10.1080%2F01932691.2011.646601).
S2CID 94333872 (https://api.semanticscholar.org/CorpusID:94333872). Retrieved
15 January 2023.
94. Pettit, Robert E. "Organic matter, humus, humate, humic acid, fulvic acid and humin: their
importance in soil fertility and plant health" (http://www.harvestgrow.com/.pdf%20web%20sit
e/Humates%20General%20Info.pdf) (PDF). Retrieved 15 January 2023.
95. Diamond, Sidney; Kinter, Earl B. (1965). "Mechanisms of soil-lime stabilization: an
interpretive review" (http://onlinepubs.trb.org/onlinepubs/hrr/1965/92/92-006.pdf) (PDF).
Highway Research Record. 92: 83–102. Retrieved 15 January 2023.
96. Woodruff, Clarence M. (1955). "The energies of replacement of calcium by potassium in
soils" (https://www.ipipotash.org/uploads/pdf/review/30_1956_1.pdf) (PDF). Soil Science
Society of America Journal. 19 (2): 167–171. Bibcode:1955SSASJ..19..167W (https://ui.adsa
bs.harvard.edu/abs/1955SSASJ..19..167W).
doi:10.2136/sssaj1955.03615995001900020014x (https://doi.org/10.2136%2Fsssaj1955.03
615995001900020014x). Retrieved 15 January 2023.
97. Fronæus, Sture (1953). "On the application of the mass action law to cation exchange
equilibria" (https://doi.org/10.3891%2Facta.chem.scand.07-0469). Acta Chemica
Scandinavica. 7: 469–480. doi:10.3891/acta.chem.scand.07-0469 (https://doi.org/10.3891%
2Facta.chem.scand.07-0469).
98. Bolland, Mike D. A.; Posner, Alan M.; Quirk, James P. (1980). "pH-independent and pH-
dependent surface charges on kaolinite" (https://www.researchgate.net/publication/2372946
35). Clays and Clay Minerals. 28 (6): 412–418. Bibcode:1980CCM....28..412B (https://ui.ads
abs.harvard.edu/abs/1980CCM....28..412B). doi:10.1346/CCMN.1980.0280602 (https://doi.o
rg/10.1346%2FCCMN.1980.0280602). S2CID 12462516 (https://api.semanticscholar.org/Co
rpusID:12462516). Retrieved 15 January 2023.
99. Chakraborty, Meghna (8 August 2022). "What is cation exchange capacity in soils?" (http://w
ww.soilminerals.com/Cation_Exchange_Simplified.htm). Retrieved 15 January 2023.
100. Silber, Avner; Levkovitch, Irit; Graber, Ellen R. (2010). "pH-dependent mineral release and
surface properties of cornstraw biochar: agronomic implications" (https://www.academia.edu/
24532141). Environmental Science and Technology. 44 (24): 9318–23.
Bibcode:2010EnST...44.9318S (https://ui.adsabs.harvard.edu/abs/2010EnST...44.9318S).
doi:10.1021/es101283d (https://doi.org/10.1021%2Fes101283d). PMID 21090742 (https://pu
bmed.ncbi.nlm.nih.gov/21090742). Retrieved 15 January 2023.
101. Dakora, Felix D.; Phillips, Donald D. (2002). "Root exudates as mediators of mineral
acquisition in low-nutrient environments" (https://www.researchgate.net/publication/2252657
45). Plant and Soil. 245: 35–47. doi:10.1023/A:1020809400075 (https://doi.org/10.1023%2F
A%3A1020809400075). S2CID 3330737 (https://api.semanticscholar.org/CorpusID:333073
7). Archived (https://web.archive.org/web/20190819123707/http://www.plantstress.com/articl
es/min_deficiency_i/root_exudates.pdf) (PDF) from the original on 19 August 2019.
Retrieved 15 January 2023.
102. Brown, John C. (1978). "Mechanism of iron uptake by plants" (https://booksc.me/book/93180
43/764ac6). Plant, Cell and Environment. 1 (4): 249–257. doi:10.1111/j.1365-
3040.1978.tb02037.x (https://doi.org/10.1111%2Fj.1365-3040.1978.tb02037.x). Retrieved
29 January 2023.
103. Donahue, Miller & Shickluna 1977, p. 114.
104. Singh, Jamuna Sharan; Raghubanshi, Akhilesh Singh; Singh, Raj S.; Srivastava, S. C.
(1989). "Microbial biomass acts as a source of plant nutrient in dry tropical forest and
savanna" (https://www.researchgate.net/publication/236941524). Nature. 338 (6215): 499–
500. Bibcode:1989Natur.338..499S (https://ui.adsabs.harvard.edu/abs/1989Natur.338..499
S). doi:10.1038/338499a0 (https://doi.org/10.1038%2F338499a0). S2CID 4301023 (https://a
pi.semanticscholar.org/CorpusID:4301023). Retrieved 29 January 2023.
105. Szatanik-Kloc, Alicja; Szerement, Justyna; Józefaciuk, Grzegorz (2017). "The role of cell
walls and pectins in cation exchange and surface area of plant roots" (https://booksc.me/boo
k/65260543/5df35d). Journal of Plant Physiology. 215: 85–90.
doi:10.1016/j.jplph.2017.05.017 (https://doi.org/10.1016%2Fj.jplph.2017.05.017).
PMID 28600926 (https://pubmed.ncbi.nlm.nih.gov/28600926). Retrieved 29 January 2023.
106. Donahue, Miller & Shickluna 1977, pp. 115–116.
107. Hinsinger, Philippe (2001). "Bioavailability of soil inorganic P in the rhizosphere as affected
by root-induced chemical changes: a review" (https://www.researchgate.net/publication/225
852665). Plant and Soil. 237 (2): 173–95. doi:10.1023/A:1013351617532 (https://doi.org/10.
1023%2FA%3A1013351617532). S2CID 8562338 (https://api.semanticscholar.org/CorpusI
D:8562338). Retrieved 29 January 2023.
108. Gu, Baohua; Schulz, Robert K. (1991). "Anion retention in soil: possible application to
reduce migration of buried technetium and iodine, a review" (https://www.osti.gov/servlets/pu
rl/5980032). doi:10.2172/5980032 (https://doi.org/10.2172%2F5980032). S2CID 91359494
(https://api.semanticscholar.org/CorpusID:91359494). Retrieved 29 January 2023.
109. Lawrinenko, Michael; Jing, Dapeng; Banik, Chumki; Laird, David A. (2017). "Aluminum and
iron biomass pretreatment impacts on biochar anion exchange capacity" (https://www.acade
mia.edu/90757446). Carbon. 118: 422–30. doi:10.1016/j.carbon.2017.03.056 (https://doi.org/
10.1016%2Fj.carbon.2017.03.056). Retrieved 29 January 2023.
110. Sollins, Phillip; Robertson, G. Philip; Uehara, Goro (1988). "Nutrient mobility in variable- and
permanent-charge soils" (https://lter.kbs.msu.edu/docs/robertson/Sollins_et_al._1988_Bioge
ochemistry.pdf) (PDF). Biogeochemistry. 6 (3): 181–99. doi:10.1007/BF02182995 (https://do
i.org/10.1007%2FBF02182995). S2CID 4505438 (https://api.semanticscholar.org/CorpusID:
4505438). Retrieved 29 January 2023.
111. Sanders, W. M. H. (1964). "Extraction of soil phosphate by anion-exchange membrane" (http
s://doi.org/10.1080%2F00288233.1964.10416423). New Zealand Journal of Agricultural
Research. 7 (3): 427–31. doi:10.1080/00288233.1964.10416423 (https://doi.org/10.1080%2
F00288233.1964.10416423).
112. Lawrinenko, Mike; Laird, David A. (2015). "Anion exchange capacity of biochar" (https://ww
w.researchgate.net/publication/280973853). Green Chemistry. 17 (9): 4628–36.
doi:10.1039/C5GC00828J (https://doi.org/10.1039%2FC5GC00828J). S2CID 52972476 (htt
ps://api.semanticscholar.org/CorpusID:52972476). Retrieved 29 January 2023.
113. Robertson, Bryan. "pH requirements of freshwater aquatic life" (https://www.waterboards.ca.
gov/centralvalley/water_issues/basin_plans/ph_turbidity/ph_turbidity_04phreq.pdf) (PDF).
Retrieved 6 June 2021.
114. Chang, Raymond, ed. (2010). Chemistry (https://www.academia.edu/44394574). Chemistry -
Chang 12Ed (12th ed.). New York, New York: McGraw-Hill. p. 666. ISBN 9780078021510.
Retrieved 6 June 2021.
115. Singleton, Peter L.; Edmeades, Doug C.; Smart, R. E.; Wheeler, David M. (2001). "The many
ways of making anionic clays" (https://doi.org/10.1007%2FBF02708799). Journal of
Chemical Sciences. 113 (5–6): 671–680. doi:10.1007/BF02708799 (https://doi.org/10.100
7%2FBF02708799). S2CID 97507578 (https://api.semanticscholar.org/CorpusID:9750757
8).
116. Läuchli, André; Grattan, Steve R. (2012). "Soil pH extremes" (https://www.researchgate.net/p
ublication/269112359). In Shabala, Sergey (ed.). Plant stress physiology (1st ed.).
Wallingford, United Kingdom: CAB International. pp. 194–209.
doi:10.1079/9781845939953.0194 (https://doi.org/10.1079%2F9781845939953.0194).
ISBN 978-1845939953. Retrieved 13 June 2021.
117. Donahue, Miller & Shickluna 1977, pp. 116–117.
118. Calmano, Wolfgang; Hong, Jihua; Förstner, Ulrich (1993). "Binding and mobilization of
heavy metals in contaminated sediments affected by pH and redox potential" (https://www.re
searchgate.net/publication/234056281). Water Science and Technology. 28 (8–9): 223–235.
doi:10.2166/wst.1993.0622 (https://doi.org/10.2166%2Fwst.1993.0622). Retrieved 13 June
2021.
119. Ren, Xiaoya; Zeng, Guangming; Tang, Lin; Wang, Jingjing; Wan, Jia; Liu, Yani; Yu,
Jiangfang; Yi, Huan; Ye, Shujing; Deng, Rui (2018). "Sorption, transport and biodegradation:
an insight into bioavailability of persistent organic pollutants in soil" (http://ee.hnu.edu.cn/__l
ocal/E/E3/44/F76DCA19501AE153573A22D4C29_17709BE2_110161.pdf) (PDF). Science
of the Total Environment. 610–611: 1154–1163. Bibcode:2018ScTEn.610.1154R (https://ui.
adsabs.harvard.edu/abs/2018ScTEn.610.1154R). doi:10.1016/j.scitotenv.2017.08.089 (http
s://doi.org/10.1016%2Fj.scitotenv.2017.08.089). PMID 28847136 (https://pubmed.ncbi.nlm.ni
h.gov/28847136). Retrieved 13 June 2021.
120. Ponge, Jean-François (2003). "Humus forms in terrestrial ecosystems: a framework to
biodiversity" (https://www.academia.edu/20508983). Soil Biology and Biochemistry. 35 (7):
935–945. CiteSeerX 10.1.1.467.4937 (https://citeseerx.ist.psu.edu/viewdoc/summary?doi=1
0.1.1.467.4937). doi:10.1016/S0038-0717(03)00149-4 (https://doi.org/10.1016%2FS0038-07
17%2803%2900149-4). S2CID 44160220 (https://api.semanticscholar.org/CorpusID:441602
20). Retrieved 13 June 2021.
