You are on page 1of 49

1.3.

Chemical Composition of Earths Crust


About 95 chemical elements are known to exist in the earths crust. The possible
combinations of such a large number of elements make more than 2000 minerals already
recognized to be present in soils. However, relatively few elements and minerals are of real
importance for agricultural purpose.
The solid crust of the earth (Lithosphere) is made up largely of eight elements (O, Si,
Al, Fe, Ca, Na, K, Mg). In fact, two elements, oxygen and silicon compose some 75% of it.
On the other hand, many of the elements important in the growth of plants and animals occur
in very small quantities. These elements and their compounds are not evenly distributed
throughout the earths surface. For example, in some places phosphorous compounds are so
concentrated that they are mined, whereas in many other areas there is a deficiency of
phosphorous for maximum plant growth.

Table 1.1. The Composition of Lithosphere


Sr. Composition Composition
Elemental Oxides
No. (%) (%)
1 O 47.3
2 Si 27.7 SiO2 59.07
3 Al 7.8 Al2O3 15.22
4 Fe 4.5 Fe2O3 3.1
FeO 3.7
5 Ca 3.5 CaO 5.1
6 K 2.5 K2O 3.11
7 Na 2.5 Na2O 3.71
8 Mg 2.2 MgO 3.45
9 P 0.1 P2O5 0.3
10 S 0.1
11 Ti 0.5
12 H 0.2
13 C 0.2
14 Mn 0.1
15 Others 0.8

1.4. Elements Required in Plant Nutrition


Protoplasm is common to plant and animal life. Animals however, must rely on other
animals or plants as a source of food so that they may continue to live. In the final analysis,
animals are completely dependent on the plant kingdom for life. The protoplasm of green
plants however, can exist and increase independently of any animal life. All that is required is
a supply of water, carbon dioxide and several mineral elements to make the green plant in
light a completely self-sufficient organism. The raw materials that the plant consumes in the
manufacture of its own tissues are of great importance in the production of crop plants.
-2-
1.5. Criteria of Essentiality of Elements in Plant Nutrition
Plants absorb mineral elements from the rooting medium; but the presence in a plant
of any particular element is of no proof that the element is essential to its development.
Arnon (1953) of California has made the following points in this regard:
(1) A deficiency of the element makes it impossible for the plant to complete the
vegetative or reproductive stage of its life.
(2) The deficiency symptom of the element in question can be prevented or corrected
only by supplying the element.
(3) The element is directly involved in the nutrition of the plant, quite apart from its
possible effect in correcting some microbiological or chemical condition in the soil or
culture medium.
Results of more recent work indicate that point 2 may be too rigid and found to be
arguable.
e.g. 1. In some species, vanadium can substitute completely for the molybdenum.
2. Bromine may substitute for chlorine though at a higher concentration.
3. Iodine, which until recently has not been considered essential by Arnons criteria has
been shown to increase the yield of crops such as sugar beets, celery, turnip, and
table beets.

1.6. Essential Elements for Plant Growth


Carbon, hydrogen, oxygen, nitrogen, phosphorus, and sulphur are the elements
usually found in proteins. In addition to these six, there are fourteen other elements which are
essential to the growth of some plant or plants; calcium, magnesium, potassium, iron,
manganese, molybdenum, copper, boron, zinc, chlorine, sodium, cobalt, vanadium, and
silicon. (The essentiality of sodium, cobalt, vanadium and silicon are still yet to be confirmed.
Not all are required for all plants, but all have been found to be essential to some.)

1.7. Contents of Chemical Elements in Soils


The proportion of chemical elements in soil are greatly changed by chemical
weathering and leaching and this is markedly reflected in the composition in the fine fractions
of the alluvial or B horizons even before much change is noted in the composition of the
whole soil mass. Oxides of iron, aluminum, and titanium increase markedly even on a whole
soil basis when nearly all of the minerals have undergone pedogeochemical transformations.

-3-
In the following sections the discussion will be made for a selected number of
elements usually found in soils not in order of merit for their importance to plant nutrition but
in order of merit for their abundance in most soils.

1.7.1. Silicon
The silicon content, expressed as percentage of SiO2, ranges from 50 to 70% of many
soils, which averages about the same as for the earths crust. In sandy soils, the SiO2 content
rises as in sandstone. The silicon percentage is decreased with the increases of organic matter
in organic soils and with the increase of CaCO3 as in marl.
Silicon is gradually lost from many soils especially in humid tropical soils which have
been subjected to intensive weathering and leaching. It can be clearly seen in the following
table.

Table 1.2. Comparisons between the chemical compositions of lateritic soil and black
cotton soil
Lateritic Soil Black cotton soil
(top layer) (Mandalay Soil)
%
SiO2 20.33 69.76
Al2O3 11.68 10.45
Fe2O3 41.60 4.48
CaO 0.59 1.38
MgO 0.88 1.0
K2O 0.90 0.71
Na2O 0.29 0.28
P2O5 0.21 0.05
SO3 0.24 0.20
Soil organic matter 5.00 0.75

1.7.2. Aluminum
After oxygen and silicon, aluminum is the most abundant element in the earths crust
and in the majority of rocks and soils. The aluminum content of soils, expressed on the basis
of Al2O3, frequently is in the range of 2 to 12%. The content ranges up to 20 to 60 percent in
the highly weathered soils and laterites. These examples are extreme and should not be taken
as a measure of the majority of the soils of the humid tropics. Aluminum occurs mainly in the
aluminosilicate minerals, feldspars, amphiboles, pyroxene, and layer silicates.
As silicon is depleted and aluminum is enriched, the molar ratio, SiO2/Al2O3 in soil
colloids falls from over 4 in colloids high in layer silicate to less than 1 in colloids high in
allophone.

-4-
1.7.3. Iron
The iron content, expressed as percentages of Fe2O3, make up 1 to 6 percent of many
soils, which is comparable to the composition in the earths crust (about 7%) and the
composition in various rocks (1 to 7 percent). Iron is subject to increase in concentration
through soil development processes, as reflected by contents of 10 to 15 percent in many soil
colloids. In latosols and laterites, the Fe2O3content is frequently 20 to 80 percent. Soils
having the higher percentages are approaching the iron content of low-grade iron ore.
High stability of soil aggregates and high soil porosity are usually associated with
high iron oxide contents. Under poor drainage the iron becomes reduced and in the presence
of organic matter is frequently mobilized. It can be removed from the soil profile to a lower
position in the landscape, but frequently becomes concentrated as an iron-rich layer in the
lower portion of the soil profile.
The principle form of iron in soils is generally as hydrous oxides, but iron freely
enters the 2:1 and 2:2 layer silicate structures of soils, including biotite, vermiculite and
montmorillonite. Smaller percentages of iron occur in soils as pyroxene and amphiboles.

1.7.4. Titanium
The titanium content, expressed as TiO2, makes up 0.2 to 1 percent in many soils of
temperate regions. This value is similar to the range for sedimentary and igneous rocks.
The titanium of soils occurs primarily as fine grained crystals of free TiO2 (rutile and
anatase) and FeTiO3 (ilmonite). Since titanium oxide minerals are relatively resistant to
weathering the titanium content of soils tend to increase as other elements are leached away.
The TiO2 content of highly weathered soils tends to run 2 percent or more on a dry-soil basis.
That of laterites in various localities increases into the range of 5 to 20 percent or more,
approach the range of titanium ores. Not all latosols and laterites are extremely high in TiO2.

1.7.5. Calcium
The calcium content, expressed as CaO, is generally in the low ranges of about 1
percent in soils except when calcium occurs in carbonate or sulphate form. The content of
calcium in soils is thus low, compared to that in igneous and sedimentary rocks, the CaO
content of which averages about 5 percent. Limestone average 43 percent CaO. The decrease
reflects the fact that carbonates and sulphates of calcium leach readily from the upper
horizons of well drained soils except under arid climates.
Calcium carbonate occurs in many soils, in lower horizons. This form of calcium is
found in soil horizons at lesser depths as one transverse from more humid to more arid
-5-
regions. The black earth soils of the world, including such broad groups as Chernozems and
Rendzinas in North and South Dakota, Texas, Alabama, South Africa, Central India, Russia
and Australia frequently contain 40 to 50 percent CaCO3 in the subsoil horizons. Younger
soils in the humid region may contain calcium carbonate, as calcite (CaCO3) or dolomite
(MgCO3CaCO3) inherited from the parent rock.
Pedogenic CaCO3 is mainly the mineral calcite, precipitated when leaching water
carries Ca(HCO3)2 downward to certain depth from which the water is removed by roots or is
evaporated. Crystalline gypsum (CaSO42H2O) also is accumulated in soils of semi-arid
regions. A few hundred parts of calcium phosphate per million of soil occur in nearly neutral
to calcareous soils.

1.7.6. Magnesium
In well-leached soils, magnesium is found chiefly in minerals, such as biotite, augite,
hornblende, and montmorillonite. The magnesium content of soils expressed as MgO
frequently is less than 1 percent in non-calcareous soils. Dolomite CaCO3MgCO3 and Mg
substituted calcite (Ca, Mg )CO3 occur in substantial quantities in some soils, inherited from
parent rocks. Soil contents of 2 percent or more of MgO in carbonate form occur in Brown,
Chestnut, and Chernozen soils of the semiarid parts of the world.

1.7.7. Potassium
The potassium content, expressed as K2O, ranges between 0.05 to 3.5 percent for
mineral soils. On a percentage basis, this is somewhat lower than the average content of
potassium in igneous rocks, but somewhat higher than that in some sedimentary rocks.
The proportion of the total potassium in soils held in soluble and exchangeable forms
is usually relatively small. The majority of it resides in potassium bearing feldspars and
micas. The high potassium primary silicate mineral are muscovite, biotite, orthoclase and
microcline; but other micas, feldspars, and other minerals may contain substantial amount of
potassium.
The chemical relationships between the various forms of soil potassium can be
expressed by the following equation.

