You are on page 1of 265

GUY SELA

FERTILIZATION
AND IRRIGATION
THEORY AND BEST PRACTICES

2021 EDITION
About the Author
Guy Sela was born in 1973, in Rehovot, one of the first
agricultural towns in Israel. Mr. Sela graduated, with a BSc.
degree from the Hebrew University as an agronomical
engineer with specialization and expertise in irrigation,
fertilization, and water treatment.
His work in large scale, nationally and internationally, enabled
him to lead comprehensive research programs on crop
nutrition and irrigation using cutting edge agricultural
technologies.
Guy spent many years researching and experimenting ways to
improve agricultural production and bring progress also to developing countries.
He observed many growers who mainly rely on trial and error and estimation and past
experience. He realized that the broad misuse of fertilizers, water and pesticides is a
global phenomenon which results in decreased yields, waste, and damage to crops
and the environment.
This led him to establish and lead his innovative venture Smart Fertilizer between the
years 2008-2017. In 2018 Guy established Cropaia, an ag and water consulting firm,
dedicated to knowledge dissemination and in 2019 he founded yieldsApp, a startup
company dedicated to innovation in agriculture.

Guy Sela is an internationally renowned speaker in in conferences, symposiums, and


online webinars, which are brought to thousands of growers. This has allowed him to
share his knowledge, expertise world-wide.

1
Table of Contents
Chapter 1 ............................................................................................................................. 5
The Essential Nutrients ......................................................................................................... 5
Plant Nutrients - Introduction ................................................................................................. 6
Nitrogen ................................................................................................................................ 9
Potassium ........................................................................................................................... 14
Phosphorus......................................................................................................................... 18
Calcium ............................................................................................................................... 22
Magnesium ......................................................................................................................... 25
Sulfur .................................................................................................................................. 28
Iron ..................................................................................................................................... 31
Manganese ......................................................................................................................... 35
Zinc ..................................................................................................................................... 38
Copper ................................................................................................................................ 41
Boron .................................................................................................................................. 43
Chloride .............................................................................................................................. 47
Silicon ................................................................................................................................. 51
Plant Tissue Analysis .......................................................................................................... 54
Identifying Nutrient Disorders .............................................................................................. 57
Chapter 2 ........................................................................................................................... 61
The Irrigation Water .......................................................................................................... 61
Irrigation Water Quality ....................................................................................................... 62
The Electrical Conductivity .................................................................................................. 67
The pH ................................................................................................................................ 70
Water Alkalinity ................................................................................................................... 72
Hardness ............................................................................................................................ 74
Irrigation Water Analysis ..................................................................................................... 76
The Principle of Electrical Neutrality .................................................................................... 79
Chapter 3 ........................................................................................................................... 82
Soils ................................................................................................................................... 82
Soil Fertility ......................................................................................................................... 83
The Cation Exchange Capacity ........................................................................................... 86
Soil pH and Acidity .............................................................................................................. 90
Soil Salinity ......................................................................................................................... 93
Managing Soil Salinity ......................................................................................................... 96
Soil Sodicity ...................................................................................................................... 101
Soil Organic Matter ........................................................................................................... 106

2
The Soil Analysis .............................................................................................................. 110
Units on the Soil Test Report ............................................................................................ 118
How to Interpret Soil Test Results ..................................................................................... 122
Soil Test Interpretation Guide............................................................................................ 125
Raising Soil pH ................................................................................................................. 128
Quality Parameters of Liming Materials ............................................................................. 131
Soil Water Content ............................................................................................................ 134
Chapter 4 ......................................................................................................................... 137
Fertilizer Management .................................................................................................... 137
Fertilizer Recommendations Philosophies......................................................................... 138
Yield Response to Fertilizers............................................................................................. 141
Calculating Fertilizer Application Rates ............................................................................. 144
Timing of Fertilizer Application .......................................................................................... 148
Pre-plant fertilizer application ............................................................................................ 151
The Ammonium:Nitrate Ratio ............................................................................................ 153
Types of Fertilizers ............................................................................................................ 156
Urea .................................................................................................................................. 159
Compost: Benefits and Quality Parameters....................................................................... 162
Slow-release and controlled-release fertilizers .................................................................. 165
Chelated Micronutrients .................................................................................................... 167
Foliar Fertilization.............................................................................................................. 170
Chapter 5 ......................................................................................................................... 174
Fertigation and Soilless Culture .................................................................................... 174
Fertigation ......................................................................................................................... 175
Hydroponics ...................................................................................................................... 181
Criteria for a Balanced Nutrient Solution ........................................................................... 184
Calculating Nutrient Solution Formulas ............................................................................. 187
Closed Hydroponic Systems ............................................................................................. 191
Fertilizer Solubility and Compatibility ................................................................................. 194
Fertilizer Stock Solutions................................................................................................... 199
Fertilizer Injectors.............................................................................................................. 204
Calibration of Fertilizer Injectors ........................................................................................ 207
Controlling the Irrigation Water pH .................................................................................... 211
Growing Media and Their Properties ................................................................................. 215
In-house Nutrient Monitoring in Container Plants .............................................................. 220
Chapter 6 ......................................................................................................................... 224
Irrigation .......................................................................................................................... 224

3
Water Requirements of Crops ........................................................................................... 225
Irrigation Scheduling Using Soil Water Budget Approach .................................................. 228
Irrigation Scheduling Using Soil Moisture Sensing ............................................................ 231
Principles of Irrigation System Design ............................................................................... 236
Drip irrigation Systems ...................................................................................................... 241
Causes and Prevention of Emitter Clogging ...................................................................... 247
Irrigation Scheduling in Container Plants........................................................................... 251
Variable Rate Irrigation ..................................................................................................... 254
Irrigation with Desalinated Water ...................................................................................... 257
Appendix I: Conversion Tables ...................................................................................... 260
Appendix II: Nutrient Uptake by Crop ............................................................................ 262

4
Chapter 1

The Essential
Nutrients

5
Plant Nutrients - Introduction
Essential plant nutrients are elements that plants need for proper growth. Sixteen
elements are considered essential nutrients for plants. These are carbon (C), oxygen
(O), hydrogen (H), nitrogen (N), phosphorus (P), potassium (K), calcium (Ca),
magnesium (Mg), sulfur (S), iron (Fe), manganese (Mn), zinc (Zn), copper (Cu), boron
(B), molybdenum (Mo) and chlorine (Cl).
Plants absorb carbon and oxygen from the air through their leaves, as carbon dioxide
(CO2). In the photosynthesis process, they transform carbon dioxide and water into
hydrogen, carbon and oxygen. All other nutrients are absorbed through their root
system.
Plants from the Legumes family can use atmospheric nitrogen. They form a symbiotic
relationship with specific bacteria that convert atmospheric nitrogen into ammonia and
then into ammonium, which the plant can absorb. This process is called ‘nitrogen
fixation’.
The essential plant nutrients can be categorized as macronutrients, secondary
nutrients and micronutrients. This classification is based on the relative requirement
by the plant.
Macronutrients are required in relatively large quantities. Secondary nutrients are
required is lesser amounts and micronutrients are required in very small amounts. This
does not imply that micronutrients are less important to the plant. A deficiency of one
micronutrient can limit the growth of the crop to the same extent as a deficiency in
macronutrients do.

• Macronutrients include nitrogen, phosphorus, potassium, carbon, hydrogen,


oxygen
• Secondary nutrients – calcium, magnesium, sulfur
• Micronutrients – boron, iron, manganese, zinc, copper, molybdenum, chlorine

6
Common average nutrient requirements of crops are in the following ranges:
Macronutrients Common Secondary Common Micronutrients Common daily
daily uptake nutrients daily uptake uptake
Nitrogen (N) 1.5-4 kg/ha Calcium (Ca) 0.5-1.5 Iron (Fe) 20-50 g/ha
kg/ha
Phosphorus (as 0.3-0.7 Magnesium 0.2-0.5 Manganese 5-20 g/ha
P2O5) kg/ha (Mg) kg/ha (Mn)
Potassium (K) 1.5-5 kg/ha Sulfur (SO4) 0.2-0.5 Zinc (Zn) 5-10 g/ha
kg/ha
Copper (Cu) 2-8 g/ha

Nutrient uptake by plants


Plants can absorb specific ionic forms of the nutrients, as described in the table below.
In that respect, nitrogen is unique, as it can be absorbed either as an anion (NO 3-) or
a cation (NH4+). The two nitrogen forms are very different in their metabolism within
the plant and in their effect on the root system environment.

Nutrient Form/s in which it is Nutrient form name


absorbed by plants
Nitrogen (N) NO3- Nitrate
NH4+ Ammonium
Phosphorus (P) H2PO4- Dihydrogen phosphate
HPO42- Hydrogen phosphate
Potassium (K) K+ Potassium
Calcium (Ca) Ca2+ Calcium
Magnesium (Mg) Mg2+ Magnesium
Sulfur (S) SO42- Sulphate
Boron (B) H3BO3 Boric acid
Iron (Fe) Fe2+ Ferric
Fe3+ Ferrous
Manganese (Mn) Mn2+ Manganese
Zinc (Zn) Zn2+ Zinc
Copper (Cu) Cu2+ Cupric ion
Molybdenum (Mo) MoO22+ Molybdate

7
Nutrient uptake by growth stages
Plants absorb nutrients in different rates throughout their development cycle.
Generally, uptake rate is lower at the beginning of the growth cycle, increases during
fruit development and drops just before harvest. Furthermore, uptake rates of
individual nutrients vary along the growth cycle.
For example, plants require more nitrogen during the establishment and vegetative
growth stages, while potassium is required in greater amounts during the fruit set
period.
Uptake

Days from planting

N P2O5 K2O

Example nutrient uptake curve

Nutrient availability
Not all nutrients that are present in soil are available for plants. In fact, most of the
nutrients in soil are locked up in minerals or in organic matter and only a small fraction
becomes available for plant uptake.
Plant roots can absorb nutrients only from aqueous solutions. Therefore, in order for
a plant nutrient to become available to the plant, it must first be “delivered” into the soil
solution.
The processes that are responsible for that are:
1. Dissolution of soil minerals.
2. Chemical equilibrium between nutrients that are adsorbed to soil particles and
the soil solution (exchangeable cations).
3. Mineralization
Soil pH affects all these processes, and therefore plays a major role in nutrient
availability. Additional factors that affect nutrient availability include specific bacteria
that mineralize nitrogen and phosphorus, balance /between nutrients in the soil and
more.

8
Nitrogen
Nitrogen is an essential macronutrient, which all plants require for proper growth. It is
an important constituent of the chlorophyll molecule, nucleic acids, and proteins.
Nitrogen is abundant in the atmosphere and comprises about 78% of its content.
Nitrogen is also a constituent of organic matter in the soil. However, atmospheric, and
organic nitrogen cannot be directly used by plants. For plants to absorb nitrogen, it
must be first converted into ammonium (NH4+) and/or nitrate (NO3-), which are the
forms that are available for plant uptake.

How does nitrogen become available to plants?


The processes in which nitrogen is converted to available forms are fixation,
decomposition, mineralization, and nitrification.
Fixation - atmospheric nitrogen is absorbed by nitrogen-fixing soil bacteria which
convert the nitrogen into ammonia, using an enzyme called nitrogenase. The ammonia
is then converted to ammonium (NH4), which can be used by plants.
Decomposition and mineralization are biological processes carried out by soil
microorganisms, in which organic nitrogen is converted into inorganic form - ammonia
and ammonium.
R-NH2 → NH3 → NH4+
The process and its rate are affected by soil temperature and moisture. Warm soil
temperatures (20-35ºC / 68-95ºF) and moist, but aerated soils favor decomposition
and mineralization.
Nitrification - Under aerobic condition and warm temperatures, ammonium is further
oxidized to nitrate (NO3) in a process called nitrification.
Nitrification involves two reactions:
2NH4 +3O2 → 2HNO2 + 2H+ + 2H2O (carried out by Nitrosomonas).
2HNO2 + O2 → 2NO3- + 2H+ (carried out by Nitrobacter).

The processes that cause nitrogen losses from soil are leaching, denitrification, and
volatilization.
Denitrification is an anaerobic process in which bacteria reduce nitrate to nitrogen
gases, N2O and N2 that are lost to the atmosphere. This process occurs mainly in
oversaturated soils.

9
NO3- → NO2- (nitrite) → NO (nitric oxide) → N2O (nitrous oxide) → N2 (nitrogen gas)

Leaching – The nitrate form (NO3-) of nitrogen carries a negative charge. Therefore,
it does not bind to soil particles, but rather moves with soil water. When irrigation water
reach below the root zone of the plant, nitrogen might leach and become unavailable
for the crop. In addition, excess nitrogen and excess irrigation may cause severe
environmental impact, as nitrates might reach ground water.
Volatilization – Under certain soil conditions, surface-applied urea, manure, and other
ammonium-forming fertilizers, convert into ammonia gas (NH3) and volatilize. Soil
conditions that increase volatilization include wet soils, high soil pH and increased soil
temperature.

The nitrogen cycle

Nitrogen uptake by plants


As mentioned above, plants absorb nitrogen mainly as ammonium and nitrate. A
mixture of both forms is usually beneficial.
These two nitrogen forms differ in their metabolism in the plant, in which they are
converted to amino acid. Ammonium is metabolized in the roots and requires more
oxygen, while the metabolism of nitrate takes place in the leaves.

10
In addition, the uptake of ammonium and nitrate affects differently the root
environment and uptake of other nutrients. For example, chlorides compete with
nitrate on uptake, since both carry a negative charge. In the same way, potassium and
other positively charged nutrients compete with ammonium.

Nitrogen deficiency and excess


Nitrogen deficient plants exhibit poor growth. Older leaves become pale green and
smaller, as a result of reduced chlorophyll content. At a more advanced stage of the
deficiency the entire plant becomes yellow and leaves.
Excess of nitrogen promotes excessive vegetative growth, while flowering and fruit set
may delay. This results in decreased yield.

Nitrogen fertilizers
Sources of nitrogen applied in agriculture include both mineral fertilizers, as well as
organic fertilizers.

Organic nitrogen sources


Organic nitrogen fertilizers include manure, compost and other organic products, such
as blood meal, bone meal and seaweed.
Manure is organic matter derived from animal feces. For example, Guano is made of
excrements of seabirds and bats. A typical nitrogen content in Guano is about 10-16%
N.
Livestock manure typically contains 0.5-2.2% nitrogen, depending on the animal from
which it was obtained and its diet. Note that nitrogen content may be presented on a
dry weight basis or on a fresh weight basis.
Using fresh manure increases the risk of contamination of the produce as a result of
pathogens that might be present in the manure.
Compost is a decomposed organic matter, which can include plants, manure,
eggshells etc. It is more stable than manure, releases nitrogen and other nutrients
more slowly over time, improves soil structure and can suppress plant diseases and
pests.

Mineral nitrogen sources


Mineral nitrogen fertilizers contain high concentration of available nitrogen in the form
of ammonium, nitrate and urea. Fertilizers may contain one or all of these forms. The

11
fertilizer may contain only nitrogen, or nitrogen in combination with other nutrients,
such as potassium, calcium, sulfate and phosphate.
The best fertilizer source of nitrogen for a crop will depend on various factors, such as
soil properties, temperature and growth stage of the crop.
Examples of nitrogen fertilizers:
Anhydrous ammonia

• Formula: NH3
• Composition: 82% nitrogen.
• Pressurized gas, there must be injected into the soil.

Ammonium nitrate

• Formula: NH4NO3
• State: Solid
• Composition: 17.5% N-NH4, 17.5% N-NO3 (34% total nitrogen)
• Available also as a liquid, at lower concentration
• Solubility: 1920 g/liter at 20ºC

Urea-ammonium nitrate (UAN)

• Formula: a solution of urea CO(NH2)2 and ammonium nitrate NH4NO3


• Composition: 28-32% Nitrogen
• State: liquid

Urea

• Formula: CO(NH2)2
• Composition: Solid 46% Nitrogen
• State: Solid
• Solubility: 1200 g/liter at 20ºC
• Available also as a liquid, at lower concentrations

Ammonium sulfate

• Formula: (NH4)2SO4
• Composition: 21% Nitrogen, 24% S
• State: Solid

12
• Solubility: 750 g/liter at 20ºC

Mono ammonium phosphate (MAP)

• Formula: NH4H2PO4
• Composition: 12% nitrogen, 61% phosphorus as P2O5
• State: Solid
• Solubility: 364 g/liter at 20ºC

13
Potassium
Potassium is an essential plant nutrient, one of the three macro-elements – nitrogen,
potassium and phosphorus (NPK), as it is absorbed by plants in relatively large
quantities.
Potassium enhances crop yields and quality in different ways. For example, it
increases sugar content in fruits, size of vegetable crop fruits, protein content in
cereals, helps maintaining longer shelf life, improves the plant resistance to diseases
and to drought and more.
How does potassium do that?

Roles of potassium in plants


Potassium is involved in many processes in plants, from water regulation, through
production of energy:

• Regulates stomata opening/closing – in order to open the stomata,


potassium is actively pumped into the guard cells (the cells that surround the
stomata). This reduces the osmotic potential inside the cell and water enters.
Stomata close when potassium is pumped out of the guard cells.
• Influences the photosynthesis process and respiration:
o Potassium affects gas exchange (CO2 and O2) with the atmosphere, by
regulating opening and closing of the stomata.
o Potassium is involved in the synthesis of ATP (Adenosine triphosphate),
which all cells use for energy.
• Regulates and improves water uptake - potassium that accumulates in root
cells results in water entering the root.
• Activates enzymes - potassium is needed for the activation of many enzymes.
It changes the three-dimensional structure of the enzymes and, as a result, their
rate of reaction and affinity for the substrate increase.
• Required for protein metabolism. When there is no sufficient potassium
supply, protein synthesis stops.
• Required for proper uptake and use of other nutrients, such as nitrate (NO 3-),
which is required for protein synthesis. Potassium accompanies nitrate, as a
counter-ion, as it translocates within the plant.
• Strengthen cell walls.

14
Potassium deficiency in plants
Deficiency symptoms may vary among crops. However, the most common visual
symptoms of potassium deficiency include scorching and yellowing of leaf edges,
while the inner side of the leaf remains green. Leaf edges eventually become brown
and die.
Other potassium deficiency symptoms include:

• Smaller leaves.
• Poor crop yield
• Poor yield quality – size, uniformity, sugar content, protein content etc.
• Shorter shelf life.
• The crop might be more susceptible to diseases.

Potassium availability
Potassium is absorbed by plants as K+.
The availability of potassium for plants is mainly dependent on soil composition and
properties and on cultural practices.
Heavy, clay soils, have a higher cation exchange capacity (CEC) and, therefore, retain
more available potassium than light, sandy soils.
Acidic soils also have lower CEC, because H+ ions occupy the exchange sites on soil
clay particles. As a result, there is less potassium available for the plants.

Potassium in Soil
Potassium forms in soil can be classified to four categories:

• Mineral, or structural potassium


• Non-exchangeable, fixed potassium
• Exchangeable potassium
• Potassium in soil solution

The basis for the classification is potassium availability for absorption. Depending on
the type of soil and environmental conditions, potassium availability may vary.

Structural potassium – Potassium is a constituent of soil minerals and is bonded


within the crystalline structure of minerals, such as feldspars, clay minerals and micas
in the soil. Structural potassium is not available to plants. However, small quantities
are slowly released to the soil solution by long chemical weathering processes.

15
Fixed potassium – Potassium can become trapped (“fixed”) between the structural
layers of clay minerals layers. This occurs mainly during wetting and drying cycles of
the soil. Small quantities of fixed potassium can become slowly available to plants
throughout the growing season, as some of the fixed potassium is in equilibrium with
the exchangeable potassium.

Exchangeable potassium – Since potassium is a positively charged ion, it can be


adsorbed on the negatively charged surfaces of clay minerals and organic matter. The
exchangeable potassium is in equilibrium with the soil solution. When potassium is
depleted from the soil solution, as a result of plant uptake, it can be easily replenished
by the exchangeable potassium. Therefore, this form of potassium is readily available
for plants and is considered the most important pool of potassium.

Potassium in the soil solution – The soil solution is the immediate pool of nutrients
for plants. Potassium dissolved in the soil solution is, therefore, readily available for
plants. However, this pool of potassium is small and does not usually represent the
amount of potassium available to plants.

Fractions of potassium in soil

Potassium fertilizers
There are various types of potassium fertilizers available. They are all soluble,
however, some potassium fertilizers may contain insoluble compounds, such as iron
oxide.

16
Potassium chloride (Muriate of potash, MOP):

• Formula: KCl
• Composition: 60% potassium as K2O (50% K) and 45% Cl-.
• A highly soluble potassium fertilizer.
• Solubility ranges from 275 g/liter at 30°C and 229 g/liter at 5°C.
• Should not be applied to crops that chloride-sensitive or to seeds.
• Most economic source of potassium for plants.
Potassium nitrate:

• Formula: KNO3
• Composition: 13% nitrate nitrogen and 46% potassium as K 2O (38% K).
• Very soluble.
• Solubility ranges from 458 g/liter at 30°C and 133 g/liter at 5°C.
• Serves also as a source of nitrogen.
• Has a relatively high cost.
• Used mainly for greenhouse crops and in hydroponics.

Potassium sulfate (Sulfate of potash, SOP):

• Formula: K2SO4
• Composition: 52% potassium as K20 (43% K) and 54% SO42- (18% S).
• Has a relatively low solubility.
• Solubility ranges from 120 g/liter at 25°C and 80 g/liter at 5°C.
• Mainly used for chloride-sensitive crops and when sulfur fertilization is required.
Mono potassium phosphate (MKP):

• Formula: KH2PO4
• Composition: 34% potassium as K20 (28% K) and 52% phosphorus as P2O5
(22.5% P).
• Solubility ranges from 300 g/liter at 25°C and 110 g/liter at 5°C.