121. Fujii, Kazumichi (2003). "Soil acidification and adaptations of plants and microorganisms in
Bornean tropical forests" (https://doi.org/10.1007%2Fs11284-014-1144-3). Ecological
Research. 29 (3): 371–381. doi:10.1007/s11284-014-1144-3 (https://doi.org/10.1007%2Fs11
284-014-1144-3).
122. Kauppi, Pekka; Kämäri, Juha; Posch, Maximilian; Kauppi, Lea (1986). "Acidification of forest
soils: model development and application for analyzing impacts of acidic deposition in
Europe" (http://pure.iiasa.ac.at/id/eprint/2766/1/RR-87-05.pdf) (PDF). Ecological Modelling.
33 (2–4): 231–253. doi:10.1016/0304-3800(86)90042-6 (https://doi.org/10.1016%2F0304-38
00%2886%2990042-6). Retrieved 13 June 2021.
123. Andriesse, Jacobus Pieter (1969). "A study of the environment and characteristics of tropical
podzols in Sarawak (East-Malaysia)" (https://coek.info/pdf-a-study-of-the-environment-and-c
haracteristics-of-tropical-podzols-in-sarawak-ea.html). Geoderma. 2 (3): 201–227.
Bibcode:1969Geode...2..201A (https://ui.adsabs.harvard.edu/abs/1969Geode...2..201A).
doi:10.1016/0016-7061(69)90038-X (https://doi.org/10.1016%2F0016-7061%2869%299003
8-X). Retrieved 13 June 2021.
124. Rengasamy, Pichu (2006). "World salinization with emphasis on Australia" (https://doi.org/1
0.1093%2Fjxb%2Ferj108). Journal of Experimental Botany. 57 (5): 1017–1023.
doi:10.1093/jxb/erj108 (https://doi.org/10.1093%2Fjxb%2Ferj108). PMID 16510516 (https://p
ubmed.ncbi.nlm.nih.gov/16510516).
125. Arnon, Daniel I.; Johnson, Clarence M. (1942). "Influence of hydrogen ion concentration on
the growth of higher plants under controlled conditions" (https://www.ncbi.nlm.nih.gov/pmc/ar
ticles/PMC438054). Plant Physiology. 17 (4): 525–539. doi:10.1104/pp.17.4.525 (https://doi.
org/10.1104%2Fpp.17.4.525). PMC 438054 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC
438054). PMID 16653803 (https://pubmed.ncbi.nlm.nih.gov/16653803).
126. Chaney, Rufus L.; Brown, John C.; Tiffin, Lee O. (1972). "Obligatory reduction of ferric
chelates in iron uptake by soybeans" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC36611
1). Plant Physiology. 50 (2): 208–213. doi:10.1104/pp.50.2.208 (https://doi.org/10.1104%2Fp
p.50.2.208). PMC 366111 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC366111).
PMID 16658143 (https://pubmed.ncbi.nlm.nih.gov/16658143).
127. Donahue, Miller & Shickluna 1977, pp. 116–119.
128. Ahmad, Sagheer; Ghafoor, Abdul; Qadir, Manzoor; Aziz, M. Abbas (2006). "Amelioration of a
calcareous saline-sodic soil by gypsum application and different crop rotations" (https://ww
w.researchgate.net/publication/228966353). International Journal of Agriculture and Biology.
8 (2): 142–46. Retrieved 13 June 2021.
129. McFee, William W.; Kelly, J. Michael; Beck, Robert H. (1977). "Acid precipitation effects on
soil pH and base saturation of exchange sites" (https://doi.org/10.1007%2FBF00284134).
Water, Air, and Soil Pollution. 7 (3): 4014–08. Bibcode:1977WASP....7..401M (https://ui.adsa
bs.harvard.edu/abs/1977WASP....7..401M). doi:10.1007/BF00284134 (https://doi.org/10.100
7%2FBF00284134).
130. Farina, Martin Patrick W.; Sumner, Malcolm E.; Plank, C. Owen; Letzsch, W. Stephen (1980).
"Exchangeable aluminum and pH as indicators of lime requirement for corn" (https://www.re
searchgate.net/publication/250123873). Soil Science Society of America Journal. 44 (5):
1036–1041. Bibcode:1980SSASJ..44.1036F (https://ui.adsabs.harvard.edu/abs/1980SSAS
J..44.1036F). doi:10.2136/sssaj1980.03615995004400050033x (https://doi.org/10.2136%2F
sssaj1980.03615995004400050033x). Retrieved 20 June 2021.
131. Donahue, Miller & Shickluna 1977, pp. 119–120.
132. Sposito, Garrison; Skipper, Neal T.; Sutton, Rebecca; Park, Sun-Ho; Soper, Alan K.;
Greathouse, Jeffery A. (1999). "Surface geochemistry of the clay minerals" (https://www.ncbi.
nlm.nih.gov/pmc/articles/PMC34275). Proceedings of the National Academy of Sciences of
the United States of America. 96 (7): 3358–3364. Bibcode:1999PNAS...96.3358S (https://ui.
adsabs.harvard.edu/abs/1999PNAS...96.3358S). doi:10.1073/pnas.96.7.3358 (https://doi.or
g/10.1073%2Fpnas.96.7.3358). PMC 34275 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC
34275). PMID 10097044 (https://pubmed.ncbi.nlm.nih.gov/10097044).
133. Sparks, Donald L. "Acidic and basic soils: buffering" (https://lawr.ucdavis.edu/classes/ssc10
2/Section8.pdf) (PDF). Davis, California: University of California, Davis, Department of Land,
Air, and Water Resources. Retrieved 20 June 2021.
134. Ulrich, Bernhard (1983). "Soil acidity and its relations to acid deposition" (https://rd.springer.c
om/content/pdf/10.1007%2F978-94-009-6983-4_10.pdf) (PDF). In Ulrich, Bernhard;
Pankrath, Jürgen (eds.). Effects of accumulation of air pollutants in forest ecosystems
(1st ed.). Dordrecht, The Netherlands: D. Reidel Publishing Company. pp. 127–146.
doi:10.1007/978-94-009-6983-4_10 (https://doi.org/10.1007%2F978-94-009-6983-4_10).
ISBN 978-94-009-6985-8. Retrieved 21 June 2021.
135. Donahue, Miller & Shickluna 1977, pp. 120–121.
136. Donahue, Miller & Shickluna 1977, p. 125.
137. Dean 1957, p. 80.
138. Russel 1957, pp. 123–125.
139. Brady, Nyle C.; Weil, Ray R. (2008). The nature and properties of soils (https://www.research
gate.net/publication/301200878) (15th ed.). Upper Saddle River, New Jersey: Pearson.
ISBN 978-0-13-325448-8. Retrieved 27 June 2021.
140. Van der Ploeg, Rienk R.; Böhm, Wolfgang; Kirkham, Mary Beth (1999). "On the origin of the
theory of mineral nutrition of plants and the Law of the Minimum". Soil Science Society of
America Journal. 63 (5): 1055–1062. Bibcode:1999SSASJ..63.1055V (https://ui.adsabs.harv
ard.edu/abs/1999SSASJ..63.1055V). CiteSeerX 10.1.1.475.7392 (https://citeseerx.ist.psu.ed
u/viewdoc/summary?doi=10.1.1.475.7392). doi:10.2136/sssaj1999.6351055x (https://doi.or
g/10.2136%2Fsssaj1999.6351055x).
141. Knecht, Magnus F.; Göransson, Anders (2004). "Terrestrial plants require nutrients in similar
proportions" (https://doi.org/10.1093%2Ftreephys%2F24.4.447). Tree Physiology. 24 (4):
447–460. doi:10.1093/treephys/24.4.447 (https://doi.org/10.1093%2Ftreephys%2F24.4.447).
PMID 14757584 (https://pubmed.ncbi.nlm.nih.gov/14757584).
142. Dean 1957, pp. 80–81.
143. Roy, R. N.; Finck, Arnold; Blair, Graeme J.; Tandon, Hari Lal Singh (2006). "Soil fertility and
crop production" (http://www.fao.org/fileadmin/templates/soilbiodiversity/Downloadable_file
s/fpnb16.pdf) (PDF). Plant nutrition for food security: a guide for integrated nutrient
management. Rome, Italy: Food and Agriculture Organization of the United Nations. pp. 43–
90. ISBN 978-92-5-105490-1. Retrieved 27 June 2021.
144. Parfitt, Roger L.; Giltrap, Donna J.; Whitton, Joe S. (1995). "Contribution of organic matter
and clay minerals to the cation exchange capacity of soil" (https://www.researchgate.net/pub
lication/249073571). Communications in Soil Science and Plant Analysis. 26 (9–10): 1343–
1355. doi:10.1080/00103629509369376 (https://doi.org/10.1080%2F00103629509369376).
Retrieved 27 June 2021.
145. Hajnos, Mieczyslaw; Jozefaciuk, Grzegorz; Sokołowska, Zofia; Greiffenhagen, Andreas;
Wessolek, Gerd (2003). "Water storage, surface, and structural properties of sandy forest
humus horizons" (https://www.researchgate.net/publication/229970348). Journal of Plant
Nutrition and Soil Science. 166 (5): 625–634. doi:10.1002/jpln.200321161 (https://doi.org/10.
1002%2Fjpln.200321161). Retrieved 27 June 2021.
146. Donahue, Miller & Shickluna 1977, pp. 123–131.
147. Pimentel, David; Harvey, Celia; Resosudarmo, Pradnja; Sinclair, K.; Kurz, D.; McNair, M.;
Crist, S.; Shpritz, L.; Fitton, L.; Saffouri, R.; Blair, R. (1995). "Environmental and economic
costs of soil erosion and conservation benefits" (https://www.academia.edu/9512072).
Science. 267 (5201): 1117–23. Bibcode:1995Sci...267.1117P (https://ui.adsabs.harvard.edu/
abs/1995Sci...267.1117P). doi:10.1126/science.267.5201.1117 (https://doi.org/10.1126%2F
science.267.5201.1117). PMID 17789193 (https://pubmed.ncbi.nlm.nih.gov/17789193).
S2CID 11936877 (https://api.semanticscholar.org/CorpusID:11936877). Archived (https://we
b.archive.org/web/20161213065558/http://www.rachel.org/files/document/Environmental_an
d_Economic_Costs_of_Soil_Erosi.pdf) (PDF) from the original on 13 December 2016.
Retrieved 4 July 2021.
148. Schnürer, Johan; Clarholm, Marianne; Rosswall, Thomas (1985). "Microbial biomass and
activity in an agricultural soil with different organic matter contents" (https://www.academia.e
du/20647751). Soil Biology and Biochemistry. 17 (5): 611–618. doi:10.1016/0038-
0717(85)90036-7 (https://doi.org/10.1016%2F0038-0717%2885%2990036-7). Retrieved
4 July 2021.
149. Sparling, Graham P. (1992). "Ratio of microbial biomass carbon to soil organic carbon as a
sensitive indicator of changes in soil organic matter" (https://www.researchgate.net/publicati
on/248884528). Australian Journal of Soil Research. 30 (2): 195–207.
doi:10.1071/SR9920195 (https://doi.org/10.1071%2FSR9920195). Retrieved 4 July 2021.