-6-
Because of equilibrium between primary mineral forms and exchangeable and water-
soluble potassium, plants can sometime survive on soils having primary minerals in fine
particle sizes as the sole source of potassium. Of the layer-silicate clays, only mica or illite
has a substantial potassium content. A very close correlation has been found between the
specific surface of mica in soils and potassium released for plant growth. In silt loam and
clay loam of the sub-humid region, the exchangeable potassium is kept at adequate levels by
weathering release of potassium. Intensive cropping and removal of potassium in the humid
region, on sandy soils generally, results in the development of stunted growth of plants
exhibiting potassium deficiency symptom. Application of water soluble forms of fertilizer
such as KCl and K2SO4 is necessary for intensive agriculture on such areas.

1.7.8. Sodium
The sodium content, expressed as Na2O, ranges from 0.1 to 1 percent of many soils.
This is much smaller than the average of 3.7 percent in the earth's crust. The Na2O content in
sedimentary rocks ranges from 1.3 percent in shale to 0.4 percent in sandstone and 0.05
percent in limestone. The decreases in Na2O content in soils and sedimentary rocks reflect
weathering away of sodium-bearing minerals. The high sodium content of ocean waters
reflects the leaching of sodium from soil minerals.
Sodium occurs as NaCl, Na2SO4 and sometimes as Na2CO3 and other soluble salts in
saline and alkaline-saline soils, soils in which, leaching has not been intensive enough to
remove the soluble salts. Sodic soils, containing exchangeable sodium on the order of
-7-
15 percent or more of the cation-exchange capacity, tend to disperse and develop adverse
tillage properties.

1.7.9. Phosphorus
The phosphorus content of most mineral soils falls between 0.02 and 0.5 percent P, and
a general average of 0.05 percent (0.12 percent P2O5) frequently is representative of soils
compared to an average of 0.12 percent P in the earth's crust. About half the soil phosphorus
occurs in combination with organic matter of surface soils, and the remainder occurs in
mineral or inorganic combination.
Inorganic phosphorus occurs mainly as calcium phosphate in alkaline and calcareous
soils. The main calcium phosphate in soils is fluorapatite Ca10(PO4)6F2. As weathering
proceeds and the acidity develops in soils, the phosphate becomes increasingly bonded to
aluminum and iron ions released from silicate minerals by weathering.
Under reducing conditions, vivianite Fe3(PO4)28H2O forms in soils and being fairly
soluble, may be deposited in lower levels in the soil profile. Reduction of ferric ion renders
the oxides soluble and the phosphate becomes more available to plants, as to rice and paddy
culture. The phosphate which had been associated with iron may leach out of highly reduced
soils, while the relative amount of aluminum phosphate form increases in acid soils.
The solubility of phosphate from calcium aluminum and iron compounds are close to
each other in the soil pH range between pH 6 and 7. Below this pH range, the aluminum and
iron phosphate are more stable than calcium phosphate; above this range various calcium
phosphate are formed. The calcium formed include hydrooxy apatite [Ca10(PO4)6(OH)2],
dicalcium phosphate [CaHPO4] etc. Finely ground phosphate rock containing much
fluorapatite Ca10(PO4)6F2 is sometimes added to soils as an amendment. This phosphate in
acid soils is rapidly transformed into combination with aluminum and iron phosphate.
The rates of transformations of soil phosphates are very slow. Calcium phosphate is
highly unavailable at high soil pH values. Iron and aluminum phosphates of acid soils
become more available when the soil pH is raised by liming.

1.7.10. Molybdenum
The molybdenum content of soils generally ranges from 1 to 10 ppm but in certain soils
it may rise to 20 or 30 ppm or more. The element Mo occurs in soil mainly as MoO 4 ion and
undergoes fixation as basic iron and aluminum molybdates. The availability of soil
molybdenum rises with the rises in soil pH on liming. Plant deficiencies in Molybdenum

-8-
occur in soils low in this element. Plant contents of Mo may become sufficiently high as to
become toxic to animals when level of available Mo in soil is high.

1.7.11. Nitrogen
The quantity of nitrogen in surface soils generally ranges from 0.02 to 0.25 percent. It is
closely related to the amount of soil organic matter which contains approximately 5 percent
nitrogen. The available forms of nitrogen for plant nutrition are nitrates, nitrites and
exchangeable ammonium and they make up less than 1 percent of total soil nitrogen content
of mineral soils.
An appreciable fraction of the nitrogen content of subsoil and rocks occur as NH4+ ions
substituting for K+ in micas. These NH4+ ions are released when mineral fractions are treated
with hydrofluoric acid.

1.7.12. Sulphur
Sulphur is present in soils in both inorganic and organic forms. In well-leached surface
soils, much of the sulphur is combined with organic matter. Soils of humid temperate region
have 50 to 500 ppm of water-soluble sulphates and 100 to 1500 ppm total sulphate. Soil high
in free iron oxides contains SO4 ions substituting for hydroxyl in basis sulphate complexes.
This form of sulphate is gradually extractable in NaOH solutions and forms a sulphur
reservoir of slowly available sulphate.
Crops remove 10 to 30 pounds of sulphur per acre per year, and 40 to 60 pounds of
sulphur per acre appear in the leachate of soils. Only a few area shows sulphur deficiency.
Sulphur is added to soils in rain and snowfall, amounting to as much as 30 pounds of sulphur
per acre in industrial and urban areas where coal, oil and gas are burned, and as little as 5
pounds in thinly populated areas. Sulphur is added to soils in many commercial fertilizers and
in irrigation waters.

1.7.13. Selenium
Selenium is found in trace amounts in most soils. But crops grown on soils which are
exceptionally high in selenium have an excessive selenium content. Under conditions of
limited rainfall, enough Se can be present in the crops grown on these soils to produce
vegetation toxic to livestock. Certain native plants are ''accumulators'' of Se. Analysis of then
may be used as a guide to soil areas of excess selenium supply.
The element Se behaves much like Se, oxidizing to selenite on weathering, being
subject to leaching, and combining with free iron oxides to form basic selenates.
-9-
1.7.14. Boron
The total boron content of soils generally ranges from 4 to 98 ppm and may average
about 30 ppm. This is comparable to 10 ppm in igneous rocks and 4.5 ppm in sea water.
Sandy soils may contain as low as 2 to 6 ppm. The boron, soluble in 85 percent H3PO4 at
100C (The average is about 17 ppm) consists of that present in organic matter and associated
with layer silicates. Acid-soluble boron is associated with the mica fraction of sediments and
mica-derived clays of soils.
Plant availability of soil boron is most closely correlated with boiling water-soluble
boron which ranges from 0.2 to 1.5 ppm in soils of humid temperate regions. Soils of arid and
semiarid areas may contain 10 to 40 ppm. Extensive leaching of fine-textured desert soils
may be required to decrease the boron toxic levels before crop production under irrigation
can be carried out.

- 10 -
CHAPTER 2
CHEMISTRY OF SOIL COLLOIDS

2.1. The Colloidal Fraction of Soil


Physical chemistry of soils is essentially a study of the most reactive fractions of soils
or the so called colloidal phase. For agricultural purpose, two types of colloids in soils are
considered important; they are mineral colloids and organic colloids. They are chemically the
most active portions of the soil and exist in intimate mixture. Organic colloids consist of
organic residues in various states of decomposition. Mineral colloids consist mainly of
hydrous aluminum silicate clay minerals of various kinds and the hydrous oxides of iron and
aluminum.
The inorganic or mineral colloids will be initially discussed and the organic colloids
will be left for later consideration.

2.2. The Origin of Silicate Clay Minerals


Silicate clay minerals are hydrated crystalline aluminosilicates of secondary origin
and include illite, montmorillonite and kaolinite. They are commonly called silicate clays.
The hydrous oxides of aluminum and iron are commonly called oxide clays.
These secondary minerals are formed by alteration of existing minerals or by
synthesis. Their origin is related to the kinds of minerals in the soils that can be altered and
the amount and kinds of constituents that can recombine.
Micas, feldspars and ferromagnesian minerals may weather directly to silicate clay
minerals like illite, montmorillonite, kaolinite or even hydrooxide clays like gibbsite
(hydrated aluminum oxide).
The nature of weathering environment plays an important role in determining when a
given mineral will be formed.
In temperate region, illite formation is common.
Montmorillonite formation requires an abundant supply of magnesium and a neutral
or only slightly acid environment. In the temperate regions illite can weather into
montmorillonite.
Prolonged leaching of a soil can load to the development of high acidity and
conditions favorable for the alteration of montmorillonite to kaolinite. Kaolinite is often
formed directly from primary minerals in soils of the humid tropics. Kaolinite can also be

- 11 -
formed by the resilication of aluminum oxide when silica rich water is present. Although
kaolinite is very stable, it can weather to form gibbsite, Al(OH)3.

2.2.1. Size of Colloidal Fraction


The upper limit in size of the mineral colloidal particle is less than 0.001 mm in
diameter, or one micron (P). Values, as low as 0.5 micron or even 0.2 microns are commonly
accepted. Because the maximum size limit of the clay fraction of a soil is considered to be
0.002 mm in diameter (2 microns), not all the clay particles are considered strictly colloidal.
The finely divided particles of the colloidal fraction have a very high specific surface
area, which is important from the stand point of water adsorption.
They also have electrochemical charge which is important for the adsorption of
nutrient ions.

2.2.2. Shape of Colloidal Clays


Silicate clays are laminated, that is made up of layers of plates or flakes. Their
individual sizes and shapes depend upon their mineralogical composition and conditions
under which they have developed.
Some of the minerals are hexagonal, others are irregularly flake-like; still others seem
to be lath-shaped blades or even rods. The edges of some particles are cleans out while the
appearances of others is indistinctly frayed or fluffy.
In all cases, the horizontal extension of the individual particles greatly exceeds their
vertical dimension.