Additional types of fertilizers that contain potassium are available, mainly compound
fertilizers that are composed of the above straight fertilizers. These fertilizers contain
three elements and more.

17
Phosphorus
Phosphorus is an essential macro-element. It is a component of certain enzymes and
proteins, as well as of ATP (Adenosine Tri Phosphate) and nucleic acids – DNA and
RNA.

Phosphorus enhances root growth, improves flower formation, increases the


resistance of the plant to environmental stress and improves overall crop quality. It
participates in metabolic processes, like the photosynthesis, energy transfer (ATP is
a molecule responsible for storing and transferring energy in plant cells), and
degradation carbohydrates.

The amount of phosphorus available for plants is very low, compared to the total
amount of phosphorus in the soil. Therefore, in many cases, application of phosphorus
fertilizers is required in order to meet crop requirements.

Phosphorous in soil
Phosphorus is not mobile in soil. It is found in the soil in organic matter and in minerals.
Mineral forms include calcium phosphates, iron and aluminum phosphates. Organic
phosphorus is found in microbes, plant residues and humus. Organic forms include
inositol phosphates, phospholipids and nucleic acids.

Plants can only take up phosphorus dissolved in the soil solution. Most of the
phosphorus in soil phosphorus is found in minerals and organic matter and only a
small fraction of the total phosphorus in soil is available plants.

Phosphorus found in the solid phase of the soil is in an equilibrium with phosphorus in
the soil solution.

When plant roots take up phosphorus from the soil solution, phosphorus of the solid
phase is released into the soil solution in order to maintain this equilibrium.

The types of phosphorus minerals in soil are mostly determined by soil pH and by the
type and amounts of other elements.

In low soil pH, phosphorus is fixed by aluminum, iron and manganese become, while
in alkaline soils it reacts with calcium. The ideal pH range for phosphorus availability
is between 6.0 and 7.5.

18
In many soils, decomposition of organic material and crop residue contributes to
available phosphorus in the soil.

The phosphorus cycle

Phosphorus uptake by plants


Plants can absorb phosphorus from the soil solution as either HPO42- or H2PO4-. The
proportion of these in the soil solution is governed by the soil pH. As can be noted
from the curve below, HPO42- is predominant at a pH range of 7.0 to 10.0, while H2PO4-
, which is the more readily absorbed form, is predominant at pH of between 5.0 and
6.0.

Phosphorus does not readily move through the soil and, in order to absorb it, plant
roots must come into contact with it. Therefore, phosphorus fertilizers must be applied
to the root system of the crop.

19
The concentration of phosphorus in the soil solution is low, as it tends to react with
other elements in the soil, such as calcium. Uptake of phosphorus by plants is an
active process. It is absorbed in diffusion, against the concentration gradient, and
requires energy. Conditions that

Active uptake is an energy consuming process, so conditions that impede root activity,
such as low temperatures, lack of oxygen etc., also restrict phosphorus uptake.

Source: Chemical Equilibria in Soils, Willard L. Lindsay

Phosphorus deficiency
Symptoms of phosphorus deficiency first occur on lower leaves. Leaves may develop
dark green or purplish color (result of accumulation anthocyanin pigments), starting
from the edges of the leaves. This symptom typically occurs in the earlier growth
stages of the crop, when root system is not yet developed. Deficiency is more common
under conditions cool, wet soils, soils with pH <5.5 and soils low in organic matter.

Additional symptoms may include poor root development, thin stems and narrow
leaves. Proper phosphorus levels in plant tissue ranges from 0.2%-0.8%, on a dry
matter basis.

Excess phosphorus
Excess phosphorus in soil may interfere in the availability of other nutrients, such as
zinc, iron and manganese. Furthermore, excessive applications of phosphorus may
result in contamination of surface water runoff, which can cause eutrophication – an
unwanted growth of algae and aquatic weeds in the water, which restricts its use for
fisheries, drinking, industry etc.

Many growers tend to apply high rates of phosphorus, sometimes without intending to
do so. Excess phosphorus is typically a result of:

20
• Excessive applications of fertilizers that contain phosphorus, along with other
nutrients
• Repeated applications of animal manure
• Application of phosphoric acid to the irrigation water in order to reduce its pH.
Often, the phosphorus applied as a result of acidifying the water is mistakenly
not taken into consideration in the fertility program.

Common Phosphorus fertilizers


Single superphosphate

Formula: Ca(H2PO4)2 + 2CaSO4


• Composition: 16-20% P as P2O5, 18-21% Ca, 11-12% S
• Solubility: 90% soluble in water
• Suitability for fertigation: not suitable

Triple superphosphate

• Formula: Ca(H2PO4)2*H2O
• Composition: 44-52% P as P2O5, 15% Ca
• Solubility: >90%
• Suitability for fertigation: not suitable

Monoammonium phosphate
• Formula: NH4H2PO4
• Composition: 61% P as P2O5, 12% N-NH3
• Solubility: 410 g/liter at 25°C
• A highly soluble fertilizer, suitable for fertigation

Diammonium phosphate
• Formula: (NH2)HPO4
• Composition: 46% P as P2O5, 18% N-NH4
• Solubility: 588 g/liter at 20°C
• A highly soluble fertilizer, suitable for fertigation

21
Calcium
Calcium is an essential plant nutrient and has many vital roles:

• Calcium strengthen cell walls by forming calcium pectate compounds that


stabilize cell wall structures and holds adjacent cells together.
• Participates in enzymatic and hormonal processes.
• Helps in protecting the plant against heat stress - calcium improves stomata
function and participates in induction of heat shock proteins.
• Helps in protecting the plant against diseases - numerous fungi and bacteria
secret enzymes which impair plant cell wall. Stronger Cell walls, induced by
calcium, can avoid the invasion.
• Regulates growth and elongation of cells.
• Affects fruit quality.
• Regulates stomata opening and closing.

Calcium is absorbed by plants as Ca2+ and is transported within the plant in the xylem,
along with water. Uptake is passive and does not require energy. Therefore, uptake
rate is directly related to the transpiration rate.

Calcium in soil
Calcium is a positively charged ion, with a charge of 2+. Its content in soil is largely
influenced by the parent material. Calcium is a constituent of many insoluble minerals.
It readily reacts with other elements in soil, such as phosphorus and sulfate. Insoluble
calcium forms are not available for plant uptake.

Due to its positive charge, calcium readily adsorbs to the surfaces of colloidal clay
minerals and organic matter, that have a negative charge and form the exchange
complex of the soil. Calcium adsorbed on the exchange complex is referred to as
“exchangeable calcium” and is in equilibrium with the soil solution. This form of calcium
is, therefore, considered to be available for plant uptake. The level of exchangeable
calcium is evaluated and reported in soil tests for fertility.

The availability of calcium to plants is determined by factors such as the cation


exchange capacity of the soil (CEC), soil pH and the presence of other elements that
react or compete with calcium.

Soils with high pH usually contain more calcium than acidic soils, as in acidic soils
hydrogen ions replace the calcium ions adsorbed on the exchange complex of the soil.
In the same way, high levels of calcium ions replace hydrogen ions on the exchange
complex.

22
The cation exchange capacity of the soil (CEC), describes the amount of negatively
charged sites on the exchange complex. Soils that have high CEC (higher percentage
of clay minerals and organic matter) can retain more exchangeable calcium than soils
with lower CEC. In most soils, calcium occupies between 25 and 70% of the cation
exchange capacity.

Other positively charged ions, such as potassium (K+), magnesium (Mg2+) and sodium
(Na+) compete with calcium both for uptake by plants as well as for adsorption to the
exchange sites. Therefore, excess of those ions might restrict calcium uptake.
Furthermore, exchangeable calcium improves soil structure, while excess sodium
might result in destruction of the soil structure and cause soil infiltration problems.

The optimal pH range for calcium availability is 7.0 to 8.5. However, at soil pH greater
than 6.0, and when phosphorus is in excess, calcium tends to react with the
phosphorus and form insoluble minerals. The availability of both calcium and
phosphorus might then be reduced.

Calcium deficiency
Calcium deficiencies that are related to low levels of calcium in the soil are more likely
to occur in sandy, acidic soils. Calcium deficiency may also occur in soils containing
sufficient amounts of calcium. Because calcium uptake and distribution within the plant
is with the water flow, conditions that restrict water uptake will also affect the uptake
of calcium.

For example:
• Low evapotranspiration rates due to cloudy conditions, low temperatures or
high humidity.
• Water stress
• Salinity stress – salinity reduces the ability of the plant to take up water

Calcium is not mobile within the plant and cannot be remobilized from old to young
tissue via the phloem. Therefore, deficiency symptoms occur in young leaves and
in fruits. Fruits have a low transpiration rates and, therefore, are more prone to
calcium deficiencies.

Symptoms of calcium deficiency may vary between plants and may include stunted
growth, curling of younger leaves, tip burn, chlorosis, discoloration of fruits, fruits that
ripen prematurely, rotten, water-soaked areas on bottom of the fruit (blossom end rot),
dark spots on the fruit (bitter pit in apples) and stunted roots.

23
Blossom end rot in tomato Calcium deficiency in cucumber

24
Magnesium
Magnesium is an essential plant nutrient, classified as a secondary nutrient. It is an
important constituent of the chlorophyll molecule, which is a green pigment that gives
plants their green color. The chlorophyll plays an important role in the photosynthesis
process. It absorbs light energy that is then used to convert carbon dioxide and water
into sugar, named glucose, and oxygen. The sugar is used as an available energy
source for the plant.

Magnesium in soil
There are three pools of magnesium in the soil:
1. Unavailable, non-exchangeable magnesium - Magnesium that is found primary
and secondary minerals in the soil is not available for plants. It can become
available when these minerals break down and dissolve into the soil solution.
However, this is a very slow process and, therefore non-exchangeable
magnesium is considered to be unavailable for plants and is not determined in
fertility soil tests.
The principal primary minerals that contain magnesium include olivine,
pyroxene and micas. Secondary magnesium minerals include minerals such as
magnesite, dolomite and chlorites.

2. Exchangeable magnesium – Magnesium ions have a positive charge of 2+.


Therefore, it can be held by the cation exchange sites of the soil, which carry a
negative charge, i.e., on the surfaces of clay minerals and soil organic matter.
Exchangeable magnesium is in equilibrium with the soil solution and is available
for plant uptake. The availability of magnesium in the exchange complex of the
soil affects the concentration of magnesium in the soil solution and determines
the ability of the soil to replenish magnesium that was depleted from it. Soil
fertility analysis determines this fraction of the magnesium in soil.

3. Magnesium in soil solution – Free magnesium ions in the soil solution are
readily available for plant uptake. Soil salinity analysis methods, such as
saturated paste extract, determine are used to determine this form of
magnesium.

25
Magnesium availability
The availability of magnesium in soil depends on several conditions. Magnesium
availability is reduced in soils with low cation exchange capacity, in acidic soils and in
the presence of competing cations:

Soil cation exchange capacity (CEC) – soils with low CEC, such as sandy soils, have
less cation exchange sites and cannot retain much magnesium, as well as other
cations.

Soil pH – Acidic soils have more free hydrogen ions. In such soils, aluminum becomes
soluble and is released from soil minerals. The hydrogen and aluminum ions replace
magnesium on the cation exchange sites. Since magnesium has a high hydrated
radius, it is held loosely on the exchange sites of the soil and can be easily replaced
by hydrogen and aluminum ions. As a result, magnesium is released to the soil solution
and becomes susceptible to leaching.

Competing cations – High levels of competing cations such as calcium, potassium and
sodium may interfere with magnesium uptake and reduce its availability. Under low
soil pH conditions, aluminum and manganese become highly available and reduce the
availability of magnesium to plants.

Magnesium uptake by plants


Magnesium is absorbed by plants as Mg+2, which is the ionic form of magnesium in
the soil solution. Due to the to the fact that it can be transported in the phloem tissue,

26
its mobility within the plant is high and it can be translocated from old to young plant
tissues.

Magnesium reaches the roots both by mass flow, which is a passive process, and
diffusion. Mass flow is associated with the transpiration rate, while diffusion occurs
due to concentration gradient of magnesium in the soil solution, where magnesium
moves from areas of high concentration to areas of low concentration.

Magnesium deficiencies
Magnesium deficiency reduces crop yields and quality. Due to its high mobility within
the plant, magnesium deficiency symptoms appear on older leaves first. Magnesium
deficient plants tend to have a pale color. As the deficiency progresses, older leaves
may develop an interveinal chlorosis. In some crops, reddish, purple or brown lesions
will develop on the leaves. In severe deficiency, the margins of older leaves may turn
brown/reddish. The necrosis moves inward, and leaves may drop.

It was found that light intensity influences the severity and extent of magnesium
deficiency symptoms. High light intensity enhances chlorosis and necrosis in
magnesium deficient plants.

Magnesium deficiency symptoms in tobacco. Magnesium deficiency in pecan.


Photo by R.J. Reynolds Tobacco Company, Photo by Jonas Janner Hamann,
R.J. Reynolds Tobacco Company, Bugwood.org Universidade Federal de Santa Maria
(UFSM), Bugwood.org

27
Sulfur
Sulfur is an essential nutrient for plants. It is considered a secondary nutrient.
However, sulfur requirement of many crops is relatively high, similar to phosphorus. It
was found to be one of the major limiting nutrients for crop production. Therefore,
some consider it a macronutrient. Proper sulfur concentration in plant tissue is
between 0.2% and 0.6%.

Roles of sulfur in plants


Sulfur plays some vital roles in plants:

• It is a constituent of methionine and cysteine, amino acids that are components


of many proteins. These amino acids comprise up to 90% of the total sulfur in
plants.
• Iron-sulfur protein clusters (Fe-S) have a critical role in the photosynthesis
process, respiration, vitamin synthesis and nitrogen fixations in legumes.
Sulfur is essential for the metabolism of Rubisco (Ribulose-1,5-bisphosphate
carboxylase/oxygenase), an enzyme involved in atmospheric carbon fixation.
• Improves the oil content of seeds.

Sulfur in soil
Soil organic matter is the main reservoir of sulfur in soils. In fact, organic matter
contributes between 90% to 95% of the total sulfur in soil. In order to become available
to plants, sulfur must go through a mineralization process, in which sulfur is converted
into sulfate (SO42-), which is the plant available form.
Mineralization is driven by bacteria. Factors such as soil moisture, temperature
(optimum temperature of 30˚C), clay content and pH also affect the mineralization
process, as they directly affect the microbial population and organic matter
decomposition.
This process is highly dependent on the C:S and the C:N ratios in the decomposing
organic matter. High C:S ratio will result in a slow decomposition rate or in
immobilization of sulfur, as there is not enough sulfur in the organic matter to satisfy
microbial needs.
The dependency of sulfur mineralization on the C:N ratio is related to the fact that S
and N are combined in the organic matter and, therefore, their mineralization
processes depend on each other
Because of its negative charge, sulfate, the inorganic (mineral) form of sulfur, is highly
mobile in soil.

28
Immobilization of sulfur is a microbial-induced process in which mineral sulfur (i.e.
sulfate, SO42) is converted back to organic sulfur. Generally, a C:S ratio < 400 will
result in immobilization of sulfur.

Sulfur deficiencies in plants


Sulfur deficiencies will be more inclined to occur in sandy soil with 2% organic matter
or lower. It may also be prevalent in areas with heavy rainfalls.
Because of its high mobility in the soil, topsoil often has lower amounts of sulfur, while
greater levels are found at lower soil depths. Therefore, deficiencies may occur early
in the growth cycle of the crop, before the root system is fully developed.
Sulfur is immobile in plants and does not readily translocate from older leaves to
younger leaves. Therefore, sulfur deficiency symptoms tend to appear on younger
leaves first.
Common symptoms include a uniform yellowing or pale green coloring of the whole
plant. Chlorosis of younger leaves with necrotic tips may also occur. However, some
symptoms may often be crop specific and include symptoms such as leaf spotting
(potato), interveinal chlorosis of younger leaves (corn).
In some cases, sulfur deficiency might be difficult to identify, as it might resemble
nitrogen deficiency.
Plant tissue analysis is the best approach to diagnosing sulfur deficiencies.

Sulfur deficiency in peas


Photo by Mary Burrows, Montana State University, Bugwood.org

29
Sulfur fertilizers
Application of sulfur as a nutrient in the fertilization program is often overlooked. There
are several types of sulfur fertilizers. The most common ones are ammonium sulfate
(24% S), calcium sulfate (gypsum, 14-18% S), potassium sulfate (18% S),
superphosphate (12% S) and magnesium sulfate (22% S).
Fertilizers should be applied based on soil conditions and tissue analysis.

30
Iron
Iron is required for the biosynthesis of the chlorophyll molecule and functions as an
electron carrier in the respiration and photosynthesis reactions. In addition, it
participates in many enzymatic processes. Iron deficiency is a limiting factor of plant
growth.

Iron is present at high quantities in soils, but its availability to plants is usually very low,
and therefore iron deficiency is a common problem.

Iron in soil
In soils, iron oxides are formed as a result of weathering of iron-containing minerals,
such as olivine, pyroxene, and biotite. Most of the iron oxides in soil are in the form of
Fe3+ (ferric iron), which are much less soluble than Fe2+ (ferrous iron). Ferrous iron
readily oxidizes to the ferric form and precipitates out of the solution.

The solubility of Fe3+ is highly depended on soil pH. At neutral and high soil pH levels,
iron becomes insoluble and therefore, although iron is abundant in many soils, its
availability to plants is very low in soils with high pH, such as calcareous soils.

Organic matter and the activity of microorganisms in soils may improve iron
availability. Interactions between organic matter and microorganisms with iron
minerals form soluble iron compounds over a wide range of pH. Crops growing in low
organic matter soils are more susceptible to iron deficiencies.

Iron uptake by plants


Plants can absorb iron both in the ferric and ferrous forms, as well as in the chelated
form.

Iron uptake is a metabolically regulated process. Two iron uptake mechanisms were
evolved in plants: proton release mechanism and Fe3+ chelate reduction. The
purpose of both mechanisms is to make iron more soluble and available for uptake.

Proton release mechanism – Under iron deficiency conditions, plant roots release
protons into the soil solution, which reduces the pH the immediate root surrounding
and increase iron availability.

Research shows that the source of nitrogen fertilizer used can also affect iron
availability for plants. Ammonium (NH4+) is a positively charged form of nitrogen.

31
Uptake of ammonium leads to a decrease of the pH in the rhizosphere (root
surrounding). As a result, iron becomes more soluble and available to the plant.

On the contrary, the pH of the rhizosphere may increase when nitrate nitrogen (NO3-)
is applied, as roots release hydroxide ions. As a result, iron availability decreases.

Chelate reduction – In iron-deficient soils, gramineous crops (e.g., corn, rice, wheat
etc.) secrete organic substances, named phytosiderophores, which act as chelating
agents. The phytosiderophores bind iron and make it soluble and available for uptake.
Chelate reduction can also be carried out by certain bacteria in the soil.

Iron uptake is carried out primarily by young roots. Therefore, maintaining a healthy
root system is important for adequate iron uptake.

Iron deficiencies
Iron is not mobile within the plant and, therefore, deficiency symptoms appear on
younger leaves first. Leaves turn yellow and chlorotic, but leaf veins remain green.

Often, iron deficiency is not a direct result of lack of iron in the soil. A variety of soil
conditions may affect its availability for plants. Conditions such as high soil pH, high
soil moisture, low temperatures, high phosphorus and the high concentrations of
competing elements, such as zinc, calcium, and manganese, may reduce the
availability of iron to plants.

Iron deficiency in blackberry. Photo by: Guy Sela

Therefore, applying iron without considering soil conditions might not help in correcting
the deficiency.

32
Foliar application of iron can help correcting iron deficiencies in the short term. A
longer-term solution must include understanding the reason for the deficiency and
correcting the conditions that cause it.

In high-pH soils, one of the most common practices, in addition to foliar applications,
is the application of chelated iron fertilizers, such as Fe-EDTA, Fe-EDDHA and Fe-
EDDHMA. The iron chelates allow iron to be available for plants at a wide range of soil
pH levels.

Iron fertilizers
The most commonly used sources of iron are:

Iron sulfate (FeSO4) – typically contains between 20%-30% iron, depending on the
hydration state. It is the least expensive iron source. Iron sulfate can be applied as a
foliar spray or directly to soil. However, this form of iron is not available for plants at a
pH above 7.0 and it may take a soil applied iron sulfate a few years to correct iron
deficiencies.

Iron chelates. Chelates are compounds in which iron is combined with a compound
that helps keep iron available and avoids its precipitation. Several compounds are
used as chelating agents. For example, EDTA, DTPA, EDDHA, (all three are sodium
salts), amino acid, humic-fulvic acids and citrate.

Chelates differ in their stability and their ability to hold iron at different pH levels. In
addition, the susceptibility of chelates to substitution of the iron by other cations, such
as calcium and magnesium, varies among chelate types, as is shown in the following
table.

Iron chelate pH range at which Affinity to cations Suitable for…


stable
Fe-EDTA Acidic soil
Calcium-poor soil

3.5-6.0 High affinity to calcium Soilless media


Hydroponics
(Where pH monitoring is
more accurate)
Fe-DTPA Alkaline soil
4.0-7.0 Low affinity to calcium
Calcium rich
Fe-EDDHA Alkaline soil
4.0-11.0 (Where it is difficult to
effectively lower pH levels)

33
.
Iron chelate stability at different pH levels

34
Manganese
Manganese is an essential plant micronutrient. It is absorbed by plants as Mn2+.
Manganese is an immobile nutrient and, therefore, deficiency symptoms show up on
younger leaves first.
A manganese level of 20 to 40 ppm (mg kg-) in plant tissue is sufficient for most plants.
Toxicity might occur when manganese tissue levels are greater than 400 ppm.