150. Varadachari, Chandrika; Ghosh, Kunal (1984). "On humus formation" (https://doi.org/10.100
7%2FBF02182933). Plant and Soil. 77 (2): 305–313. doi:10.1007/BF02182933 (https://doi.o
rg/10.1007%2FBF02182933). S2CID 45102095 (https://api.semanticscholar.org/CorpusID:4
5102095).
151. Prescott, Cindy E. (2010). "Litter decomposition: what controls it and how can we alter it to
sequester more carbon in forest soils?" (https://doi.org/10.1007%2Fs10533-010-9439-0).
Biogeochemistry. 101 (1): 133–q49. doi:10.1007/s10533-010-9439-0 (https://doi.org/10.100
7%2Fs10533-010-9439-0). S2CID 93834812 (https://api.semanticscholar.org/CorpusID:938
34812).
152. Lehmann, Johannes; Kleber, Markus (2015). "The contentious nature of soil organic matter"
(http://www.css.cornell.edu/faculty/lehmann/publ/Nature%20528,%2060-68,%202015%20Le
hmann.pdf) (PDF). Nature. 528 (7580): 60–68. Bibcode:2015Natur.528...60L (https://ui.adsab
s.harvard.edu/abs/2015Natur.528...60L). doi:10.1038/nature16069 (https://doi.org/10.1038%
2Fnature16069). PMID 26595271 (https://pubmed.ncbi.nlm.nih.gov/26595271).
S2CID 205246638 (https://api.semanticscholar.org/CorpusID:205246638). Retrieved 4 July
2021.
153. Piccolo, Alessandro (2002). "The supramolecular structure of humic substances: a novel
understanding of humus chemistry and implications in soil science" (https://www.researchga
te.net/publication/222526145). Advances in Agronomy. 75: 57–134. doi:10.1016/S0065-
2113(02)75003-7 (https://doi.org/10.1016%2FS0065-2113%2802%2975003-7).
ISBN 9780120007936. Retrieved 4 July 2021.
154. Scheu, Stefan (2002). "The soil food web: structure and perspectives" (https://www.research
gate.net/publication/263041521). European Journal of Soil Biology. 38 (1): 11–20.
doi:10.1016/S1164-5563(01)01117-7 (https://doi.org/10.1016%2FS1164-5563%2801%2901
117-7). Retrieved 4 July 2021.
155. Foth, Henry D. (1984). Fundamentals of soil science (http://base.dnsgb.com.ua/files/book/Ag
riculture/Soil/Fundamentals-of-Soil-Science.pdf) (PDF) (8th ed.). New York, New York:
Wiley. p. 139. ISBN 978-0471522799. Retrieved 4 July 2021.
156. Ponge, Jean-François (2003). "Humus forms in terrestrial ecosystems: a framework to
biodiversity" (https://www.academia.edu/45579598). Soil Biology and Biochemistry. 35 (7):
935–945. CiteSeerX 10.1.1.467.4937 (https://citeseerx.ist.psu.edu/viewdoc/summary?doi=1
0.1.1.467.4937). doi:10.1016/S0038-0717(03)00149-4 (https://doi.org/10.1016%2FS0038-07
17%2803%2900149-4). S2CID 44160220 (https://api.semanticscholar.org/CorpusID:441602
20). Archived (https://web.archive.org/web/20160129153903/https://www.researchgate.net/p
ublication/222567430) from the original on 29 January 2016.
157. Pettit, Robert E. "Organic matter, humus, humate, humic acid, fulvic acid and humin: their
importance in soil fertility and plant health" (http://www.harvestgrow.com/.pdf%20web%20sit
e/Humates%20General%20Info.pdf) (PDF). Retrieved 11 July 2021.
158. Ji, Rong; Kappler, Andreas; Brune, Andreas (2000). "Transformation and mineralization of
synthetic 14C-labeled humic model compounds by soil-feeding termites". Soil Biology and
Biochemistry. 32 (8–9): 1281–1291. CiteSeerX 10.1.1.476.9400 (https://citeseerx.ist.psu.ed
u/viewdoc/summary?doi=10.1.1.476.9400). doi:10.1016/S0038-0717(00)00046-8 (https://do
i.org/10.1016%2FS0038-0717%2800%2900046-8).
159. Drever, James I.; Vance, George F. (1994). "Role of soil organic acids in mineral weathering
processes" (https://link.springer.com/content/pdf/10.1007%2F978-3-642-78356-2_6.pdf)
(PDF). In Pittman, Edward D.; Lewan, Michael D. (eds.). Organic acids in geological
processes. Berlin, Germany: Springer. pp. 138–161. doi:10.1007/978-3-642-78356-2_6 (http
s://doi.org/10.1007%2F978-3-642-78356-2_6). ISBN 978-3-642-78356-2. Retrieved 11 July
2021.
160. Piccolo, Alessandro (1996). "Humus and soil conservation" (https://www.researchgate.net/p
ublication/281451183). In Piccolo, Alessandro (ed.). Humic substances in terrestrial
ecosystems. Amsterdam, The Netherlands: Elsevier. pp. 225–264. doi:10.1016/B978-
044481516-3/50006-2 (https://doi.org/10.1016%2FB978-044481516-3%2F50006-2).
ISBN 978-0-444-81516-3. Retrieved 11 July 2021.
161. Varadachari, Chandrika; Ghosh, Kunal (1984). "On humus formation" (https://www.researchg
ate.net/publication/225528442). Plant and Soil. 77 (2): 305–313. doi:10.1007/BF02182933
(https://doi.org/10.1007%2FBF02182933). S2CID 45102095 (https://api.semanticscholar.or
g/CorpusID:45102095). Retrieved 11 July 2021.
162. Mendonça, Eduardo S.; Rowell, David L. (1996). "Mineral and organic fractions of two
oxisols and their influence on effective cation-exchange capacity" (https://www.researchgat
e.net/publication/250128642). Soil Science Society of America Journal. 60 (6): 1888–1892.
Bibcode:1996SSASJ..60.1888M (https://ui.adsabs.harvard.edu/abs/1996SSASJ..60.1888
M). doi:10.2136/sssaj1996.03615995006000060038x (https://doi.org/10.2136%2Fsssaj199
6.03615995006000060038x). Retrieved 11 July 2021.
163. Heck, Tobias; Faccio, Greta; Richter, Michael; Thöny-Meyer, Linda (2013). "Enzyme-
catalyzed protein crosslinking" (https://www.researchgate.net/publication/233769618).
Applied Microbiology and Biotechnology. 97 (2): 461–475. doi:10.1007/s00253-012-4569-z
(https://doi.org/10.1007%2Fs00253-012-4569-z). PMC 3546294 (https://www.ncbi.nlm.nih.g
ov/pmc/articles/PMC3546294). PMID 23179622 (https://pubmed.ncbi.nlm.nih.gov/2317962
2). Retrieved 11 July 2021.
164. Lynch, D. L.; Lynch, C. C. (1958). "Resistance of protein–lignin complexes, lignins and
humic acids to microbial attack" (https://www.nature.com/articles/1811478a0.pdf) (PDF).
Nature. 181 (4621): 1478–1479. Bibcode:1958Natur.181.1478L (https://ui.adsabs.harvard.ed
u/abs/1958Natur.181.1478L). doi:10.1038/1811478a0 (https://doi.org/10.1038%2F1811478a
0). PMID 13552710 (https://pubmed.ncbi.nlm.nih.gov/13552710). S2CID 4193782 (https://ap
i.semanticscholar.org/CorpusID:4193782). Retrieved 11 July 2021.
165. Dawson, Lorna A.; Hillier, Stephen (2010). "Measurement of soil characteristics for forensic
applications" (https://people.ok.ubc.ca/robrien/soil%20characteristics.pdf) (PDF). Surface
and Interface Analysis. 42 (5): 363–377. doi:10.1002/sia.3315 (https://doi.org/10.1002%2Fsi
a.3315). S2CID 54213404 (https://api.semanticscholar.org/CorpusID:54213404). Retrieved
18 July 2021.
166. Manjaiah, K.M.; Kumar, Sarvendra; Sachdev, M. S.; Sachdev, P.; Datta, S. C. (2010). "Study
of clay–organic complexes" (https://www.researchgate.net/publication/228867334). Current
Science. 98 (7): 915–921. Retrieved 18 July 2021.
167. Theng, Benny K.G. (1982). "Clay-polymer interactions: summary and perspectives". Clays
and Clay Minerals. 30 (1): 1–10. Bibcode:1982CCM....30....1T (https://ui.adsabs.harvard.ed
u/abs/1982CCM....30....1T). CiteSeerX 10.1.1.608.2942 (https://citeseerx.ist.psu.edu/viewdo
c/summary?doi=10.1.1.608.2942). doi:10.1346/CCMN.1982.0300101 (https://doi.org/10.134
6%2FCCMN.1982.0300101). S2CID 98176725 (https://api.semanticscholar.org/CorpusID:9
8176725).
168. Tietjen, Todd; Wetzel, Robert G. (2003). "Extracellular enzyme-clay mineral complexes:
enzyme adsorption, alteration of enzyme activity, and protection from photodegradation" (htt
p://www.vliz.be/imisdocs/publications/54440.pdf) (PDF). Aquatic Ecology. 37 (4): 331–339.
doi:10.1023/B:AECO.0000007044.52801.6b (https://doi.org/10.1023%2FB%3AAECO.0000
007044.52801.6b). S2CID 6930871 (https://api.semanticscholar.org/CorpusID:6930871).
Retrieved 18 July 2021.
169. Tahir, Shermeen; Marschner, Petra (2017). "Clay addition to sandy soil: influence of clay
type and size on nutrient availability in sandy soils amended with residues differing in C/N
ratio" (https://www.researchgate.net/publication/314221508). Pedosphere. 27 (2): 293–305.
doi:10.1016/S1002-0160(17)60317-5 (https://doi.org/10.1016%2FS1002-0160%2817%2960
317-5). Retrieved 18 July 2021.
170. Melero, Sebastiana; Madejón, Engracia; Ruiz, Juan Carlos; Herencia, Juan Francisco
(2007). "Chemical and biochemical properties of a clay soil under dryland agriculture system
as affected by organic fertilization" (https://coek.info/pdf-chemical-and-biochemical-propertie
s-of-a-clay-soil-under-dryland-agriculture-sys.html). European Journal of Agronomy. 26 (3):
327–334. doi:10.1016/j.eja.2006.11.004 (https://doi.org/10.1016%2Fj.eja.2006.11.004).
Retrieved 18 July 2021.
171. Joanisse, Gilles D.; Bradley, Robert L.; Preston, Caroline M.; Bending, Gary D. (2009).
"Sequestration of soil nitrogen as tannin–protein complexes may improve the competitive
ability of sheep laurel (Kalmia angustifolia) relative to black spruce (Picea mariana)" (https://
doi.org/10.1111%2Fj.1469-8137.2008.02622.x). New Phytologist. 181 (1): 187–198.
doi:10.1111/j.1469-8137.2008.02622.x (https://doi.org/10.1111%2Fj.1469-8137.2008.02622.
x). PMID 18811620 (https://pubmed.ncbi.nlm.nih.gov/18811620).