2.2.3. Surface Area of Colloidal Clays


All clay particles, because of their fineness of divisions, expose a large amount of
external surface. Besides their fineness, the shape of the clay particles also has a considerable
influence on the exposed surface area. The silicate clay minerals are generally plate-like in
shape, thus exposing a very large amount of surface per unit weight. For example, according
to L. D. Baver (1956) quoting Ostwald, points out that:
- a spherical particle with volume of 1 cubic centimeter (cc) expose 4.8 square
centimeter (sq cm) of surface,
- a cubical particle with a volume of 1 cc expose 6 sq cm of surface, and
- a plate-shaped particle with a volume of 1cc (with a height of 1 mm) expose 21 sq cm
of surface.

- 12 -
In addition to the fineness of the particle and their shape (i.e. plate-like structure), the
internal surface in some clay particle also play an important role in determining the reactive
surface area. The internal surface occurs between the plate-like crystal units that make up
each particle. It is illustrated in the following diagram.

Diagram 2.1. Silicate clay crystal with its sheet-like structure, its in-numerable negative
charges and its swarm of adsorbed cations and an enlarged schematic
view of the edge of the crystal illustrates the negatively charged internal
surface of this particular particle to which cations and water are attracted

2.2.4. Electronegative Charge and Adsorbed Cations of Colloidal Clays


The minute silicate-clay colloid particles, referred to as micelles (microcells),
ordinarily carry negative charges. Consequently, thousands of positively charged ions or
cations are attracted to each colloidal crystal. This gives rise to what is known as an ionic
double layer (see Diagram 2.1). The colloidal particles constitute the inner ionic layer, being
essentially a huge anion; the surfaces are highly negative in charge. The outer ionic layer is
made up of a swarm of loosely held cations which surrounds the clay particle. Thus a clay
particle is accompanied by an enormous number of adsorbed cations.
A large number of water molecules are associated with the layer of adsorbed cations.
Parts of these water molecules are carried by the adsorbed cations since most of them are
definitely hydrated. In addition, some silicate clays held numerous water molecules packed
between the plates of the clay micelle.

- 13 -
Diagram 2.5. Six silica tetrahedrons joined together to form a hexagonal ring by
sharing basal oxygen with each other.

An interlocking plane of a series of silica tetrahedra tied together by shared


oxygen atoms gives a sheet like tetrahedral layer (See Diagram 2.6.).

Single Chain

Diagram 2.6. A silica tetrahedral layer Double chain structure [n(Si2O5)]

All basal oxygen atoms of the silica tetrahedral layer are shared. The apical oxygens
of the layer are unshared and all point in the same direction. The layer has a net negative
charge and the formula is n(Si2O5).

Alumina octahedron
The alumina octahedron is an eight sided building block. It consists of a core
aluminum atom surrounded by six hydroxyls or oxygen (See Diagram 2.7.).

- 16 -
NOTE:
x That the aluminum atom is larger than the silicon.
x That the interstice formed by an octahedron is larger than the interstice formed by a
tetrahedron.
x That the size of the interstice governs the kind of atoms that can exist in these structures.
(This will be further discussed in the origin of the negative charge of clay minerals.)
Adjacent aluminum octahedra share common hydroxyls to form an octahedral layer in
a manner comparable to the sharing of oxygen in the silica tetrahedral layer.
All of the hydroxyls are shared, giving the layer a neutral charge and the formula
Al(OH)3. This is the structure of the mineral gibbsite.
These two basic layers (i.e. Silica tetrahedron and alumina octahedron), in different
stacking arrangements and combinations, provide the fundamental structural units of silicate
clays. The layers are bound to each other within the clay crystals by shared oxygen atoms.

Isomorphous substitution
The silicon in the tetrahedral layer and the aluminum in the octahedral layer can be
replaced or substituted by other ions of comparable size. The atomic radii of a number of ions
common in clays are listed in Table 2.1 to illustrate this point.

Table 2.1. Ionic Radii of Elements Common in Silicate Clays and an Indication of
which are found in the Tetrahedral and Octahedral Layers.
Ion Radius (A) Found in
Si4+ 0.41
Silica Tetrahedra
Al3+ 0.50
Fe3+ 0.64
2+
Mg 0.65
Alumina Octahedra
Zn2+ 0.70
Fe2+ 0.75
Ca2+ 0.94
Na+ 0.98 Exchange sites
K+ 1.33
O 1.45 Both layers
An angstrom unit (A) is 10-8 centimetre.

- 18 -
Note that aluminum is only slightly larger than silicon. Therefore aluminum can fit
into the centre of the tetrahedron in the place of the silicon atom. When some silicates form,
part of the silicon atoms in the layer are replaced by aluminum by a process called
isomorphous substitution. This substitution of a three-valent ion (Al3+) for one with four
valences (Si4+) results in one negative charge in a previously neutral silicate layer. The extent
of such substitution helps determine the net negative charge.
The Table 2.1 also shows that ions such as iron, zinc, and magnesium are not too
greatly different in size from aluminum. As a result these ions can fit in the place of
aluminum as the central ion in the units making up the octahedral layer. The isomorphous
substitution of the two-valent ion such as Mg2+ for three-valent Al3+ leaves unsatisfied
negative charges from the oxygen atoms in the layer. This type of substitution helps account
for the overall negative charge associated with several silicate clays.

2.2.6. Mineralogical Organization of Silicate Clays


On the basis of the number and arrangement of tetrahedral (silica) and octahedral
(alumina) layers contained in the crystal units, they may be classified into four different
groups:
(a) 1:1 (silica : alumina)-type minerals,
(b) 2:1-type minerals, which expand between crystal units,
(c) 2:1-type minerals, which do not expand between crystal units, and
(d) 2:2-type minerals.

1:1-type mineral
The crystal units of the 1:1-type minerals are made up of one silica (tetrahedral) layer
alternating with one alumina (octahedral) layer, hence, the terminology 1:1-type crystal
lattice (see Diagram 2.9).
In soils, kaolinite is the most prominent member of this group which includes
halloysite, anauxite, and dickite.

- 19 -
Isomorphous substitution of magnesium for some of the aluminum in the octahedral
sheet and also the substitution of aluminum for silicon in the tetrahedral sheet provides
montmorillionite crystals with a high net negative charge. This charge is satisfied by a swarm
of cations (H+, Al3+, Ca2+, K+, etc.) which are attracted to both the external and internal
surfaces. Montmorillonite possesses a high cation adsorption capacity perhaps 10 15 times
of kaolinite.
Montmorillonite has high plasticity and cohesion. It also has a marked shrinkage on
drying. Wide cracks are formed when soils containing a high percentage of montmorillonite
are dried out. The dry aggregates or clods are very hard, making such soils difficult to till.

(b) The Vermiculite Group


Vermiculites have structural characteristics similar to those of montmorillonite i.e.
2:1-type . However, in some vermiculite, there are more magnesium ions than aluminum in
the octahedral layer. In tetrahedral layer of vermiculites, there has been considerable
substitution of aluminum for silicon. These substitutions provide the mineral with a very high
net negative charge.
Water molecules and magnesium ions are strongly adsorbed between crystal units.
However they act more as bridges holding the units together than as wedges driving the units
apart. Therefore the degree of swelling is less for vermiculite than for montmorillonite.
The cation adsorption capacity of vermiculite exceeds that of all other silicate clays,
including montmorillonite. This is due to the high negative charge in the tetrahedral layer.
Vermiculite crystals are larger than montmorillonite crystals but are much smaller
than kaolinite crystals.

Table 2.2. Comparative Properties of Three Major Types of Silicate Clay.


Type of Clay
Property
Montmorillonite Illite Kaolinite
Size (Pm) 0.01 - 1.0 0.1 - 2.0 0.1 - 5.0
Shape Irregular flakes Irregular flakes Hexagonal
Specific surface area (m2/ g) 700-800 100-120 5-20
External surface area High Medium Low
Internal surface area Very high Medium None
Cohesion, plasticity High Medium Low
Swelling capacity High Medium Low
Cation exchange capacity (CEC) 80-100 15-40 3-15
(meq/100 g)
- 22 -
2:1-type nonexpanding minerals
There is a rather indefinite group called the hydrous micas. Illite is the most important
member of this group. Like montmorillonite, illite has a 2:1-type lattice. However, illite
particles are much larger than montmorillonite particles.
The source of charge is in the tetrahedral layer rather than the octahedral layer. About
15 percent of the tetrahedral silicon is substituted by aluminum atoms. This results in a high
net negative charge in the tetrahedral layer.
To satisfy the negative charges, potassium ions ate strongly attracted between the
crystal units. These K ions are just the right size to fit snugly into certain spaces in the
adjoining tetrahedral layer. The potassium, thereby, acts as a binding agent, preventing much
of the expansion of the crystal. Therefore, illite is relatively non-expansive.
Properties such as hydration, cation absorption, swelling, shrinkage and plasticity are
less intense in illite than in montmorillonite. Nevertheless, illite exceeds kaolinite in respect
to these properties.
The size of the illite crystals is intermediate between montmorillonite and kaolinite.

2:2-type mineral
The most important member of this silicate group is chlorites which are common in
some soils. Chlorites are basically silicates of magnesium with some iron and aluminum
present.
A typical chlorite crystal unit is composed of alternate tale (similar to a
montmorillonite crystal unit) and brucite Mg(OH)2 layers. Magnesium dominates the
octahedral position in the tale layer thus the crystal unit contains two silica tetrahedral sheets
and the two magnesium octahedral sheets, giving rise to the term "2:2" type lattice structure.
Chlorites have about the same cation exchange capacity as illite but they have less
CEC than montmorillonite and vermiculite.
Particle size and specific surface area for chlorite are about the same as for illite.
There is relatively little water adsorption between the chlorite crystal units, with a
result of relatively non-expansive nature of this mineral.

Mixture of silicate clays


In soils, several clay minerals are usually found in an intimate mixture. Minerals
having properties and composition intermediate between those of two well defined minerals
(those described earlier) will also be found. Such minerals are termed mixed layer because
- 23 -
within a given crystal, the individual crystal units may well be of more than one type. Terms
such as "Chloride-illite" and "illite-montmorillonite" are used to describe mixed-layer
minerals. In some soils they are more common than the single-structured minerals such as
montmorillonite.