Roles of manganese in plants


Manganese plays a vital role in various processes:

• Participates in the photosynthesis process


• Participates in chlorophyll production
• Activates enzymes, such as the nitrate-reducing enzyme and carbohydrate
metabolism enzymes.
• Enhances starch production (carbohydrate)
• Induces cell division and elongation
• Has a role in the biosynthesis of fatty acids

Manganese deficiency
Manganese deficiency results in reduced crop yields and quality, mainly due to
impairment of the photosynthesis process and synthesis of starch.
Deficiency symptoms begin as interveinal chlorosis of younger leaves and/or necrotic
spots.
Deficiency occurs mainly in calcareous soils, soils with high pH, soils with high organic
matter content and in poorly aerated soils.
Excess of iron might also cause manganese deficiency, as iron competes with
manganese for uptake.
In calcareous soils, a foliar application is recommended for correcting manganese
deficiency.

35
Manganese deficiency in citrus

Manganese Toxicity
When in excess, manganese damages the photosynthesis process and other
processes, such as enzyme activity. The threshold of manganese toxicity is highly
dependent on the plant species.
Toxicity symptoms include brown spots on mature leaves and chlorotic specks on
young leaves. Toxicity symptoms appear on older leaves first. Symptoms spread from
leaf borders inwards.
Manganese toxicity is major limiting factor in acidic soils.
Soil conditions that favor accumulation of toxic levels of manganese:
Soil pH lower than 5.5
Low calcium levels in the soil
Lack of oxygen as a result of excessive irrigation, poor soil drainage, soil compaction,
high precipitation.

How to treat manganese toxicity


Manganese toxicity can be treated in various ways:

• Application of magnesium
• Application of organic matter
• Application of lime for soils with low pH
• Prevent fluctuations in soil moisture level

36
Manganese in soil and its availability to plants
The reactions of manganese in soil are complex. The two major factors that affect
manganese availability are pH and redox conditions. Other factors include soil organic
matter, microbial activity, soil temperature and seasonal variations affect its availability
to plants.
The most soluble form of manganese is Mn2+. Other oxidation states form low-
solubility compounds, such as MnO2, Mn2O3, Mn3O4.
Soil pH – Solubility of manganese increases at lower soil pH. Manganese is available
in soil pH lower than 7.0. At soil pH lower than 5.5, manganese toxicity might occur.
At a higher soil pH, low-solubility manganese compounds form and manganese
solubility is reduced. Furthermore, at high soil pH, a higher rate of manganese adsorbs
to soil particles and, as a result, its availability to plants decreases.
Soil moisture - Dry soil conditions also decrease manganese availability. On the other
hand, manganese availability increases in waterlogged soils, due to the reduction of
manganese oxides. Rapid change in manganese may occur, depending on the soil
moisture status.
Microorganisms – Redox reactions carried by microorganisms greatly affect
manganese availability to plants.
Soil temperature – Higher soil temperature increases manganese availability, as
manganese is reduced to the Mn2+ soluble form.
Soil organic matter – Organic matter forms complexes with manganese and reduces
its availability. In fact, manganese deficiency is more common in soils with high organic
matter content than in alkaline soils.
It has been shown that plant roots also affect manganese availability by reducing and
releasing Mn+2 from insoluble manganese compounds. However, the mechanisms of
such reactions are not yet fully understood.

Manganese fertilizers
Manganese sulfate – 26-28% Mn.
Manganese oxide – 41-68% Mn. For acidic soils only, as it has a low solubility.
Manganese EDTA – 12-13%.
Manganese chloride – MnCl2

37
Zinc
Zinc (Zn) is an essential micronutrient for plant growth. It is required by plants in small
quantities, yet it is required for proper plant development.
Zinc is an important constituent of several enzymes and proteins. In addition, it is
required for the activation of many enzymes. Zinc plays an important role in metabolic
processes such as internode elongation and synthesis of the auxin growth hormone.
Plant tolerance against environmental stress conditions, such as heat and salt stress,
and to defense mechanisms against pests and diseases, are also promoted by zinc.
The mobility of zinc within the plant is variable and greatly depends on the availability
of zinc in the soil. If the zinc supply is adequate, it can be more readily translocated
from older to younger leaves. However, translocation is restricted under conditions of
zinc deficiency.

Zinc deficiency in bean (photo by Howard F. Schwartz, Colorado State University, Bugwood.org)

Zinc in soil
Zinc deficiency is widespread in many soils and crops. Soil conditions under which
zinc deficiency is most likely to occur include:
• Soil with a high pH level. Availability of zinc is usually low under pH of 6.0.
• Soil with a high level of phosphorus and/or silicon
• Low soil temperature
• Limed soils or calcareous soil
• Anaerobic, waterlogged conditions result in precipitation of zinc minerals.

38
• Extremely low or high organic matter content, such as in peat soils.

Zinc disorders
Zinc deficient soils are widespread worldwide. Zinc deficiency results in reduced crop
yields and causes nutritional health problems to humans. It is estimated that up to 2
billion people globally suffer from zinc deficiency.
Zinc deficiency is common in major food crops, such as wheat, maize and rice, but
also affects vegetables, fruits, and other crops.
Visual symptoms of zinc deficiencies may not always occur. It was shown that zinc
deficiency can result in up to 20% reduction in yield, even before any deficiency
symptoms become visible.
The visual symptoms associated with zinc deficiencies are relatively easy to identify
and may include:
• Distorted leaves
• Stunted growth
• Interveinal chlorosis, while the main veins remain green
• Necrotic spots on the leaves
• Small, narrow leaves that curl upwards
• Reduced internode length
Chlorosis and necrosis often appear on middle leaves, although initial symptoms
appear on younger leaves first.
Soil and tissue analyses are necessary in order to determine whether zinc is deficient.
Without proper testing, it is often difficult identify a hidden deficiency, where no visual
symptoms are present.
Because yield reduction occurs well before visual deficiencies occur, regular soil and
plant tissue testing is recommended.

DTPA-extraction test is one of the most used methods to test soil zinc levels.

Interpretation of zinc levels in plant tissue of various crops

Crop Plant Part Time of sampling Deficient Adequate High


-------------------ppm------------------
Potato Most recently matured leaves 8-10 inches tall plants <30 30-60 >60
Snap bean Most recently matured trifoliate Prior to bloom <20 20-40 >40
Corn Most recently matured leaves 30 inches tall <25 25-40 >40
Tomato Most recently matured leaves First flower <25 30-60 >60
Cotton Most recently matured leaves A week before bloom <20 20-200 >200

39
Zinc toxicity is a rare condition, because most soils either have a deficient or normal
zinc levels. However, toxicity symptoms may occur in acidic soils or in soils that were
contaminated by excessive fertilizers application, pesticides, sewage sludge, mines
and other human activities.
At low soil pH levels, the solubility of zinc increases, and it might reach toxic levels.
Leaf concentrations of above 150 mg/kg of were associated with zinc toxicity in many
crops. Leafy vegetables and legumes are among the crop species that are sensitive
to zinc phytotoxicity.
Toxicity symptoms include stunted growth, reduced yields, browning of roots and
chlorosis in young leaves. The chlorosis is associated with induced deficiency of other
micronutrients.

Correction of zinc deficiencies


Zinc fertilizers can be applied to correct zinc levels in soil. The most commonly used
zinc fertilizers are:
• Zinc sulfate (ZnSO4) – contains 21-36% zinc.
• Zinc chelates (Zn-EDTA, Zn-DTPA) - typically contain 14% zinc
• Zinc oxide (ZnO)- contains70-80% zinc
The most efficient way to correct zinc deficiencies is to use both soil and foliar
applications.

40
Copper
Copper (Cu) is an essential plant nutrient, classified as a micronutrient. It has a role in
various important processes in plants. For example:
• Copper is a structural element in numerous proteins
• Essential for the photosynthesis process - plastocyanin is a copper protein that
acts as an electron carrier in the electron transport chain of photosynthesis.
• Essential for the respiration process
• Activates several enzymes
• Participates in cell wall metabolism

Copper uptake by plants and its availability in soil


Copper exists in soils as Cu2+ and most of the copper is absorbed by the plant as Cu 2+.
Once absorbed, it accumulates mainly in the roots. Its concentration in plant tissue
ranges from 5 to 20 ppm and in soil from 2 to 100 ppm (mg/kg). However, most of the
copper in the soil is not available for plants.
Availability of copper increases in soil pH below 7.0 and decreases at a higher pH,
due to fixation to soil clay minerals. Copper tends to easily bind to organic matter.
Therefore, soil organic matter reduces copper availability to plants and despite their
low pH, copper deficiency might occur on acidic soils if the organic matter content of
the soil is high.
Copper can also be adsorbed to iron and manganese oxides and precipitate with
carbonate and phosphate minerals. These reactions reduce copper availability to
plants.

Copper deficiency symptoms in plants


Copper is an immobile nutrient and, therefore, deficiency symptoms appear on the tips
of young leaves first. Symptoms then extend to leaf margins.
Symptoms vary between different crops and may include:

• Twisted young leaves


• Young leaves may become bluish-green
• Chlorosis between the veins of young leaves
• Compact appearance of the entire plant
• Dropping of mature leaves
• Delayed flowering
• Wilting

41
Copper toxicity in plants
Although copper is an essential micronutrient, excess of copper might be toxic to
plants. It might inhibit plant growth by causing an oxidative damage to cells and
interfering with the photosynthesis process. When in excess, copper may also replace
the magnesium (Mg2+) in the chlorophyll molecule and impair the photosynthesis
process.
In addition to the direct toxicity, excess of copper may also cause antagonistic
interactions with other nutrients. Therefore, excess of copper may result in deficiency
of nutrients such as molybdenum, iron, manganese and zinc.
Copper toxicity in plants depends mainly on plant species and soil conditions. Copper
toxicity in alkaline soils is less likely to occur than in acidic soils.
Toxicity symptoms include:
• Interveinal chlorosis
• Necrosis
• Stunting
• Inhibited root growth
• Inhibited shoot growth

42
Boron
Boron (B) is an essential micronutrient. Therefore, it is required by plants in order to
complete their life cycle.

The average concentration of boron in soil is about 30 mg/kg. However, only 1-5% are
available for plants. Among plant nutrients, the range between boron deficiency and
toxicity is very narrow. For example, for boron extracted with hot water, a concentration
lower than 5 ppm is considered deficient, while concentrations higher than 1.5 ppm
may already become toxic to sensitive crops.

Roles of boron in plants


Boron takes part in several processes in plants, including:

• Participates in cell wall synthesis by forming diester bridges between pectins


• Increases pollen quality
• Participates in transport of sugars and in carbohydrates metabolism
• Participates in RNA synthesis
• Participates in auxin metabolism (hormone that promotes cell elongation)
• Plays a role in the regulation of membrane functions

Boron uptake by plants


Like calcium, boron is taken up by plants along with water. It flows towards the roots
by mass flow and its transport within the plant is governed by transpiration. Boron
uptake is passive, and it moves upward in the plant through the xylem vessels, while
it’s mobility in the phloem is limited. Therefore, boron supply to plants greatly depends
on factors such as transpiration rate and water availability in the soil.

Boron is taken up by plants primarily as boric acid (H3BO3) and to some extent as
borate (H2BO3-). It is the only nutrient that plants absorb as an uncharged molecule,
rather than as an ion. This small molecule easily passes through the root cell
membranes.

Boron in soil and its availability to plants


Most of the boron in the soil is adsorbed on soil particles, mainly on iron and aluminum
oxides, clay minerals, calcium carbonate, magnesium oxides and organic matter.

In the soil solution it exists as boric acid and borate, the plant-available forms.

43
Boron availability to plants is affected by soil-related factors, climate conditions and
agricultural practices.

Soil organic matter - is often the largest pool of boron. Boron is released from soil
organic matter as it decomposes. Boron deficiency occurs mainly in low organic matter
soils.

Drought conditions - decrease boron availability its mass flow towards the roots is
restricted and transpiration rate decreases.

Soil pH – Boron is most available in slightly acidic soil, at a pH range between 5.0-6.5.
Boron fixation increases at high soil pH. In highly acidic soils, boron adsorbs to
aluminum and iron oxides.

Wet, cold soils – reduce mineralization rates as a result of reduced microbial activity.

Excess irrigation or high precipitation – Boron is not strongly held by soil particles and
tends to leach. Boron in soil solution moves readily with water. Therefore, highly
leached soils are often deficient in boron.

Soil texture – coarse-textured soils often contain less boron than fine-textured soils,
because boron can be easily leached to below the root zone in sandy soils.

The irrigation water – groundwater used for irrigation may contain boron. A boron
concentration of 0.2-0.3 ppm is sufficient for most plants. Concentrations of above 0.5
ppm require attention and proper irrigation management, in order to avoid
accumulation of excess boron in the root zone.

Boron disorders in plants


Symptoms of boron deficiency and toxicity vary among plants.

Boron is immobile in most plants and, therefore, deficiency symptoms appear on


younger leaves and terminal buds first.

44
Boron deficiency in strawberry
Photo by Natalia Teixeira Schwab

Symptoms include:
• Yellowish-white chlorosis of the tips of younger leaves
• Short internodes
• Stunted roots
• Reduced pollen tube growth
• Reduced flowering
• Malformed leaves
• Hollow stem

Boron toxicity in apple


Source: University of Georgia Plant Pathology, University of Georgia, Bugwood.org

45
The most common boron toxicity symptoms include brown leaf tips, interveinal necrotic
spots on older leaves, curled leaves, and reduced growth.

Common boron fertilizers

Fertilizer Formulation % boron


Borax Na2B4O7·10H2O 11%

Boric acid H3BO3 17,5%

Solubor Na2B8O13·4H2O 20%

46
Chloride
Chloride is often associated with soil salinity and toxicity symptoms in plants. However,
at the same time, it is an essential element that fulfills many functions in plants. Plants
require small quantities of chloride and absorb it from soil solution as a Cl - ion.
Common chloride range in plant tissue is between 2 and 20 ppm.

Crops differ both in their chloride requirements as well as in their tolerance to chloride
toxicity. Some crops, such as strawberry, avocado and almonds are very sensitive to
chlorides, while others, such as celery, sugar beet and wheat benefit from application
of chloride-containing fertilizers.

Functions of chloride in plants include:


• Regulation of water balance and osmotic regulation of stomatal openings
• Chloride has a role in the photosynthesis process
• There is evidence that it can suppress some plant diseases

When chloride concentration in plant tissue reaches toxic levels, yield and quality of
the crop are adversely affected.

Chloride deficiency in wheat. Photo by Mary Burrows, Montana State University, Bugwood.org

Chloride toxicity
Common toxicity symptoms often start with a marginal leaf chlorosis, that turns into
necrosis and leaf margins become scorched. Growth is reduced and leaves may be

47
smaller. In severe toxicity, leaves might fall off. Symptoms will usually appear on older
leaves first.

Chloride toxicity may resemble other salt-related toxicities, such as sodium toxicity
and, therefore, it might be difficult to distinguish between chloride toxicity and salinity
damage.

Overhead irrigation with water containing high concentration of chloride might also
cause chloride toxicity, as chloride can be also absorbed by plant leaves.

Chloride toxicity in tobacco. Photo by R.J. Reynolds Tobacco Company , R.J. Reynolds Tobacco Company,
Bugwood.org

Chloride sources
Atmospheric deposition is a main source of chloride in coastal areas, as sea spray
adds chloride to the atmosphere. The concentration of chlorides in rainwater
decreases with the distance from sea. Therefore, chloride deficiency may occur mainly
in areas that are far from the sea.

The main natural sources of chlorides in groundwater is seawater intrusion in coastal


areas and rainwater. Groundwater in coastal, arid and semi-arid areas will usually
contain high levels of chlorides. In areas far from the sea, the main source of chlorides
in groundwater is related to human activity, such as application of fertilizers, use of
water softeners etc.

48
Chloride in soil
Chloride concentration in the irrigation water and irrigation management practices
greatly affect chloride concentration in soil.

For example, an irrigation of 20 m3/ha/day, with water containing 50 mg/liter of


chlorides will result in an application of 1 kg/ha/day, or 365 kg/ha of chlorides per year.

kg/ha/year of added chloride = (50 x 20 x 365) / (1000) = 365

Chloride exists in soil as a chloride anion. It does not adsorb to soil components and
does not form insoluble salts. Therefore, its mobility in soil is high and it readily moves
with the water in the soil.

Chloride in irrigation water


Excess chlorides must be leached below the root zone of the crop, depending on the
chloride level that the crop can tolerate.

Chloride moves readily throughout the soil and, therefore, provided that the initial
chloride level in the water is tolerable by the crop, applying a certain amount of excess
water to leach chlorides can help in avoiding its accumulation in the root zone.

For most crops a chloride concentration of up to 140 mg/l in the irrigation water is
manageable. Water containing chloride concentration of less than 70 mg/l of chloride
are safe for most crops, provided that proper irrigation management practices are
applied.

Chloride classification of irrigation water


Chloride (ppm) Effect on Crops
Below 70 Generally safe for all plants.
70-140 Sensitive plants show injury
141-350 Moderately tolerant plants show injury
Above 350 Can cause severe problems

49
Susceptibility ranges for crops to foliar injury from saline sprinkler water
Source: Colorado State University
Cl concentration (mg/L) causing foliar injury
Cl concentration <175 175-350 351-700 >700
Apricot Pepper Alfalfa Sugar beet
Plum Potato Barley Sunflower
Tomato Corn Sorghum

Fertilizers containing chloride


Several fertilizer types contain chloride:
Muriate of potash (KCl) – 47% chloride, 53% potassium
Calcium chloride (CaCl2) – 50% chloride, 28% calcium
Ammonium chloride (NH4Cl) – 66% chloride, 25% nitrogen
Magnesium chloride (MgCl2) – 75% chloride, 25% magnesium

50
Silicon
After oxygen, silicon (Si) is the second most abundant element in earth’s crust and is
a constituent of most minerals. In soil solution, silicon is mainly present as monomeric
silicic acid (H4SiO4), which is also the form in which it is taken up by plants.

Plant tissue contains silicon. The concentration of silicon varies considerably among
plant species, ranging from 0.1 to 10% Si on a dry weight basis.

Silicon is an essential element for plants. Deficiencies can cause various abnormalities
in plant growth. However, it has not been recognized as an essential nutrient for plant
growth, mainly because it is not involved in the metabolism of the plant.

Several benefits of silicon to plants are now recognized. These benefits of silicon vary
among plant species and are expressed mainly under stress conditions.

Such benefits include:

• Increased resistance to pests and diseases


• Increased photosynthesis
• Alleviation of heavy metals toxicity
• Better nutrient balance
• Increased drought and frost tolerance
• Improves soil structure
• Increases the availability of other nutrients, such as Ca, P, S, Zn and Cu

Below are some examples of diseases that are suppressed or have reduced severity
when silicon is applied:

• Blast disease (Magnaporthe grisea) in rice


• Fusarium wilt in banana
• Powdery mildew in barley
• Pythium in cucumber
• Pink rot in melon
• Gray leaf spot in ryegrass
• Anthracnose and angular leaf spot in beans
• Powdery mildew in grapes

Silicon can be applied to the soil or by foliar applications. Foliar applications have been
reported to be effective against powdery mildew in several crops.

51
Enhanced resistance to diseases and pests
The enhanced resistance to diseases is attributed to the accumulation of silicon in the
epidermal system. To effectively infect the plant, the pathogen must penetrate some
physical barriers, such as the cuticle and the epidermis. The silicon layer that is formed
under the cuticle provides another physical barrier against pathogens and strengthen
the plant.

In addition to the physical mechanism, silicon induces the formation and accumulation
of compounds that suppresses plant diseases, such as phenolic compounds and
phytoalexins, and activates defense enzymes.

Silicon can suppress various pests, such as rice leaf folder, stem borers, green
leafhoppers and brown planthoppers, armyworms in corn, whitefly in beans, mites and
more.

The mechanisms in which silicon enhances the resistance to pests are also comprise
of physical and biochemical mechanisms. Silicon hardens and roughens plant tissue,
which makes it more difficult for pests to feed and damage their mouthparts.

In addition, silicon enhances the natural defense mechanisms of the plant against
pests, either direct defense or indirect defense. For example, it was found that silicone
enhances the induced production of chemicals that attract the natural enemies of the
pest (indirect defense).

Effect on nutrient availability


Silicon competes with nutrients such as phosphorus for binding to soil particles and
to iron and manganese. As a result, more phosphorus remains available for plant
uptake. The availability of iron and manganese, however, may decrease.

Tolerance to abiotic stress conditions


Water stress – The accumulation of silicon under the cuticle can reduce transpiration
rate. This may result in better tolerance of plants to drought conditions.

Temperature stress – leakage of electrolytes from plant cells is an indication of a


damage made to the cell membrane. Both heat and freezing stress conditions may
cause such a damage to cell membranes. There is evidence that silicon improves the
stability of the membranes and alleviate the damage caused by temperature stress. In
addition, silicon stimulates the production of antioxidants and heat shock proteins.

Silicon was also found to enhance plant resistance to radiation damage.

52
Salinity stress – Sodium uptake is one of the main causes of salinity stress. High
sodium concentration adversely affects plant metabolism. Silicon reduces sodium
uptake and accumulation and restricts its mobility within the plant.

53
Plant Tissue Analysis
Plant tissue analysis is a quantitative measurement of the nutrient content in plant
tissue, typically leaves and/or petioles. It is performed in the lab and used to diagnose
nutrient disorders at different stages of development of the crop and to give fertilizer
recommendations. Often, nutrient deficiencies can be diagnosed even before visual
symptoms occur.