172. Fierer, Noah; Schimel, Joshua P.; Cates, Rex G.; Zou, Jiping (2001). "Influence of balsam
poplar tannin fractions on carbon and nitrogen dynamics in Alaskan taiga floodplain soils" (h
ttps://www.academia.edu/12814037). Soil Biology and Biochemistry. 33 (12–13): 1827–
1839. doi:10.1016/S0038-0717(01)00111-0 (https://doi.org/10.1016%2FS0038-0717%280
1%2900111-0). Retrieved 18 July 2021.
173. Peng, Xinhua; Horn, Rainer (2007). "Anisotropic shrinkage and swelling of some organic
and inorganic soils" (https://doi.org/10.1111%2Fj.1365-2389.2006.00808.x). European
Journal of Soil Science. 58 (1): 98–107. doi:10.1111/j.1365-2389.2006.00808.x (https://doi.o
rg/10.1111%2Fj.1365-2389.2006.00808.x).
174. Wang, Yang; Amundson, Ronald; Trumbmore, Susan (1996). "Radiocarbon dating of soil
organic matter" (https://escholarship.org/content/qt6b14h4bv/qt6b14h4bv.pdf) (PDF).
Quaternary Research. 45 (3): 282–288. Bibcode:1996QuRes..45..282W (https://ui.adsabs.ha
rvard.edu/abs/1996QuRes..45..282W). doi:10.1006/qres.1996.0029 (https://doi.org/10.100
6%2Fqres.1996.0029). S2CID 73640995 (https://api.semanticscholar.org/CorpusID:736409
95). Retrieved 18 July 2021.
175. Brodowski, Sonja; Amelung, Wulf; Haumaier, Ludwig; Zech, Wolfgang (2007). "Black carbon
contribution to stable humus in German arable soils" (https://www.academia.edu/33858429).
Geoderma. 139 (1–2): 220–228. Bibcode:2007Geode.139..220B (https://ui.adsabs.harvard.e
du/abs/2007Geode.139..220B). doi:10.1016/j.geoderma.2007.02.004 (https://doi.org/10.101
6%2Fj.geoderma.2007.02.004). Retrieved 18 July 2021.
176. Criscuoli, Irene; Alberti, Giorgio; Baronti, Silvia; Favilli, Filippo; Martinez, Cristina; Calzolari,
Costanza; Pusceddu, Emanuela; Rumpel, Cornelia; Viola, Roberto; Miglietta, Franco
(2014). "Carbon sequestration and fertility after centennial time scale incorporation of
charcoal into soil" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3948733). PLOS ONE. 9
(3): e91114. Bibcode:2014PLoSO...991114C (https://ui.adsabs.harvard.edu/abs/2014PLoS
O...991114C). doi:10.1371/journal.pone.0091114 (https://doi.org/10.1371%2Fjournal.pone.0
091114). PMC 3948733 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3948733).
PMID 24614647 (https://pubmed.ncbi.nlm.nih.gov/24614647).
177. Kim, Dong Jim; Vargas, Rodrigo; Bond-Lamberty, Ben; Turetsky, Merritt R. (2012). "Effects of
soil rewetting and thawing on soil gas fluxes: a review of current literature and suggestions
for future research" (https://www.researchgate.net/publication/307827983). Biogeosciences.
9 (7): 2459–2483. Bibcode:2012BGeo....9.2459K (https://ui.adsabs.harvard.edu/abs/2012BG
eo....9.2459K). doi:10.5194/bg-9-2459-2012 (https://doi.org/10.5194%2Fbg-9-2459-2012).
Retrieved 3 October 2021.
178. 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" (https://www.academia.edu/20165844). Geoderma. 147 (1–2): 23–33.
Bibcode:2008Geode.147...23W (https://ui.adsabs.harvard.edu/abs/2008Geode.147...23W).
doi:10.1016/j.geoderma.2008.07.010 (https://doi.org/10.1016%2Fj.geoderma.2008.07.010).
hdl:10261/82461 (https://hdl.handle.net/10261%2F82461). Retrieved 25 July 2021.
179. Minayeva, Tatiana Y.; Trofimov, Sergey Ya.; Chichagova, Olga A.; Dorofeyeva, E. I.; Sirin,
Andrey A.; Glushkov, Igor V.; Mikhailov, N. D.; Kromer, Bernd (2008). "Carbon accumulation
in soils of forest and bog ecosystems of southern Valdai in the Holocene" (https://www.resea
rchgate.net/publication/225229436). Biology Bulletin. 35 (5): 524–532.
doi:10.1134/S1062359008050142 (https://doi.org/10.1134%2FS1062359008050142).
S2CID 40927739 (https://api.semanticscholar.org/CorpusID:40927739). Retrieved 25 July
2021.
180. Vitousek, Peter M.; Sanford, Robert L. (1986). "Nutrient cycling in moist tropical forest" (http
s://www.researchgate.net/publication/234150505). Annual Review of Ecology and
Systematics. 17: 137–167. doi:10.1146/annurev.es.17.110186.001033 (https://doi.org/10.11
46%2Fannurev.es.17.110186.001033). S2CID 55212899 (https://api.semanticscholar.org/C
orpusID:55212899). Retrieved 25 July 2021.
181. Rumpel, Cornelia; Chaplot, Vincent; Planchon, Olivier; Bernadou, J.; Valentin, Christian;
Mariotti, André (2006). "Preferential erosion of black carbon on steep slopes with slash and
burn agriculture" (https://www.academia.edu/14788543). Catena. 65 (1): 30–40.
Bibcode:2006Caten..65...30R (https://ui.adsabs.harvard.edu/abs/2006Caten..65...30R).
doi:10.1016/j.catena.2005.09.005 (https://doi.org/10.1016%2Fj.catena.2005.09.005).
Retrieved 25 July 2021.
182. Paul, Eldor A.; Paustian, Keith H.; Elliott, E. T.; Cole, C. Vernon (1997). Soil organic matter in
temperate agroecosystems: long-term experiments in North America. Boca Raton, Florida:
CRC Press. p. 80. ISBN 978-0-8493-2802-2.
183. "Horizons" (https://soilsofcanada.ca/soil-formation/horizons.php). Soils of Canada. Archived
(https://web.archive.org/web/20190922153041/https://soilsofcanada.ca/soil-formation/horizo
ns.php) from the original on 22 September 2019. Retrieved 1 August 2021.
184. Frouz, Jan; Prach, Karel; Pizl, Václav; Háněl, Ladislav; Starý, Josef; Tajovský, Karel;
Materna, Jan; Balík, Vladimír; Kalčík, Jiří; Řehounková, Klára (2008). "Interactions between
soil development, vegetation and soil fauna during spontaneous succession in post mining
sites" (https://www.researchgate.net/publication/223699609). European Journal of Soil
Biology. 44 (1): 109–121. doi:10.1016/j.ejsobi.2007.09.002 (https://doi.org/10.1016%2Fj.ejso
bi.2007.09.002). Retrieved 1 August 2021.
185. Kabala, Cezary; Zapart, Justyna (2012). "Initial soil development and carbon accumulation
on moraines of the rapidly retreating Werenskiold Glacier, SW Spitsbergen, Svalbard
archipelago" (https://www.academia.edu/31221217). Geoderma. 175–176: 9–20.
Bibcode:2012Geode.175....9K (https://ui.adsabs.harvard.edu/abs/2012Geode.175....9K).
doi:10.1016/j.geoderma.2012.01.025 (https://doi.org/10.1016%2Fj.geoderma.2012.01.025).
Retrieved 1 August 2021.
186. Ugolini, Fiorenzo C.; Dahlgren, Randy A. (2002). "Soil development in volcanic ash" (http://
www.airies.or.jp/attach.php/6a6f75726e616c5f30362d32656e67/save/0/0/06_2-09.pdf)
(PDF). Global Environmental Research. 6 (2): 69–81. Retrieved 1 August 2021.
187. Huggett, Richard J. (1998). "Soil chronosequences, soil development, and soil evolution: a
critical review" (https://www.academia.edu/2116704). Catena. 32 (3): 155–172.
Bibcode:1998Caten..32..155H (https://ui.adsabs.harvard.edu/abs/1998Caten..32..155H).
doi:10.1016/S0341-8162(98)00053-8 (https://doi.org/10.1016%2FS0341-8162%2898%2900
053-8). Retrieved 1 August 2021.
188. De Alba, Saturnio; Lindstrom, Michael; Schumacher, Thomas E.; Malo, Douglas D. (2004).
"Soil landscape evolution due to soil redistribution by tillage: a new conceptual model of soil
catena evolution in agricultural landscapes" (https://www.academia.edu/22300477). Catena.
58 (1): 77–100. Bibcode:2004Caten..58...77D (https://ui.adsabs.harvard.edu/abs/2004Cate
n..58...77D). doi:10.1016/j.catena.2003.12.004 (https://doi.org/10.1016%2Fj.catena.2003.12.
004). Retrieved 1 August 2021.
189. Phillips, Jonathan D.; Marion, Daniel A. (2004). "Pedological memory in forest soil
development" (https://www.srs.fs.usda.gov/pubs/ja/ja_phillips004.pdf) (PDF). Forest Ecology
and Management. 188 (1): 363–380. doi:10.1016/j.foreco.2003.08.007 (https://doi.org/10.101
6%2Fj.foreco.2003.08.007). Retrieved 1 August 2021.
190. Mitchell, Edward A.D.; Van der Knaap, Willem O.; Van Leeuwen, Jacqueline F.N.; Buttler,
Alexandre; Warner, Barry G.; Gobat, Jean-Michel (2001). "The palaeoecological history of
the Praz-Rodet bog (Swiss Jura) based on pollen, plant macrofossils and testate
amoebae(Protozoa)" (https://www.academia.edu/31915005). The Holocene. 11 (1): 65–80.
Bibcode:2001Holoc..11...65M (https://ui.adsabs.harvard.edu/abs/2001Holoc..11...65M).
doi:10.1191/095968301671777798 (https://doi.org/10.1191%2F095968301671777798).
S2CID 131032169 (https://api.semanticscholar.org/CorpusID:131032169). Retrieved
1 August 2021.
191. Carcaillet, Christopher (2001). "Soil particles reworking evidences by AMS 14C dating of
charcoal" (https://www.researchgate.net/publication/238379602). Comptes Rendus de
l'Académie des Sciences, Série IIA. 332 (1): 21–28. Bibcode:2001CRASE.332...21C (https://
ui.adsabs.harvard.edu/abs/2001CRASE.332...21C). doi:10.1016/S1251-8050(00)01485-3
(https://doi.org/10.1016%2FS1251-8050%2800%2901485-3). Retrieved 1 August 2021.
192. Retallack, Gregory J. (1991). "Untangling the effects of burial alteration and ancient soil
formation" (https://www.researchgate.net/publication/234148901). Annual Review of Earth
and Planetary Sciences. 19 (1): 183–206. Bibcode:1991AREPS..19..183R (https://ui.adsab
s.harvard.edu/abs/1991AREPS..19..183R). doi:10.1146/annurev.ea.19.050191.001151 (http
s://doi.org/10.1146%2Fannurev.ea.19.050191.001151). Retrieved 1 August 2021.