2.2.7. Sources of the Negative Charge on Silicate Clays

Negative charges on exposed crystal edges


There are at least two sources of the negative charges associated with silicate clay
particles. The first involves unsatisfied valences at the broken edges of the silica and alumina
sheets. Also, the flat external surfaces of minerals such as kaolinite have some exposed
oxygen and hydroxyl groups which act as negatively charged sites. These groups are attached
to silicon and aluminum atoms within their respective sheets. Especially at high pH, the
hydrogen of these hydroxyls dissociates slightly and the colloidal surface is left with a
negative charge carried by the oxygen. The loosely held oxygen is readily exchangeable. The
situation is illustrated in the Diagram 2.11.

Si Si

Diagram 2.11. A broken edge of a kaolinite crystal, showing oxygen as the source of
negative charge. At high pH values the hydrogen ions tend to be hold
loosely and can be exchanged for other cations.

The presence of thousands of such groups gives the kaolinite clay particles a definite
electronegativity.
In acid soils, the hydrogen is tightly hold and not readily replaced by other cations.
Therefore this type of electronegativity is known as pH-dependent charge.

- 24 -
In most cases, the properties of the amorphous mineral do not differ greatly from
those of crystalline materials. The amorphous mineral has a high cation exchange capacity,
which is pH dependent.

2.5. Organic Soil ColloidsHumus

Diagram 2.14. Schematic presentation of metal-humic substances complexation


(Stevenson, 1982). 1, Electrostatic interaction; 2, inner-sphere
complexation; 3, weak: water bridging.

Humus has a colloidal organization of a highly charged anion (micelle) surrounded by


a swarm of adsorbed cations, as shown in Diagram 2.13.

Diagram 2.13. Colloidal organization of a humus colloid. Note the phenolic hydroxyl
groups are attached to the aromatic rings ( OH); and carboxyl
groups (COOH) are bonded to other carbon atoms in the central unit.
- 27 -
Some important properties of organic colloids should be noted in composition with
the inorganic colloids.
1. The humus micelle is composed basically of carbon, hydrogen, and oxygen. (Silicate
clays are composed basically of aluminum silicon, and oxygen.)
2. The cation exchange capacity of humus per unit weight is very high (It is higher than
that of monmorillonite.)
3. The humus micelle is not crystalline. (Most of silicate clays are crystalline.)
4. The humus is dynamic, i.e., it is being formed and destroyed rapidly. (Silicate clays
are more stable, i.e., they are being formed and destroyed more slowly.)
5. The charge on humus colloids is pH dependent. (The charge for part of the silicate
clays is pH dependent.)
6. The major sources of negative charge are thought to be carboxylic (COOH) and
phenolic ( OH) groups. (The major sources of negative charge on clay colloids
are exposed crystal edges and isomorphous substitution.) (See Diagram 2.13)
7. Because of their complexity, specific structure of humus colloids is not clearly
understood yet. (Specific structure of clay colloids are more or less understood.)

2.6. Anion Exchange


It should be noted that some clay minerals exhibit positive as well as negative charge.
This makes possible anion exchange between surface hydroxyl units and anions such as
phosphate, sulphate, chloride and nitrate. This anion exchange phenomenon will be
discussed in more detail in Chapter 3.

- 28 -
CHAPTER 3
CATION AND ANION EXCHANGE PHENOMENA

Ion exchange is the reversible process by which cations and anions are exchanged
between solid and liquid phases. This exchange also occurs between solid phases which are
in close contact with each other. Ion exchange is generally considered as the most important
of all the processes occurring in a soil.
The ion-exchange property of a soil is due almost entirely to the clay and silt fractions
(<20 P) and the organic matter. The colloidal material of the soil is considered most
important for this exchange process.

3.1. Cation Exchange


Cation exchange is one of the most important of soil reactions. The adsorption of one
or more cations by a colloid and the accompanying release of one or more ions held by the
colloid is termed cation exchange.

3.1.1. Simple Example of Cation Exchange Phenomenon


A substantial amount of organic and inorganic acids is formed when organic matter
decomposes. As a result, hydrogen ions are generated. They tend to replace the exchangeable
ions like calcium which are adsorbed on the colloid. This situation can be shown as follows:

2+ + 2+
Ca Micelle + 2H H+ Micelle + Ca
H+

The reaction takes place fairly rapidly. The interexchange of calcium and hydrogen is
chemically equivalent. If the hydrogen ion concentration were decreased for any reason or the
calcium ions are increased by adding a calcium-containing limestone, the adjustment would
be to the left in response to mass action.
If the hydrogen ions are increased or if the calcium ions are removed by leaching, the
adjustment would be to the right.

- 29 -
3.1.2. Important Factors Governing the Cation Exchange Process in Soil
The efficiency of cation exchange process is determined by a number of factors such
as:
(1) Relative concentration or numbers of the ions,
(2) The number of charges on the ion, and
(3) The speed of movement of the different ions.
The first factor is an application of the well-known chemical law of mass action
(sometimes known as the law of chemical equilibrium). Guldberg and Waage in 1867 stated
the law of mass action as the velocity of chemical reaction is proportional to the product of
the active masses (concentrations) of the reacting substances.
The second factor, the greater the number of charges carried by the ion the greater is
its efficiency, other factors being the same.
For the third factor, the speed or activity of an ion is primarily a function of its size.
However, the degree of hydration should also be considered. When an ion associates itself
with water molecules, its size becomes larger due to the shell of water. Consequently the
speed of its activity is much reduced.

3.1.3. Cation Exchange under Natural Conditions


In humid region surface soils, the reactions similar to the following example
commonly occur.

Micelle Micelle

NOTE:
For the salt of simplicity, it is assumed that the number of calcium (Ca), aluminum (Al),
hydrogen (H), and other metallic cations (M) are in the ratio 40, 20, 20 and 20 per micelle,
respectively. The metallic cations are considered monovalent.

It is important to note that where sufficient rainfall is available to leach the calcium,
the reaction tends to go toward the right that is, hydrogen ions are entering and calcium and
other bases (M) are being forced out of the exchange complex into the soil solution. As long
as bases are removed and the hydrogen ion concentration is not correspondingly reduced, the
adjustment will continue to the right.

- 30 -
In regions of low rainfall, the calcium and other salts are not easily leached from the
soil. Thus the removal of bases from the exchange sites is prevented, and the pH of the soil is
kept neutral or alkaline.
It is important to note that the interaction of climate, biological processes, and cation
exchange helps determine the properties of soil.

3.1.4. Influence of Lime and Fertilizer on Cation Exchange


Cation exchange reactions are reversible. If some form of limestones or other basic
calcium compound is applied to an acid soil, the reverse of the replacement stated in section
3.1.3 occurs.
The active calcium ions replace the hydrogen and other cations by mass action. As a
result, the colloidal micelle becomes higher in exchangeable calcium and lower in adsorbed
hydrogen. And the pH of the soil is naturally raised.
When a soil is treated with an application of a fertilizer containing KCl, an exchange
such as follows may occur (M is considered to be monovalent):

Micelle Micelle

Some of the added K pushed its way into the colloidal complex and forces out
equivalent quantities of calcium, hydrogen, and other elements which appear in the soil
solution.
The adsorption of the added potassium is considered to be advantageous because a
nutrient so held remains largely in an available condition but is less subject to leaching.

3.1.5. Cation Exchange Capacity (CEC)


Cation exchange capacity (CEC) is also termed as cation adsorption capacity. It is
expressed in terms of equivalents or more specifically, as milliequivalents per 100 grams of
soil.
The term equivalent is defined as 1 gram atomic weight of hydrogen or the amount
of any other ion that will combine with or displace this amount of hydrogen. For monovalent
ions such as Na, K, NH4, and Cl, the equivalent weight and atomic weight are the same since
they can replace or react with one H ion. Divalent cations such as Ca and Mg can take the

- 31 -
place of two H ions. Therefore, their atomic weight must be divided by 2 to obtain the
equivalent weight.
The milliequivalent weight of a substance is one thousandth of its atomic weight.
Since the equivalent weight of hydrogen is about 1 gram, the term milliequivalent (meq) may
be defined as 1 milligram of hydrogen or the amount of any other ion that will combine with
or displace it.
Thus, if a clay has a cation exchange capacity of 1 milliequivalent (i.e.1 meq per 100
gram), it is capable of exchanging 1 mg of H+ to 100,000 mg of clay, or 10 parts per million
(10 ppm) of H+. Therefore, an acre-furrow slice of such a clay weighing 2 million pounds
could adsorb 20 pounds of exchangeable hydrogen or its equivalent.
The milliequavalent method of expression is so convenient and so commonly used
that it is important to be familiar with it.
It is well to note this term "equivalent". It indicates that other ions also may be
expressed in terms of milliequivalent. For example, consider calcium. This element has an
atomic weight of 40 compared to 1 for hydrogen. Each Ca2+ ion has two charges and is thus
equivalent to 2 H+ ions. Therefore, the amount of Ca required to displace 1 mg of hydrogen
is 40/2 or 20 mg. This is the weight of 1 meq of calcium.
If 100 grams of certain clay is capable of exchanging a total of 250 mg of calcium, the
cation exchange capacity is 250/20, or 12.5 meq per 100 grams.
The milliequivalent method of expression can be converted easily to practical field
terms. For instance, 1 meq of hydrogen can be replaced on the colloids by 1 meq of CaCO3
contained in ordinary limestone. The molecular weight of CaCO3 (100) contains 2 equivalent
weights. Since the amount of CaCO3 needed is only 1 meq weight, 100/2 = 50 mg will be
needed to replace 1 mg of hydrogen.
Since 1 meq of H+ per 100 grams is the same as 20 pounds of H+ per acre-furrow slice
(as explained earlier) , 1 meq of CaCO3 per 100 grams is the same as 20 50 = 1000 pounds
of CaCO3 per acre-furrow slice. Thus, 1/2 ton of CaCO3 per acre-furrow slice has the
potential of 1 meq of H+ per 100 grams of soil.

