Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Republic of the Philippines

CAGAYAN STATE UNIVERSITY


Carig Campus, Tuguegarao City
College of Engineering

Course CHEM111E (Chemistry for Engineers)


Title of the
Unit IV. The Chemistry of the Environment – Soil Chemistry
Module
Module No. Module 4 - 3
At the end of this module, the student shall be able:
Learning a. To determine the different components of soil.
Objectives b. To distinguish the chemical reactions in the soil system.
c. To characterize the soil by its acidity and salinity.
I. Introduction
II. Soil Components
Content III. Reaction Processes in Soil
IV. Soil Acidity
V. Soil Salinity

Contributor/s Engr. Alva P. Durian

Introduction
Soils exist at an interface between the hydrosphere, atmosphere, biosphere, and lithosphere, and soil
solutions are greatly influenced by interactions with these other spheres. The interaction of Earth’s subsystems
within soil results in a mixture of solid, liquid, gas, and biota, called the pedosphere.

Soil is a complex mixture of inorganic and organic solids, air, water, solutes, microorganisms, plant, and
other types of biota that influence each other. Air and water weather rocks to form soil minerals and release
ions; microorganisms catalyze many soil weathering reactions; and plant roots absorb and exude inorganic and
organic chemicals that change the distribution and solubility of ions. Soil chemistry studies these chemical
processes in soils; specifically, chemical reactions, species, and transformations within and between solid, gas,
and liquid phases. Understanding speciation of solids and chemicals in soils is key to predicting soil properties
and how they will interact with plants, microbes, and animals. Chemical reactions in soils often lead to changes
between solid, liquid, and gas states that dramatically influence the availability of chemicals for plant uptake and
losses from soil that in turn are important aspects of fate and transport of nutrients and contaminants in the
environment.

Units of Soil Chemistry


Table 1. Parameters used in soil chemistry
Parameter Unit Unit conversion
Land area Hectare, Ha 1 Ha = 104 m2
Volume Cubic meter, m3 1 L = 10-3 m3
Specific surface area Square meters per kilogram m2•kg-1
Conductance or 1 S•m-1 = 1 ohm-1 = 1 mho;
Siemens per meter, S•m-1
electric conductivity 1 S•m-1 = 10 mho•cm-1 ; 1 dS•m-1 = 1 mho•cm-1
Amount of ion charge Moles charge mol(+) or mol(-)
Molar, M 1 M = 1 mol•L-1 = 103 mol•m-3 = 1 mmol•m-3
Concentration
Normality, N mol charge•L-1

Chemical Engineering Department Module 4 - 3: Page 1


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

Millimoles charge per kilogram 1 mol(+)•kg-1 = 10 mmol(+)•kg-1 = 1 cmol•kg-1


Cation exchange capacity
solid 1 mol(+)•kg-1 = 1 milliequivalent•100 g-1
Total exchangeable acidity Moles charge mol(+) or mol(-)
Sodium adsorption ratio Moles charge per volume mol(+) L-1 or mol(-) L-1
Total dissolved solids Mass per volume mg L-1

Soil Components
The elements in rock minerals at Earth’s surface are the starting materials for soils and contain the essential
elements from which soil and life evolved. Concentrations of elements in soils depend on the soil formation
factors, especially parent material, weathering processes, and biologically driven fluxes of elements. Most of the
mass of soil is comprised of oxygen, followed by silicon, aluminum, carbon, and iron. Elemental contents of
soils are highly variable, especially for nutrients such as nitrogen, sulfur, and phosphorus.

Figure 1. Percent mass composition of elements is soil. (Source: DG Strawn, 2019)

A. Essential Elements
Essential elements, commonly referred to as essential nutrients, are the elements from which plants, animals,
and humans evolved. Essential nutrients, in range of limit, are required for an organism to complete its life
cycle. When a nutrient is lower than the lower limit requirement, organisms suffer deficiency. While when some
essential nutrients are higher than the upper limit, it becomes toxic. Boron and selenium are micronutrient in
soils that has a narrow range between deficiency and toxicity essential for plant and animal, respectively. Some
micronutrients are only needed for animals, and not in plants. Most plant nutrients are also nutrients for animals.
However, animals require some micronutrients that plants do not.

Most essential elements are present as ions in the soil solution, and flow into the plant as it absorbs water.
Plants obtain hydrogen, carbon, and oxygen from air, but soils have pore space for O2 and CO2 movement
between plant roots and the atmosphere, and supply CO2 to the atmosphere through the decay of organic matter
by soil microorganisms. Animals derive most of their essential elements from plants. The ability of plants to
supply these elements to animals depends on a combination of factors: availability of the ions in the soil
solution, plant selectivity at the soil solution–root interface, and ion translocation from root to plant top.
Animals have evolved in the presence of soils and plants, and thus do not typically suffer microelement
deficiencies. However, essential elements in natural systems are occasionally too high or too low for animals
because plants can tolerate a much wider range of elemental concentrations than animals.

Chemical Engineering Department Module 4 - 3: Page 2


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

Figure 2. Periodic table of elements, not including lanthanides and actinide, showing nutrients and common contaminants.
(Source: DG Strawn, 2019)
The elements of primary interest for soil chemistry exhibit a wide range of chemical behavior that causes
highly varied mobility and bioavailability. The elements of common interest are grouped, as shown in Table 2,
according to general chemical properties and reactivity in soil.