193. Bakker, Martha M.; Govers, Gerard; Jones, Robert A.; Rounsevell, Mark D.A. (2007). "The
effect of soil erosion on Europe's crop yields" (https://doi.org/10.1007%2Fs10021-007-9090-
3). Ecosystems. 10 (7): 1209–1219. doi:10.1007/s10021-007-9090-3 (https://doi.org/10.100
7%2Fs10021-007-9090-3).
194. Uselman, Shauna M.; Qualls, Robert G.; Lilienfein, Juliane (2007). "Contribution of root vs.
leaf litter to dissolved organic carbon leaching through soil" (https://www.academia.edu/344
75958). Soil Science Society of America Journal. 71 (5): 1555–1563.
Bibcode:2007SSASJ..71.1555U (https://ui.adsabs.harvard.edu/abs/2007SSASJ..71.1555U).
doi:10.2136/sssaj2006.0386 (https://doi.org/10.2136%2Fsssaj2006.0386). Retrieved
8 August 2021.
195. Schulz, Stefanie; Brankatschk, Robert; Dümig, Alexander; Kögel-Knabner, Ingrid; Schloter,
Michae; Zeyer, Josef (2013). "The role of microorganisms at different stages of ecosystem
development for soil formation" (https://doi.org/10.5194%2Fbg-10-3983-2013).
Biogeosciences. 10 (6): 3983–3996. Bibcode:2013BGeo...10.3983S (https://ui.adsabs.harva
rd.edu/abs/2013BGeo...10.3983S). doi:10.5194/bg-10-3983-2013 (https://doi.org/10.5194%2
Fbg-10-3983-2013).
196. Gillet, Servane; Ponge, Jean-François (2002). "Humus forms and metal pollution in soil" (htt
ps://www.academia.edu/45705588). European Journal of Soil Science. 53 (4): 529–539.
doi:10.1046/j.1365-2389.2002.00479.x (https://doi.org/10.1046%2Fj.1365-2389.2002.00479.
x). S2CID 94900982 (https://api.semanticscholar.org/CorpusID:94900982). Retrieved
8 August 2021.
197. Bardy, Marion; Fritsch, Emmanuel; Derenne, Sylvie; Allard, Thierry; do Nascimento, Nadia
Régina; Bueno, Guilherme (2008). "Micromorphology and spectroscopic characteristics of
organic matter in waterlogged podzols of the upper Amazon basin". Geoderma. 145 (3):
222–230. Bibcode:2008Geode.145..222B (https://ui.adsabs.harvard.edu/abs/2008Geode.14
5..222B). CiteSeerX 10.1.1.455.4179 (https://citeseerx.ist.psu.edu/viewdoc/summary?doi=1
0.1.1.455.4179). doi:10.1016/j.geoderma.2008.03.008 (https://doi.org/10.1016%2Fj.geoderm
a.2008.03.008).
198. Dokuchaev, Vasily Vasilyevich (1967). "Russian Chernozem" (https://fr.scribd.com/doc/2068
59253/Russian-Chernozem). Jerusalem, Israel: Israel Program for Scientific Translations.
Retrieved 15 August 2021.
199. IUSS Working Group WRB (2022). "World Reference Base for Soil Resources, 4th edition"
(https://www3.ls.tum.de/boku/?id=1419). IUSS, Vienna.
200. Sambo, Paolo; Nicoletto, Carlo; Giro, Andrea; Pii, Youry; Valentinuzzi, Fabio; Mimmo, Tanja;
Lugli, Paolo; Orzes, Guido; Mazzetto, Fabrizio; Astolfi, Stefania; Terzano, Roberto; Cesco,
Stefano (2019). "Hydroponic solutions for soilless production systems: issues and
opportunities in a smart agriculture perspective" (https://www.ncbi.nlm.nih.gov/pmc/articles/P
MC6668597). Frontiers in Plant Science. 10 (123): 923. doi:10.3389/fpls.2019.00923 (https://
doi.org/10.3389%2Ffpls.2019.00923). PMC 6668597 (https://www.ncbi.nlm.nih.gov/pmc/arti
cles/PMC6668597). PMID 31396245 (https://pubmed.ncbi.nlm.nih.gov/31396245).
201. Leake, Simon; Haege, Elke (2014). Soils for landscape development: selection,
specification and validation. Clayton, Victoria, Australia: CSIRO Publishing. ISBN 978-
0643109650.
202. Pan, Xian-Zhang; Zhao, Qi-Guo (2007). "Measurement of urbanization process and the
paddy soil loss in Yixing city, China between 1949 and 2000" (http://www.cern.ac.cn/ftp/030
1%20Measurement%20of%20urbanization%20process%20and%20the%20paddy%20soil%
20loss%20in%20Yixing%20city,%20China%20between%201949%20and%202000).pdf)
(PDF). Catena. 69 (1): 65–73. Bibcode:2007Caten..69...65P (https://ui.adsabs.harvard.edu/a
bs/2007Caten..69...65P). doi:10.1016/j.catena.2006.04.016 (https://doi.org/10.1016%2Fj.cat
ena.2006.04.016). Retrieved 15 August 2021.
203. Kopittke, Peter M.; Menzies, Neal W.; Wang, Peng; McKenna, Brigid A.; Lombi, Enzo (2019).
"Soil and the intensification of agriculture for global food security" (https://doi.org/10.1016%2
Fj.envint.2019.105078). Environment International. 132: 105078.
doi:10.1016/j.envint.2019.105078 (https://doi.org/10.1016%2Fj.envint.2019.105078).
ISSN 0160-4120 (https://www.worldcat.org/issn/0160-4120). PMID 31400601 (https://pubme
d.ncbi.nlm.nih.gov/31400601).
204. Stürck, Julia; Poortinga, Ate; Verburg, Peter H. (2014). "Mapping ecosystem services: the
supply and demand of flood regulation services in Europe" (http://docs.gip-ecofor.org/public/
Sturck_et_al_2014.pdf) (PDF). Ecological Indicators. 38: 198–211.
doi:10.1016/j.ecolind.2013.11.010 (https://doi.org/10.1016%2Fj.ecolind.2013.11.010).
Retrieved 15 August 2021.
205. Van Cuyk, Sheila; Siegrist, Robert; Logan, Andrew; Masson, Sarah; Fischer, Elizabeth;
Figueroa, Linda (2001). "Hydraulic and purification behaviors and their interactions during
wastewater treatment in soil infiltration systems" (https://www.academia.edu/17525373).
Water Research. 35 (4): 953–964. Bibcode:2001WatRe..35..953V (https://ui.adsabs.harvard.
edu/abs/2001WatRe..35..953V). doi:10.1016/S0043-1354(00)00349-3 (https://doi.org/10.101
6%2FS0043-1354%2800%2900349-3). PMID 11235891 (https://pubmed.ncbi.nlm.nih.gov/1
1235891). Retrieved 15 August 2021.
206. Jeffery, Simon; Gardi, Ciro; Arwyn, Jones (2010). European atlas of soil biodiversity (https://o
p.europa.eu/en/publication-detail/-/publication/7161b2a1-f862-4c90-9100-557a62ecb908).
Luxembourg, Luxembourg: Publications Office of the European Union. doi:10.2788/94222 (h
ttps://doi.org/10.2788%2F94222). ISBN 978-92-79-15806-3. Retrieved 15 August 2021.
207. De Deyn, Gerlinde B.; Van der Putten, Wim H. (2005). "Linking aboveground and
belowground diversity" (https://www.researchgate.net/publication/7080980). Trends in
Ecology and Evolution. 20 (11): 625–633. doi:10.1016/j.tree.2005.08.009 (https://doi.org/10.
1016%2Fj.tree.2005.08.009). PMID 16701446 (https://pubmed.ncbi.nlm.nih.gov/16701446).
Retrieved 15 August 2021.
208. Hansen, James; Sato, Makiko; Kharecha, Pushker; Beerling, David; Berner, Robert;
Masson-Delmotte, Valerie; Pagani, Mark; Raymo, Maureen; Royer, Dana L.; Zachos, James
C. (2008). "Target atmospheric CO2: where should humanity aim?" (https://benthamopen.co
m/contents/pdf/TOASCJ/TOASCJ-2-217.pdf) (PDF). Open Atmospheric Science Journal. 2
(1): 217–231. arXiv:0804.1126 (https://arxiv.org/abs/0804.1126).
Bibcode:2008OASJ....2..217H (https://ui.adsabs.harvard.edu/abs/2008OASJ....2..217H).
doi:10.2174/1874282300802010217 (https://doi.org/10.2174%2F1874282300802010217).
S2CID 14890013 (https://api.semanticscholar.org/CorpusID:14890013). Retrieved
22 August 2021.
209. Lal, Rattan (11 June 2004). "Soil carbon sequestration impacts on global climate change
and food security" (http://www.tinread.usarb.md:8888/jspui/bitstream/123456789/1067/1/soil
_carbon.pdf) (PDF). Science. 304 (5677): 1623–1627. Bibcode:2004Sci...304.1623L (https://
ui.adsabs.harvard.edu/abs/2004Sci...304.1623L). doi:10.1126/science.1097396 (https://doi.
org/10.1126%2Fscience.1097396). PMID 15192216 (https://pubmed.ncbi.nlm.nih.gov/15192
216). S2CID 8574723 (https://api.semanticscholar.org/CorpusID:8574723). Retrieved
22 August 2021.
210. Blakeslee, Thomas (24 February 2010). "Greening deserts for carbon credits" (https://www.r
enewableenergyworld.com/om/greening-deserts-for-carbon-credits/#gref). Orlando, Florida,
USA: Renewable Energy World. Archived (https://web.archive.org/web/20121101011735/htt
p://www.renewableenergyworld.com/rea/news/article/2010/02/greening-deserts-for-carbon-c
redits) from the original on 1 November 2012. Retrieved 22 August 2021.
211. Mondini, Claudio; Contin, Marco; Leita, Liviana; De Nobili, Maria (2002). "Response of
microbial biomass to air-drying and rewetting in soils and compost" (https://www.academia.e
du/5321925). Geoderma. 105 (1–2): 111–124. Bibcode:2002Geode.105..111M (https://ui.ad
sabs.harvard.edu/abs/2002Geode.105..111M). doi:10.1016/S0016-7061(01)00095-7 (https://
doi.org/10.1016%2FS0016-7061%2801%2900095-7). Retrieved 22 August 2021.
212. "Peatlands and farming" (https://www.countrysideonline.co.uk/food-and-farming/protecting-th
e-environment/peatlands-and-farming). Stoneleigh, United Kingdom: National Farmers'
Union of England and Wales. 6 July 2020. Retrieved 22 August 2021.
213. van Winden, Julia F.; Reichart, Gert-Jan; McNamara, Niall P.; Benthien, Albert; Sinninghe
Damste, Jaap S. (2012). "Temperature-induced increase in methane release from peat bogs:
a mesocosm experiment" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3387254). PLoS
ONE. 7 (6): e39614. Bibcode:2012PLoSO...739614V (https://ui.adsabs.harvard.edu/abs/201
2PLoSO...739614V). doi:10.1371/journal.pone.0039614 (https://doi.org/10.1371%2Fjournal.
pone.0039614). PMC 3387254 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3387254).
PMID 22768100 (https://pubmed.ncbi.nlm.nih.gov/22768100).
214. Davidson, Eric A.; Janssens, Ivan A. (2006). "Temperature sensitivity of soil carbon
decomposition and feedbacks to climate change" (https://www.woodwellclimate.org/wp-cont
ent/uploads/2015/09/DavidsonetalNature.06.pdf) (PDF). Nature. 440 (7081): 165–173.