Tissue analysis can also be useful to diagnose inconsistencies in the field. Samples
from a healthy crop can be compared with samples from sections of the field where
the crop show nutrient deficiency or toxicity symptoms.

Plant tissue analysis provides information about the nutritional status of the crop at the
specific time when samples were taken. Both sampling time and the sampled plant
parts are extremely important, as nutrient concentrations vary between plant parts and
between different growth stages.

Often, the information revealed in plant tissue analysis could not be obtained from soil
analysis. While soil analysis provides information on the availability of nutrients in the
soil, tissue analysis gives an indication of the actual nutritional status of the crop,
hence of nutrient uptake by the plant.

Plant tissue levels depend on various factors. In some situations, such as occurrence
plant disease, certain weather conditions, pests, excess or deficiency of water etc.,
nutrients that are in adequate levels in the soil may be found deficient in plant tissue.
The effect of foliar nutrient applications would also not be reflected in soil analysis. For
effective monitoring, it is recommended to sample both soil and plant tissue.

Sampling
The time of sampling and the tissue sampled depend on the purpose of performing
the analysis.

1) For diagnosing nutritional disorders - If the purpose is to diagnose nutrient


disorders, then samples should be taken as soon as the issue is identified. For
example, when the purpose is to compare healthy leaves with leaves that show
symptoms of nutrient disorders.

2) For nutrient recommendations - Samples should be taken based on a schedule


that is specific to the crop and is based on research. Guidelines are provided
for each crop. These guidelines include the time of sampling, which part of the
plant to sample and, often, also the recommended number of samples to take.

54
• Take samples from plants that represent the field.
• All sampled plants must be at the same growth stage.
• For most crops, the most recently matured leaves are taken as a sample.
• Collect samples in a zigzag pattern.
• Do not take samples from the edges of the field.
• Sample size should usually be between 20 and 50 leaves or 100 -200 petioles,
depending on the crop and the specific guidelines for it.

Examples of sampling guidelines:

Number of plants to
Crop Sampling time Plant part to sample
sample
Before and during 3rd-6th leaf from the
Potato 15-20
flowering growing point
Terminal leaves from
Citrus Late season 25-40
not-fruiting stems
Corn 6-16 inches tall Whole plant 20-30
Tassel to silk, 3-6 First fully developed
15-25
feet tall leaf from top
Leaf opposite and
Silking 15-20
below primary ear

Interpretation of plant tissue analysis


Three main concepts are used for interpreting tissue analysis results: the critical level
concept, sufficiency range and the DRIS (Diagnosis and Recommendation Integrated
System) index.

The critical level concept - A reference point is defined for each nutrient. Results lower
than the critical level indicate deficiency. At values higher than the critical level there
will be no response in yield if the nutrient is added.

Sufficiency range - Ranges for deficiency, sufficiency and excess are given for each
nutrient.

An example for potato is given below.

55
Potato sufficiency ranges – Most recently matured leaves - First bloom
%
N P K Ca Mg S
Low <3.0 <0.2 <3.0 <0.6 <0.25 <0.2
Adequate 3.0-4.0 0.2-0.5 3.0-5.0 0.6-2.0 0.25-0.6 0.2-5.0
High >4.0 >0.5 >5.0 >2.0 >0.6 >5.0

ppm
B Fe Mn Zn Cu Mo
Low <20 <40 <30 <30 <5 0.1
Adequate 20-30 40-150 30-100 30-60 5-10 0.1-0.2
High >30 >150 >100 >60 >10 --

The DRIS index - The index is based on the ratios between nutrients in the plant tissue.
The ratio between nutrients is calculated, and then compared with the ratio at a
maximum crop yield.

The index is either positive or negative. The closer the index of a specific nutrient is to
zero, the closer is the plant to adequate nutritional balance. A positive value indicates
that the nutrient is not deficient, while the more negative the DRIS index is, the greater
the deficiency is. Values between -15 and +15 are considered an appropriate level.

Units of expression
For the DRIS method, values are expressed as a ratio and, therefore, have no units.
For the critical level and sufficiency range, the concentrations of N, P, K, Ca, Mg and
S are usually expressed as a percentage, based on dry weight.
The concentrations of B, Fe, Mn, Zn, Cu, Mo and Cl are usually expressed in ppm
(mg/kg), based on dry weight.

56
Identifying Nutrient Disorders
Plants with nutrient deficiencies or excess show specific symptoms, depending on the
severity of the disorders. In addition to soil and plant tissue analysis, being able to
identify nutrient disorders is an important part of nutrient management. However,
accurate diagnosis requires knowledge, experience and skill.

• Symptoms are not identical among plant species. The same nutrient disorder
may display different symptoms in different crops.

• Not all visual symptoms are a result of nutrient disorders. Some symptoms, that
are directly related to diseases, pests and spray damage, may resemble
nutrient deficiencies or toxicities.

• Nutrient deficiency symptoms may occur as a result of other stress factors that
limit nutrient uptake. For example, drought stress, excess irrigation, root
disease etc.

Therefore, in order to effectively solve the problem, it is essential to identify the cause
of the symptoms.
Deficiencies and toxicities can be characterized by the following features:
Chlorosis – Yellowing of leaves. The entire plant may become yellow, or just individual
leaves. Chlorosis may be interveinal or cover the entire leaf.
Necrosis – Death of plant tissue. The affected tissue usually becomes brown or black.
Necrosis can occur on leaf tips or edges, or as spots on leaves and on fruits.
Discoloration or color change – Changes in leaves and stem color, that are not
chlorosis. Leaves and stems may turn purple, reddish, discolored (whitish), dark green
etc.
Distortions – Abnormal growth of leaves or fruits.
The form and location of symptoms is key to identifying the problem.

Location of symptoms
The mobility of nutrients in the phloem determines where on the plant their deficiency
or toxicity symptoms occur. A mobile nutrient can be translocated from older to
younger leaves, while immobile nutrients cannot. Therefore, visual deficiency
symptoms of mobile nutrients will appear on older (lower) leaves first. When in
shortage, mobile nutrients are delivered to younger, growing tissue. In contrast,
deficiency symptoms of immobile nutrients appear on younger leaves.

57
Deficiency symptoms by nutrient

Nutrient Most common deficiency symptoms


Mobile nutrients (symptoms appear on older leaves first)
Nitrogen General chlorosis - leaves are pale green or
yellow, stunted growth
Phosphorus Dark green leaves, purple coloration, poor
root development, stunted growth
Potassium Chlorosis and necrosis starting at leaf edges
and moving inwards, smaller leaves
Magnesium Interveinal chlorosis, necrotic spots
Molybdenum Chlorotic and necrotic leaf margins,
interveinal chlorosis
Zinc Interveinal mottling, bronze/brown necrosis,
smaller leaves (younger leaves are also
smaller)

Immobile nutrients (symptoms appear on younger leaves first)


Sulfur General chlorosis of younger leaves (the
whole plant might become yellow), necrotic
leaf tips
Calcium Curled leaves, necrotic leaf edges, stunted
roots. Crop-specific symptoms include
blossom end rot, bitter pit, fruit splitting
Boron Abortion of growing points, curled and brittle
leaves, short internodes
Iron Interveinal chlorosis that develops to general
chlorosis of younger leaves
Manganese Interveinal chlorosis that develops to
interveinal chlorosis
Copper Twisted leaves, chlorosis, bluish coloration

Causes of nutrient disorders


Nutrient disorders are not necessarily a result of deficient or excess levels of the
nutrient in the soil or in water. Disorders may also be caused by factors that affect the
availability of the nutrient. Identifying the cause of a nutrient deficiency and correcting
it is often more efficient than applying more of the nutrient.

Possible causes of nutrient disorders include:


• Root diseases - Might be a cause of nutrient deficiencies, as they limit nutrient
uptake.
• Excess irrigation / rainfall – Result in lack of oxygen and imped the activity of
the root system.
• Water shortage – Flow of nutrients toward the roots is slowed down when water
is deficient. Furthermore, some nutrients, such as calcium, move within the

58
plant together with water and their deficiency can be directly related to water
deficiency.
• pH – Affects many processes in soil and water and, therefore, the availability of
nutrients.
• Competition between nutrients – Excess of one nutrient might lead to deficiency
of another, as a result of antagonistic reaction between nutrients and
competition.

Eliminating other causes of symptoms


As mentioned above, it is essential to identify the cause of the symptoms. Therefore,
it is important to scout the field for pests and diseases and be familiar with the
symptoms of the major pests and diseases of the crop, as well as with the conditions
that favor their occurrence.

Plant diseases and pests – Diseases might cause symptoms such as chlorosis and
necrosis, that might be misidentified as nutrient deficiencies. Look for signs that might
be directly related to plant diseases or for other disease-related symptoms. For
example, while symptoms of downy mildew may include chlorosis, the signs of the
disease also include whitish-greyish growth on the underside of the leaves. Identifying
those signs can rule out nutrient deficiency as a cause of the chlorosis.

Viruses cause symptoms such as distortions and chlorosis, that can be confused with
nutrients deficiencies.

Some pests may also cause symptoms. For example, sucking pests, like aphids, can
cause yellowing and distortions. Thrips cause leaves to become distorted.

Spray damage - Under certain conditions, such as spray drift, hot temperatures etc.,
materials applied by spray might cause symptoms, such as necrosis, chlorosis and
distortions, that might resemble nutrient deficiencies. Keeping records of pesticides
and herbicides that were applied can help in determining whether symptoms might be
a result of a spray application.

Salinity stress - Some nutrient deficiency symptoms can be confused with salt injury.
For example, potassium deficiency may occur as marginal necrosis. Testing the soil
and/or the irrigation water can help determining whether salinity is an issue.

59
Nitrogen deficiency in corn. Potassium deficiency in soybean.
Photo by R. L. Croissant, Photo by Daren Mueller, Iowa State University, Bugwood.org
Bugwood.org

Phosphorus deficiency in corn. Calcium deficiency in tobacco.


Photo by R. L. Croissant, Bugwood.org Photo by R.J. Reynolds Tobacco Company ,
R.J. Reynolds Tobacco Company, Bugwood.org

60
Chapter 2

The Irrigation Water

61
Irrigation Water Quality
The quality and composition of the irrigation water may affect plant development, soil
structure and the irrigation system itself. Irrigation water quality refers mainly to the
chemical composition of the water, or more specifically, to the mineral composition of
water.

Some physical and biological properties, such as turbidity and presence of algae,
bacteria or viruses, also determine the suitability of the water for irrigation.

However, removing turbidity by filtration, for example, and pathogens, by disinfection,


is usually more feasible than removing minerals from the water.

The quality criteria for irrigation water are entirely different than the criteria for drinking
water. Furthermore, the quality criteria may vary among crops, as different crops have
different susceptibilities to certain minerals or properties of the water.

The parameters, or chemical properties, that determine water quality for irrigation are:

• Water pH
• Water salinity
• Water hardness
• Water alkalinity
• SAR ((Sodium Adosrption Ratio) - The ratio between sodium to calcium and
magnesium
• The concentration of specific minerals

The above parameters above are all obtained from the water mineral composition.
These will be briefly explained here-under.

Water sources
Knowing the source of your irrigation water can help you evaluate its quality for
irrigation.

Irrigation quality of groundwater may be very different than the quality of surface water.
Groundwater generally contain higher levels of dissolved salts (minerals) than surface
water, while surface water may contain higher levels of turbidity and biological
impurities.

The reason for this is that mineral rocks, which surround the groundwater, break down
and dissolve into the water and constitute the source of dissolved salts in the water.
Surface water, on the other hand is exposed to the outside environment and to runoff.

62
Mineral salts that are commonly found in natural water sources:

Mineral / ion Name Comments

Ca2+ Calcium Essential plant nutrient. Used for calculating SAR.

Mg2+ Magnesium Essential plant nutrient. Used for calculating SAR

SO42- Sulfate Essential plant nutrient

HCO32- Bicarbonate Used for calculating alkalinity

Used for calculating SAR. Might harm plants when


Na+ Sodium
concentration > 50ppm. Crops differ in their susceptibility.

Might harm crops at concentrations in water > 100 ppm. Crops


Cl– Chloride
differ in their susceptibility.

Essential plant nutrient. At concentrations > 1ppm might result


Fe compounds Iron
in iron bacteria and clogging of irrigation system parts

Essential plant nutrient. Might harm crops at concentration > 0.5


B compounds Boron
ppm. Crops differ in their susceptibility.

F compounds Fluoride Can cause phytotoxicity at concentrations > 1.0ppm

Essential plant nutrient. Only low concentrations are usually


K+ Potassium
found in natural water sources.

Essential plant nutrient. Reaches water sources as result of


NO3– Nitrate
agricultural activity – leaching and runoff of fertilizers.

63
The pH of the irrigation water
The pH of the Irrigation water influences the solubility of mineral salts. Minerals that
are not dissolved are not available for plants, as plants can only uptake minerals from
an aqueous solution – directly from water or from the soil solution.

Most nutrients are available at a pH range of 5.5-6.5.

Because of its infinite volume, it is very difficult to impossible to affect soil pH by


controlling the irrigation water pH. Therefore, adjusting water pH is important in the
following cases:

• To avoid clogging of emitters (e.g., in drip irrigation) by mineral precipitation. For


example, calcium carbonate.
• In hydroponics and soilless media, where irrigation water pH directly affects the
availability of nutrients.
• When frequent irrigation is applied to the soil. In such case, water pH may affect
nutrient uptake.

Water salinity
Salinity is one of the water parameters that growers are most familiar with and relate
to the quality of the irrigation water. Salinity level that is too high reduces the plant’s
ability to absorb water. This might result in reduced yields, wilting, burned leaves and
other symptoms.

Water salinity is measured as TDS (Total Dissolved Salts) or as electrical conductivity


(EC). Both relate to the total concentration of dissolved salts in the water.

Water salinity interpretation, according to Ayers and Westcot, 1985, is given in the
following table:

Units No restriction Slight to moderate Severe

ECw ds/m <0.7 0.7 – 3.0 > 3.0


(EC of the irrigation
water)

TDS mg/l <450 450 – 2,000 >2,000

64
Water hardness
Water hardness is basically the sum of the concentrations of calcium and magnesium
in the water, expressed as ppm (parts per million) of CaCO3. Calcium and magnesium
are both essential plant nutrients and adequate concentration of them in the water is
beneficial.

However, when water hardness that is too high, precipitation of calcium and
magnesium salts might occur in the irrigation system, damage it or reduce its
efficiency. Hardness that is too low might cause corrosion in the irrigation system.

Alkalinity
Alkalinity is measure of the ability of the water to resist changes in pH. It is calculated
as the sum of carbonic acid (H2CO3), bicarbonates (HCO3–) and carbonates (CO32-) in
the water.

It is considered an important irrigation water quality parameter, as it is much more


difficult to lower the pH of water with high alkalinity, than lowering the pH of water with
low alkalinity, even if they both have the same initial pH level. This may affect the
availability of many nutrients.

Alkalinity, like hardness, is expressed as ppm of CaCO3.

SAR – Sodium adsorption ratio


SAR is an irrigation water quality parameter that helps to estimate the potential of
sodium in the water to adsorb to soil particles, in relation to calcium and magnesium.
Soil that is irrigated with water with SAR values of 10 and more, might lose its structure
and infiltration capacity. This is particularly true for soils with relatively high
concentration of clay. The soil tends to crack when dry and swell when wet. This
results in poor germination of seeds, root damage and poor aeration.

SAR is calculated in the following way:

Where all concentrations are expressed in meq/L, not in ppm.

65
Salinity and SAR have an opposite effect on soil structure and infiltration rate. The
relation between water salinity, SAR and soil infiltration rate is described in the
following diagram.

Source: FAO, Adapted from Rhodes Rhoades 1977; and Oster and Schroer 1979

Concentration of specific elements


As indicated in the table above, different crops have different susceptibilities to certain
elements. Sometimes, the difference is only a few parts per million.

For example, while tomato will tolerate irrigation water with boron concentration of up
to 6 ppm, citrus will suffer boron toxicity at concentrations of above 0.5 ppm.

Therefore, the suitability of the water for irrigation and its quality may sometimes be
determined based on one element only.

66
The Electrical Conductivity
The electrical conductivity of an aqueous solution is its ability to conduct an electrical
current.

The electrical conductivity is an important parameter used to estimate the level of


dissolved salts in water and soil, hence, it is used as an index of water salinity. Salts
include mineral plant nutrients that occur naturally in soil and water, fertilizers applied
and other dissolved minerals that are not plant nutrients and might even be harmful to
plants.

Distilled water does not conduct electricity. When salts dissolve in water, they
dissociate to into ions – positively charged ions, called cations, and negatively charged
ions, called anions. This gives the water the ability to conduct electricity.

A common parameter used to describe the total amount of the dissolved salts in a
solution is the TDS (Total Dissolved Solids). TDS is the sum of the concentrations of
all the dissolved ions, measured in milligrams per liter or ppm (parts per million), where
1 milligram/liter = 1 ppm.

The electrical conductivity is, therefore, related to the TDS and is used as a measure
of total amount of dissolved salts. Generally, the higher the concentration of dissolved
salts the higher the electrical conductivity.

However, above a certain concentration of ions in the solution, the electrical


conductivity does not increase with the concentration of ions. This is a result of
formation of ion pairs, or counterions, that weaken each other’s electrical charge.

EC vs. TDS

67
The effect of EC on plant growth
Plant roots constantly interact with their surroundings, whether it is the soil solution or
nutrient solution (in case of hydroponic production). The salt concentration in this
aqueous medium - and therefore, the electrical conductivity of that medium - greatly
affects plant growth.
High salinity affects plants in three ways:

• Restricted water uptake. Water uptake by plants is regulated by the


transpiration rate and the osmotic pressure in root cells. Water flows through
root cell membranes from low-solute concentration to high solute concentration.
Therefore, high salt concentration in the soil solution or in the water, reduces
water availability to plants.

• Toxicity of specific ions – although the electrical conductivity level does not
provide information on the presence of specific ions in the solution, high
electrical conductivity of the soil solution or nutrient solution can usually imply
high concentration of particular ions, which are potentially toxic to the plant. For
example, high concentration of chlorides, sodium, boron etc.

• Competition - High salt content of the irrigation water or water in soil may result
in nutrient imbalances. The high concentration of ions increases the chance for
competition between ions for uptake. For example, high concentration of
chlorides might restrict nitrate uptake, calcium competes with potassium etc.

Not all plants respond to salinity in the same way. Some are more susceptible than
others. The literature describes the threshold soil EC level for individual crops, above
which yield is reduced.

EC units and conversion factors


The most common unit used to measure electrical conductivity in agriculture is
deciSiemens per meter (dS/m).

Other units include μS/cm (microSiemens/cm), mS/cm, mmho/cm (milliMhos/cm) and


µmho/cm (Moh/cm; a resistance unit).

Conversion factors are as follows:

1 dS/m = 1 mS/cm = 1 mmho/cm = 1 1000 μS/cm = 1000 µmho/cm

TDS can be estimated from the electrical conductivity using the following equation:
68
TDS (ppm) = 0.64 X EC (in μS/cm) = 640 X EC (in dS/m)

This equation gives an estimate only and actual TDS may vary significantly from the
calculated value, depending on the composition of the water solution.

Most EC agricultural standards are given for readings at 25ºC. However, the EC
reading may vary depending on the temperature. The higher the temperature the
higher the electrical conductivity. As a rule of thumb, the electrical conductivity of water
increases by 2-3% for each increase of 1ºC in temperature.

69
The pH
pH is measure of the level of acidity or alkalinity of a solution. The pH value is related
to the concentration of hydrogen ions (H+) in the water. The lower the concentration of
hydrogen ions in the water is, the higher the pH.

The relationship is given as pH = -log(H+), where the concentration of H+ is given in


mol/L.

For example, a concentration of 10-6 mol/L of hydrogen ions in a solution will result in
a pH of 6.0: pH = -log (10-6) = 6.

Because the pH scale is logarithmic, a change of pH unit results in a tenfold change


in the concentration of hydrogen ions. pH 5.0 is, therefore, 100 times more acidic than
pH 6.0: 10-5 / 10-6 = 100.

The pH scale ranges from 0 to 14, where a pH below 7.0 is considered acidic (higher
concentration of hydrogen ions), pH 7.0 is neutral and a pH greater than 7.0 is basic.

pH affects many biological, chemical, and physical processes both in water and soil,
hence its importance. It affects the availability of nutrients, solubility of minerals and
fertilizers, activity of microorganisms, soil structure, efficiency of water disinfection
processes and more.

The ideal pH range for plants is 5.5-6.5.

A high pH of the irrigation water might result in nutrient deficiencies and could cause
clogging of emitters:

Nutrient deficiencies - Micronutrients, such as iron and manganese become


unavailable to plants at pH greater than 7.0.

Clogging of emitters - Precipitation of minerals, such as calcium carbonates, calcium


sulfate and phosphates, is a common cause of emitter clogging in irrigation systems
at high irrigation water pH.

Lowering the pH of the irrigation water is usually feasible using acid. The amount of
acid to add depends on the acid type, initial pH of the water and its alkalinity.

A low pH, below 5.5, might cause toxicity of micronutrients such as iron and
manganese and might also cause irreversible damage to growing roots.

70
A low irrigation water pH is usually a result of adding too much acid, or using water
that is unstable, with low carbonate content, such as desalinated water.

It’s worth noting that adjusting the pH of the irrigation water is not likely to affect the
pH of the soil. This is because soils have a high buffering capacity (an ability to resist
changes in pH via various reactions).