Table 2. Ions of major interest in soil chemistry and their common species
Ion Remarks
Exchangeable cations They are retained by the negative charge of soil minerals and organic matter.
Ca2+, Mg2+, Na+, K+, NH4+, Al3+ It occur as exchangeable cations in soils; these ions are easily manipulated
by liming, irrigation, or acidification; exchangeable Al 3+ is characteristic of
acid soils; productive agricultural soils are rich in exchangeable Ca 2+; NH4+,
and K+; K+ can become fixed in clay minerals.
Soluble anions They are present in lower concentrations than the major cations except in
NO3 –, SO42–, Cl–, coarse‐textured and strongly saline soils; sulfate and nitrate are important
H2CO3, HCO3 –, nutrient sources for plants; sulfate, chloride, and bicarbonate salts
Se6+ as SeO42– accumulate in saline and alkaline soils; selenite (SeO42–) anion is more
soluble than selenite (SeO32–).
Poorly soluble anions They are strongly retained by soils and present at low concentrations in soil
– 2–
H2PO4 , HPO4 , H3BO3, H4BO4 , –
solution (typically <10–5 M). It often occur as minerals, salts, or sorbed to
Si(OH)4, MoO42–, Se4+ as HSeO3 – mineral surfaces; borates are the most soluble of the group; molybdate and
silica are more soluble at high pH; phosphate is more soluble at neutral or
slightly acid pH; selenite is more reduced and less mobile than selenate.
Poorly soluble metal cations Its concentration in soil solutions are typically much less than alkali and
Al3+, Cu2+, Zn2+, Ni2+, Co2+ alkaline earth cations. As silica and other ions leach during weathering,
Fe(OH)2 +, Fe2+, Fe3+, insoluble hydroxides accumulate in soils; iron and manganese are more
Mn4+, Mn3+ Mn2+ soluble in waterlogged or reduced soils; availability increases with
increasing soil acidity; metals are complexed by soil organic matter (SOM)
and strongly adsorbed on mineral surfaces.
Toxic ions These ions are present in soil at concentrations that pose toxicity risks. Often
Al3+, not readily soluble; Al3+ is a hazard to plants, the others are of more concern
As(V): H2AsO4 –, HAsO42– to animals; Cd2+ is relatively available to plants; As and Cr oxyanions
As(III): H2AsO3 –, HAsO32– increase in solubility with pH; As(III) is more soluble than As(V); cation
Cr(VI): CrO42– contaminants are less available to plants with increasing pH; many
Cd2+, Hg2+, Pb2+ contaminants form poorly soluble sulfides in reduced soils.
Biogeochemical cycling elements These elements are primarily produced by plants, fixation by microbes, and
C4– to C4+ (CH4 to CO2) microbial driven mineralization. Soil biochemistry revolves around the
oxidation state changes of soil carbon, nitrogen, and sulfur compounds;

Chemical Engineering Department Module 4 - 3: Page 3


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

N3– to N5+ (NH4+, N2, NO2, NO3–) nitrogen occurs in oxidation states from N(‐III) to N(V); molecular oxygen
O2–, O2 is the main electron acceptor; nitrate, and sulfate are electron acceptors
S2– to S6+ (H2S to SO42–) when oxygen supply is low.

B. Inorganic Components
Inorganic solids in soils comprise mixtures of various types of minerals existing as rocks, sand, silt, and
clays. The inorganic components of soils represent more than 90% of the solid components that include both
primary and secondary minerals which range in size (particle diameter) from clay-sized colloids (<0.002mm) to
gravel (>2 mm) and rocks. A mineral can be defined as a natural inorganic compound with definite physical,
chemical, and crystalline properties. A primary mineral is one that has not been altered chemically since its
deposition and crystallization from molten lava. Primary minerals occur primarily in the sand (2–0.05mm
particle diameter) and silt (0.05–0.002mm particle diameter) fractions of soils but may be found in slightly
weathered clay-sized fractions. A secondary mineral is one resulting from the weathering of a primary mineral;
either by an alteration in the structure or from reprecipitation of the products of weathering (dissolution) of a
primary mineral. The secondary minerals are primarily found in the clay fraction of the soil but can also be
located in the silt fraction.
Table 3. Common primary and secondary minerals in soils (Source: DL Sparks, 2003)

C. Organic Components
There is no unique structure to soil organic matter (SOM). SOM are fractions of organic material and the
products in the stages of its degradation. SOM consists of nonhumic and humic substances. Non-humic
substances have recognizable physical and chemical properties which include carbohydrates, lipids (fatty acid,
fat, waxes, resins) and amino acids. While humic substance display a complex chemical structure with high
molecular weight, have hydrophilic character and acid properties. SOM contents range from 0.5% to 5% on a
weight basis in the surface horizon of mineral soils to 100% in organic soils. It improves soil structure, water-
holding capacity, aeration, and aggregation. The main constituents of SOM are C (52–58%), O (34–39%), H
(3.3–4.8%), and N (3.7–4.1%). Other prominent elements in SOM are P and S. SOM has a high specific surface
(as great as 800–900 m2 g–1) and a cation exchange capacity (CEC) that ranges from 150 to 300 cmol kg–1. Due
to the high specific surface and CEC of SOM, it is an important sorbent of plant macronutrients and
micronutrients, heavy metal cations, and organic materials such as pesticides.