Bibcode:2006Natur.440..165D (https://ui.adsabs.harvard.edu/abs/2006Natur.440..165D).
doi:10.1038/nature04514 (https://doi.org/10.1038%2Fnature04514). PMID 16525463 (https://
pubmed.ncbi.nlm.nih.gov/16525463). S2CID 4404915 (https://api.semanticscholar.org/Corp
usID:4404915). Retrieved 22 August 2021.
215. Abrahams, Pter W. (1997). "Geophagy (soil consumption) and iron supplementation in
Uganda" (https://doi.org/10.1046%2Fj.1365-3156.1997.d01-348.x). Tropical Medicine and
International Health. 2 (7): 617–623. doi:10.1046/j.1365-3156.1997.d01-348.x (https://doi.or
g/10.1046%2Fj.1365-3156.1997.d01-348.x). PMID 9270729 (https://pubmed.ncbi.nlm.nih.go
v/9270729). S2CID 19647911 (https://api.semanticscholar.org/CorpusID:19647911).
216. Setz, Eleonore Zulnara Freire; Enzweiler, Jacinta; Solferini, Vera Nisaka; Amêndola, Monica
Pimenta; Berton, Ronaldo Severiano (1999). "Geophagy in the golden-faced saki monkey
(Pithecia pithecia chrysocephala) in the Central Amazon" (https://www.academia.edu/26464
333). Journal of Zoology. 247 (1): 91–103. doi:10.1111/j.1469-7998.1999.tb00196.x (https://d
oi.org/10.1111%2Fj.1469-7998.1999.tb00196.x). Retrieved 22 August 2021.
217. Kohne, John Maximilian; Koehne, Sigrid; Simunek, Jirka (2009). "A review of model
applications for structured soils: a) Water flow and tracer transport" (http://www.pc-progress.c
om/documents/jirka/ko-ko_sim_2008_jcontamhydrol.pdf) (PDF). Journal of Contaminant
Hydrology. 104 (1–4): 4–35. Bibcode:2009JCHyd.104....4K (https://ui.adsabs.harvard.edu/ab
s/2009JCHyd.104....4K). CiteSeerX 10.1.1.468.9149 (https://citeseerx.ist.psu.edu/viewdoc/s
ummary?doi=10.1.1.468.9149). doi:10.1016/j.jconhyd.2008.10.002 (https://doi.org/10.1016%
2Fj.jconhyd.2008.10.002). PMID 19012994 (https://pubmed.ncbi.nlm.nih.gov/19012994).
Archived (https://web.archive.org/web/20171107005433/http://www.soil.tu-bs.de/lehre/Maste
r.Monitoring/2009/Daten/5_Literatur/A%20review%20of-Koehne-2009.pdf) (PDF) from the
original on 7 November 2017. Retrieved 22 August 2021.
218. Diplock, Elizabeth E.; Mardlin, Dave P.; Killham, Kenneth S.; Paton, Graeme Iain (2009).
"Predicting bioremediation of hydrocarbons: laboratory to field scale" (https://coek.info/pdf-pr
edicting-bioremediation-of-hydrocarbons-laboratory-to-field-scale-.html). Environmental
Pollution. 157 (6): 1831–1840. doi:10.1016/j.envpol.2009.01.022 (https://doi.org/10.1016%2
Fj.envpol.2009.01.022). PMID 19232804 (https://pubmed.ncbi.nlm.nih.gov/19232804).
Retrieved 22 August 2021.
219. 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" (https://www.academia.edu/15598352). Environmental
Science and Technology. 42 (22): 8374–8380. Bibcode:2008EnST...42.8374M (https://ui.ads
abs.harvard.edu/abs/2008EnST...42.8374M). doi:10.1021/es801703k (https://doi.org/10.102
1%2Fes801703k). hdl:11383/8693 (https://hdl.handle.net/11383%2F8693). PMID 19068820
(https://pubmed.ncbi.nlm.nih.gov/19068820). Retrieved 22 August 2021.
220. 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)" (https://www.researchgate.net/publication/225752596).
Journal of Soils and Sediments. 9: 23–32. doi:10.1007/s11368-008-0046-9 (https://doi.org/1
0.1007%2Fs11368-008-0046-9). S2CID 129416733 (https://api.semanticscholar.org/CorpusI
D:129416733). Retrieved 22 August 2021.
221. Johnson, Dan L.; Ambrose, Stanley H.; Bassett, Thomas J.; Bowen, Merle L.; Crummey,
Donald E.; Isaacson, John S.; Johnson, David N.; Lamb, Peter; Saul, Mahir; Winter-Nelson,
Alex E. (1997). "Meanings of environmental terms" (https://www.researchgate.net/publicatio
n/240784159). Journal of Environmental Quality. 26 (3): 581–589.
doi:10.2134/jeq1997.00472425002600030002x (https://doi.org/10.2134%2Fjeq1997.00472
425002600030002x). Retrieved 29 August 2021.
222. Oldeman, L. Roel (1993). "Global extent of soil degradation" (https://library.wur.nl/WebQuery/
wurpubs/fulltext/299739). ISRIC Bi-Annual Report 1991–1992. Wageningen, The
Netherlands: International Soil Reference and Information Centre(ISRIC). pp. 19–36.
Retrieved 29 August 2021.
223. Sumner, Malcolm E.; Noble, Andrew D. (2003). "Soil acidification: the world story" (https://pdf
-drive.com/pdf/Zdenko20Rengel20-20Handbook20of20Soil20Acidity2028Books20in20Soil
s2C20Plants2C20and20the20Environment292028200329.pdf#page=16) (PDF). In Rengel,
Zdenko (ed.). Handbook of soil acidity. New York, NY, USA: Marcel Dekker. pp. 1–28.
Retrieved 29 August 2021.
224. Karam, Jean; Nicell, James A. (1997). "Potential applications of enzymes in waste
treatment" (https://www.researchgate.net/publication/30002097). Journal of Chemical
Technology & Biotechnology. 69 (2): 141–153. doi:10.1002/(SICI)1097-
4660(199706)69:2<141::AID-JCTB694>3.0.CO;2-U (https://doi.org/10.1002%2F%28SICI%2
91097-4660%28199706%2969%3A2%3C141%3A%3AAID-JCTB694%3E3.0.CO%3B2-U).
Retrieved 5 September 2021.
225. Sheng, Guangyao; Johnston, Cliff T.; Teppen, Brian J.; Boyd, Stephen A. (2001). "Potential
contributions of smectite clays and organic matter to pesticide retention in soils" (https://ww
w.academia.edu/4875079). Journal of Agricultural and Food Chemistry. 49 (6): 2899–2907.
doi:10.1021/jf001485d (https://doi.org/10.1021%2Fjf001485d). PMID 11409985 (https://pub
med.ncbi.nlm.nih.gov/11409985). Retrieved 5 September 2021.
226. Sprague, Lori A.; Herman, Janet S.; Hornberger, George M.; Mills, Aaron L. (2000). "Atrazine
adsorption and colloid‐facilitated transport through the unsaturated zone" (https://lmecol.evs
c.virginia.edu/pubs/73-Sprague_JEQ2000.pdf) (PDF). Journal of Environmental Quality. 29
(5): 1632–1641. doi:10.2134/jeq2000.00472425002900050034x (https://doi.org/10.2134%2
Fjeq2000.00472425002900050034x). Retrieved 5 September 2021.
227. Ballabio, Cristiano; Panagos, Panos; Lugato, Emanuele; Huang, Jen-How; Orgiazzi,
Alberto; Jones, Arwyn; Fernández-Ugalde, Oihane; Borrelli, Pasquale; Montanarella, Luca
(15 September 2018). "Copper distribution in European topsoils: an assessment based on
LUCAS soil survey" (https://doi.org/10.1016%2Fj.scitotenv.2018.04.268). Science of the
Total Environment. 636: 282–298. Bibcode:2018ScTEn.636..282B (https://ui.adsabs.harvar
d.edu/abs/2018ScTEn.636..282B). doi:10.1016/j.scitotenv.2018.04.268 (https://doi.org/10.10
16%2Fj.scitotenv.2018.04.268). ISSN 0048-9697 (https://www.worldcat.org/issn/0048-9697).
PMID 29709848 (https://pubmed.ncbi.nlm.nih.gov/29709848).
228. Environment, U. N. (21 October 2021). "Drowning in Plastics – Marine Litter and Plastic
Waste Vital Graphics" (http://www.unep.org/resources/report/drowning-plastics-marine-litter-
and-plastic-waste-vital-graphics). UNEP - UN Environment Programme. Retrieved 23 March
2022.
229. Le Houérou, Henry N. (1996). "Climate change, drought and desertification" (http://www7.na
u.edu/mpcer/direnet/publications/publications_l/files/LeHouerou_1996.pdf) (PDF). Journal
of Arid Environments. 34 (2): 133–185. Bibcode:1996JArEn..34..133L (https://ui.adsabs.harv
ard.edu/abs/1996JArEn..34..133L). doi:10.1006/jare.1996.0099 (https://doi.org/10.1006%2Fj
are.1996.0099). Retrieved 5 September 2021.
230. Lyu, Yanli; Shi, Peijun; Han, Guoyi; Liu, Lianyou; Guo, Lanlan; Hu, Xia; Zhang, Guoming
(2020). "Desertification control practices in China" (https://doi.org/10.3390%2Fsu12083258).
Sustainability. 12 (8): 3258. doi:10.3390/su12083258 (https://doi.org/10.3390%2Fsu120832
58). ISSN 2071-1050 (https://www.worldcat.org/issn/2071-1050).
231. Kéfi, Sonia; Rietkerk, Max; Alados, Concepción L.; Pueyo, Yolanda; Papanastasis, Vasilios
P.; El Aich, Ahmed; de Ruiter, Peter C. (2007). "Spatial vegetation patterns and imminent
desertification in Mediterranean arid ecosystems" (https://www.researchgate.net/publication/
232801317). Nature. 449 (7159): 213–217. Bibcode:2007Natur.449..213K (https://ui.adsabs.
harvard.edu/abs/2007Natur.449..213K). doi:10.1038/nature06111 (https://doi.org/10.1038%2
Fnature06111). hdl:1874/25682 (https://hdl.handle.net/1874%2F25682). PMID 17851524 (ht
tps://pubmed.ncbi.nlm.nih.gov/17851524). S2CID 4411922 (https://api.semanticscholar.org/
CorpusID:4411922). Retrieved 5 September 2021.
232. Wang, Xunming; Yang, Yi; Dong, Zhibao; Zhang, Caixia (2009). "Responses of dune activity
and desertification in China to global warming in the twenty-first century" (https://www.resear
chgate.net/publication/229103975). Global and Planetary Change. 67 (3–4): 167–185.
Bibcode:2009GPC....67..167W (https://ui.adsabs.harvard.edu/abs/2009GPC....67..167W).
doi:10.1016/j.gloplacha.2009.02.004 (https://doi.org/10.1016%2Fj.gloplacha.2009.02.004).
Retrieved 5 September 2021.