71
Water Alkalinity
Alkalinity is the ability of the water to neutralize acidity. It refers to the presence of
carbonates (CO32-), bicarbonates (HCO3-) and hydroxides (OH-) and expressed in
mg/L (ppm) as calcium carbonate (CaCO3). It is also known as “the buffering
capacity of the water”, as it gives the water the ability to resist changes in pH upon
the addition of acid.

Alkalinity = HCO3- + CO3 2- + (OH-)

The alkalinity is usually a result of weathering of carbonate rocks and the dissolution
of carbon dioxide. Once acid is added to water that contains alkalinity, the hydrogen
ions of the acid react primarily with the carbonate species.

H2CO3 ↔H+ + HCO3- pH = 6.37 + log (HCO3-/H2CO3)

HCO3- ↔ H+ + CO32- pH = 10.33 + log (CO32-/HCO3-)

As can be seen in the figure above, carbonates (CO32-) exist in water at pH above 8.3
and at a pH of 4.5 it is certain that all the carbonates were converted to carbonic acid
and the water does no longer have buffering capacity.

Therefore, the more alkalinity there is, the more acid is required in order to lower the
water pH.

72
Once the alkaline compounds are consumed, any addition of acid will result in an
immediate change of the pH, as the hydrogen ions, contributed by the acid, remain
free in the solution.

For example, let’s compare the following two water sources:

Water source 1: pH 6.5, 120 mg/L HCO3-

Water source 2: pH 7.5, 80 mg/L HCO3-

Which water source requires more acid in order to lower the pH to 5.5?

Intuitively, one would say that water source 2 requires more acid because its pH is
higher.

However, it can be seen from the titration curves below that water source 1 requires
more acid in order to lower its pH to the desired level. The reason for that is the higher
level of bicarbonates in water source 1 results in a higher buffering capacity, hence
stronger resistance to changes in pH.

73
Hardness
Water hardness is defined as the sum of the divalent metallic ions in the water. The
main contributors to hardness are Calcium and Magnesium. Additional contributors to
the hardness of the water include Iron (Fe+2), Strontium (Sr+2), Zinc (Zn+2), Manganese
(Mn+2) and other ions. However, their concentrations are usually significantly lower
than the concentration of calcium and magnesium.

In most cases, summing up the calcium and magnesium in the water gives an
adequate hardness measure.

What are the effects of water hardness?


• Hard water might cause scale deposition in water distribution and irrigation
systems
• It reduces the efficiency of heat exchangers
• Might cause corrosion
• Might cause scaling in membrane filtration systems

In adequate concentrations, calcium and magnesium have a positive effect on plants.


They are both considered to be essential nutrients, and their deficiency might cause
reduced yields and quality problems. Therefore, in irrigation water, a certain level of
hardness is favorable.

Types of hardness
Temporary hardness vs. permanent hardness
There are two types of hardness – temporary hardness and permanent hardness.
Temporary hardness - also called ‘Carbonate hardness’. This type of hardness refers
to the calcium and magnesium carbonates and bicarbonates in the water. Heating the
water or reacting it with lime removes this hardness. CO2 is released as gas, and
precipitates of insoluble calcium carbonate and/or magnesium hydroxide form.
Ca(HCO3)2 ➔ CaCO3↓ + CO2↑ + H2O.
Mg(HCO3)2 ➔ Mg(OH)2↓ + 2CO2 ↑

Permanent hardness, also referred to as ‘non-carbonate hardness, is the hardness


due to the presence of calcium or magnesium sulfates, chlorides and nitrates. For
example, calcium sulfate, magnesium chloride etc.
Removal of permanent hardness is done by using lime with soda ash. Excess lime
must be used when magnesium hardness is high.

74
Units
The most common unit for expressing the level of hardness in the water is ‘mg/L as
CaCO3.
The concentrations of calcium and magnesium are expressed as equivalent of CaCO 3.
When the concentrations of calcium and magnesium in the water are known, hardness
can be calculated, using the following formula:
Water hardness = 2.5Ca + 4.1Mg
Where calcium and magnesium are measured in ppm (1ppm=1 mg/L).
For example, water with 50 ppm Ca and 15 ppm magnesium will have hardness of:
2.5x50 + 4.1x15 = 186.5 mg/L CaCO3
Additional units include:
dGH – ‘Degrees of General Hardness’ or ‘German Degrees’.
1 dGH = 17.484 mg/L CaCO3
Grains per gallon: 1 gpg = 17.1 mg/L CaCO3
French degrees ⁰fH:
1 French degree = 10 mg/L CaCO3

Classification of water hardness:

mg/l CaCO3 Classification

< 60 mg/l Soft

60 – 120 mg/l Medium-hard

120 – 180 mg/l Hard

> 180 mg/l Very hard

75
Irrigation Water Analysis
Water quality affects many aspects of crop production, including crop yield, quality,
design of the irrigation system, required treatment (e.g., filtration), fertilizer application
and scheduling, operational aspects and more.
It is, therefore, essential to test the source water before any decision is made.

What to test for


The water testing lab will need to know which tests it is expected to perform.
The decision on which parameters to test depends on variables such as the water
source, crop grown, type of irrigation system and general available information of the
quality of other water sources in the same area.
Water quality parameters can be classified into chemical, physical, and biological
parameters.
For irrigation water, a chemical analysis of the water, i.e. an analysis of the mineral
composition of the water, is always required. The exact set of elements to be tested
depends on the source of the water and to the crop’s sensitivity to specific elements.
The most common set of parameters to test includes EC, pH and the following
elements:
Common concentrations in Desirable range
groundwater (mg/L) (mg/L)
Potassium (K+) 0-10 5-10
Calcium (Ca2+) 10-150 60-120
Magnesium (Mg2+) 5-30 10-30
Sulfur (S-SO42-) 5-30 30-80
Boron (B) 0-2 <0.5
Iron (Fe) 0-3 <1.0
Manganese (Mn) 0-2 <0.8
Bicarbonates (HCO3-) 30-500 30-120
Sodium (Na+) 5-130 0-50
Chlorides (Cl-) 0-350 <120

In areas where there is an extensive human activity, such as agriculture and industry,
a certain level of nitrates may also be found in the water. In volcanic areas, fluoride
may be present in concentrations that may require attention.
Note that even a single element that is present at a concentration that is too high might
restrict the use of the water for irrigation. For example, a boron concentration that
exceeds the tolerance threshold of the crop.

76
In addition, since all the elements in the water are dissolved salts, a high level of one
ion may indicate a high level of another. For example, if the level of chlorides is high,
it is most likely that the level of sodium would also be high.
Water treatment is required if the level of a specific ion or of the total soluble salts
exceeds the threshold of the crop. However, removing salts from water is an expensive
treatment. Available treatments include desalination and ion-exchange.
The chemical analysis of the irrigation water is also used to assess clogging potential
of in drip irrigation systems or foggers, as a result of precipitation of minerals (e.g.,
calcium carbonate, calcium sulfate etc.), or of slime formed by iron or sulfur bacteria.
Testing the physical properties of the irrigation water is required mostly in the planning
phase of the irrigation system. The purpose of the physical test is to avoid clogging of
emitters and to design the filtration system, if required. Parameters that are measured
in a physical test include the total suspended solids (TSS), particle size analysis,
turbidity, and temperature.
A physical analysis is more commonly performed when surface water is used, as
surface water is more exposed to physical contaminations and may contain sand and
clay particles.
Testing for biological contaminants, such as bacteria, viruses and algae is not a
common practice. It is usually performed when sensitive crops are grown or when it is
necessary to know whether a specific pathogen is present in the water.

Sampling guidelines
The way the water is sampled can significantly affect the results. Therefore, it is
important to follow the sampling guidelines.

• Water sample must represent the irrigation water to be used.


• A glass or plastic bottle/container should be used.
• The container should be rinsed multiple times with the sampled water.
• For biological analyses, a glass bottle is preferred. The bottle must be sterilized
prior to taking the sample. It is recommended to use a sterilized bottle provided
by the lab. Do not rinse the bottle.
• The container should be labeled clearly and right after sampling.
• The container should be sealed properly.
• The sample needs to be placed inside an ice cooler box, then sent to the lab
as soon as possible.

77
Sampling frequency
The frequency in which one should take water samples for analysis depends on the
expected variability of the water quality. Seasonal variability may occur, especially if
the source water is surface water.
In soilless culture and greenhouse production, where fertigation is used, it is a
common practice to field-test the EC (electrical conductivity) and pH of the irrigation
water on a weekly basis.

78
The Principle of Electrical Neutrality
The principle of electrical neutrality states that all aqueous ionic solution are electrically
neutral. The equivalent amounts of positively charged ions (cations) are equal to the
equivalent amounts of negatively charged ions (anions).

The main reason for checking the balance of electrical charges in a solution is to
validate water or nutrient solution test results. A difference in the ionic balance that is
outside the range of ±5% usually indicates that there is an error in the lab analysis or
that one or more of the significant ions that may affect the balance were not tested.

In most cases, and for simplicity purposes, it is sufficient to calculate the balance using
only the major cations and anions, omitting the micronutrients, as their low
concentrations does not significantly affect the balance.

Common major cations Common major anions


Calcium (Ca2+) Chloride (Cl-)
Magnesium (Mg2+) Bicarbonate (HCO3-)
Sodium (Na+) Sulfate (SO42-)
Potassium (K+) Nitrate (NO3-)
Ammoniacal nitrogen (NH4+) Phosphate (PO43-)

Not all the ions mentioned in the above table are necessarily present in water,
depending of the nature of the water sample and its source. For example, natural water
sources are not likely to contain ammoniacal nitrogen.

If it is assumed that a certain ion is present in the water at such a concentration that
might affect the balance, but it was not tested, then the balance calculation will not
give a proper result.

In order to calculate the balance, the concentrations of all cations and anions must
first be converted to units of milliequivalent per liter (meq/L). This unit represents the
number of charges of each ion. The concentration of each ion in meq/L can be
calculated by multiplying the number of millimoles per liter of each ion by its charge.

Example:

Calculate the concentration of magnesium in meq/L if the magnesium concentration


in the solution is 30 ppm.

The molecular weight of magnesium is 24 mg/mmol.


30 ppm = 30 mg/L = 30 mg/L / 24 mg/mmol = 1.25 mmol.

79
Magnesium carries a charge of +2. Therefore:
1.25 mmol Mg = 1.25 x 2 = 2.5 meq/L

In the same way, 50 ppm of calcium are also 2.5 milliequivalents. The molecular weight
of calcium is 40 grams/mole and it carries a positive charge of +2:
2 X 50 / 40 = 2.5 meq/L

Therefore, in terms of electrical charges and in the same water volume, 30 grams of
magnesium ions have the same number of electrical charges as 50 grams of calcium.

As mentioned above, any aqueous solution must be electrically neutral. This includes
nutrient solutions, to which fertilizers are added. This implies that dissolving fertilizer
in water does not imbalance the solution and, therefore, each fertilizer that is dissolved
in water always adds an equal number of equivalents of cations and anions to the
solution.

For example, magnesium sulfate contains approximately 11% N (as nitrate-nitrogen)


and 9.6% Mg. Let’s assume that 100 grams of magnesium sulfate are dissolved in
water. These 100 grams will contain 11 grams nitrate-nitrogen and 9.6 grams
magnesium.

The nitrate-nitrogen must first be converted to the ionic form of nitrate: 11 x 4.43 =
48.73 grams NO3-.

The molecular weight of NO3- is 62 grams/mole and it carries a charge of -1. Therefore,
the number of equivalents is 1 X 48.73 / 62 = 0.79 eq

The ionic form of magnesium is Mg2+, its molecular weight is 24.3 grams/mole, and it
carries a charge of +2. Therefore, the number of equivalents of magnesium is: 2 x 9.6
/ 24 = 0.79 eq.

As can be seen from the above example, the addition of fertilizer adds exactly the
same number of equivalents of cations (Mg2+) and anions (NO3-) and does not change
the ionic balance of the solution.

Electroneutrality in water – example


Calculate the cation-anion balance for the following water analysis.
Parameter Result
EC 0.51 ds/m
pH 7.58
N-NO3 ND
HPO4 ND
K 4.30 mg/L

80
Ca 40.08 mg/L
Mg 13.13 mg/L
Na 41.93 mg/L
HCO3 122.0 mg/L
S-SO4 6.40 mg/L
Cl 98.0 mg/L
B 0.024 mg/L
ND=Not detected

To calculate the balance, first divide the elements to two groups – cations and anions.
Cations: K+, Ca2+, Mg2+, Na+
Anions: NO3, HPO42-, HCO3-, SO42- , Cl-

Then, convert all values to meq/L


mg/L Molecular weight Charge Conversion factor meq/L
(mg/mmol)
Cations
K 4.3 39 +1 1/39 = 0.025 0.11
Ca 40.08 40 +2 1/(40/2) = 0.05 2.00
Mg 13.13 24.3 +2 1/(24.3/2) = 0.082 1.07
Na 41.93 23 +1 1/23 = 0.043 1.80
Sum = 4.98
Anions
NO3 0 62 -1 1/62 = 0.016 0
HPO4 0 96 -2 1/(96/2) = 0.02 0
HCO3 122 61 -1 1/61 = 0.016 1.952
SO4 6.4 x 3 = 19.2 96 -2 1/96 = 0.010 0.192
Cl 98 35.4 -1 1/35.4 = 0.028 2.744
Sum = 4.88

Note that when doing the balance, all elements must be converted to the form in which they
are present in water. In the example above, sulfur was given in its elemental form (S) and,
therefore, it must first be converted to sulfate (SO42-).

81
Chapter 3

Soils

82
Soil Fertility
The soil is the surface layer of earth's crust in which plants grow and where biological
activity occurs.
The soil consists of three phases: solid, liquid and gas.
The solid phase consists of minerals and organic matter. Most soils contain between
1-6% organic matter and 94-99% minerals.
Nutrients can be held on the surfaces of mineral particles and organic matter. They
can be also released from organic matter, as it decomposes.

The liquid phase refers to the water held by the soil, as water fills the spaces between
soil particles. The liquid phase forms the soil solution, which contains dissolved
minerals.
The gaseous phase contains the same components as air – oxygen, nitrogen and
carbon dioxide. It contains more carbon dioxide than air and is in balance with the
atmospheric air.

The mineral composition of the soil


Soil fertility refers to the ability of the soil to retain and provide nutrients to the crop.
The mineral composition of the soil is the main factor that determines soil fertility.
Minerals can be classified into two main groups: primary minerals and secondary
minerals.
Primary minerals are formed during the original crystallization of molten rocks
(magma). They have not undergone chemical changes since their original formation.
Primary minerals constitute the sand and silt fractions of the soil.
Secondary minerals are formed at lower temperatures and pressures, by weathering
of the primary minerals. They usually constitute the clay fraction of the soil.
Soil is formed by weathering of parent material, such as bedrock mineral sediments,
marine sediments volcanic ash and organic deposits. The types of parent materials
and the conditions under which they weather determine the properties of the soil that
is formed.
For example, soils formed from granite are often sandy and not fertile, while basalt,
under wet conditions breaks down to form fertile, clay soils.

83
Soil texture
The soil texture defines the particle size distribution of the soil, where:
• Sand – particle diameter of 2 - 0.05mm
• Silt – particle diameter of 0.05 - 0.002 mm
• Clay – particle diameter less than 0.002 mm

Because of their small size, the surface area of the clays is much larger than that of
silt and sand particles. Therefore, there is a larger surface area where reactions can
occur.

Cation Exchange Capacity


Clay minerals are negatively charged due to isomorphic substitutions and hydroxyl
(OH-) groups on the edges of clay particles and organic matter.
Isomorphic substitution is the replacement of an ion by another ion with a similar size.
If the replacing ion has a lower positive charge than the ion replaced, then the net
charge of the particle will be negative. In clays Si4+ is replaced by Al3+, which can be
also replaced by Mg2+.
These negative charges are sites that can adsorb and release cations (positively
charged ions) and, therefore, important nutrients, such as K+, Mg2+, Ca2+ and NH4+,
but also harmful elements, such as sodium.
The total number of negative sites available in the soil is referred to as the Cation
Exchange Capacity (CEC) of the soil. The cations adsorbed to the soil particles can
be exchanged with other cations present in the soil solution.

84
A soil with a high CEC is more fertile than a soil with a low CEC, as it can retain more
nutrients.

Soil pH
Soil pH greatly affects soil fertility:
1. Nutrient availability is governed primarily by the pH
2. Soil CEC is affected by soil pH. High soil pH increases the CEC of the soil,
while a low pH decreases it

Soil organic matter


Soil fertility is also influenced, in several ways, by the organic matter content in the
soil:

• Organic matter is an important nutrient source for plants. Nitrogen, phosphorus


and other nutrients become available to plants as it decomposes.

• It contributes both to the CEC of the soil, as well as to the anion exchange
capacity of the soil (AEC). Soil AEC is much lower than CEC in most soils.

• It has chelating properties, that increase nutrient availability to plants.

• It improves soil structure

Maintaining soil fertility


Nutrients are depleted from soil by plant uptake, leaching, erosion etc. Practices to
maintain soil fertility, under intensive agriculture, include applying fertilizers and/or
organic amendments, crop rotation, reducing tillage and using cover crops.

85
The Cation Exchange Capacity
The cation exchange capacity of the soil is the total amount of exchangeable cations
that the soil can hold on the surface of soil particles. Clay minerals and organic matter
have negative charges on their surfaces, called ‘exchange sites’ or ‘exchange
complex’. In order to balance the negative charges, positively charged ions (cations)
adsorb to the exchange sites by electrostatic forces. These cations are in equilibrium
with the soil solution, hence are referred to as ‘exchangeable cations.

The most common exchangeable cations are calcium (Ca2+), magnesium (Mg2+),
potassium (K+), sodium (Na+), aluminum (Al3+) and hydrogen (H+). Cations such as
NH4+ (ammoniacal nitrogen), Fe2+ (iron), Mn2+ (manganese) and Cu2+ (copper) are
present at much lower amounts.

The CEC of a soil is closely related to the fertility of the soil, as it provides the soil a
reservoir of nutrients. Nutrients in the soil solution are just a small portion of the
available nutrients. When plants take up nutrients such as potassium, calcium and
magnesium from the soil solution, these nutrients can be replenished by the
exchangeable cations. Soils with a high CEC can hold more nutrients than soils with
a low CEC.

The CEC of the soil depends on soil texture, the type of minerals in the soil,
concentration of organic matter and soil pH.

Due to their structure and particle size, fine-textured soils have a higher CEC than
coarse-textured soils. Due to their small size, clays have a much larger surface area
than coarse-textured soils. Therefore, clay soils can retain a larger amount of nutrients
than sandy and silty soils and are considered more fertile. In order to provide the crop
with adequate nutrient rates, sandy soils require smaller and more frequent fertilizer
applications, compared to clay soils.

86
Units of expression
The cation exchange capacity of the soil is measured in units of meq/100g. This is the
number of exchange sites in each 100 grams of soil.

1 meq/100g = mmol/100g X the charge of the ion.

For example, 1 mmol/100g of calcium will occupy twice the exchange sites as 1
mmol/100 of potassium, because the charge of calcium is +2, while the charge of
potassium is +1.

Typical CEC of soil minerals and soil types

Mineral CEC (meq/100g)


Clays
Kaolinite 3-15
Illite 15-40
Montmorillonite 80-100
Organic Matter 100-300
Soil Texture
Sand 1-5
Loamy Sand to Sandy Loam 5-10
Loam 5-15
Clay Loam 15-30
Clay 25-40

Why do soil surfaces have a negative charge?


As mentioned above, the source of negative charges are clay minerals and organic
matter.

Clays are secondary minerals, formed as a result of weathering of primary minerals.


They consist of two fundamental structures:
1. Sheets of silicon-oxygen tetrahedrons - SiO4
2. Octahedral sheets

Tetrahedron Octahedron

87
The negative charge is formed as a result of isomorphous substitutions, in which ions
within the mineral structure are replaced by other ions with a similar size. For example,
replacement of Si4+ by Al3+ and Al3+ by Mg2+. The replacement of an ion by another ion
with a lower positive charge, results in a net negative surface charge. Isomorphous
substitutions are structural and permanent and occur during the weathering process
of minerals.

Isomorphous substitution

Additional sources for negative charge exchange sites are hydroxyl groups (OH -) on
edges of clay particles and hydroxyl or carboxyl groups in organic matter.
The organic matter contributes to the CEC of soils, as the CEC of colloidal humus is
very high due to carboxyl and hydroxyl functional groups. Between 20 and 70% of the
CEC of many soils is due to colloidal humic substances.

Carboxyl and hydroxyl groups

Hydroxyl and carboxyl groups can release or bond hydrogen ions, depending on soil
pH. At low soil pH the negative charge of the hydroxyl or carboxyl groups is
neutralized, while a high soil pH will result in more sites with negative charges.
Therefore, for the same soil, CEC will increase in high pH and will be lower in low pH.

Determination of CEC

88
Several methods are used to determine soil CEC. The measured CEC may be
different than the actual CEC in the field. The CEC at field pH is referred to as
“effective CEC” (CECe), while buffered solutions are often used to determine the CEC.
For example, barium chloride (BaCl2) is used for determining CEC. However,
ammonium acetate buffered at pH 7.0 or Mehlich 3 buffered at pH 8.2 are widely used
to determine the CEC under standardized conditions.

At pH greater than 7.5 or when lime has been recently applied, the CEC measurement
is often overestimated.