The quantity of SOM in a soil depends on the five soil-forming factors: time, climate, vegetation, parent
material, and topography. These factors vary for different soils, and thus SOM accumulates at different rates
and, therefore, in varying quantities. The accumulation rate of SOM is usually rapid initially, declines slowly,
and reaches an equilibrium level varying from 110 years for fine-textured parent material to as high as 1500

Chemical Engineering Department Module 4 - 3: Page 4


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

years for sandy materials. The equilibrium level is attributed to organic acids that are produced which are
resistant to microbial attack, the stability of humus due to its interactions with polyvalent cations and clays, and
low amounts of one or more essential nutrients such as N, P, and S which limit the quantity of stable humus that
can be synthesized by soil organisms.

Reaction Processes in Soil


The figure below, Figure 3, is a comprehensive view of the six reactions in a soil system. The soil chemical
may change via one of these reactions to re-establish equilibrium. The properties such as soil pH, microbial
activity, soil pore space, mineral composition, organic matter composition, chemical concentration, and soil
texture influence the reactivity and speciation of chemicals in soils.

The reactivity of soils greatly influences the composition of the soil solution. Ion availability in solution is
renewed by soil reactions that add ions back to the soil solution after they are depleted. Main sources of ions to
soil solution are mineral weathering, organic matter decay, rain, irrigation waters, fertilization, and release of
ions adsorbed by clays and organic matter in soils. Reaction processes in soil includes sorption/desorption,
precipitation/dissolution, immobilization/mineralization, complexation/dissociation, gas
dissolution/volatilization, and oxidation/reduction.

Figure 3. Chemical reactions in soil (Source: DG Strawn et al. 2019)

A. Sorption/Desorption
Sorption and desorption reactions describe association and release of a chemical from a particle (minerals,
SOM, or a biological cell surface). Often, sorption reactions are termed adsorption, which implies accumulation
of a substance or material at an interface between the solid surface and bulk solution, and is not forming a three-
dimensional network of atoms on the surface (called surface precipitation). One example is the adsorption of a
sodium ion onto a clay mineral surface.

The release of the potassium from the clay in the equation is a desorption reaction. Together, the adsorption and
desorption reactions depicted in the equation are example of a cation exchange reaction.

B. Precipitation/Dissolution
These reactions describe the change in a chemical from solution to the solid state, where a new solid is
formed from solution constituents. Dissolution is the reverse of precipitation, where ions from the solid are
released to the solution. An example of a precipitation–dissolution reaction is the formation of the calcium
carbonate mineral calcite in soils:

Chemical Engineering Department Module 4 - 3: Page 5


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

( )

where the forward reaction is a dissolution reaction, and the reverse is a precipitation reaction. Note in this
reaction the (s) is placed on the calcite to indicate that it is a solid and not an aqueous complex while ions are in
aqueous phase.

C. Immobilization/Mineralization
Immobilization and mineralization reactions are generally biologically mediated. Immobilization refers to
the uptake of chemical into the cellular structure of an organism, such as a microbe, fungi, or plant. The
chemical within the organism is considered a biologically formed molecule (biomolecule). An example is the
uptake of nitrate from soil solution into a plant, where it is utilized as a cellular metabolite to produce amino
acids, such as glutamate (C5O4H6NH3), which are components of proteins:

In this reaction, the C5O1H9 is simply an element placeholder for cellular compounds to provide the reactants
needed for the stoichiometry to balance and does not represent a molecular species in the cell.

Mineralization implies degradation, release, or conversion of a chemical to a form that is no longer a


biomolecule. Products of mineralization reactions are inorganic chemicals and degraded organic chemicals.
Degradation of organic nitrogen in glutamate to ammonium is an example of a mineralization reaction:

This is a summary reaction describing complete degradation of the glutamate biomolecule produced in the
previous reaction to produce ammonium, carbon dioxide, and protons. The reaction is carried out by microbes in
soils.

D. Complexation/Dissociation
These reactions describe interactions of two or more chemicals or aqueous ions. Protonation and
deprotonation (gain and loss of H+ ions) are specific types of complexation and dissociation involving
acceptance and loss of a proton by an acidic ion or molecule. Hydrolysis is a dissociation reaction in which H+
is released from a water molecule. Carboxylic acid, a common functional group on soil organic matter, is a weak
acid that deprotonates between approximately pH 3 and 6:

The R indicates the rest of the organic compound that the carboxylic acid functional group is attached. The
acidity of the carboxylic acid functional group depends on the composition of the rest of the organic molecule.

Complexation reactions change the valence and molecular properties of chemicals in soil solutions, thus
changing the chemical’s solubility, plant availability, and transport through the soil. Aqueous complexation of
ions occurs in soil solution and changes the concentrations of free ions (noncomplexed ions). For example, the
inorganic ligand bicarbonate readily forms aqueous complexes in solution with dissolved metal cations:

The ZnHCO3+ aqueous complex would occur in the soil solution instead of the free hydrated Zn 2+ ion.

E. Gas dissolution/Volatilization
Dissolution and volatilization of gases in soils refers to reactions occurring between the soil atmosphere and
the soil solution – specifically, transfer of gaseous chemicals into the aqueous phase, and the reverse. Since this
reaction involves movement of gas into and out of liquid water, it is different from condensation and
vaporization of a pure liquid to gas, and vice versa. An example of gas dissolution in soil solution is the reaction

Chemical Engineering Department Module 4 - 3: Page 6


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

of carbon dioxide (gas) with water to form carbonic acid (aqueous).