233. Yang, Dawen; Kanae, Shinjiro; Oki, Taikan; Koike, Toshio; Musiake, Katumi (2003). "Global
potential soil erosion with reference to land use and climate changes" (https://www.oieau.or
g/eaudoc/system/files/documents/38/191115/191115_doc.pdf) (PDF). Hydrological
Processes. 17 (14): 2913–28. Bibcode:2003HyPr...17.2913Y (https://ui.adsabs.harvard.edu/
abs/2003HyPr...17.2913Y). doi:10.1002/hyp.1441 (https://doi.org/10.1002%2Fhyp.1441).
S2CID 129355387 (https://api.semanticscholar.org/CorpusID:129355387). Retrieved
5 September 2021.
234. Sheng, Jian-an; Liao, An-zhong (1997). "Erosion control in South China" (https://coek.info/pd
f-erosion-control-in-south-china-.html). Catena. 29 (2): 211–221.
Bibcode:1997Caten..29..211S (https://ui.adsabs.harvard.edu/abs/1997Caten..29..211S).
doi:10.1016/S0341-8162(96)00057-4 (https://doi.org/10.1016%2FS0341-8162%2896%2900
057-4). ISSN 0341-8162 (https://www.worldcat.org/issn/0341-8162). Retrieved 5 September
2021.
235. Ran, Lishan; Lu, Xi Xi; Xin, Zhongbao (2014). "Erosion-induced massive organic carbon
burial and carbon emission in the Yellow River basin, China" (https://bg.copernicus.org/articl
es/11/945/2014/bg-11-945-2014.pdf) (PDF). Biogeosciences. 11 (4): 945–959.
Bibcode:2014BGeo...11..945R (https://ui.adsabs.harvard.edu/abs/2014BGeo...11..945R).
doi:10.5194/bg-11-945-2014 (https://doi.org/10.5194%2Fbg-11-945-2014).
hdl:10722/228184 (https://hdl.handle.net/10722%2F228184). Retrieved 5 September 2021.
236. Verachtert, Els; Van den Eeckhaut, Miet; Poesen, Jean; Deckers, Jozef (2010). "Factors
controlling the spatial distribution of soil piping erosion on loess-derived soils: a case study
from central Belgium" (https://lirias.kuleuven.be/retrieve/109942). Geomorphology. 118 (3):
339–348. Bibcode:2010Geomo.118..339V (https://ui.adsabs.harvard.edu/abs/2010Geomo.1
18..339V). doi:10.1016/j.geomorph.2010.02.001 (https://doi.org/10.1016%2Fj.geomorph.201
0.02.001). Retrieved 5 September 2021.
237. Jones, Anthony (1976). "Soil piping and stream channel initiation" (https://booksc.eu/book/2
0668631/3ac27a). Water Resources Research. 7 (3): 602–610.
Bibcode:1971WRR.....7..602J (https://ui.adsabs.harvard.edu/abs/1971WRR.....7..602J).
doi:10.1029/WR007i003p00602 (https://doi.org/10.1029%2FWR007i003p00602). Retrieved
5 September 2021.
238. Dooley, Alan (June 2006). "Sandboils 101: Corps has experience dealing with common
flood danger" (https://web.archive.org/web/20080418185527/http://www.hq.usace.army.mil/c
epa/pubs/jun06/story8.htm). Engineer Update. US Army Corps of Engineers. Archived from
the original (http://www.hq.usace.army.mil/cepa/pubs/jun06/story8.htm) on 18 April 2008.
239. Oosterbaan, Roland J. (1988). "Effectiveness and social/environmental impacts of irrigation
projects: a critical review" (http://www.waterlog.info/pdf/irreff.pdf) (PDF). Annual Reports of
the International Institute for Land Reclamation and Improvement (ILRI). Wageningen, The
Netherlands. pp. 18–34. Archived (https://web.archive.org/web/20090219070320/http://water
log.info/pdf/irreff.pdf) (PDF) from the original on 19 February 2009. Retrieved 5 September
2021.
240. Drainage manual: a guide to integrating plant, soil, and water relationships for drainage of
irrigated lands (https://www.usbr.gov/tsc/techreferences/mands/mands-pdfs/DrainMan.pdf)
(PDF). Washington, D.C.: United States Department of the Interior, Bureau of Reclamation.
1993. ISBN 978-0-16-061623-5. Retrieved 5 September 2021.
241. Oosterbaan, Roland J. "Waterlogging, soil salinity, field irrigation, plant growth, subsurface
drainage, groundwater modelling, surface runoff, land reclamation, and other crop
production and water management aspects" (http://www.waterlog.info). Archived (https://we
b.archive.org/web/20100816225219/http://www.waterlog.info/) from the original on 16
August 2010. Retrieved 5 September 2021.
242. Stuart, Alexander M.; Pame, Anny Ruth P.; Vithoonjit, Duangporn; Viriyangkura, Ladda;
Pithuncharurnlap, Julmanee; Meesang, Nisa; Suksiri, Prarthana; Singleton, Grant R.;
Lampayan, Rubenito M. (2018). "The application of best management practices increases
the profitability and sustainability of rice farming in the central plains of Thailand" (https://ww
w.researchgate.net/publication/314091782). Field Crops Research. 220: 78–87.
doi:10.1016/j.fcr.2017.02.005 (https://doi.org/10.1016%2Fj.fcr.2017.02.005). Retrieved
12 September 2021.
243. Turkelboom, Francis; Poesen, Jean; Ohler, Ilse; Van Keer, Koen; Ongprasert, Somchai;
Vlassak, Karel (1997). "Assessment of tillage erosion rates on steep slopes in northern
Thailand" (https://www.academia.edu/17993140). Catena. 29 (1): 29–44.
Bibcode:1997Caten..29...29T (https://ui.adsabs.harvard.edu/abs/1997Caten..29...29T).
doi:10.1016/S0341-8162(96)00063-X (https://doi.org/10.1016%2FS0341-8162%2896%290
0063-X). Retrieved 12 September 2021.
244. Saleth, Rathinasamy Maria; Inocencio, Arlene; Noble, Andrew; Ruaysoongnern, Sawaeng
(2009). "Economic gains of improving soil fertility and water holding capacity with clay
application: the impact of soil remediation research in Northeast Thailand" (https://ageconse
arch.umn.edu/record/53064/files/RR130.pdf) (PDF). Journal of Development Effectiveness.
1 (3): 336–352. doi:10.1080/19439340903105022 (https://doi.org/10.1080%2F19439340903
105022). S2CID 18049595 (https://api.semanticscholar.org/CorpusID:18049595). Retrieved
12 September 2021.
245. Semalulu, Onesmus; Magunda, Matthias; Mubiru, Drake N. (2015). "Amelioration of sandy
soils in drought stricken areas through use of Ca-bentonite" (https://www.ajol.info/index.php/
ujas/article/download/141752/131487). Uganda Journal of Agricultural Sciences. 16 (2):
195–205. doi:10.4314/ujas.v16i2.5 (https://doi.org/10.4314%2Fujas.v16i2.5). Retrieved
12 September 2021.
246. International Water Management Institute (2010). "Improving soils and boosting yields in
Thailand" (http://www.iwmi.cgiar.org/Publications/Success_Stories/PDF/2010/Issue%202%
20-%20Improving_soils_and_boosting_yields_in_Thailand.pdf) (PDF). Success Stories (2).
doi:10.5337/2011.0031 (https://doi.org/10.5337%2F2011.0031). Archived (https://web.archiv
e.org/web/20120607030912/http://www.iwmi.cgiar.org/Publications/Success_Stories/PDF/2
010/Issue%202%20-%20Improving_soils_and_boosting_yields_in_Thailand.pdf) (PDF)
from the original on 7 June 2012. Retrieved 12 September 2021.
247. Prapagar, Komathy; Indraratne, Srimathie P.; Premanandharajah, Punitha (2012). "Effect of
soil amendments on reclamation of saline-sodic soil" (https://www.researchgate.net/publicati
on/267202667). Tropical Agricultural Research. 23 (2): 168–176. doi:10.4038/tar.v23i2.4648
(https://doi.org/10.4038%2Ftar.v23i2.4648). Retrieved 12 September 2021.
248. Lemieux, Gilles; Germain, Diane (December 2000). "Ramial chipped wood: the clue to a
sustainable fertile soil" (https://www.healthy-vegetable-gardening.com/support-files/rcw-the-
clue-to-a-sustainable-fertile-soil.pdf) (PDF). Université Laval, Département des Sciences du
Bois et de la Forêt, Québec, Canada. Retrieved 12 September 2021.
249. Arthur, Emmanuel; Cornelis, Wim; Razzaghi, Fatemeh (2012). "Compost amendment of
sandy soil affects soil properties and greenhouse tomato productivity" (https://www.academi
a.edu/31660161). Compost Science and Utilization. 20 (4): 215–221.
doi:10.1080/1065657X.2012.10737051 (https://doi.org/10.1080%2F1065657X.2012.107370
51). S2CID 96896374 (https://api.semanticscholar.org/CorpusID:96896374). Retrieved
12 September 2021.
250. Glaser, Bruno; Haumaier, Ludwig; Guggenberger, Georg; Zech, Wolfgang (2001). "The 'Terra
Preta' phenomenon: a model for sustainable agriculture in the humid tropics" (https://www.re
searchgate.net/publication/12032464). Naturwissenschaften. 88 (1): 37–41.
Bibcode:2001NW.....88...37G (https://ui.adsabs.harvard.edu/abs/2001NW.....88...37G).
doi:10.1007/s001140000193 (https://doi.org/10.1007%2Fs001140000193). PMID 11302125
(https://pubmed.ncbi.nlm.nih.gov/11302125). S2CID 26608101 (https://api.semanticscholar.o
rg/CorpusID:26608101). Retrieved 12 September 2021.
251. Kavitha, Beluri; Pullagurala Venkata Laxma, Reddy; Kim, Bojeong; Lee, Sang Soo; Pandey,
Sudhir Kumar; Kim, Ki-Hyun (2018). "Benefits and limitations of biochar amendment in
agricultural soils: a review" (https://booksc.eu/book/72239607/440436). Journal of
Environmental Management. 227: 146–154. doi:10.1016/j.jenvman.2018.08.082 (https://doi.
org/10.1016%2Fj.jenvman.2018.08.082). PMID 30176434 (https://pubmed.ncbi.nlm.nih.gov/
30176434). S2CID 52168678 (https://api.semanticscholar.org/CorpusID:52168678).
Retrieved 12 September 2021.
252. Hillel, Daniel (1992). Out of the Earth: civilization and the life of the soil. Berkeley, California:
University of California Press. ISBN 978-0-520-08080-5.
253. Donahue, Miller & Shickluna 1977, p. 4.
254. Columella, Lucius Junius Moderatus (1745). Of husbandry, in twelve books, and his book
concerning trees, with several illustrations from Pliny, Cato, Varro, Palladius, and other
antient and modern authors, translated into English (https://catalog.hathitrust.org/Record/005
783003). London, United Kingdom: Andrew Millar. Retrieved 19 September 2021.
255. Kellogg 1957, p. 1.
256. Ibn al-'Awwam (1864). Le livre de l'agriculture, traduit de l'arabe par Jean Jacques Clément-
Mullet (https://catalog.hathitrust.org/Record/009953450). Filāḥah.French. (in French). Paris,
France: Librairie A. Franck. Retrieved 19 September 2021.
257. Jelinek, Lawrence J. (1982). Harvest empire: a history of California agriculture. San
Francisco, California: Boyd and Fraser. ISBN 978-0-87835-131-2.