3.1.6. The Effect of Soil pH on CEC


Cation exchange in most soil increases with pH. At very low pH value, only the
"permanent" charges of the clays and a small protion of the charges of organic colloids hold
ions that can be exchanged. Majority of organic colloids exchange sites and those of
inorganic colloids, hydrogen and other ions are held so tightly as to resist replacement.

- 32 -
As the pH is raised, the hydrogen held by the inorganic and organic colloids becomes
ionized and is replaceable. The result is an increase in the cation exchange capacity.
In most cases the cation exchange capacity is determined at a pH of 7.0 or above. This
means that it includes most of those charges dependent on pH as well as the more or less
permanent ones.

3.1.7. Determinations of Cation Exchange Capacity of Soil


The cation exchange capacity determination involves measuring the total quantity of
negative charges per unit weight of the material. There are a number of methods to determine
the CEC of soil samples. They can be roughly divided into the following groups.
(a) A common method is to saturate the soil with a given cation; the total amount
adsorbed will be exchange again and determined.
(b) Another common method is also to saturate the soil with an extraneous cation. The
exchanged cations originally adsorbed on the soil sample will be measured. The sum
of the exchanged cations gives the CEC of the soil.
(c) Method using radio-active-isotopes can also be carried out. This method has the
advantage of giving very accurate results. However, it may need special skill and
equipment to be able to deal with the radio-active materials.

CEC determination by Bascomb's method


Only one of the most common and convenient methods will be discussed here. It is
named Bascomb's method. It involves using barium as index ion. This method can be used
for all kinds of soil including calcareous or organic samples.

Theory of CEC determination by Bascomb's method


The soil sample is saturated with Ba at a pH 8.1.

Micelle Micelle

The adsorbed Ba++ is determined indirectly. A given amount of MgSO4 is added.


The Mg++ cations exchange the Ba++ cations which will precipitate as BaSO4.
- 33 -
Micelle Micelle

The Mg left in the solution is back titrated using standard EDTA (Ethylene diamine
tetra acetic acid) solution.
The difference between the added Mg and the left over Mg in the solution will give
the amount of Mg which replaces the Ba cations adsorbed on soil colloids.

Procedure for CEC determination by Bascomb's method


Weigh exactly 5 g of 2 mm air-dry soil sample into a pre-weighed centrifuge bottle.
Note the weight of the bottle + soil (W1).
Treat the soil with approximately 100 ml of buffered barium chloride reagent. The
bottle is gently shaken for 1 hour and is later centrifuged at 1500 rpm for 15 minutes. The
supernatant liquid is discarded. (For non-calcareous soils, this step may be omitted).
Treat the soil with another 100 ml of buffered barium chloride reagent overnight,
centrifuge at 1500 rpm and discard the supernatant liquid.
Add approximately 200 ml of distilled water and shake for a few minutes to break the
soil cake. Centrifuge and discard the supernatant liquid. Weigh the bottle with contents (i.e.
soil and water molecules sticking to the soil particles) (W2) g.
Pipette into the bottle 100 ml of magnesium sulphate solution and shake the bottle for
2 hours. Centrifuge at 1500 rpm for 15 minutes. Transfer the clear liquid immediately to a
stoppered flask.
Take 5 ml portion of the above clear supernatant liquid. Add 10 ml of buffer solution
and 2 drops of Eriochrome Black T indicator. Titrate with a standard EDTA solution until the
color changes from clear blue to reddish purple. The titre value is noted (A1 ml).
Aliquot of 5 ml of the original magnesium sulphate solution are also titrated under the
same conditions as above. The titre value is also noted (B ml).

Calculation for CEC determination by Bascomb's method


W1 g = wt. of centrifuge bottle + 5 g of soil
W2 g = wt. of centrifuge bottle + 5 g of soil + water
(W2 W1) g = wt. of water sticking to/ 5 g of the soil particles.

- 34 -
A1 ml of std. EDTA = 5 ml of supernatant liquid.
B ml of std. EDTA = 5 ml of original MgSO4 solution.
The volume A1 ml of the EDTA solution used to titrate the soil extract must be
corrected for the effect of the volume of liquid retained by the centrifuged soil after washing
with water.
Corrected volume A2 ml = A1 (100 + W2 W1)/ 100 ml
The CEC of soil is calculated using the following equation.
CEC = 8 (B A2) meq/ 100 g soil.
3.2. Anion Exchange
Under certain conditions hydrous oxides of iron and aluminum, allophane and even
kaolinite show evidence of having positive charges on their crystal surfaces. These charges
are thought to have two sources:
(a) the protonation of adding of hydrogen ions to hydroxyl groups on the edge of
these minerals; and
(b) the exchange of the hydroxyl group for other anions, such as phosphate. These
mechanisms can be illustrated by the appropriate equations. First the protonation
reaction can be shown as follows.

At high pH values (left) the hydrogen ion tends to dissociate from the oxygen, leaving
a negative charge on the surface. As the pH is lowered, an additional hydrogen ion associates
with the hydroxyl, leaving a net positive charge. This charge will attract anions such as
H2PO4, SO42, NO3, Cl. Those ions can exchange with each other, giving rise to the term
anion exchange.
The second mechanism that shows a positive charge on certain colloids is that in
which phosphates and similar ions exchange for hydroxyl:

This reaction also occurs primarily at low pH values, the negative hydroxyl ion being
replaced from positively charged aluminum ion in the crystal.
By one of the two of these mechanisms of anion exchange, SO42, Cl, and NO3 ions
as well as phosphates become adsorbed by allophane, the hydrous oxides, and to a degree by
- 35 -
the kaolinite group. The reactions are most significant for phosphates since these ions tend to
be quite tightly adsorbed. These reactions emphasize once again the importance of the type of
clay in determining the basic reactions in soils.

- 36 -
CHAPTER 4
SOIL REACTIONS

Soils may be acid, neutral or alkaline in reaction. Each set of chemical conditions in
soil produces corresponding degree of soil reaction. Each degree of soil reaction affects plant
growth in certain way due to either depressed solubility of some elements or to an increased
solubility of others. The chemical condition which produces different degrees of soil reaction
may be favorable to the growth of some crops and affects the growth of other crops to a great
extant. Therefore, soil reaction is considered important in crop-production and soil-
management practices.
Soil reaction may be considered as symptom of the particular chemical conditions
which caused it. Therefore, it may be used to indicate the possible effect of these conditions
on plant growth.
It is also useful in diagnosing the fertility of soils.

4.1. Soil Acidity


In humid regions, there is a natural tendency for soils to become increasing acid. This
process is accelerated when soils are under cultivation. Most of the leguminous crops do not
thrive on soils of strong acidity and the yields of many other crops are decreased by strong
acidity. Therefore addition of lime on acid is becoming a common practice in the agriculture
of humid regions.
At least 95 percent of the acidity of the mineral soils is due to exchangeable hydrogen
and/or aluminum associated with the colloidal material. The positively charged cations are
adsorbed on the surface of colloids. These cations dissociate to some extent and their places
on the colloids are taken up by H ions from the water. This can be shown as follow:

Micelle Micelle

Water molecules H2O dissociate into positively charged H+ ions and negatively
charged OH ions. H+ ions are held close to the micelle and OH ions go into soil solution.
When the number of H+ ions and the number of OH ions in the solution around the
micelle are equal a neutral reaction exists. When the number of H+ ions exceeds the number
of OH ions, the reaction of the soil is acid.

- 37 -
Other factors also contribute to soil acidity to some extent.
Traces of inorganic acids such as HNO3, HNO2, H2SO4, H3PO4, etc. may be present in
soil as a result of some chemical and biological reactions.
Soils containing mineral like FeS2 may produce H2SO4 to a considerable extent.
CO2 is present in most soils and with H2O it produces H2CO3.
Other organic acids such as acetic, citric, and oxalic are liberated during the decay of
fresh plant tissue.
Although these organic and inorganic acids contribute to the acidity of the soil to
some extent the reaction is due primarily to the relative proportion of H+ ions and other
cations on the colloidal complex.

4.1.1. Expression of Soil Acidity


It is customary to base the expression of acidity on the concentration of the
dissociated H+ ion which is known as active acidity. This tendency has given rise to use of
the ''pH'' values. The pH scale is derived as follows.
Water is a neutral substance with equal concentration of H+ and OH ions. The
ionization constant of water (Kw) is 1014.
 
   

As concentration of H2O (Conc. H2O) may be regarded as constant and almost equal
to the molar concentration per liter of water it may be ignored and the equation may be
written as follow:
  

As concentrations of H+ and OH are equal each of them must be 107g/liter of water.


Considering only the concentration of the H+ ion, because it is the cause of the acid
reaction we have:
Conc. H+ = 107 g/liter = 0.0000001 gram H+ ions per liter
The reciprocal of this is 1/0.0000001 = 10,000,000.
The log of 10,000,000 is 7. Pure water is said to have a pH of 7.
As water is neutral any solution with a pH of 7 is said to be neutral.
Therefore the definition of the pH value is the logarithm of the reciprocal of the H+
ion concentration at 25 C.

pH = log

- 38 -
If the concentration of H+ ions in a solution were 0.000001 gram per liter (106), the
pH would be 6. As the concentration of H+ ions in the solution is 10 times that in water, it is
clear that a solution with a pH 6 is 10 times as acid as a solution with pH 7. It should never be
forgotten that the pH values are logarithms.
As the product of the concentration of the H+ and OH ions is always equal to 1014,
the concentration of the OH can be calculated if the pH value of the solution is known.
Thus pH 7 indicates a concentration of H+ ion, [H+] = 107g/liter and a concentration
of OH ion, [OH] = 107g/liter
Kw = [H+] [OH] = 107g/liter 107 g/liter = 1014
A pH of 6 indicates a concentration of H+ ion, [H+] = 106g/liter and a concentration of OH
ion, [OH] = 108 g/liter
Kw = [H+] [OH] = 106g/liter 108 g/liter = 1014
On the other hand a pH of 8 would indicate a concentration of H+ ion [H+] =108g/lit and a
concentration of OH ion, [OH]= 106g/liter
Kw = [H+] [OH] = 108g/liter 106 g/liter = 1014
Therefore, OH ion concentration, [OH]= 106g/liter at pH 8 is 10 times higher than
the OH ion concentration, [OH]= 107g/liter at pH 7 or neutral. This solution therefore is
alkaline. Any pH value above 7 indicates alkaline condition.