( ) ( )

Oxygen, nitrogen, and sulfur also have important gas dissolution and volatilization reactions. Ammonium is
a common ion in soil solution; however, when it deprotonates, it forms ammonia that will volatilize:
( ) ( ) ( )

The reverse reaction is used to dissolve ammonia gas (anhydrous ammonia) into soil solution to produce
ammonium ions for soil fertilization. The reaction predicts that adding anhydrous ammonia to soils would cause
the pH of the soil solution to increase because the ammonia would protonate to ammonium, thereby consuming
protons; or hydrolyzing water and releasing hydroxide ions.

F. Oxidation/Reduction
The gain and loss of electrons from an element cause a change in oxidation state. Often, redox reactions
result in changes in the physical state or molecular structure, and thus may be combined with other reaction
types. Many of the most reactive redox environments are those that exist at the interface between oxidized and
reduced zones (the redox interface). At the redox interface, chemical and mineral speciation is continuously
changing. The redox interface is influenced by fluxes of temperature, water, nutrients, and gases through the
soil. Oxygen fluxes into and out of the soil system are especially important because O 2 readily accepts electrons
in abiotic and biotic processes and is the most common electron acceptor used for metabolism.

1. Electron donors in nature


In soils, carbon compounds in roots, microbes, dead plant matter, and SOM are the major electron
donors. The half‐reaction for the oxidation of theoretical soil organic matter is

The reaction proposes that nine electrons come from oxidizing one mole of SOM to CO2 (C4+). This
reaction is facilitated by microorganisms and other organisms (e.g. earthworms), who use the oxidation
reaction to obtain electrons for respiration. SOM also contains amino (‐NH2) and sulfhydryl (-SH) groups,
which are also electron donors.

Inorganic electron donors in soils typically occur in much smaller amounts than organic compounds, and
include sulfide (S2–), sulfur (S0), Fe2+, Mn2+, and Mn3+, and ammonia (N3–). Oxidation of inorganic
chemicals can occur biotically or abiotically. Organisms (called chemotrophs) can use electrons by oxidizing
these elements. Nitrification is an example of an important soil process in which the chemotrophic microbes
utilize the electrons in ammonia as an energy source.

Oxidation by Nitrosomonas bacteria


Oxidation by Nitrobacter
Nitrification reaction

2. Electron acceptors in nature


In soils, O2 diffuses through pores to plant roots and soil microbes, where it can be utilized as a terminal
electron acceptor (TEA). Soils and waters that have available O2 are called oxic while soils that have no
available O2 are called anoxic. When O2 availability is low, soil microorganisms utilize the oxidized states of
nitrogen, sulfur, iron, manganese, and other elements as TEAs. The general order of preference for TEAs
based on energy produced is:

Chemical Engineering Department Module 4 - 3: Page 7


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

The exact order varies, depending on the species of the electron acceptors and environmental conditions.
In addition, pH is a major environmental factor that influences the relative order of the preference for TEAs.

Figure 4. Correlation of soil redox potential with change in TEA, water, O2, metabolism and redox status category
(Source: DG Strawn et al. 2019)
The most common secondary electron acceptors in soils include iron and manganese oxides, sulfate, and
oxidized forms of nitrogen. Reducing half‐reactions for these species are
( ) ( )
( ) ( )
( ) ( )
( ) ( )
( ) ( )
( ) ( )

Reduction of secondary electron acceptors produces compounds that are unstable in the presence of
oxygen and convert back to the oxidized species through either abiotic or biotic processes. The accumulation
rate of SOM reflects the difference of the rates of organic matter addition vs. oxidation rates. The rate of
addition is equivalent to the rate of net photosynthesis. The oxidation rate is governed by temperature and by
the rate of oxygen supply. Spatial variation of O 2 concentrations in soils provides zones of varying redox
conditions, or redox gradients, where microbial communities establish their own niches based on the
availability of TEAs.

Soil Acidity
Approximately 30% of Earth’s soils are acidic. Soil acidification can be detrimental for agriculture because
it decreases availability of anionic nutrients, causes cationic nutrients to be leached from the soil profile, and
causes Al3+ and Mn2+ toxicity. Forest soils and wetland soils are often naturally acidic. pH, the negative
logarithm of the activity of hydrogen cations in a solution, affects speciation and availability of many chemicals
in soils and is often considered the master variable for characterizing soil chemical behavior.

Chemical Engineering Department Module 4 - 3: Page 8


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

Figure 5. The soil-pH scale (Source: DG Strawn et al. 2019)

A. Soil acidification
Soil acidification occurs when acids are added to soils, or bases are lost. Soil acidification is a continuum of
reactions and include natural fluxes (e.g., weathering followed by leaching), and human‐caused fluxes to the soil
(fertilization and harvest) that cause net changes in the active, exchangeable, and reserve acidity components.
Fluxes alter the biogeochemical cycling of carbon, nitrogen, or sulfur by changing either their concentrations or
species. Reactions important for soil acidification are classified as either net proton sources to soil solution or
net proton sinks that remove protons from soil solution.