89
Soil pH and Acidity
The degree of soil acidity or alkalinity is expressed in pH units. Soil pH affects many
processes in soils, including nutrient availability, activity of microorganisms and root
development.
The pH of natural soils typically ranges from 4.5 to 8.0. Soils with pH between 6.5 and
7.0 are considered to be neutral, soils with pH above 7.5 are alkaline and soils with
pH lower than 6.5 are acidic. Most plant nutrients become available at slightly acidic
pH of 5.8 to 6.5.
At a soil pH lower than 5.2, nutrients such as calcium, magnesium, nitrogen,
phosphorus and boron may become unavailable for plants, while the solubility and
availability of trace elements, such as iron, aluminum, manganese, zinc and copper
increases significantly and they may become toxic.
In highly acidic soils, the mineralization process of organic matter is slowed down and
can even stop completely, because microbial activity declines under low pH
conditions. This results in a decreased availability of nitrogen and phosphorus.
In high pH soils, micronutrient deficiencies become common.
Phosphorous availability is reduced in both highly acidic soils, with pH lower than 5.5,
as well as in alkaline soils with pH greater than 7.5. In acidic soils, phosphorus reacts
with iron and aluminum and becomes unavailable, while in alkaline soils, phosphorus
reacts with calcium.
Alkaline soils are characterized by presence of calcium, magnesium, and sodium
carbonates. At soil pH levels between 7.2 and 8.2 soil pH is dominated by calcium and
magnesium carbonates and the soil is called a ‘calcareous soil’. At a soil pH above
8.2, the soil is dominated by sodium carbonates, as they become highly soluble. High
sodium levels relative to calcium and magnesium may negatively affect soil structure.
The availability of nutrients at different pH levels is described in the following chart.
The thicker the bar, the more available the nutrient is.

90
Nutrient availability vs. pH

There are three pools of acidity in the soil: active acidity, exchangeable acidity and
residual acidity.
The active acidity is the concentration of free hydrogen ions (H+) in the soil solution.
Soil pH is a measure of the active acidity of the soil.
The exchangeable acidity refers to hydrogen and aluminum ions retained on the
exchange complex of the soil, i.e. on the surfaces of soil colloids. It is the potential
acidity of the soil, as adsorbed Aluminum (Al3+) and Hydrogen (H+) ions, along with
calcium (Ca+2), magnesium, (Mg+2), potassium (K+) and sodium (Na+), are in
equilibrium with the soil solution.
The higher the cation exchange capacity of the soil, the greater its buffering capacity
is, as the soil can retain a larger amount of hydrogen and aluminum ions.
The buffering capacity of the soil refers to its ability to resist changes in pH. Since the
active acidity is in equilibrium with the exchangeable acidity, adsorbed aluminum and
hydrogen ions can replenish aluminum and hydrogen that were removed from the soil
solution. Therefore, soils with high CEC and low pH will require more lime in order to
increase the pH to the desired level.

While calcium, magnesium, potassium and sodium are considered base cations,
aluminum and hydrogen are considered acidic.

91
Aluminum is considered acidic due to the hydrolysis reactions it goes through in the
soil solution. The hydrolysis of aluminum generates hydrogen ions, i.e. active acidity.

Al3+ + H2O = Al(OH)2+ + H+

Al(OH)2+ + H2O = Al(OH)2+ + H+

Al(OH)2+ + H2O = Al(OH)3 + H+

Al(OH)3 + H2O = Al(OH)4- + H+

Residual acidity is associated with aluminum and hydrogen ions that are bound
to soil colloids, but not in an exchangeable form.

Determination of soil pH
There are several methods for determining soil pH. All methods measure the active
acidity, i.e. hydrogen ions in the soil solution, but each method will provide different
results for the same soil sample. Therefore, in order to properly interpret the results, it
is important to understand the difference between the methods.
Methods based on water extraction:

• pH of the saturated paste extract

• pH of 1:2 extract (1 part soil, 2 parts water)

• pH of 1:5 extract
The more water is used for the extraction, the higher the measured pH would be,
because the addition of water dilutes the hydrogen ions in the extract.
In methods that use only water for the extraction, hydrogen ions that are bound to soil
particles remain bound and are not released into the solution.
Methods that use chemical agent for the extraction:
In order to obtain a result that better represents the field conditions, a diluted solution
of potassium chloride (1.0M KCl) or calcium chloride (0.01M CaCl2) are commonly
used. The salt concentration in the extracting solutions intends to represent the salt
concentration in the soil solution. The potassium or calcium in the extracting solutions
replace some of hydrogen ions bound to soil particles and, therefore, pH measured
using these methods is usually closer to the actual pH of the soil.
Readings are 0.5 to 1.5 units lower than pH measured in water, due to the higher
concentration of hydrogen ions in the solution.

92
Soil Salinity
The salinity of soil refers to the amount of salts in the soil. Soil degradation by
salinization is a serious problem that affects world agriculture. About 20% of irrigated
soils are affected by salts and for important crops, the average yields are only between
20% and 50% of their potential yield.

Two types of salinity can be defined: primary (natural) salinity and secondary salinity.

Primary salinity is a result salt accumulation by natural processes, over a long period
of time. For example, salt deposition from rain, rock degradation, capillary rise of
groundwater and salt intrusion from the sea.

Secondary salinity is caused as a result of human activities. For example, irrigation


management, fertilizer application and inadequate soil drainage conditions.

Measurement of soil salinity


Salinity can be estimated by measuring the electrical conductivity (EC) of an extracted
soil solution.

It is important not to confuse between different types of electrical conductivity.


Readings change greatly when measured in a different medium or different soil
moisture levels.

ECw – electrical conductivity of the irrigation water.

ECe – electrical conductivity of the saturated paste extract of the soil. The
saturated paste extract is an extract taken at a specific moisture status of the
soil. All EC thresholds related to crop tolerance to salinity are given as ECe.

ECe ≈ ECw x 1.3 to 1.5, depending on soil texture, irrigation frequency etc.

1:2 or 1:5 EC extract - the electrical conductivity measured in an extract of 1:2


or 1:5 soil:water ratio.

To convert EC obtained by the 1:2 or he 1:5 extracts to ECe, or to compare


between the values obtained by these methods, regression equations must be
used.

93
ECb or ECa - the bulk electrical conductivity (also named apparent Electrical
conductivity) of the soil, which is the electrical conductivity of bulk soil, i.e. the
soil water and air. This is a different parameter than ECe.
Example: Soil salinity classification and EC values in ECe and 1:2 Extract EC

Classification ECe (ds/m) 1:2 Extract EC (ds/m)


Negligible salinity affect,
0-2 <0.4
most crops will grow well
Only sensitive crops will be
2-4 0.4-1.6
negatively affected
Moderately salt-tolerant
4-8 1.6-2.4
crops are affected
Strong saline conditions.
Only high salt-tolerant crops
8-16 2.4-3.2
will grow well. Yields of most
crops will reduce.
Only a few high salt-tolerant
>16 >3.2
crops will grow.

Effect of soil salinity on plant growth


Soil salinity is one of the most common factors that affect crop yields. Yield reduction
and damages to the crop are related to restricted water uptake due to high osmotic
potential, toxicity of specific ions and competition between ions for uptake.

The salinity tolerance threshold varies among crops and is given in the literature as
threshold ECe.

Yield response to salinity


100
90
80
Relative yield (%)

70
60 Rice
50 Soybean
40 Tomato
30
Grape
20
10 Strawberry
0
0 2 4 6 8 10 12 14 16
Root zone salinity (ECe, ds/m)

94
Yield decreases linearly as salinity increases above the threshold level of the crop.
The yield response to salinity can be estimated using the following equation (Matt and
Hoffman, 1977):

Y = 100 – B (ECe – A) ECe > A

Y = 100 ECe ≤ A

Where

Y Relative yield (%)

A The salinity threshold of the crop (ECe in units of ds/m)

B Yield reduction per unit increase in ECe (%)

Onion field affected by high salinity


Photo by Howard F. Schwartz, Colorado State University, Bugwood.org

95
Managing Soil Salinity
Several strategies can be applied for managing salinity. Some of them can be applied
prior to planting and others throughout the growth of the crop.

Salinity is only a problem if it affects the crop. Therefore, the first step in salinity
management is to identify its cause and evaluate the potential effect on the crop. In
some cases, it may be more economical to grow a different crop or different variety,
that are less sensitive to salinity, or choose a different location, than to apply any
salinity management strategy.

Salinity assessment
Performing soil salinity tests and water analysis gives the best indication of the extent
of the salinity problem. It is recommended to test for both general indices, such as EC
and pH, as well as for specific elements, such as chlorides, sodium, bicarbonates,
calcium and magnesium. In the soil salinity analysis, soluble salts are measured in a
solution extracted from the soil, most commonly in the saturated paste extract.

Soil and water tests can help understand the source of salinity and the measures that
must be taken. For example, reclaiming a saline soil requires good quality water. If the
salinity of the water exceeds the salinity threshold of the crop, reclamation measures
may not be efficient.

The required number of tests to be performed and their frequency depend on the initial
findings and on the salt sensitivity of the crop.

Irrigation system type and design


In irrigated soils, salts concentrate at the root zone of the crop as the crop takes up
water in the transpiration process.

Both type and design of the irrigation system affect salt distribution in soil. Soluble
salts move in soil with water and the distribution of salt accumulation in soil is usually
very similar to the distribution of the irrigation water.

Different irrigation system types have different water distribution patterns and,
therefore, in irrigated soils, salt distribution and accumulation largely depends on the
irrigation system type and design.

In sprinkler irrigation or flooding, salts move downward, while in drip irrigation systems,
salts move both laterally and vertically. The depth and width of the wetted area depend
on the emitter flow rate and soil properties. The soil near the drip line is wetter than
the soil at the periphery of the wetted area.

96
Salts may accumulate in between emitters and with distance from the dripline. In sub-
drip irrigation (SDI), salts accumulate also near the surface of the soil. Therefore, the
distance between dripper and laterals must be adequate in order to ensure uniform
irrigation and leaching of salts.

Leaching
Leaching refers to the application water in such an amount that is sufficient to remove
excess salts from the root zone of the crop to deeper soil layers. It is one of the main
strategies to avoid accumulation of salts or reclaim saline soils.

In reclamation leaching, a sufficient amount of good-quality water is applied in order


to bring the salinity level in the root zone to below the salinity threshold of the crop.
Best effect is achieved with sprinkler irrigation. In order for the reclamation to be
successful, the soil must be well drained.

The amount of water to be applied depends on the water quality and soil properties.
In theory, a leaching curve can be created for each specific soil. The leaching curve
describes the correlation between the fraction of the initial salt concentration remaining
in the soil and the depth of leached water per unit of depth soil. However, obtaining
the leaching curve is complex and, therefore, not a common practice.
A more practical approach consists of using simple models or rules of thumb.

For example, Hanson et al., 2006 proposed the following equation:

Dw = (k x Ds x ECei) / ECed

Where:

Dw = Required depth of water for leaching (mm or inches)


Ds = Depth of soil to be reclaimed (mm or inches)
k = 0.45 for organic soils, 0.3 for fine-textured soils and 0.1 for coarse-textured soils
ECei = Initial soil salinity (ds/m in the saturation extract)
ECed = Desired soil salinity (ds/m in the saturation extract)

As a rule of thumb, a unit depth of water will remove 80% of salts from a unit depth of
soil. For example, 200mm of water will remove 80% of accumulated salts from a soil
layer depth of 20 cm.

It is recommended to do the leaching at a time when evapotranspiration rates are low


(e.g., in the winter), in order to avoid salts that were leached from moving back up as
a result of high evapotranspiration rates.

97
Note that in saline-sodic soils, salts must be leached only after exchangeable sodium
is replaced with calcium. Otherwise, soil structure might be further damaged.

Maintenance leaching is a practice used to avoid salinity from accumulating in the


root zone and exceeding the threshold of the crop. In this practice, irrigation is applied
more frequently, and excess water is applied with each irrigation, more than the
amount required by the crop.

The amount of water applied in each irrigation can be calculated using the following
equation:

Dw = (∑ETc)/(1-LR)

Where

Dw = water depth to apply (mm or inches)


∑ETc = Crop evapotranspiration from last irrigation (mm or inches)
LR = the leaching requirement (fraction)

The leaching requirement is defined as:

LR = Depth of water leached / Depth of water applied

LR can be calculated using the following equation.

LR = ECw/(5ECet - ECw)

Where:

ECw = The EC of the irrigation water (ds/m)


ECet = The salinity threshold tolerated by the crop, measured as EC of the soil
saturation extract (ds/m).

Example:
The average ETc of an onion crop is 4mm/day
The EC of the irrigation is 1.0 ds/m
The salinity threshold of onion is 1.2 ds/m

What is water amount required in order to ensure leaching?

LR = 1.0/(5x1.2 – 1.0) = 0.2

Dw = 4/(1 – 0.2) = 5 mm/day


98
Irrigation interval
Knowing the total amount of irrigation water to be applied may not be sufficient for
maintaining adequate salinity in the root zone. Due to the lateral movement of water
in fine-textured soils, greater amounts of water, at larger intervals are needed in order
to leach salts out of the root zone, compared to coarse-textured soils.

Fertilizer application
The application of fertilizers directly affects the salinity level in the soil. Fertilizer
application rates must be adequate and determined based on a soil analysis, in order
to avoid excess fertilizer.

In addition, some fertilizers contain elements that plants do not take up or take up in
very small amounts. At high concentrations, these elements may become toxic to the
crop. For example, potassium muriate (KCl) should be avoided for crops that are
sensitive to chlorides.

The effect of different fertilizers on soil salinity can be evaluated comparing the salt
index of different fertilizers. The salt index is the proportion of the increase in osmotic
pressure produced by the fertilizer, relative to the osmotic pressure produced by an
equal weight of NaNO3, taken as 100.

It is recommended to select fertilizers with lower index for crops that are sensitive to
salinity or for soils that are already saline.

Fertilizer Salt index


Potassium chloride 1116.3
Ammonium nitrate 104.7
Sodium nitrate 100
Urea 75.4
Potassium nitrate 73.6
Ammonium sulfate 69
Calcium nitrate 52.5
Potassium sulfate 46.1
Magnesium sulfate 44
Diammonium phosphate 34.2
Monopotassium phosphate 29.9
Triple superphosphate 10.1
Calcium sulfate 8.1
Single superphosphate 7.8

99
Additional practices to avoid salt build-up
Soil leveling: Improves water distribution and eliminates salt accumulation in sections
of the field.
Application of organic matter amendments: Maintain a good structure, increase water
retention capacity and facilitate water movement through the soil profile.
Application of plant stimulants: The energy stress of plants can be reduced with the
application of amino acids and bio-stimulants.

100
Soil Sodicity
Sodic soils are defined as soils that contain an exchangeable sodium percentage
(ESP) higher than 15. Sodium can be a constituent of the total salinity. However, it
has a unique effect on the soil structure.

Soil structure is the arrangement of soil particles. When soil particles clump together
and aggregate, they create a stable structure, in a process called flocculation. This
allows for proper water movement in the soil, adequate air exchange and easier
growth and penetration of roots into the soil.

Soil dispersion occurs when soil particles detach from each other and soil structure is
destroyed. Loose particles block the pores of the soil. Therefore, sodic soils tend to
swell when wet harden and crack when dried. A saline crust may occur on soil surface.
As a result, soil permeability is decreased, oxygen availability in the root zone is limited
and root growth is restricted. This adversely affects crop growth and yield. Soil
dispersion may also result in increased erosion because plants die and leave a bare
soil, which is more susceptible to erosion.

Flocculation and dispersion are phenomena related to the clay fraction of the soil. Clay
particles have negative charges on their surfaces. In order to neutralize the negative
charges, cations from the soil solution are attracted to the surfaces of the clay particles
and an electric double layer is formed. Because negative charges repulse each other,
the extent to which the electrical charge is neutralized determines the strength of the
repulsion forces. When repulsion forces are high, soil particles disperse and vice
versa.

Therefore, the properties of the exchangeable cations in the soil affect the flocculation
and dispersion processes. The higher the valency of a cation and the smaller is its
hydrated radius, the greater is its power to flocculate the soil. Cations that carry a
higher charge are more effective in neutralizing negative charges, and the smaller the
cation is, the easier it is for it to get closer to the surface of the clay particle.

101
Catíon Hydrated radius (ºA)
Potassium (K+) 3.3
Sodium (Na+) 3.6
Calcium (Ca2+) 4.1
Magnesium (Mg2+) 4.3
Aluminum (Al3+) 4.8

Sodium is less effective in neutralizing negative charges on clay surfaces, while


calcium and magnesium neutralize surface charges more effectively. Therefore,
indices expressing the ratio between sodium, calcium and magnesium were
developed.

Dispersion. Sodium is adsorbed to soil particles Flocculation. More calcium is adsorbed to soil particles

The indices used to assess sodium risk are:

• SAR of the irrigation water


• SAR of the soil solution
• ESP (Exchangeable Sodium Percentage)

SAR was already previously discussed in the chapter on water quality. It is calculated
for the irrigation water and in the soil salinity test, in the following way:

𝑁𝑎
𝑆𝐴𝑅 =
√𝐶𝑎 + 𝑀𝑔
2

102
Where the concentrations of Na, Ca and Mg are in meq/L. Water with SAR values
greater than 10 may cause severe damage to soil.

ESP is defined as the amount of sodium adsorbed to soil particles, expressed


as a percentage of the CEC.

𝐸𝑥𝑐ℎ𝑎𝑛𝑔𝑒𝑎𝑏𝑙𝑒 𝑁𝑎
𝐸𝑆𝑃 = 𝑥 100
𝐶𝐸𝐶

Where both exchangeable Na and CEC are expressed in units of meq/100g.

As mentioned above, sodic soils have an ESP value >15.

Example:
Consider the following soil analysis. Calculate ESP and SAR.

Fertility analysis Salinity analysis


EC 3.2 ds/m 3.2 ds/m
CEC 18.4 meq/100g
Potassium 1.15 meq/100g 1.2 meq/L
Calcium 12 meq/100g 7.0 meq/L
Magnesium 3.42 meq/100g 3.92 meq/L
Sodium 1.39 meq/100g 5.3 meq/L

1.39
𝐸𝑆𝑃 = 𝑥 100 = 7.55%
18.4

For calculating SAR, the salinity analysis must be used:

5.3
𝑆𝐴𝑅 = = 2.27
√7 + 3.92
2

Sodic soils have a pH between 8.5 and 10, as a result of the high quantities of sodium
carbonate. At this pH level, calcium and magnesium precipitate out of the soil solution
and sodium becomes more dominant on the exchange complex.

Effect of salinity on sodium potential damage


Salinity negatively affects crop yields. However, it induces flocculation and
aggregation of soil particles and alleviates the negative effect of sodium on soil
structure. The high concentration of cations in the soil solution helps neutralizing the
negative charges on clay surfaces and reduces the thickness of the electric double
layer.

103
Soil classification

Soil physical
Classification EC (ds/m) Soil pH SAR ESP condition
Normal < 4.0 6.5-7.5 <13 < 15 Good
Saline > 4.0 < 8.5 < 13 < 15 Normal to poor
Sodic <4.0 > 8.5 > 13 > 15 Very poor
Saline-sodic >4.0 < 8.5 > 13 > 15 Poor

Reclamation of sodic soils


The reclamation process of sodic soils can take years. It involves the application of
soil amendments, such as gypsum or acid forming amendments.

The purpose of adding the soil amendment is to replace the adsorbed sodium with
calcium and improve soil structure. Only after the application, sodium can be leached
out of the soil, using good quality water. An attempt to leach the sodium before it is
replaced with calcium, will result in further damage to soil structure, as salts will leach
out of the soil profile, while sodium remains adsorbed to soil particles. The reduced
salinity may, therefore, cause soil dispersion.

Small, repeated applications of gypsum are preferred over one large application.

The reactions that take place upon the application of soil amendments to recover sodic
soils:

Gypsum:
Na-Soil + H2O + CaSO4•2H2O → Ca-Soil + Na2SO4 + H2O

Sulfuric acid (only for calcareous soils)


2H2O + H2SO4 + CaCO3 → CaSO4•2H2O + H2O + CO2↑
Na-Soil + H2O + CaSO4•2H2O → Ca-Soil + Na2SO4 + H2O

Sulfur - acid-forming amendment:


2S + 3O2 + 2H2O → 2H2SO4
2H2O + H2SO4 + CaCO3 → CaSO4•2H2O + H2O + CO2↑
Na-Soil + H2O + CaSO4•2H2O → Ca-Soil + Na2SO4 + H2O

The rate of gypsum to apply can be calculated using the following formula:

104
𝑡𝑜𝑛𝑠 8.61 𝑥 (𝐸𝑆𝑃𝑎 − 𝐸𝑆𝑃𝑡) 𝑥 𝐶𝐸𝐶 𝑋 𝐷 𝑋 𝐷𝑏
𝐺𝑦𝑝𝑠𝑢𝑚 ( )=
ℎ𝑎 1000 𝑒

Where
ESPa = Actual ESP of the soil
ESPt = Target ESP
D = Depth of soil layer for which reclamation is required.
Db = Bulk density of the soil
e = Gypsum application efficiency.

And the formula of gypsum is CaSO4*2H2O.

For amendments other than gypsum, divide the gypsum rate, resulted from the above
formula, by the gypsum equivalent given in the following table.

Amendment Gypsum equivalent


Gypsum (CaSO4*2H2O) 1.00
Calcium chloride (CaCl2*2H2O) 0.85
Sulfuric acid (H2SO4) 0.57
Iron sulfate (FeSO4*7H2O) 1.62
Aluminum sulfate (Al2(SO4)3*18H2O) 1.29
Sulfur (S) 0.19

105
Soil Organic Matter
Soil organic matter refers to the remains of plants, animals and microbes, in different
decomposition stages. It consists of:

1. Living soil organisms and plant residues


2. Partially decomposed organic matter (detritus)
3. Fully decomposed, stable organic matter, also referred to as Humus

Most soils contain 1-6% organic matter.