4.1.2. Types of Soil Acidity


For a soil to give an acid reaction there must be an excess of H+ ion over OH ions in
the solution surrounding the colloidal complex. The concentration of these H+ ions can be
readily measured and constitutes the active acidity of the soil. The concentration of these
dissociated H+ ions is in equilibrium with these adsorbed on the complex.
Supposing that addition of lime neutralizes a portion of active H+ ions or that leaching
removes a portion of active H+ ions from soil, more H+ ions will immediately dissociate from
the colloidal complex to restore the equilibrium. The H+ ions on the micelle therefore
constitute a reserve supply which is known as the reserve or potential soil acidity.

- 39 -
4.1.3. Range of Reaction in Soils
Range of soil reaction or soil pH can be roughly divided into the following groups:
Extremely acidic less than pH 4.5
Very strongly acidic from pH 4.5 to 5.1
Strongly acidic from pH 5.1 to 5.5
Moderately acidic from pH 5.6 to 6.0
Slightly acidic from pH 6.1 to 6.5
Neutral from 6.6 to 7.4
Mildly alkaline from 7.5 to 8.0
Strongly alkaline from pH 8.1 to 9.0
Very strongly alkaline higher than pH 9.0

The pH 4.0 is usually the lower limit for soils of humid region. It is unusual to find a
soil with a pH of more than 7.5 or 8.0 in humid regions.
However, in arid regions, where soluble salts of sodium (Na2CO3) may accumulate, a
pH of 9.5 to 11.0 is sometimes attained.
The usual pH of agricultural soils in the humid region ranges approximately from 5.0
to 6.8.
All soils do not have the same pH for a given percentage base saturation, because the
exchangeable hydrogen does not ionize with equal case from the different clay minerals and
organic colloids.
The bonding strengths for hydrogen-
(1) of a site originating from isomorphous substitution
(2) of a site originating from the broken bonds at the edges of particles
(3) of a site originating from an exposed hydroxyl
(4) of a site originating from a carboxyl group are different from one another.
This is the basic explanation of why soils containing high percentage of kalonite clays
have a higher pH than those with 2:1 clays at a given percentage base saturation.
Calcareous soils contain free carbonates of calcium and magnesium. These salts give
alkaline reaction when hydrolyzed and produce a maximum pH of 8.3.
A pH of higher than 8.3 indicates the presence of exchangeable sodium that
hydrolyses to form NaOH, a very strong base. Such soils are termed alkali. Soils containing
high percentage of soluble salts, tend to have a neutral reaction and such soils are termed
saline.

- 40 -
4.1.4. Source of Hydrogen Ions
Adsorbed aluminum ions
Under acid conditions, much aluminum becomes soluble. They are present in the form
of aluminum or aluminum hydroxy cations. They are adsorbed by the permanent charges of
clay minerals.
The adsorbed aluminum is in equilibrium with aluminum ions in the soil solution.

Micelle

The aluminum ions in the soil solution are then hydrolyzed in a manner such as the
following;
     
The H+ ions thus released give a very low pH value in the soil solution.

Adsorbed hydrogen ions


Adsorbed hydrogen ions concentration is in equilibrium with those in the soil
solution. A simple equation to show the release of adsorbed hydrogen to the soil solution can
be written as follows:

Micelle

Adsorbed aluminum hydroxyl ions


In soil, the aluminum cannot exist as Al3+ ions. They are converted to aluminum
hydroxy ions by reactions as follows:
   

   


Aluminum hydroxy ions
The actual aluminum hydroxy ions found in soils are more complex than those shown
above. Formulas such as [Al6(OH)12]6+ and [Al10(OH)22]8+ with the possibility of ring
configurations have been postulated.

- 41 -
Some of positively charged aluminum hydroxy ions adsorbed and act as exchangeable
cations. They are in equilibrium with those in soil solution.
In the soil solution they are able to produce hydrogen ions by the following hydrolysis
reactions:

4.1.5. Conditions in Acid Soil Which Affect Plant Growth


Soil activity is an indication of existing chemical conditions in a soil. A deficiency of
certain nutrients, such as boron, may occur in some soils with a high pH, whereas in other
neutral or alkaline soils of the arid regions an excess of boron may be present. The conditions
which affect plant growth in acid soils will be discussed.

Effect of H+ ions on plants


Plant growth is usually affected in soils of very low pH which contain high
concentration of H+ ion. However, experimental results have shown that in a nutrient solution
with a concentration of H+ ions equivalent to that of strongly acid soil, plant grow much
better than in the acid soil itself. In other words, when separated from the accompanying
conditions which exist in an acid soil, the H+ ion concentration is no longer harmful to many
plants. Therefore it becomes necessary to look for some cause in addition to H+ ion
concentration for the detrimental effect of acid soils on plant growth.

Soluble iron, aluminum, and manganese


The quantity of soluble iron, aluminum, and manganese in many soils increases as
soil acidity increases. High ferrous iron concentration is known to be toxic to many plants.
Soluble aluminum also has a high toxicity for many commonly grown crops.
It is generally believed that toxic concentration of aluminum and manganese are a
major factor for the poor growth of plants in strongly acid soils.

Phosphorus availability in acid soils


The chemical elements such as iron and aluminum combine with phosphate ions in
acid soils to form compounds of low solubility. In soil of pH lower than 5.0, complex

- 42 -
phosphate of iron and aluminum may be formed. These complex iron and aluminum
phosphates have a very low solubility and do not supply sufficient phosphorus for plants.
Under strongly acid conditions when soluble phosphate fertilizers are applied to the
soil, they are converted into the insoluble forms. Therefore, on these soils, the efficiency of
applied phosphates is greatly reduced.
Calcium supply in acid soils
Soil reaction becomes acidic only when the supply of basic ions, such as Ca, on the
soil colloids is depleted. It has been generally accepted that a nutrient deficiency of calcium is
a factor in the poor growth of plants on many acid soils.

Activity of microorganisms
The microorganisms which convert ammonia into nitrite and nitrite into nitrate are
sensitive to acidity.
One group of bacteria (Azotobactor) which utilizes atmospheric nitrogen without
association with a legume does not function properly below a pH of 6.0. Symbiotic bacteria
which grow in conjunction with several legumes, such as alfalfa, clover, soybeans, etc. are
badly affected by acidity.

Variation in tolerance of plants for acidity


The existence of plants in a given environment is mainly due to their tolerance. The
intensity of acidity which plants will tolerate is also influenced by the supply of available
nutrients and of moisture. When other conditions are favorable for the growth of a plant, it
will thrive in a much more acid soil than it would if the nutrient or moisture supply were
limited.
Certain plants e.g. blueberry and rhododendron are very highly tolerant for acidity.
Other plant such as velvet bean, corn, sorghum, millet, tobacco, strawberry, peach, etc. are
also highly tolerant for acidity.

4.1.6. Methods of Determining Soil Reaction


Measurement of pH is becoming increasingly important in conducting soil tests. It is
often used as a tool in liming and similar problems. The current methods used for pH
determinations are easy and rapid.

- 43 -
Dye or indicator methods
The dye methods are very simple and easy, but the results may not be very accurate.
The method consists of using certain indicators. Many dyes change colors with an increase or
decrease of pH. It therefore becomes possible (within the range of the indicator) to estimate
the approximate hydrogen ion concentration of a solution. By using a number of dyes either
separately or mixed, a range of pH from 3 to 8 is easily covered.
To carry out such a pH determination on soil, the sample is first saturated with the
dye. After standing for a few minutes, a drop of the liquid is taken and observed in thin layer.
By the use of a suitable color chart, the approximate pH value may be ascertained. The
indicator or dye method is accurate within about 0.2 pH unit, when the determination is
carefully and correctly done.

Electrometric method
The most accurate method of determining soil pH is by a pH meter. In this
electrometric method, the hydrogen concentration of the soil solution is balanced against a
standard hydrogen electrode. The potential may be measured by using a potentiometer. The
potential established is proportional to the concentration of H+ ions.
Although the instrument gives very accurate results, its use requires a skilled operator
having the knowledge of complicated mechanism how a pH meter works.

4.2. Soil Alkalinity


In arid regions, the annual rainfall is very low (usually lower than 20 inches per year).
The extensive leaching is therefore prevented and the soil is usually saturated with high
percentage of bases. In arid region a normally developed soil profile contains a calcium
carbonate accumulation (usually in the C horizon). The lower the rainfall, the nearer the
surface this calcium carbonate layer will be.
As a result, these soils may have alkaline or neutral reactions.

4.2.1. Halomorphic Soils


When the evapotranspiration exceeds the precipitation, and when the leaching process
is reached, soluble salts accumulates in the surface horizon of soil profile in arid regions.
These soils are termed halomorphic soils and can be divided into three different groups: (1)
saline soils, (2) saline-alkali soils (sometime known as saline-sodic soils), and (3) alkaki soils
(sometime known as sodic soils).

- 44 -
Saline soils
These soils contain a sufficient concentration of neutral soluble salts to interfere with
the growth of most plants. The electrical conductivity of a saturated extract (ECe) is greater
than 4 mmhos/ cm.
Less than 15 percent of the cation exchange capacity of those soils is occupied by
sodium ions and the pH usually is below 8.5. This is because the soluble salts present are
mostly neutral and because only a small amount of exchangeable sodium is present.
Such soils are sometimes called white alkali soils because the surface crust is usually
light in color.
The excess soluble salts are mostly chlorides and sulphates of sodium, calcium, and
magnesium. These salts can be easily leached out of these soils, with no appreciable rise in
soil pH. However, care must be taken to be certain that the leaching water is low in sodium.