258. de Serres, Olivier (1600). Le Théâtre d'Agriculture et mesnage des champs (https://gallica.b
nf.fr/ark:/12148/bpt6k738381/f1.image) (in French). Paris, France: Jamet Métayer. Retrieved
19 September 2021.
259. Virto, Iñigo; Imaz, María José; Fernández-Ugalde, Oihane; Gartzia-Bengoetxea, Nahia;
Enrique, Alberto; Bescansa, Paloma (2015). "Soil degradation and soil quality in western
Europe: current situation and future perspectives" (https://doi.org/10.3390%2Fsu7010313).
Sustainability. 7 (1): 313–365. doi:10.3390/su7010313 (https://doi.org/10.3390%2Fsu70103
13).
260. Van der Ploeg, Rienk R.; Schweigert, Peter; Bachmann, Joerg (2001). "Use and misuse of
nitrogen in agriculture: the German story" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC60
84271). Scientific World Journal. 1 (S2): 737–744. doi:10.1100/tsw.2001.263 (https://doi.org/
10.1100%2Ftsw.2001.263). PMC 6084271 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6
084271). PMID 12805882 (https://pubmed.ncbi.nlm.nih.gov/12805882).
261. "Van Helmont's experiments on plant growth" (https://www.bbc.co.uk/bitesize/clips/zpgb4w
x). BBC World Service. Retrieved 19 September 2021.
262. Brady, Nyle C. (1984). The nature and properties of soils (https://archive.org/details/naturepr
operties00brad_0) (9th ed.). New York, New York: Collier Macmillan. ISBN 978-0-02-
313340-4. Retrieved 19 September 2021.
263. Kellogg 1957, p. 3.
264. Kellogg 1957, p. 2.
265. de Lavoisier, Antoine-Laurent (1777). "Mémoire sur la combustion en général" (http://www.ac
ademie-sciences.fr/pdf/dossiers/Franklin/Franklin_pdf/Mem1777_p592.pdf) (PDF).
Mémoires de l'Académie Royale des Sciences (in French). Retrieved 19 September 2021.
266. Boussingault, Jean-Baptiste (1860–1874). Agronomie, chimie agricole et physiologie,
volumes 1–5 (https://archive.org/details/8TSUP364_1) (in French). Paris, France: Mallet-
Bachelier. Retrieved 19 September 2021.
267. von Liebig, Justus (1840). Organic chemistry in its applications to agriculture and physiology
(https://archive.org/details/organicchemistry00liebrich). London: Taylor and Walton.
Retrieved 19 September 2021.
268. Way, J. Thomas (1849). "On the composition and money value of the different varieties of
guano" (https://www.biodiversitylibrary.org/item/37078#page/220/mode/1up). Journal of the
Royal Agricultural Society of England. 10: 196–230. Retrieved 19 September 2021.
269. Kellogg 1957, p. 4.
270. Tandon, Hari L.S. "A short history of fertilisers" (https://web.archive.org/web/2017012321424
1/http://www.tandontech.net/fertilisers.html). Fertiliser Development and Consultation
Organisation. Archived from the original (http://www.tandontech.net/fertilisers.html) on 23
January 2017. Retrieved 17 December 2017.
271. Way, J. Thomas (1852). "On the power of soils to absorb manure" (https://biodiversitylibrary.o
rg/page/45583402). Journal of the Royal Agricultural Society of England. 13: 123–143.
Retrieved 19 September 2021.
272. Warington, Robert (1878). Note on the appearance of nitrous acid during the evaporation of
water: a report of experiments made in the Rothamsted laboratory (https://books.google.com/
books?id=NlISAQAAMAAJ). London, United Kingdom: Harrison and Sons. Retrieved
19 September 2021.
273. Winogradsky, Sergei (1890). "Sur les organismes de la nitrification" (https://gallica.bnf.fr/ark:/
12148/bpt6k30663/f1087?lang=EN) [On the organisms of nitrification]. Comptes Rendus
Hebdomadaires des Séances de l'Académie des Sciences (in French). 110 (1): 1013–1016.
Retrieved 19 September 2021.
274. Kellogg 1957, pp. 1–4.
275. Hilgard, Eugene W. (1907). Soils: their formation, properties, composition, and relations to
climate and plant growth in the humid and arid regions (https://www.biodiversitylibrary.org/bi
bliography/24461). London, United Kingdom: The Macmillan Company. Retrieved
19 September 2021.
276. Fallou, Friedrich Albert (1857). Anfangsgründe der Bodenkunde (https://web.archive.org/we
b/20181215223343/http://digital.slub-dresden.de/fileadmin/data/321768043/321768043_tif/j
pegs/321768043.pdf) (PDF) (in German). Dresden, Germany: G. Schönfeld's Buchhandlung.
Archived from the original (http://digital.slub-dresden.de/fileadmin/data/321768043/3217680
43_tif/jpegs/321768043.pdf) (PDF) on 15 December 2018. Retrieved 15 December 2018.
277. Glinka, Konstantin Dmitrievich (1914). Die Typen der Bodenbildung: ihre Klassifikation und
geographische Verbreitung (in German). Berlin, Germany: Borntraeger.
278. Glinka, Konstantin Dmitrievich (1927). The great soil groups of the world and their
development (http://reader.library.cornell.edu/docviewer/digital?id=chla3055800#mode/1up).
Ann Arbor, Michigan: Edwards Brothers. Retrieved 19 September 2021.

Sources
  This article incorporates text from a free content work. Licensed under Cc BY-SA 3.0 IGO (license
statement/permission (https://commons.wikimedia.org/wiki/File:United_Nations_Environment_Programme
_Drowning_in_Plastics_%E2%80%93_Marine_Litter_and_Plastic_Waste_Vital_Graphics.pdf)). Text
taken from Drowning in Plastics – Marine Litter and Plastic Waste Vital Graphics​(https://www.unep.org/r
esources/report/drowning-plastics-marine-litter-and-plastic-waste-vital-graphics), United Nations
Environment Programme.

Bibliography
Donahue, Roy Luther; Miller, Raymond W.; Shickluna, John C. (1977). Soils: An Introduction
to Soils and Plant Growth (https://archive.org/details/soilsintroductio00dona). Prentice-Hall.
ISBN 978-0-13-821918-5.
"Arizona Master Gardener" (http://ag.arizona.edu/pubs/garden/mg/soils/soils.html).
Cooperative Extension, College of Agriculture, University of Arizona. Retrieved 27 May
2013.
Stefferud, Alfred, ed. (1957). Soil: The Yearbook of Agriculture 1957 (https://archive.org/strea
m/yoa1957#page/n18/mode/1up). United States Department of Agriculture.
OCLC 704186906 (https://www.worldcat.org/oclc/704186906).
Kellogg. "We Seek; We Learn (https://archive.org/stream/yoa1957#page/n17/mode/1u
p)". In Stefferud (1957).
Simonson. "What Soils Are (https://archive.org/stream/yoa1957#page/n34/mode/1up)".
In Stefferud (1957).
Russell. "Physical Properties (https://archive.org/stream/yoa1957#page/n49/mode/1up)".
In Stefferud (1957).
Dean. "Plant Nutrition and Soil Fertility (https://archive.org/stream/yoa1957#page/n100/
mode/1up)". In Stefferud (1957).
Russel. "Boron and Soil Fertility (https://archive.org/stream/yoa1957#page/n145/mode/1
up)". In Stefferud (1957).

Further reading
Soil-Net.com (http://www.soil-net.com/) Archived (https://web.archive.org/web/20080710061
716/http://www.soil-net.com/) 10 July 2008 at the Wayback Machine 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.
David R. Montgomery, Dirt: The Erosion of Civilizations, ISBN 978-0-520-25806-8
Faulkner, Edward H. 1943. Plowman's Folly. New York, Grosset & Dunlap. ISBN 0-933280-
51-3
LandIS Free Soilscapes Viewer (https://web.archive.org/web/20080705133103/http://www.la
ndis.org.uk/soilscapes) Free interactive viewer for the Soils of England and Wales
Jenny, Hans. 1941. Factors of Soil Formation: A System of Quantitative Pedology (https://we
b.archive.org/web/20130225050838/http://soilandhealth.org/01aglibrary/010159.Jenny.pdf)
Logan, W.B. 1995. Dirt: The ecstatic skin of the earth. ISBN 1-57322-004-3
Mann, Charles C. September 2008. " Our good earth" National Geographic Magazine
"97 Flood" (https://web.archive.org/web/20080624040143/http://www.mvm.usace.army.mil/R
eadiness/97flood/flood.htm). USGS. Archived from the original (http://www.mvm.usace.army.
mil/Readiness/97flood/flood.htm) on 24 June 2008. 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 (Matching suitable forage species to soil type) (https://web.archive.org/web/200608280
63956/http://forages.oregonstate.edu/is/ssis/main.cfm?PageID=3), Oregon State University
Gardiner, Duane T. "Lecture 1 Chapter 1 Why Study Soils?" (https://web.archive.org/web/20
180209052922/http://jan.ucc.nau.edu/~doetqp-p/courses/env320/lec1/Lec1.html). ENV320:
Soil Science Lecture Notes. Texas A&M University-Kingsville. Archived from the original (htt
p://jan.ucc.nau.edu/~doetqp-p/courses/env320/lec1/Lec1.html) on 9 February 2018.
Retrieved 7 January 2019.
Janick, Jules. 2002. Soil notes (https://web.archive.org/web/20050317030248/http://www.hor
t.purdue.edu/newcrop/tropical/lecture_06/chapter_12l_R.html), Purdue University
LandIS Soils Data for England and Wales (http://www.landis.org.uk/) Archived (https://web.ar
chive.org/web/20070716033248/http://www.landis.org.uk/) 16 July 2007 at the Wayback
Machine 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
Short video explaining soil basics (https://www.theguardian.com/environment/video/2019/jul/
11/its-time-we-stopped-treating-soil-like-dirt-video)
The Soil Water Compendium (soil water content sensors explained) (http://www.edaphic.co
m.au/soil-water-compendium/)
Global Soil Partnership (http://www.fao.org/globalsoilpartnership/en/)
FAO Soils Portal (http://www.fao.org/soils-portal/en/)
World Reference Base for Soil Resources (http://www.fao.org/ag/agl/agll/wrb/)
ISRIC – World Soil Information (ICSU World Data Centre for Soils) (http://www.isric.org/)
World Soil Library and Maps (http://www.isric.org/explore/library)
Wossac the world soil survey archive and catalogue (http://www.wossac.com/)
Canadian Society of Soil Science (http://csss.ca/)
Soil Science Society of America (https://www.soils.org/)
USDA-NRCS Web Soil Survey (https://websoilsurvey.nrcs.usda.gov/app/HomePage.htm)
European Soil Portal (http://eusoils.jrc.ec.europa.eu/) (wiki)
National Soil Resources Institute UK (http://www.cranfield.ac.uk/sas/nsri)
Plant and Soil Sciences eLibrary (http://passel.unl.edu/)
Copies of the reference 'Soil: The Yearbook of Agriculture 1957' in multiple formats (https://ar
chive.org/details/yoa1957)

Retrieved from "https://en.wikipedia.org/w/index.php?title=Soil&oldid=1166122963"

You might also like