Process Net Reaction


Carbon Cycle
Organic acid production (SOM)
Degradation of organic acid (decarboxylation) ( )
Nitrogen Cycle
Organic N mineralization (ammonification)
Nitrification

B. Sources of soil acidity


Soil acidity sources are acid rain, agronomic practices, and mine spoil and acid sulfate soils. Acid vapors,
primarily sulfuric (H2SO4) and nitric (HNO3), form in the atmosphere as a result of the emission of sulfur
dioxide (SO2) and nitrogen oxides from natural and anthropogenic sources such as burning of fossil fuels
(source of sulfur gases) and the exhaust from motor vehicles (source of nitrogen oxides). Acid rain can cause
leaching of nutrient cations such as Ca 2+, Mg2+, and K+ from the soil, resulting in low pH and the solubilization
of toxic metals such as Al3+ and Mn2+. This can cause reduced soil biological activity such as ammonification
(conversion of NH4+ to NO3–) and reduced fixation of atmospheric N2 by leguminous plants and can also reduce
nutrient cycling.

In most cases, the amount of soil acidification that occurs naturally or results from agronomic practices is
significantly higher than that occurring from acid rain. For example, if one assumes annual fertilizer application
rates of 50–200 kg N ha–1 to soils being cropped, soil acidification from the fertilizer would be 4–16 times
greater than acidification from acid rain in highly industrialized areas. However, on poorly buffered soils, such
as many sandy soils, acid rain could increase their acidity over time.

Chemical Engineering Department Module 4 - 3: Page 9


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

Mine spoil and acid sulfate soils have very low pH due to the oxidation of pyrite. Mine spoil soils are
common in surface-mined coal areas, and acid sulfate soils occur in marine flood plains in temperate and
tropical areas. The complete oxidation of pyrite (FeS 2) produce sulfuric acid.
Acidity from rain

Acidity from agronomic practices (fertilizer application)

( ) Acidity from surface-mine coal areas

The high concentrations of sulfuric acid cause pH as low as 2 in mine spoil soils and <4 in acid sulfate soils.

C. Forms of Soil Acidity


The extreme acid produced moves into drainage and floodwaters, corrodes steel and concrete, and causes
dissolution of clay minerals, releasing soluble Al. The main form of acidity in mineral soils are associated with
Al, which can be exchangeable/extractable, nonexchangeable, or precipitated as an array of solid phases such as
gibbsite, or nordstrandite. Only in acidic soils with a pH < 4 and in soils high in organic matter does one find
major quantities of exchangeable H+.

Exchangeable acidity is the amount of the total cation exchange capacity (CEC) due to H+ and primarily
3+
Al . As a proportion of the total acidity its quantity depends on the type of soil (e.g., type and quantity of soil
components) and the percentage of the CEC composed of exchangeable bases such as Ca 2+, Mg2+, K+, and Na +,
or the percentage base saturation. Aluminum oxides appear to be more effective in reducing the quantity of
exchangeable acidity than iron oxides.

D. Measurement of Soil Acidity


1. H+ activity of aqueous solutions is measured using either a pH probe or colorimetric analysis. In the
laboratory, soil pH measurement methods use different ratios of soil and solution, and different soil‐
wetting solutions (salt or deionized water), all of which may affect the measured pH. Thus, it is important
to note the pH measurement method used when measuring soil pH.

2. The concentration of base cations in soil solutions is an indirect measure of the ability of a solution to
neutralize acid, which is equivalent to a solution’s alkalinity or acid neutralization capacity. Soils with
high concentrations of base cations are typically alkaline. Fluxes of base cations into a soil create
alkalinity, and fluxes out of a soil are associated with soil acidification. The relationship between base
cations and soil acidity and alkalinity is complex, and involves both solution and solid phase reactions.

2.1 Exchangeable base cations


When base cations are removed from solution by precipitation, absorption by plants and microbes,
or leaching, the lost base cations are replenished by desorption of base cations from the soil’s
exchangeable base cations. The presence or absence of base cations on exchange sites greatly influences
exchangeable acidity. The percent of exchange sites occupied by exchangeable base cations is the
percent base‐cation saturation (%BS):
[ ] [ ] [ ] [ ]

where [XM+] indicates the concentration of base cation (M+) on the exchange site (X), and CEC is
the cation exchange capacity. The pH used for CEC measurement must be specified because CEC
increases with pH. The denominator includes any additional charge (CEC) generated by SOM and
oxide‐mineral complexes between the initial soil pH and the reference pH (7 or 8.2) that the CEC is

Chemical Engineering Department Module 4 - 3: Page 10


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

measured. Since neither exchangeable Al3+ nor exchangeable H+ is appreciable above pH 5.5, the CEC
above this pH should be 100% base saturated. However, % BS in soils in the pH range 5.5 to 8.2 are
often well below 100%. %BS is useful for soil classification purposes, and for empirical liming
recommendations. From the standpoint of soil chemical properties and reactions, %BS is considered an
acidity index useful for relative comparisons of soil pH and buffering properties.

Learning Activity: A soil of pH 5 may have 5 mmol(+) kg−1 of exchangeable bases (Ca2+, Mg2+, K+,
and Na+), and mmol(+) kg−1 of exchangeable acidity for a total cation exchange capacity of 6 mmol(+).