Saline-alkali soils
These soils are sometimes known as saline-sodic soils. These soils contain
appreciable quantities of neutral soluble salts and enough adsorbed sodium ions to seriously
affect most plants.
Although more than 15 percent of the total exchange capacity of these soils is
occupied by sodium, their pH is likely to be below 8.5. This is because of the repressive
influence of the neutral soluble salts, as in the saline soils previously described.
The electrical conductivity of a saturated extract is more than 4 mmhos/ cm.
But unlike the saline soils, leaching will markedly raise the pH of saline-alkaline
soils, unless calcium or magnesium salt concentrations are high in the soil or in irrigation
water. This is because the exchangeable sodium, once the neutral soluble salts are removed,
readily hydrolyses and thereby sharply increases the hydroxyl ion concentration of the soil
solution.
It is important to note that these sodium ions destroy the soil structure and at the same
time, sodium toxicity to plants is increased.

Alkaline soils
These soils do not contain a large amount of neutral soluble salts. The bad effects are
mainly due to the toxicity of the sodium as well as of the hydroxyl ions. The high pH is
largely due to the hydrolysis of sodium carbonate, which occurs as follows:

- 45 -
The resulting hydroxyl ions give pH values of 10 and above.
Also, the sodium complex undergoes hydrolysis as follows:

Micelle Micelle
The exchangeable sodium occupies more than 15 percent of the total exchangeable
capacity of these soils. These exchangeable sodium ions are free to hydrolyze because the
concentration of neutral soluble salts is low.
The electrical conductivity of a saturated extract is less than 4 mmhos/cm.
Consequently, the pH is above 8.5, often rising as high as 10.0.
Due to the deflocculating influence of the sodium, such soils usually are in poor
physical condition.
Because of the extreme alkalinity resulting from the Na2CO3 present, the surface of
alkali soils is usually discolored by the dispersed humus carried upward by the capillary
water. Therefore, these alkaline or sodic soils are sometimes known as black alkali soils.

4.2.2. Growth of Plants on Halomorphic Soils


Saline and saline-alkali soils influence plants largely because of their high soluble
salts concentration. It is common knowledge that when a water solution containing a
relatively large amount of dissolved salts is brought into contact with a plant cell, it will
cause a shrinkage of the protoplasmic lining. This action is known as plasmolysis. The
phenomenon is due to the osmotic movement of the water, which passes from the cell toward
the more concentrated soil solution. The cell then collapses.
Alkali soils, dominated by active sodium, have undesirable effects on plants in three
ways:
(1) caustic influence of the high alkalinity induced by the sodium carbonate and
bicarbonate,
(2) toxicity of the bicarbonate and other anions, and
(3) The adverse effects of the active sodium ions on plant metabolism and nutrition.

4.2.3. Management of Saline and Alkali Soils


There are 3 general ways in which saline and alkali soils may be handled in order to
avoid injuring effects to plants. They are (1) eradication, (2) conversion and (3) control
methods.

- 46 -
Eradication methods
The methods used to free the soil of excess salts are (a) under drainage and (b)
leaching or flushing.
Heavy and repeated application of water is made in irrigated regions. The salts which
become soluble are then leached and drained off. The irrigation water should be relatively
free of sodium salts.
This leaching method is very good for treating saline soils, which contain mostly
neutral salts.
Leaching of saline-alkali and alkali soils may intensify the alkalinity, because of the
removed of neutral soluble salts. This allows an increase in the percent sodium saturation
thereby increasing the concentration of hxdroxyl ions in the soil solution.
This may be avoided by converting the toxic sodium carbonate and bicarbonate to
sodium sulphate by treating the soil with heavy application of gypsum or sulphur. Then
leaching of soil can be done more satisfactorily.

Conversion methods
Gypsum is often used for treating alkali soils. The gypsum reacts with Na2CO3 salt
and also with the adsorbed Na on the soil colloids as follows:

Micelle Micelle

Several tons of gypsum per acre are usually necessary. The soil must be kept moist to
improve the reaction. This treatment is followed by leaching with sodium free irrigation
water.
Sulphur can also be used to change harmful sodium carbonate to less harmful sulphate
form. When applied to soil, sulphur is oxidized to form sulphuric acid. The reactions of
sulphuric acid with sodium carbonate or adsorbed sodium take place as follows:

Micelle Micelle

- 47 -
Sulphuric acid not only transform sodium carbonate in to less harmful sodium
sulphate (because it is a mild neutral salt), but also reduce the intense alkalinity to a
considerable extent.
Control methods
The retardation of evaporation from the surface of the soil not only saves moisture,
but also prevents the translocation upward of soluble and harmful salts into the root zone.
This method can be very expensive when dealing with large acreages.
The method of irrigation is also important. Frequent and light irrigations is more
advantageous than heavy application of water at less frequent intervals.
The timing of irrigation is extremely important on salty soils. Since young seedlings
are especially sensitive to salts, irrigation often proceeds or follows planting to move the salts
downward.
The use of salt-resistant crops is another way of managing saline and alkali soils.
Sugar beets, cotton, sorghum, barley, rye, sweet clover, and alfalfa are found to be salt
resistant and should be chosen to grow on salty soils.
The use of farm manures also alleviates the alkalinity of soil temporarily. This
situation will help the less resistant crops such as alfalfa before the root system is well
established. Once the plant is growing vigorously, it may maintain itself in spite of salt
concentration that may develop later.

- 48 -
CHAPTER 5
SOIL ORGANIC MATTER

Although organic matter is present in soil in comparatively small quantity, it


influences physical, chemical and biological properties of soils to a great extent. It accounts
for at least half the cation exchange capacity of soils. It is also responsible perhaps more than
any other single factor for the stability of soil aggregates. Furthermore, it supplies energy and
body-building constituents for the microorganisms.

5.1. Sources of Soil Organic Matter


The original source of the soil organic matter is plant tissue. Under natural conditions,
the tops and roots of trees, shrubs, grasses, and other native plants annually supply large
quantities of organic residues. Plant residues are decomposed and digested by soil organisms
of many kinds. The decomposed plant materials become part of the underlying soil horizons
by infiltration or by physical incorporation. Thus, higher plant tissues are the primary sources
not only of food for the various soil organisms but also of organic matter.
Animals are considered secondary source of organic matter. As they attack or break
down the plant tissues, they contribute waste products and leave their own bodies when they
die. Animals such as earth worms, centipedes, and ants, also play an important role in the
translocation of plant residues.

5.2. Composition of Plant residues


The moisture content of plant residues varies from 60 to 90%, and averages about
75%. On the weight basis, the dry matter is mostly carbon and oxygen with less than 10%
each of hydrogen and inorganic elements (ash).
However, on an elemental basis (number of atoms of the elements), hydrogen
predominates. There are 8 hydrogen atoms for every 3.7 carbon atoms and 2.5 oxygen atoms.
These three elements dominate the bulk of organic tissue in the soil.
Even though more than 90% of the dry matter is carbon, hydrogen, and oxygen, the
other elements such as nitrogen, sulphur, phosphorus, potassium, calcium and magnesium
play a vital role in plant nutrition and in meeting microorganism body requirements.

- 49 -
5.3. General Composition of Compounds
The composition of the green plant materials usually added to the soils can be
summarized as in the following flow charts.

The types of the compounds which are usually found in the dry matter of green plant
materials can be briefly summarized as follows:

Carbohydrates range in complexity from simple sugars to celluloses. Fats and oils are
glycerides of fatty acids such as butyric, stearic and oleic. Lignins are complex compounds
which are very resistant to decomposition. Of the various groups, the crude proteins are the
most complicated compounds which carry not only C, H and O but also N, S, Fe and P. They
are compounds of high molecular weight.

5.4. Decomposition of Organic Compounds


Organic compounds vary greatly in their rate of decomposition. They may be listed in
terms of their ease of decomposition as follows:

- 50 -
1. Sugars, starches, and simple proteins Rapidly Decomposed
2. Crude proteins
3. Hemicelluloses
4. Celluloses
5. Lignin, fats, waxes, etc. Very slowly decomposed

The rate of the decomposition decreases from the top to the bottom of the list. Thus,
sugars and water soluble proteins are examples of readily available energy sources for soil
organisms. Lignins are very resistant source of food, although they eventually supply much
total energy.
When organic tissue is added to soil, three general reactions takes place.
1. The material undergoes enzymatic oxidation with carbon dioxide, water and heat
as the major products.

2. The essential elements such as N, P and S are released and/or immobilized by a


series of specific reactions relatively unique for each element.
3. Compounds resistant to microbial action are formed either from compounds in the
original plant tissues or by microbial synthesis.

5.5. Energy of Soil Organic Matter


Microorganisms in soil obtain various substances from soil organic matter for their
tissue synthesis and also for the required energy.
Organic matter contains considerable potential energy, a large proportion of which is
readily transferable to other latent forms or is liberated as heat.
Plant tissue has a great heat value approximating 4 or 5 kilocalories per gram of air
dry substance. For example, the application of 10 tons of farm manure containing 5,000
pounds of dry matter would mean an addition of 11 million kilocalories of latent energy. A
soil containing 4% of organic matter carries from 150 to 180 million kilocalories of potential
energy per acre-furrow slice. This is equivalent in heat value perhaps 20 to 25 tons of
anthracite coal.

- 51 -
Of this large amount of energy carried by the soil organic matter, only a part is used
by soil organisms. The remainder is left in the residues or is dissipated as heat. This heat loss
represents a large and continual removal of the energy from the soil.

5.6. Rate of Energy Loss from Soil


It was calculate that 1 million kilocalories per acre were lost annually from the low-
producing soil, and about 15 million kilocalories were dissipated from the more productive
soil, which was receiving liberal supply of farm manure.