Benefits of organic matter


Organic matter influences many of the physical, chemical and biological properties of
the soil.

• Increases soil CEC (Cation Exchange Capacity)


• Improves soil structure and water holding capacity
• Contributes nutrients, mainly nitrogen, phosphorus and sulfur. This is a result of
nutrient mineralization
• Improves the absorption of nutrients by plants (chelation of micronutrient)
• Increases soil buffer capacity (resistance to changes in pH)
• Improves soil biodiversity

Microorganisms break down organic matter to small molecules until a stable humus is
formed. The stable humus does not contribute much nutrients to soil, because it is
already decomposed and stable.

The chemical composition of the humus varies depending on the different types of
organisms and plant residues from which it was formed.

The same factors that affect the activity of microorganisms in soil also affect the
decomposition of soil organic matter and formation of humus:

• Soil pH – decomposition of organic matter is slowed down at soil pH that is too


acidic or alkaline.
• Temperature – decomposition rate of organic matter is increased in higher
temperatures.
• Soil moisture – biological activity requires moisture and oxygen. Therefore,
optimal conditions for organic matter decomposition will occur at around field
capacity. Soil saturation conditions, on the other hand, delay microbial activity.
• Soil texture – the breakdown of organic matter is usually better in light-textured
soils, than in clay soils because the small clay particles and small pores “protect”
the organic matter from microorganisms.

106
The CEC of colloidal humus is very high, much higher than the CEC of clay minerals.
This is due to the negative charges it carries in the form of functional groups of
carboxyl (-COOH) and hydroxyl (-OH).

Therefore, soils with high organic matter content are more fertile than soils with low
organic matter content.

In fact, between 20 and 70% of the CEC of many soils is due to colloidal humic
substances.

Clay mineral / Organic


matter CEC, meq/100g

Kaolinite 3 – 15

Illite 20 – 40

Montmorillonite 60 – 100

Humus 100 – 300

The CEC of soil organic matter depends on the pH of the soil. Higher pH results in
higher CEC. Hydrogen ions H+ are released from the carboxyl groups and more
negative sites are formed on the surface of the organic colloids.

Nutrient mineralization
The main nutrient contributed by soil organic matter is nitrogen. Nitrogen becomes
available for plants in the mineralization process, in which organic nitrogen is
converted to inorganic nitrogen, which is available for plants. This process is
performed by microorganisms.

107
Several models to estimate the amount of nitrogen release as a result of mineralization
of organic nitrogen were developed. However, due to the complexity of nitrogen
dynamics in soil, many assumptions must be done. Therefore, further research is
required, and new models are being developed.

A relatively simple model was proposed by Stanford and Smith (Stangord, G.


and Smith, S. J. 1972. Nitrogen mineralization potentials of soils. Soil Sci. Soc. Am.
Proc. 36:465472).

The proposed equation is:

Nt = N0(1 – e -kt)

where Nt is cumulated N mineralized (mg kg-1)

t – time (wk)

N0 – the potentially mineralizable N

k – rate constant (wk-1)

Rough estimates of the amount of nitrogen released from organic matter are given in
the following table.

Estimated amount of nitrogen released


% organic matter in soil (lbs/acre/year)

1% 15-25

2% 25-55

3% 55-70

Determination of soil organic matter


The organic matter content of soils is evaluated by measuring the organic carbon in
the soil. The two main methods used are the Walkley & Black acid wet oxidation
method and the mass loss on ignition method. The first method is most accurate for
soils with less than 2% organic matter, while the mass loss on ignition method is best
for soils with more than 6% organic matter.

108
Walkley & Black’s method involves oxidizing the organic matter in the soil by
potassium dichromate (K2Cr2O7) and sulfuric acid. The mass loss of ignition method
is based on weight loss as a result of combustion of organic matter on applying high
temperatures.
Organic carbon content can be converted to organic matter using the following
conversion factor:

OM (%) = Total Organic Carbon (%) x 1.72

109
The Soil Analysis
Soil testing is a key tool in fertilizer management. It reduces guesswork and helps in
minimizing environmental pollution. When done and used properly, soil testing can
save the grower money and increase crop yields.

Thirteen of the thirteen essential plant nutrients are taken up by plants from the soil.
Imbalance of nutrients may result in nutrient deficiencies, toxicities, and competition
between nutrients. This may result not only in reduced crop yields and quality, but
even cause severe damages to the crop.

Therefore, in order to achieve optimal yields and quality, nutrients in soil and other soil
conditions must be monitored.

Fertilizer recommendations can be given based on the soil test results. The application
of fertilizers replenishes nutrients that were depleted from the soil and help maintaining
a proper nutrient balance.

General fertilizer recommendations can provide a rough estimate of the type and rate
of fertilizer to apply. However, this often results in misapplication of fertilizers and might
end up being inefficient and costly.

Without knowing which nutrients are available in the soil, at what rates and what is
their availability, fertilizer application is merely a guess.

Successful soil testing requires:

• Proper sampling
• Using the adequate extraction methods by the laboratory
• Correct interpretation of the results

Soil test can determine parameters such as soil texture, pH, available nutrients,
organic matter and salinity-related parameters.

Collecting soil samples


The reliability of the soil test depends first and foremost on the sample taken. Since
only a very small part of the field is being tested, the soil sample must represent the
field or the plot from which it was taken, as closely as possible.

110
Sampling of field areas that are near roads, trees, manure piles, or any other factor
that might locally affect soil properties and composition, should be avoided.

1. The best soil sampling practice is to take several sub-samples from the field
and mix them together to form one representative sample. It is recommended
to sample at least each 10 hectares (about 20 acres), depending on the
uniformity of the field and crop. Note that in fields with high variability of soil
conditions, each soil zone must be sampled separately, and samples should
not be mixed.

2. Take the samples before applying any organic fertilizer to the soil. Use clean
tools and a clean container.

3. From each plot, take 10 to 20 sub-samples and mix them.

4. Soil samples should be taken from the depth where the active root system is,
usually to a depth of 15-20cm.

5. Place the mixed sub-samples in a plastic bag. About 0.5 – 1 Kg of soil is


required for soil testing.

6. Before sending the bag to the lab, identify the sample properly, by indicating
on the bag the plot name, crop, date of sampling and type of analysis needed.

Soil parameters
Soil parameters tested in soil analysis for agriculture can be classified into structural
properties, soil fertility parameters and soil salinity parameters:

Structural Soil fertility and related parameters Soil salinity parameters


properties
Soil texture Available nutrients (Common nutrients tested in Electrical conductivity
and type fertility test: phosphorus, potassium, calcium, (EC)
magnesium, sulfur, iron, manganese, zinc, and
copper).
Soil bulk Cation Exchange Capacity (CEC) Sodium Adsorption Ratio
density (SAR) – a parameter used
to evaluate the effect of
sodium on soil structure.
Soil pH, soil acidity
Percent organic O(M) matter or organic carbon

111
Soil structural properties are usually tested once, before starting the agricultural
activity on that soil. Soil fertility and salinity are tested more frequently.

Nutrient availability and soil testing


Different nutrients in soil are potentially available to plants, and the purpose of soil
testing is to estimate them as accurately as possible.

Soil consists of three phases which exist in equilibrium – solid phase, liquid phase and
gas phase. The most important equilibrium exists between the solid phase and the
liquid phase (soil solution). Plants absorb nutrients from the soil solution. However, the
amounts of nutrients in the soil solution are insufficient to sustain plant growth.

When nutrients are depleted from the soil solution, they are replenished from the solid
phase of the soil. This occurs through reactions such as dissolution, desorption,
oxidation, reduction, hydrolysis or microbial mineralization reactions.

Testing the soil solution alone will underestimate the amount of nutrients available to
the plant, while measuring the total amount of nutrients present in the soil will result in
an overestimation of the nutrients available. This is because a large portion of them is
strongly bound to soil particles or exists in a form which is not available to plants.

Therefore, to determine the elements that are available for plants in soil, the soil test
must quantify a relationship between the solid phase and the soil solution.

Why is it difficult to estimate the available


nutrients?
Soil chemical tests are mostly empirical from several reasons:

• In the sampling and testing procedure, the soil loses its physical structure,
and soil particles are in close contact with the chemical used for the extraction
of the nutrients. This close contact optimizes desorption/dissolution reactions
and, therefore, may result in overestimate or underestimate of the amount of
nutrients that can reach the plant roots.

• The physical accessibility of nutrients to plant roots affects their availability to


plants. Immobile nutrients, such as P and Zn, are less available, thus the
quantity of nutrient extracted usually differs from the quantity actually available
to the plants.

• Plants vary in their capability to absorb nutrients, as they have different


root architecture and activity.
112
• Other environmental factors, such as soil water status affect both root growth
and nutrient availability.

• Soil properties, such as pH, mineral composition and nutrient buffering


capacity also affect the capacity of the soil to release nutrients into the soil
solution.

Soil testing methods


Soils can be tested for multiple parameters, using a variety of analytical methods.
Different soil testing labs may provide different results for the same soil sample. This
is mostly related to the fact that different labs may use different extraction methods.

An extraction method is the procedure that the lab uses in order to determine the level
of a certain nutrient or element in the soil. It consists of the analytical procedure and
chemicals that are used in order to extract the nutrient/element from the soil sample.

In the extraction process, elements that are adsorbed to soil particles move into the
extracting solution. The concentration of the extracted element in the extracting
solution is then measured.

Note that the extraction method does not refer to the instrument or method that was
used to measure the concentration of the element in the solution, but to the procedure
used to move elements from the solid phase of the soil into the solution.

Therefore, “ICP”, “Flame photometer”, “Spectrometer”, “Colorimeter” etc. are the


instruments used, and cannot be considered the extraction method.

Extraction methods for each nutrient were developed for specific soil conditions and,
sometimes, for specific crops. Different extraction methods may be used to determine
the availability of the same nutrient, under different soil conditions.

For example, three common testing methods for phosphorus are Bray-1 (using
0.025M HCl and 0.03M NH4F), Olsen (using 0.5M NaHCO3) and Mehlich 3 (using
0.015M NH4F, 0.013M NH4NO3). The Olsen method is more suitable for calcareous
soils, while Bray phosphorus and Mehlich are suitable for soils with pH<7.0.

Each method extracts a different amount of the nutrient/element from a given soil
sample. The reason for that is that the amount of the extracted element does not
represent its total amount in the soil and often not even the actual amount of available
nutrient. The test results provide empirical values that must then be interpreted
correctly.

113
This concept is illustrated in the following figure:

Developing an extraction method requires an extensive research and field trials. It


involves finding correlations between the amount of nutrient extracted and the
probability of the crop responding to the application of fertilizer. This concept is
illustrated in the following chart. Two different extraction methods for a certain nutrient
were used on the same soil sample. The correlation between the nutrient amount that
was extracted from the soil and the crop uptake was investigated.

114
As can be seen from the chart above, extractant 2 extracted higher amounts of the
nutrient than extractant 2. Moreover, it is clear from the chart that the correlation
between the extracted amount and crop uptake is better for extractant 1 and, therefore,
it is likely that extractant 1 is more suitable for that specific soil.

Note that factors such as stirring time, reagent dilution, shaking rate etc. may also
affect results and are part of the extraction method.

Examples of some extraction methods are given in the table below.

Soil:
Extraction Nutrients Agitation
Extracting solution Solution Reference Comment
method extracted time
ratio

Bray 1 P HCl 0.025N, NH4F 1:7 1 min. Bray and Not suitable
0.03N Kurtz, for high
1945 pH soils
(>6.8)
Olsen P NaHCO3 0.5N; pH 1:20 30 min. Olsen at Suitable for
8.5 Al, 1954 calcareous
soils with
pH>6.5
Bray 2 P HCl 0.1N, NH4F 1:7 1 min. pH>7.5,
0.03N adsorbed
forms of P
Mehlich 1 P, K, Ca, HCl 0.05N, H2SO4 1:4 5 min.
Mg 0.025N
Ammonium K, Ca, Mg NH4OAc 1N; pH 7.0 1:10 20 min.
acetate

115
Nitric acid K HNO3 1N 1:10 15 min.
Potassium Ca, Mg KCl 1N 1:10 10 min.
chloride
Modified P, K NHCO3 0.5N, EDTA 1:10 30 min. Hunter,
Olsen 0.01M, superfloc 1977
127; pH
8.5

It is recommended to make sure that the lab to which soil samples are sent uses the
appropriate extraction methods.

Soil salinity vs. fertility


Labs often do not indicate whether they tested the soil for salinity or fertility and
growers must figure that out from the extraction methods used. In most cases, the soil
analysis report includes parameters related to salinity together with parameters related
to fertility.

Salinity test can be used for giving fertilizer recommendations for soils with low CEC
only. In medium-high CEC, nutrient levels in the salinity test cannot be used for that
purpose.

Soil fertility test results provide information on the nutrients available to plants.
Results are given in units such as ppm or meq/100g, where ppm in this case
refers to mg/kg.

Soil salinity test uses only water for the extraction, without the use of chemicals. Soil
is mixed with a specific amount of water and the solution is then tested. Common
methods include the saturated paste extract, 1:2 extract (1 volume soil: 2 volumes of
water) and 1:5 extract.

The purpose of the salinity test is to provide a measure of the concentration soluble
salts in the soil. It helps to evaluate the conditions in which plants grow – salinity and
specific toxicities.

The level of certain elements that might be toxic to the plants if their threshold is
exceeded. Usually boron, chlorides, and sodium.

Calcium and magnesium are also tested in the salinity test, as they are used for
calculating the SAR.

Results are reported in units of ppm, mmol/L or mg/L.

116
Note that while in the fertility soil test ppm refers to mg/kg, in salinity test it refers to
mg/L. This is because in the fertility test measures the nutrients adsorbed to the soil
particles, while the salinity test refers to nutrients in the soil solution.

117
Units on the Soil Test Report
Soil labs may report soil test results in different units. Often, growers may find this
confusing, as their interpretation criteria they have available may use other units.
Furthermore, for the same soil sample, different labs may return results that seem to
be significantly different, even if they used the same extraction methods.
The most commonly used units to express the concentration of extracted elements
are ppm, mg/kg, mg/L, mmol/L, meq/100g, meq/L, cmol+/kg, lbs/acre and kg/ha.

PPM
PPM stands for Parts Per Million. In the same way percent means out of a hundred,
ppm means out of a million. On the soil test report, it may refer to a milligram of certain
chemical per one kilogram of soil or milligram of chemical per one liter of soil solution
extract.

PPM may refer to different ratios, depending on the extraction method used and on
the type of the tested element. For elements that are extracted from the exchange
complex, using a chemical extractant, 1 ppm refers to mg/kg (and 1ppm = 1 mg/kg).
When the element is determined in the soil solution, using just water for the extraction
(saturated paste extract, 1:2 extract etc.), ppm usually refers to mg/L.

The reason is that in chemical extraction, the amount of the element available in the
soil, but attached to soil particles, is determined, while extraction with water is used to
determine the concentration of the element in the soil solution.

kg/ha and lbs/acre


Some labs will report nutrient levels in kg/ha or lbs/acre. These units in the soil test
report are, in fact, obtained from the ppm (mg/kg) values. To calculate the amount of
a nutrient or an element in a specific area, both the bulk density of the soil and the soil
layer depth which the test results represent must be known. Most labs use 20cm (8
inches) or 30 cm (12 inches) as a default depth for which the amount of element is
calculated.

Example:
A soil test shows an Olsen phosphorus level of 15 ppm. How many kilograms of Olsen
phosphorus are there in a one-hectare field? Assume a soil bulk density of 1.35 ton/m3
and a layer depth of 20 cm.

Assuming 20 cm soil layer depth, the volume of the soil layer is:
10,000 m2 x 0.2 m = 2,000 m3 (1 ha = 10,000 m2)

Using the bulk density, the layer’s weight can be calculated:

118
2,000 m3 x 1.35 ton/m3 = 2,700 tons.

2,700 tons = 2,700,000 kg

15 ppm = 15 mg/kg

And in 2,700,000 kg:

15 mg kg-1 X 2,700,000 kg / 1,000,000 mg kg-1 = 40.5 kg

Therefore, for a soil with bulk density of 1.35 ton/m3 and layer depth of 20 cm, 15ppm
= 40.5 kg/ha
Note that the converting ppm to kg/ha or lbs/acre does not require using the molecular
weight or the charge of the element. For example, for the same soil bulk density and
soil depth, the calculation above would be the same for 15 ppm of potassium, calcium
or any other element.

The equation for converting ppm to kg/ha in metric units is:

kg/ha = ppm x LD x BD / 10

Where:

LD = layer depth in cm
BD = bulk density in ton/m3

Converting kg/ha to lbs/acre:

1 kg = 2.20 lbs (pounds)


1 ha = 2.47 acres

Therefore,
1 kg/ha = 2.20/2.47 lbs/acre = 0.89 lbs/acre

And the equation to convert ppm to lbs/acre in imperial units is:

Lbs/acre = ppm x LD x BD / 275

Where:
LD = layer depth in inches
BD = bulk density in lbs/ft3

119
Therefore, 15 ppm phosphorus in a soil layer depth of 8 inches and soil bulk density
of 84 lbs/ft3 are:

15 x 8 x 84 / 275 = 36.6 lbs/acre

meq/100g and cmol(+)/kg


Meq/100g stands for milliequivalents per 100 grams of soil. 1 meq/100g = 1
cmol(+)/kg, where cmol(+)/kg is the abbreviation for centimoles per kilogram. These
units are used to report the cation exchange capacity (CEC) of the soil and the
amounts of exchangeable cations that can occupy the exchange sites (potassium,
calcium, magnesium, sodium, aluminum, ammonium, and hydrogen).

For CEC, meq/100g unit refers to the number of negatively charged exchange sites
and, when it comes to the exchangeable cations, it refers to the number of exchange
sites that each of them occupies. Therefore, 1 meq/100g of potassium will occupy the
same number of exchange sites as 1 meq/100 of calcium, magnesium and any of the
other exchangeable cations.

Because meq/100g refers to the exchange sites, this unit express a number of
charges.

The following equation is used to convert between ppm and meq/100g:

ppmi x zi
meq/100g =
10 x Mw

Where:

ppmi is the concentration of the cation in ppm, Zi is the charge of the cation and Mwi
is the molecular weight of the cation in mg/mmol.

Example:

How many exchange sites do 1200 ppm of calcium occupy?

Ca in meq/100g = 1200 x 2 / (10 x 40) = 6 meq/100g

This conversion can be explained as follows:

1200 ppm Ca = 1200 mg/kg Ca

The molecular weight of Ca is 40 mg/mmol. Therefore:

120
1200 𝑚𝑔/𝑘𝑔
= 30 𝑚𝑚𝑜𝑙/𝑘𝑔
40 𝑚𝑔/𝑚𝑚𝑜𝑙

There are 30 millimoles (30 x 6 x 1020) calcium ions in each kilogram of soil or 3
millimoles in 100 grams. Each calcium ion carries a positive charge of +2, and the
number of exchange sites it can occupy is, therefore 3 x 2 = 6.

Meq/100g = mmoli/100g X zi

mmol/L
This unit is used for expressing the concentration of elements in the soil solution.

1 mmol/L = ppm / Mw

Where ppm refers to mg/L and Mw is the molecular weight of the element in mmol/mg

Example:

Convert 100 ppm Mg2+ to mmol/L

The molecular weight of magnesium is 24.3 mmol/mg, therefore:

100 / 24 = 4.11 mmol/L

121
How to Interpret Soil Test Results
Soil test results might be confusing to read and interpret. One of the main reasons for
that is the lack of standardization. As mentioned above, different labs may return
different results for the same soil sample.

For example, one lab may return a potassium test result of 200 ppm, while another lab
would return a result of 50 ppm, depending on the extraction method used.

Because the results of the analysis vary depending on the method used, their
interpretation is mostly qualitative.

A classification of 'Low', 'Adequate', 'High', 'Excessive' etc. is used in order to interpret


the results. A low classification indicates a high probability for crop response to
fertilizer application, high and excessive indicate low or no probability for crop
response. Crop response refers to increase in yield and/or quality.

For example, the classification for potassium extracted with ammonium acetate at pH
7.0, is:

Low < 175 ppm


Adequate – 175-280 ppm
High – 280-800 ppm
Excessive - >800 ppm

In order to be able to correctly interpret the soil test results, it is imperative to know
which extraction method the lab used for each element and make sure that the
methods used are suitable for the specific soil conditions.

Often labs provide this information on their soil test report. However, if the extraction
methods that were used are not indicated on the report, it is advised to ask to ask for
it.

Some labs may also provide reference interpretation of the results. It should be noted,
though, that the interpretation references may vary among labs.

In conclusion, testing the soil may provide essential information about the soil salinity
and fertility and for making fertilizer management and irrigation decisions. However,
the results must be viewed in a broader context, considering the specific crop, field
conditions and the constraints mentioned above.

122
Understanding the units on your soil test report
The different units used in soil test reports is an additional source for confusion. Labs
may express the soil test results in different units.

The most used units for describing nutrient content in the soil are given in the following
table:

Weight-based units Volume-based units Amount/area units


mg/kg mg/L kg/ha
ppm in soil (=mg/kg) ppm in a solution (=mg/L) lbs./acre
meq/100g meq/L
C+mol/kg (=meq/100g)

Both the weight-based units and amount/area units refer to nutrients or elements that
were extracted from the soil phase of the soil. Volume-based units are used for
elements that were tested in the soil solution.

ppm – parts per million


meq/100g – milliequivalents per 100 grams of soil
C+mol/kg – centimole per kilogram
Meq/L – milliequivalents per liter

You may need to do some unit conversions. For example, if the lab provided the soil
test results in one set of units and the reference ranges for the interpretation are given
in other units.