The %BS at pH 5 is
( )
( )
( )

At pH 7 the CEC is 8 mmol(+) kg−1, the %BS at pH 7 is


( )
( )
( )

2.2 Total exchangeable acidity


Total exchangeable acidity (TEA) is defined as the amount of exchangeable H+ and Al3+, measured
in mmol(+) kg−1 soil:
[ ] [ ]
( )

where [Al3+] and [H+] are the concentrations of aluminum and hydrogen ions, respectively. BS is the
fraction of base saturated cation on the ion exchange sites (%BS/100). For a given pH, CEC is fixed,
thus base saturation and TEA are inversely related. The exchangeable acidity represents the resupply
capacity of soil acidity. Soils with high %BS tend to have low TEA and are alkaline. Soils with low
%BS tend to have high TEA and are acidic.

E. Managing Acidic Soils (Liming Soils)


A major challenge for managing acidic soils is to estimate the quantity of lime required to raise the soil pH
to a certain level. Both exchangeable and titratable acidity will be neutralized during the slow titration. Aqueous
solution pH measurements represent the activity of hydronium ions (H3O+), and not the H+ ion activity. Protons
are very reactive and have extremely low activity in aqueous solution.

An important effect of lime is to provide hydroxyl ions that convert exchangeable Al 3+ to Al(OH)3
(gibbsite). Increased quantities of soluble and exchangeable Ca2+ and Mg2+ are byproducts of liming, which
serve to displace exchangeable acidity, and may be beneficial to plants, such as legumes, having high calcium
requirements. These can be observed in weathering reaction in the formation of kaolinite and gibbsite.

Dissolution of feldspar

( ) Precipitation for kaolinite


( ) Weathering reaction

The calculation of CaCO3 lime required from a titration is based on production of two hydroxide ions per
mole of CaCO3:
( )

Chemical Engineering Department Module 4 - 3: Page 11


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

Learning Activity: If a soil titration, indicates that 2.0 mmol of OH− is consumed per 100 g of soil for each
unit increase in pH, then 5 Mg CaCO3 per ha (30 cm) is required for every unit increase in pH desired. How
much CaCO3 is needed to increase the pH?

( ) ( )

This is the apparent lime‐based buffer capacity (BC) of the soil, which is the amount of lime required to add
to soil per pH‐unit increase. The calculation requires the density of the soil (assumed 4.5 × 109 g soil per ha(30
cm)), and assumes that pure CaCO3 that provides 2 mmol OH− per mmol CaCO3 is used as the lime amendment
(and that it completely dissolves). Corrections for actual density of the soil, and type of amendment are required
for accurate predictions of buffer capacity of a given soil. The lime requirement (LR), in Mg ha−1 is calculated
using the soil buffer capacity:
( )

where pHoptimal is the soil pH required for the specified cropping system, pHcurrent is the pH of the topsoil, and BC
is the calculated buffer capacity of the soil.

Acidity neutralization by field‐liming is typically incomplete because of incomplete mixing and slow
reaction times. The lime dissolution reaction rates vary inversely with pH, limestone particle size, and solubility
of the liming agent. Hence, the laboratory‐based lime requirement value is often further multiplied by a
conversion factor to better estimate the amount of lime needed to achieve a given field pH. Such a conversion
factor is regionally specific, and dependent on the type of lime. The usual procedure to estimate lime
requirement is to add a pH buffer solution to the soil, measure the amount of buffer consumed or the resulting
pH of the soil‐buffer suspension, and calibrate results with field‐lime requirements for similar soils from the
same geographical area.

Learning Activity: A pH change of 0.1 unit from the initial buffer pH might correspond to 1 Mg limestone
ha−1, which corresponds to a rate of 10 Mg limestone ha−1 for a full unit pH change of the buffer (1 Mg ha −1 ×
10). If a soil has an initial pH of 5.5, the buffer solution has an initial pH of 6.8, and the final mixture has a pH
of 6.3 (pHdesired – pHbuffer = 0.5). What lime requirement for this soil?

Using a calibration curve from the titration of one soil can be used to estimate the lime requirements of other
soils from the same geographic region if soil texture and measurements of initial soil pH are incorporated in an
empirical model, but the predicted lime rates will be less precise.

Soil Salinity
Soil salinity includes soluble salts in soil water and salt solids in the soil. Soil salinity and sodicity can have
a major effect on the structure of soils. Soil structure, or the arrangement of soil particles, is critical in affecting
permeability and infiltration. Infiltration refers to the “downward entry of water into the soil through the soil
surface”. If a soil has high quantities of Na + and the electric conductivity is low, soil permeability, hydraulic
conductivity, and the infiltration rate are decreased due to swelling and dispersion of clays and slaking of
aggregates. Typically, soil infiltration rates are initially high, if the soil is dry, and then they decrease until a
steady state is reached. Swelling causes the soil pores to become more narrow, and slaking reduces the number
of macropores through which water and solutes can flow, resulting in the plugging of pores by the dispersed

Chemical Engineering Department Module 4 - 3: Page 12


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

clay.

A. Sources of Soluble Salts


The major sources of soluble salts in soils are weathering of primary minerals and native rocks, residual
fossil salts, atmospheric deposition, saline irrigation and drainage waters, saline groundwater, seawater
intrusion, additions of inorganic and organic fertilizers, sludges and sewage effluents, brines from natural salt
deposits, and brines from oil and gas fields and mining. As primary minerals in soils and exposed rocks weather
the processes of hydrolysis, hydration, oxidation, and carbonation occur and soluble salts are released. The
primary source of soluble salts is fossil salts derived from prior salt deposits or from entrapped solutions found
in earlier marine sediments.