5.7. Simple Decomposition Products


When soil organic matter is decomposed, simple products are produced. Carbon
dioxide and water appear immediately. Other products, like nitrogen, accumulate only after
the peak of the vigorous decomposition is over and the general purpose decay organisms have
diminished in numbers.
The simple products resulting from the activity of the soil microorganisms may be
listed as follows;
Carbon CO2, CO32, HCO3, CH4, elemental carbon
Nitrogen NH4+, NO2, NO3, gaseous nitrogen
Sulphur S, H2S, SO32, SO42, CS2
Phosphorus H2PO4, HPO42
Others H2O, O2, H2, H+, OH, K+, Ca2+, Mg2+, etc.

5.8. The Carbon Cycle


Carbon is a common constituent of all organic matter. Consequently, its movements
during the microbial digestion of plant tissue are extremely significant. Much of the energy
acquired by the fauna and flora within the soil comes from the oxidation of carbon. As a
result, its oxide is evolved continuously and in large amounts. The various changes of this
element within and without the soil are collectively known as the carbon cycle.
As the compounds in plant residues are digested CO2 is given off. Under optimum
conditions, as much as 100 pounds of CO2 per acre per day may be evolved, 20 to 30 pounds
being more common. The CO2 of the soil ultimately escapes in a large degree to the
atmosphere, where it may again be used by plants, thus completing the cycle.
A smaller amount of CO2 reacts in the soil, producing carbonic acid (H2CO3) and the
carbonates and bicarbonates of calcium, potassium, magnesium and other bases. The
bicarbonates are readily soluble and may be lost in drainage or used by higher plants.
- 52 -
5.9. Simple Products Carrying Nitrogen
Ammonium salts are the first in organic nitrogen compounds produced by microbial
digestion. Protein split up into amino acids and similar nitrogenous materials which readily
yield ammonium compounds by enzymatic hydrolysis. These complex transformations are
brought about by a large number of heterotrophic organisms such as bacteria, fungi, and
actinomycetes. The ammonium ion is readily available to microorganisms and most of higher
plants.
If conditions are favorable, ammonium ions are subject to ready bacterial oxidation, to
form nitrite and nitrate by two special purpose organisms as follows:

The autotrophic bacteria obtain energy by these reactions, which produce nitrates that
are assimilated by plants. The hydrogen ions that are produced by this reaction show that
nitrification tents to result in an increase in soil acidity. This effect is of even greater
important when dealing with commercial fertilizers containing ammonium salts such as
(NH4)2SO4. Extra increments of lime are often added to counteract this acidifying effect.
Under certain conditions, the reduction of nitrates and nitrites takes place in soils, and free
nitrogen or oxides of nitrogen may be evolved. This reduction reaction is serious since these
nitrogen gases ordinarily are relatively inert and are not recombined into compounds useful to
higher plants. The cultivated soils lose considerable amount of nitrogen in this way.

5.10. Simple Sulphur Containing Products


Many organic compounds contain sulphur, which appears in simple forms as it
undergoes decomposition. General purpose heterotrophic organisms apparently simplify the
complex organic compounds. The simplified sulphur by-products are then subjected to
oxidation by special autotrophic bacteria.

The organisms involved obtain energy by the transfer and the sulphur is left as
sulphate.
- 53 -
5.11. Mineralization of Organic Phosphorus
A large portion of the soil phosphorus is in organic form. Organic phosphorus
compounds are mineralized by soil microorganisms, that is, they are changed to inorganic
combinations. The particular forms depend to a considerable degree upon soil pH. As the pH
goes up from 5.5 to 7.5, the available phosphorus changes from H2PO4 to HPO42. Both of
these forms are available to higher plants. Since the small amount of phosphorus held in
complex mineral combinations in soils usually is very slowly available, the organic sources
mentioned above become especially important.
It is important to note that the maintenance of soil organic matter at normal or even at
high levels will not solve the soil phosphorus problems. Most field soils need liberal
application of phosphate fertilizers. Since soil microorganisms utilize phosphorus quickly
becomes part of the soil organic matter. Thus, this phosphorus is held in an organic condition
and is later mineralized by microbial activity.

5.12. Humus
As organic tissue is incorporated into a moist warm soil, it is immediately attacked by
a host of different soil organisms, and first intermediate substances and finally the simple and
soluble products are formed.

5.13. Humus Formation


As the decomposition of organic tissue occurs, two major kinds of organic
compounds tend to remain in the soil.
(a) resistant compounds of higher plant origin, such as oil, fats, waxes, and especially
lignin, and
(b) new compounds such as polysaccharides and polyuronides, which are synthesized
by microorganisms and held as part of their tissue.
The resultant product, newly formed humus, is quite resistant to further microbial
attack. Its nitrogen and other essential nutrients are thereby protected from ready solubility
and dissipation.

5.14. Protein-clay Combination


Another means of stabilizing nitrogen in soil is through the reaction between certain
clays and proteinaceous substances and other nitrogen compounds. Clays with expanding
lattice, such as montmorillonite seem to possess such property. The adsorbed protein seems
to be protected against rapid decomposition by soil microorganisms.
- 54 -
5.15. Definition of Humus
Humus is a complex and rather resistant mixture of brown or dark brown amorphous
and colloidal substances modified from the original tissue or synthesized by the various soil
microorganisms.
Although it is exceedingly variable and heterogeneous, it possesses properties that
distinguish it sharply from the original parent tissues and from the simple products that
develop during its synthesis.

5.16. Direct Influence of Organic Compounds on Higher Plants


Certain organic nitrogen compounds such as alanine and glycine (amino acids) can be
absorbed directly by higher plants, often rather readily.
Growth promoting substances, vitamin like compounds, hormones, etc. which are
present in soil organic matter, stimulate both higher plants and microorganisms.
On the other hand, some soil organic compounds may be harmful. As an example,
dihydroxystearic acid, which is toxic to higher plants, was isolated from some soils.
However, it should be noted that such compounds are merely products of unfavorable soil
conditions. When such conditions are corrected, the toxic matter disappears. Apparently,
good drainage and tillage, lime and fertilizer applications reduce the probability of organic
toxicity.

5.17. Influence of Soil Organic Matter on Soil Properties


A brief review of the most obvious influences of soil organic matter on soil properties
is as follows:
1. Effect on soil color brown to black
2. Influence on physical properties
a. Granulation encouraged
b. Plasticity, cohesion, etc. reduced
c. Water-holding capacity increased
3. High cation adsorption capacity
a. Two to thirty times as great as mineral colloids
b. Accounts for 30 to 90 percent of the adsorbing power of mineral soils.
4. Supply and availability of nutrients
a. Easily replaceable cations present
b. Nitrogen, phosphorus and sulphur held in organic forms
c. Extraction of elements from minerals by acid humus.
- 55 -
5.18. Carbon-Nitrogen Ratio
Carbon makes up a large and rather definite proportion of soil organic matter, and the
carbon to nitrogen ration of soils is found to be fairly constant.
The ratio of carbon to nitrogen in the organic matter of the furrow slice of arable soils
commonly ranges from 8:1 to 15:1, the median being between 10:1 and 12:1. In a given
climatic region, little variation is found in this ratio, at least in similarly managed soils. The
variation in C:N ratios seem to be correlated in a general way with climatic conditions
especially temperature and the amount and distribution of rainfall. For instance, the C:N ratio
tend to be lower in soils of arid regions than in those of humid regions when annual
temperatures are about the same.
C:N ratio is also lower in warmer regions than in cooler ones if the rainfalls are of
about the same magnitude.
C:N ratio for the sub-soils is also narrower in general than for the corresponding
surface layers.
C:N ratio in plant material is variable, ranging from 20:1 to 30:1 in legumes and farm
manure to as high as 100:1 in certain straw residues.
C:N ratio of the bodies of microorganisms is not only more constant but much
narrower, ordinarily falling between 4:1 and 9:1. Bacterial tissue in general is somewhat
richer in protein than fungi and consequently has a narrower ratio.
The most organic matter entering the soil carry large amounts of carbon and
comparatively small amounts of total nitrogen; that is, their carbon-nitrogen ratio is wide, and
the C:N values for soils are in between those of higher plants and the microbes.

5.19. Factors Affecting Soil Organic Matter and Nitrogen


5.19.1. Climate
Climate, especially temperature and rainfall, influence the amounts of nitrogen and
organic matter found in soil.
The decomposition of organic matter is accelerated in warm climates, a lower rate of
decomposition is found in cool regions. Under the uniform moisture conditions and similar
vegetation, the average total organic matter and nitrogen increase from two to three times for
each 10C fall in mean annual temperature.
Effective soil moisture also exerts a very positive control upon the accumulation of
organic matter and nitrogen in soils. Under comparable conditions, the nitrogen and organic
matter increases as the effective moisture becomes greater. At the same time, the C:N ratio
becomes wider.
- 56 -
Therefore, it can be noted that the organic situation in a soil is an expression for both
of temperature and precipitation plus other factors as well.

5.19.2. Natural Vegetation


Grasslands dominate the sub-humid and semi-arid areas, while trees dominate in
humid regions. Organic matter is higher in soils developed under grasslands than under
forests.

5.19.3. Soil Texture


A sandy soil usually carries less organic matter and nitrogen than one of a finer
texture. This is probably because of the lower moisture content and the more ready oxidation
occurring in sandy soils.

5.19.4. Soil Drainage


Poorly drained soils having high moisture retention and relatively poor aeration are
generally much higher in organic matter and nitrogen than in better drained soils.

5.19.5. Cropping
Cropped land contains much lower nitrogen and organic matter than comparable
virgin soils. This is because much of the plant materials are removed for human or animal
food and relatively little portion is returned to the soil in cultivated areas.
Soil tillage breaks up the organic residues and brings them into easy contact with soil
organisms, thereby increasing the rate of decomposition.
Considerable loss of soil organic matter and nitrogen is usually observed with
cropping.
Crop rotation, especially if legumes are included, helps to maintain soil organic
matter.
Soils that are kept highly productive by supplemental application of fertilizers, lime,
and manures are found to contain a higher organic matter content than a comparable less
productive soil. The amount of root and top residues to be returned to the soil are dependent
upon the level of soil productivity. However, even the most productive tilled field soil will
contain considerably lower organic matter than a virgin soil.
XXXXXXXXXX

- 57 -

You might also like