B. Characterization of salinity in soil and water


The parameters determined to characterize salt-affected soils depend primarily on the concentrations of salts
in the soil solution and the amount of exchangeable Na + on the soil (sodicity). Exchangeable Na + is determined
by exchanging the Na+ from the soil with another ion such as Ca 2+ and then measuring the Na+ in solution by
flame photometry or spectrometry.

1. Sodium adsorption ratio (SAR)


Cation exchange selectivity equations are used to predict the distribution of Na +, Ca2+, and Mg2+ on the
exchange sites. The Gapon exchange equation is the most commonly used cation exchange selectivity
equation used to evaluate soil salinity because it is relatively simple and has been shown to be accurate for
many salt‐affected soils. The Gapon formulation of the Na +–Ca2+ exchange reaction is

The Gapon exchange constant for this reaction is


[ ][ ]
[ ][ ]

[ ] [ ]
[ ] [ ]

where brackets on aqueous species indicate concentrations (mol L−1 or mmol L−1), and exchanger phase
concentrations are equivalent cation charge (mol(+) kg−1 or mmol(+) kg−1). This shows that the ratio of Na +
to Ca2+ on the exchange sites (left‐hand side) is a function of the ratio of the cations in the aqueous solution
multiplied by the exchange selectivity coefficient (K G) (right‐hand side). The range of KG is typically 0.008
to 0.016 mmol−1/2 L1/2 for alkali soils; a value of 0.015 is commonly used for soils with ESR < 30. If K G is
known, the Gapon equation can be used to predict the distribution of Na + and Ca2+ between the solid and
solution. The presence of Mg2+in most alkaline soil solutions complicates prediction because the ternary Na +
- Ca2+- Mg2+ exchange-reaction equilibrium state requires a more complex model than the Gapon equation.
To simplify the calculation of Na+, Ca2+, and Mg2+ on the soil particles, the exchange behavior of Ca 2+ and
Mg2+ is assumed to be similar. Thus,
[ ] [ ]
[ ] [ ] ([ ] [ ])
Exchanger phase Solution phase

This equation allows prediction of the relative distribution of Na + on soil’s exchange sites using the
concentrations of Na+, Mg2+, and Ca2+ in solution and KG as defined above. The ratio of the solution

Chemical Engineering Department Module 4 - 3: Page 13


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

composition is called the sodium adsorption ratio (SAR):


[ ]
([ ] [ ])

when the units of charge equivalent per volume (mmol(+) L−1) are used instead of molar concentrations, the
Ca2+ and Mg2+ concentrations in equation are divided by two. The units of SAR are mmol(+) 1/2 L−1/2,. SAR is
a prediction of how sodicity of irrigation water will impact the behavior of soils irrigated with the water.

2. Exchangeable sodium percentage (ESP)


The sodium concentration on the exchange sites can be directly measured, and is typically reported as the
exchangeable sodium percentage (ESP), which is based on the total Ca 2+, Mg2+, and Na+ cation charge
occupation:
[ ]
([ ] [ ] [ ])

where solid‐phase concentrations are in units of charge per mass (mmol(+) kg−1). In a system where Na+,
Mg2+, and Ca2+ are the main cations, the denominator is equal to the cation exchange capacity
[ ]

this relationship is useful because it relates the amount of Na + on the exchange site to CEC, a commonly
reported soil parameter. If ESP values are above 30%, the exchangeable sodium ratio (ESR) can be used to
predict ESP:
[ ]
([ ] [ ])

As shown, ESR can be predicted from solution composition (SAR) and an exchange coefficient (K G). Thus,

this relationship allows ESP to be predicted from the SAR of irrigation water or saturated extract solution
and an exchange coefficient (substituting ESR with SAR and K G). Thus, exchangeable sodium percentage
can be computed from cation composition of an irrigation water or saturated paste extract.

Learning Activity: An irrigation water will be used to irrigate field of citrus, which has high sensitivity for
sodium damage. Thus, the farmer desires to maintain soil ESP as low as possible. The irrigation water
contains 2.08 mmol L−1 of Ca2+, 0.71 mmol L−1 of Mg2+, 5.96 mmol L−1 of Na+. Assuming a KG of 0.015
and that the soil pore water will have the same composition as the irrigation water, the ESP can be calculated
from the SAR and ESR:
[ ]
( )
([ ] [ ]) ( )

Thus, the ESP predicted from the irrigation water is relatively low and is suitable for irrigating the citrus
crop.

Chemical Engineering Department Module 4 - 3: Page 14


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

3. Total dissolved solids


Early appraisals of the salinity of irrigation waters were in terms of amounts of total dissolved solids
(TDS). The TDS are determined by evaporating a known volume of water to dryness and the solids residue
remaining are weighed. However, this measurement is variable since in a particular sample various salts
exist in varying hydration states, depending on the amount of drying. The presence of hygroscopic water in
the resultant salt mixtures causes TDS values to strongly depend on the drying conditions and the salt type.
Concentrations of salts in most irrigation waters are less than 1000 mg L −1 TDS. Groundwater used for
irrigation is usually higher in TDS than surface waters. For comparison, TDS of seawater is about 35 000
mg L−1. TDS is a useful parameter for measuring the osmotic potential, –τo, an index of the salt tolerance of
crops. For irrigation waters in the range of 5-1000 mg L–1 TDS, the relationship between OP and TDS is

( )

The TDS (in mg L–1) can also be estimated by measuring an extremely important salinity index, EC, to
determine the effects of salts on plant growth. The TDS may be estimated by multiplying EC (dS m–1) by
640 (for EC between 0.1 and 5.0 dS m–1) for lesser saline soils and a factor of 800 (for EC > 5.0 dS m–1) for
hypersaline samples. To obtain the total concentration of soluble cations (TSC) or total concentration of
soluble anions (TSA), EC (dS m–1) is usually multiplied by a factor of 0.1 for mol L–1 and a factor of 10 for
mmol L–1.

4. Electrical conductivity (EC)


Salinity of a solution is directly related to its electrical conductivity (EC). The EC is based on the
concept that the electrical current carried by a salt solution under standard conditions increases as the salt
concentration of the solution increases. A number of EC values can be expressed according to the method
employed: ECe, the EC of the extract of a saturated paste of a soil sample; EC p, the EC of the soil paste
itself; ECw, the EC of a soil solution or water sample; and ECa, the EC of the bulk field soil. The EC and
temperature of the extract are measured using conductance meters/cells and thermometers and EC 298 is
calculated using below equation:

where ft is a temperature coefficient that can be determined from the relation ( )


and t is the temperature at which the experimental measurement is made in degrees Kelvin.

Marion and Babcock (1976) developed a relationship between EC w (dS m–1) to total soluble salt
concentration (TSS in mmol L–1) and ionic concentration (C in mmol L–1), where C is corrected for ion pairs.
If there is no ion complexation, TSS = C. The equations of Marion and Babcock (1976) are

This is applicable to 15 dS m–1, which covers the range of ECe and EC w for slightly to moderately saline
soils. Griffin and Jurinak (1973) also developed an empirical relationship between EC w and ionic strength (I)
at 298 K that corrects for ion pairs and complexes

where ECw is in dS m–1 at 298 K.

To measure the EC of a solution, it is placed between two electrodes of constant geometry separated by
a known distance. An electrical potential is applied across the electrodes, and the resistance (R) of the
solution between the electrodes is measured. The electrical current varies directly with the total
concentration of dissolved salts (ions in solution). The resistance of a conducting material (e.g., a salt
solution) is inversely proportional to the cross-sectional area (A) and directly proportional to the length (L)

Chemical Engineering Department Module 4 - 3: Page 15


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

of the conductivity cell that holds the sample and the electrodes. Specific resistance (Rs) is the resistance of a
cube of a sample volume 1 cm on edge. Since most commercial conductivity cells are not this large, only a
portion of Rs is measured. This fraction is the cell constant (K = R/Rs). The reciprocal of resistance is
conductance (C). It is expressed in reciprocal ohms or siemens (formerly mhos). When the cell constant is
included, the conductance is converted, at the temperature of the measurement, to specific conductance or
the reciprocal of the specific resistance. The EC of the saturation extract of the soil measures the soil’s
salinity, and the EC of irrigation water measures the water’s salinity. The specific conductance is the EC,
expressed as

The absolute value of the conductance in a solution is a result of the solution’s salt concentration and the
geometry of the electrode cell. The effects of electrode geometry are embodied in the cell constant, which is
related to the distance between electrodes and their cross‐sectional area.

Learning Activity: Calibration might yield a cell constant of 2.0 cm−1; a test solution measuring 2000 Ω
resistance (conductance of 1/2000 Ω−1 or 0.0005 siemens) has a conductivity of

For soil extracts in the EC range from 3 to 30 dS m−1, the osmotic potential (OP) is
( ) ( )

The osmotic pressure or osmotic potential measures the tendency of water to diffuse across a membrane
against a salinity gradient and indicates the effects of salinity on plant growth since plant roots are
semipermeable membranes (they allow water, but not salt to enter the roots).

The traditional classification of salt‐affected soils uses soluble salt concentrations or electrical conductivities
of extracted soil solutions, and the ESP of the soil. The EC dividing line for most plants between saline and non-
saline soils was established at 4 dS m−1 for water extracts from saturated soil pastes. Salt‐sensitive plants,
however, can be affected in soil with saturation extract ECs of 2 to 4 dS m−1. The relationships of soil salinity
parameters to EC, SAR, and ESP are shown in below figure.

Chemical Engineering Department Module 4 - 3: Page 16


Republic of the Philippines
CAGAYAN STATE UNIVERSITY
Carig Campus, Tuguegarao City
College of Engineering

Figure 6. Classification of soil salinity and sodicity in relation to EC, SAR, ESP, and pH (Source: DG Strawn et al. 2019)

Self-Assessment Activities:
Give one method or treatment of soil decontamination and explain its advantage and limitation.

References:
Sparks, D.L. (2019). Fundamentals of Soil Chemistry. DOI 10.1002/9781119300762.wsts0025
Sparks, D.L. (2003). Environmental Soil Chemistry (2nd Ed.). Academic Press, UK. ISBN 0-12-656446-9
Strawn, D.G., Bohn, H.L., & O’Connor, G.A. (2020). Soil Chemistry (5th Ed.). John Wiley & Sons, Ltd., USA.
ISBN 9781119515159

Chemical Engineering Department Module 4 - 3: Page 17

You might also like