As noted in the table above, 1 mg/kg = 1 ppm in soil, 1 mg/L = 1 ppm in solution.

To convert mg/kg to meq/100g use the following formula:

1 meq/100g = mg/kg x (charge of the ion) / (Mw of the ion) / 10

Where Mw is the molecular weight of the ion.

Example:

The soil test level for Mg2+ is 300 mg/kg. What is the Mg2+ level in meq/100g?

The molecular weight of magnesium is 24 g/mol and it has a charge of 2+. Therefore,
300 x 2 / 24 / 10 = 2.5 meq/100g

123
To convert mg/L to meq/L use the following formula:

1 meq/L = mg/L x (charge of the ion) / Mw

Example:
Calcium (Ca2+) level in the tested 3 meq/L in the salinity test. How many ppm
calcium does this water contain?
Calcium carries a positive charge of 2 and its molecular weight is 40 mg/mol,
therefore:

3 = mg/L x 2 / 40
mg/L Ca2+ = 3 x 40 / x = 60 mg/L = 60 ppm

124
Soil Test Interpretation Guide
Phosphorus (P), ppm
Extraction method Low Moderate High Excessive
Olsen <10 10-30 30-40 >40
Saturated paste <2.5 2.5-10 11-15 >15
Bray 1 <20 20-40 40-100 >100

Potassium (K), ppm


Extraction method Low Moderate High Excessive
Ammonium acetate <150 150-250 251-800 >800
Mehlich 3 <35 35-65 66-130 >130
Saturated paste <80 80-195 196-300 >300

Calcium (Ca), ppm


Extraction method Low Moderate High Excessive
Ammonium acetate <1000 1,000-2,000 >2,000
Mehlich 3 <500 501-655 656-835 >835
Saturated paste <100 100-400 401-500 >500

Magnesium (Mg), ppm


Extraction method Low Moderate High Excessive
Ammonium acetate <60 60-180 >180
Mehlich 3 <20 20-65 66-135 >135
Saturated paste <30 30-120 121-180 >180

Boron (B), ppm


Extraction method Low Moderate High Excessive
Hot water <0.5 0.5-2.0 2.0-6.0 >6.0
Saturated paste <1.0 1.0-1.5 >1.5
Calcium phosphate <0.5 0.5-1.0 >1

125
Iron (Fe), ppm
Extraction method Low Moderate High Excessive
DTPA <2.5 2.5-5.0 5.0-10 >10
Mehlich 3 <50 50-100 >100

Manganese (Mn), ppm


Extraction method Low Moderate High Excessive
DTPA <1.0 1.0-2.5 2.5-4.0 >4
Mehlich 3 <30 30-200 >200

Zinc (Zn), ppm


Extraction method Low Moderate High Excessive
DTPA <0.5 0.5-0.75 0.75-1.0 >10
Mehlich 3 <1.0 1.0-2.0 >2.0

Copper (Cu), ppm


Extraction method Low Moderate High Excessive
DTPA <0.2 0.2-0.3 0.3-0.6 >12
Mehlich 3 <1.6 1.6-4.5 >4.5

Sodium (Na), ppm


Extraction method Low Moderate High Excessive
Ammonium acetate <50 50-200 200-300 >300
Saturated paste <45 45-115 116-160 >160

Sulfur (S), ppm


Extraction method Low Moderate High Excessive
KCl 40 <5 5-12 13-20
Saturated paste <160 160-480 481-640 >640

126
Chloride (Cl), ppm

Extraction method Low Moderate High Excessive


Saturated paste <175 175-530 531-1240 >1240

Soil Electrical Conductivity


Category 1:1 Extract EC Saturated Paste
(ds/m) ECe (ds/m)
Low (Non-saline) 0.01-0.45 0.0-2.0
Low (Slightly saline) 0.45-1.5 2.1-4.0
Medium (Moderately saline) 1.51-2.9 4.01-8.0
High (Strongly saline) 2.91-8.5 8.01-16.0
Very high (Very strongly saline) >8.5 >16.0

Desirable range of exchangeable cations


Cation Range
Calcium 65%-80%
Magnesium 10%-20%
Potassium 3%-8%
Sodium <6%
Aluminum <1%

127
Raising Soil pH
Soil acidity is a major cause of reduced yields. Toxicity of micronutrients such as
manganese, iron and aluminum may occur, while nutrients such as potassium, calcium
and magnesium become unavailable. Furthermore, low soil pH impedes beneficial
microbial activity in the soil.

Acidification can be caused by several factors. Some are natural causes and others
are a result of an intensive agricultural activity.

Parent material – The rock from which the soil was formed affects its acidity. Soils
developed from acidic parent material, such as granite, are likely to be more acidic
than soils developed from calcareous parent material.

Rainfall – Excess rainfall leaches basic cations (calcium, magnesium, potassium) out
from the soil. These cations are replaced by hydrogen, which makes soil more acidic.

Acid rain is precipitation that is more acidic than normal and, therefore it accelerates
the acidification of soils. It contains sulfuric and nitric acids that are formed as a result
of sulfur dioxide and nitrogen dioxide emissions from industry, vehicles and power-
generating plants.

Decomposition of organic matter – The decomposition process of organic matter


produces CO2 that reacts with water to form carbonic acid.

Application of certain types of fertilizers – The nitrification process of ammonium


nitrogen yields hydrogen ions. Therefore, ammonium-based fertilizers have an
acidifying effect. Sulfur fertilizers also have a strong acidifying effect.

Removal of basic nutrients with the harvest – Crops take up basic cations from soil
and excrete protons in order to maintain electroneutrality. The basic cations are
removed from the field at harvest and the soil acid neutralizing capacity is decreased.

The most common method to raise soil pH is by applying agricultural lime (calcium
carbonate) or equivalent liming materials. The reaction of lime in soil is represented
by:

128
Once lime is added to an acidic soil, calcium replaces the exchangeable acidity, i.e.
the hydrogen ions on the exchange complex, while the carbonate reacts with hydrogen
ions in the soil solution to form carbon dioxide and water.

Because the optimum pH range varies among crops, the decision whether a lime
application is required depends on the crop and the actual pH of the soil.

Example optimum pH range for several crops

Crop Optimum soil pH


Alfalfa 6.6-7.0
Blueberries 4.5-5.0
Corn 5.8-6.2
Sweet potatoes 5.0-5.5
Peanut 5.0-7.5
Asparagus 5.5-7.5
Beets 5.5-7.5
Tomato 6.0-6.5
Soybeans 6.6-7.0
Sugar beets 6.5-7.0

The process of raising soil pH to the desired level may take two to three years. Lime
is applied to the soil surface before planting and should preferably be incorporated into
the soil.

Amount of lime to apply


The amount of lime to apply depends on the ability of the soil to resist changes in pH
(or buffering capacity). Labs measure a parameter known as the ‘buffer pH’. A buffer
solution with a pH of 7.5 is added to the soil extract and the pH of the mix is reported
on the soil test report.
A high buffer pH indicates that the buffering capacity of the soil is low (the addition of
the buffer solution increased the pH considerably) and vice versa. The higher the
buffer pH is, the higher the lime application rate should be.
The following equation can be used to calculate the lime requirement:
Lime requirement (lbs/acre) = 2000 X [ EA X [(target pH - soil pH)/(6.6- soil pH)]
Where EA is the exchangeable acidity.
If not given in the soil test report, the exchangeable acidity can be estimated from the
buffer pH:

129
Target pH Exchangeable acidity
>=7 62.6- (2.33 x soil pH) - (7.13 x buffer pH)
>=6.5 51.8- (2.77 x soil pH) - (5.38 x buffer pH)
up to 6 39.3- (3.69 x soil pH) - (2.83 x buffer pH)

Example:
Soil pH: 4.5
Target pH: 6.5
Buffer pH: 7.1

Exchangeable acidity = 51.8 – (2.77 x 4.5) – (5.38 x 7.1) = 1.137


Lime requirement = 2000 X [ EA X [(target pH - soil pH)/(6.6- soil pH)] =
2000 x [1.137 x (6.5-4.5)/(6.6-4.5) = 2165 lbs/acre

Raising soil pH using potassium carbonates


Potassium carbonate (K2CO3) and potassium bicarbonate (KHCO3) are water soluble
and, therefore can be applied through drip irrigation systems. They can be used when
it is necessary to raise soil pH in the rhizosphere and lime cannot be logistically
applied.

Due to their high potassium content, the contribution of potassium must be accounted
for in the fertilizer program. Ammonium fertilizers should be avoided, because the
ammonium is converted to ammonia which might volatilize.

130
Quality Parameters of Liming Materials
In addition to agricultural lime, there are several other types of liming materials
available. Liming materials differ in their ability to neutralize soil acidity, in
effectiveness and in their economic benefits.
The two factors that determine the quality of liming materials are:
1. Purity
2. Particle size

Purity
The purity of a liming material is also referred to as calcium carbonate equivalent
(CCE). It is the total neutralizing value (TNV) of the liming material, expressed as
calcium carbonate equivalent, on a dry weight basis. Pure calcium carbonate has a
CCE of 100% and is the standard against which other liming materials are measured.
Other materials may have CCE that is lower or higher than 100. The higher the CCE,
the greater the capacity of the liming material to neutralize soil acidity.
For example, hydrated lime, Ca(OH)2 has a CCE of 136. Therefore, 100 kg of hydrated
lime have the same neutralizing value as 136 kg of pure calcium carbonate.
The CCE of many materials is lower than 100 because they contain impurities such
as silica, alumina and other constituents that do not have neutralizing power.

CCE of common liming materials

Liming Material CCE

Calcium silicate (CaSiO3) 60-90

Calcium carbonate (CaCO3) 100

Dolomitic lime (CaCO3•MgCO3) 108

Quicklime, Calcium oxide (CaO) 179

Hydrated lime CaOH2 136

Particle size
The particle size distribution of the liming material affects its effectiveness. Due to their
larger surface area, small particles react more quickly in soil than coarser particles
and, therefore, are more effective in neutralizing acidity.

131
Particle size distribution is described by the percentage of particles that can pass
through sieves of different sizes. Sieve size is expressed in ‘mesh’, which refers to the
number of openings in one square inch. For example, a 50-mesh sieve has fifty
openings in one linear inch (0.3 mm holes).
Generally, lime particles that pass a 100-mesh sieve react quickly in soil. Particles that
pass a 60-100 mesh sieve react within 1-2 years and particles that pass a 20-mesh
sieve react in 2-3 years. Lime particles that do not pass an 8-mesh sieve are
considered ineffective.
Ideally, particle size distribution will allow the liming material to have both immediate
reaction as well as a longer-term reaction.

ENV – Effective Neutralizing Value


Lime materials are often regulated, but quality standards may slightly differ between
countries or states.
The Effective Neutralizing Value (ENV) is an index that combines both purity and
particle size distribution. The percentage of each particle size range is multiplied by
an effectiveness rating. ENV is used for comparing prices of liming materials and for
calculating its application rate. The latter is achieved by dividing the recommended
lime application rate by ENV, expressed as a fraction.
For example, if the recommended application rate of agricultural lime is 2 tons/acre
and a liming material has an ENV of 70%, then the required application rate of that
material would be 2/0.7 = 2.85 tons/acre
ENV can be calculated as follows:
A x 0.4 + B x 0.8 + C x 1.0
ENV = x CCE
100
Where:
A is the percentage of particles passing a 20-60 mesh sieve. 40% of them are
assumed to react in the first year
B is the percentage of particles passing a 60-100 mesh sieve. 80% of them are
assumed to react in the first year.
C is the percentage of particles passing a 100-mesh sieve. 100% of them are assumed
to react in the first year.
Particle size distribution is indicated on the label of the liming material.
Example:
A liming material has the following properties:

132
CCE 85%
% passing 20-mesh sieve 95%
% passing 60-mesh sieve 88%
% passing 100-mesh sieve 70%

A = 95-88 = 7%
B = 88-70 = 18%
C = 70%

7 x 0.4 + 18 x 0.8 + 70 x 1.0


ENV = x 85 = 74.12%
100

As mentioned above, ENV can be used for comparing prices of liming materials. For
example:
Liming material A: ENV 60%, Price $20/ton
Price per ton of effective lime: 20/0.60=$33.33
Liming material B: ENV 90%, Price $25/ton

Price per ton of effective lime: 25/0.90 =$27.7


Although material B has a higher price, it is more effective than material A and,
therefore, its price per ton of effective lime is lower.

133
Soil Water Content
Water is held in soil in the pores between the soil particles, therefore the maximum
amount of water that a specific soil can potentially hold is equal to its porosity (total
volume of pores).

There are three types of soil water: Gravitational water, capillary water and
hygroscopic water. Each type is affected by different forces that act on the water in
soil

Gravitational water is the water that moves through the soil by the force of gravity
and drains. Gravitational water moves in the larger pores of the soil and drains quickly.
Hygroscopic water is a thin layer of water, in a vapor form, held tightly to soil particles
by surface forces. Hygroscopic water is not available for plants.

Capillary water is the water that is held inside soil pores against gravity. The capillary
forces that hold the water inside the pores is a result of the ratio between adhesion
and cohesion forces. Adhesion is the tendency of water molecules to stick to other
surfaces, and cohesion is the tendency of water molecules to stick one to the other.
Capillary forces are stronger when the adhesion is greater than cohesion. Adhesion is
stronger in smaller pores.

Fine-textured soils (clay and clay-loam soils) have a larger porosity compared to
coarse-textured soils (sand). Therefore, they can hold more water than sandy soils.
However, a large portion of the water held in fine-textured soils is not available to
plants. This is because the pores in fine-textured soils are small and hold water more
tightly.

To absorb soil water, plants must overcome the forces that hold the water in soil pores.

Different soil moisture conditions were defined. These moisture conditions give an
indication of the availability of water to plants.

Soil moisture content – a percent volume of water in the soil at a given moment.
In the lab, a known volume of soil is dried and % soil moisture content is calculated in
the following way: % moisture content = (weight of wet soil – wet of dry soil)/(weight of
dry soil) X 100

Saturation – all soil pores are filled with water. This is not an ideal condition for plants,
as plat roots need air.

134
Field capacity – this is the moisture content of the soil after drainage has stopped.
The large pores, that cannot hold the water against gravity are filled with air. By
definition, it is the water content retained in the soil at -0.33 bar.

This is considered to be the ideal moisture condition for plants, as water in this
condition is easily available.

However, in certain soils, maintaining the soil in field capacity may result in lack of
oxygen to the root system or in development of stem and root diseases.

Permanent Wilting point – this is the moisture content of the soil at which plants
cannot absorb the water, as it is held tightly in soil pores. By definition, this is the water
content of the soil at tension of -15 bar.

The difference field capacity and wilting point is the Available Water. When converted
to water amount, it is known as Total Available Water.

For example:

Field capacity: 20%

Wilting point: 13%

Root system depth: 20cm

Calculate the Total Available Water (TAW):

Solution:

% Available water = 20-13=7%

TAW= 0.07 x 1000 mm/m x 0.2m = 14mm (140 m3/ha).

Water content at field Water content at wilting


capacity (mm/mm) point (mm/mm)

Coarse sand 0.1 0.05

Sand 0.15 0.07

Loamy sand 0.18 0.07

135
Sandy loam 0.20 0.08

Loam 0.25 0.10

Silt loam 0.30 0.12

Silty clay loam 0.38 0.22

Clay loam 0.40 0.25

Silty clay 0.40 0.27

Clay 0.40 0.28

Readily Available Water (RAW) – As water is depleted from the soil, the remaining
water becomes more difficult to extract and, at a certain point, the hydraulic
conductivity drops down and water flow towards the roots decreases significantly.

The Readily Available Water is the water that can be easily extracted by the plant. It
is the moisture content of the soil between field capacity and a refill point, which is
obtained by multiplying the Total Available Water by a fraction called “depletion
fraction” (p). The depletion fraction is crop specific. For many crops, the depletion
fraction is set to 0.5 (50%).

136
Chapter 4

Fertilizer
Management

137
Fertilizer Recommendations Philosophies
There are many considerations to take into account when giving nutrient
recommendations. These include the crop type, variety, soil analysis, water quality,
plant tissue analysis, weather conditions, irrigation management, soil type and more.
The goal is to meet the nutrient requirements of the crop.

Because the agricultural ecosystem is complex, different philosophies were developed


and the fertilizer recommendation may vary among crop advisors, even for the same
field.

Generally, there are five common nutrient recommendation philosophies. Some have
a solid scientific basis and others are yet under debate: the Buildup and Maintenance
concept, SLAN (Sufficiency Level of Available Nutrient), Maintenance, BSCR (Base
Cation Saturation Ratio) and the Quantitative method.

The Buildup and Maintenance


In this approach, the goal is to raise the level of nutrients in the soil and maintain their
level for the next few years. After reaching a critical level, the level of nutrients in the
soil is maintained at, or above, that level. In this approach, high rates of fertilizers are
often used until the critical point for each nutrient is reached. Once this level is
reached, lower rates of fertilizers are applied, replenishing the nutrients that were
taken up by the crop.

The critical level is the value of the soil analysis, required to achieve 90% of the
potential yield. Above the critical level, the yield response to fertilizer is not likely and
is less than 10%. Often the ‘high’ category of the soil test is taken as the critical level.
Beyond the critical level, the nutrient removal data for the crop is used.

Building soil nutrient levels may take years to achieve. Soil testing is recommended in
order to ensure that critical levels are not exceeded.

Fertilizer costs are higher, using this approach, and profitability in the first years is,
therefore, lower. The risk of reduced yield due to deficiency is low. However, if not
applied carefully, nutrients may reach toxicity levels in soil.

138
SLAN (Sufficiency Level of Available Nutrient)
Using the SLAN approach, nutrients are applied based on soil test categories such as
Low, Adequate, High and Excessive. The goal is to provide the crop the right amount
of nutrients it requires for the present growing season, while minimizing costs and
maximizing profits.
For example, if the level of a certain nutrient in the soil analysis is categorized as ‘low’,
the recommendation will be to apply a fertilizer rate that provides a nutrient rate higher
than the nutrient extraction by the crop.
If the level of the nutrient is categorized as ‘high’, the assumption is that the crop will
not respond to an application of the nutrient and, therefore, the recommendation would
be not to apply that nutrient.
This approach is widely accepted among universities and laboratories.

The Maintenance approach – nutrient removal


This concept uses the nutrient removal by the crop as a basis for recommendations.
The term ‘nutrient removal’ must not be confused with ‘nutrient uptake’. While ‘nutrient
uptake’ refers to the total amount of nutrients required by the crop, ‘nutrient removal’
refers only to amount of nutrients in the harvested organs (grain, fruit, fodder etc.).
This approach is suitable only when the level of nutrients in the soil is adequate. It is
more suitable for orchards and trees, in which a significant amount of the nutrients
taken up by the crop is recycled and returned to the soil.
Because this approach does not take into account the total uptake of nutrients by the
crop, it might result in depletion of nutrients from the soil and in deficiencies.

BCSR (Base Cation Saturation Ratio)


This approach is based on an assumption that there is an ideal ratio between
exchangeable cations. The percentage of the CEC (Cation Exchange Capacity) of the
soil that is occupied by each major cation is calculated, where both the CEC and the
cations are in units of meq/100g.

K + Ca + Mg + Na
% Base saturation = X 100
CEC
K
% K saturation = X 100
CEC

139
Ca
% Ca saturation = X 100
CEC
Mg
% Mg saturation = X 100
CEC
Na
% Na saturation = X 100
CEC

Generally, according to this approach, calcium level should be 65-85% of the CEC,
Magnesium 6-12% and potassium 2-5%.
Although BCSR is popular in some places, it is still in debate and there is not enough
scientific basis to support it. The interpretation varies depending on the soil type and
CEC.
Often, it results in recommendation for excessive application of nutrients. For example,
in calcareous soils, in which Ca:K ratio is high, using the BCSR approach may result
in a too high potassium recommendation.

The Quantitative approach


Nutrient values obtained in soil analysis are considered as absolute values.
For example, if the potassium requirement of the crop is 120 lbs./acre and the soil
analysis indicates a level of 80 lbs./acre in the soil, the recommendation, using this
approach, would be to apply 40 lbs./acre.
Using this approach may lead to considerable mistakes in fertilizer recommendations.
For example:
Assume a soil test result of 800 ppm calcium, extracted with ammonium acetate.
For a soil with a bulk density of 1.2 ton/m3 and 20 cm sampling depth:
Ca = 1.2 x 0.2 x 800 x 10 = 1920 kg/ha
The calcium requirement, for most crops, ranges from 50 to 300 kg/ha only. Therefore,
according to the quantitative approach, there is an excess of calcium in the soil.
However, this is not the case, as according to the soil test interpretation, a level of 800
ppm of calcium is considered to be low.

140
Yield Response to Fertilizers
Crop yield is affected by many factors, including genetic, environmental and crop
management factors. A distinction must be made between the actual yield of the crop
and its yield potential. The yield potential depends mainly on the genetics of the
specific crop, and it is usually determined under specific environmental conditions. It
is assumed that water and nutrients are not limiting, and that other stress factors, such
as pests, weeds and diseases are effectively controlled.
The actual yield, on the other hand, will depend mainly on nutrition, irrigation, other
crop management practices and the stress factors mentioned above.
The use of fertilizers significantly contributes to yield increase. However, excess
fertilizer may result in yield reduction and negatively affect the environment.
The ‘yield response to fertilizer’ curve describes the relation between the amount of
fertilizer applied and the relative yield, given there are no other limiting factors.
Generally, yield will increase with the addition adequate rates of fertilizers. At first
linearly and then at a decreasing rate, until it reaches the critical level, where the yield
is at its maximum potential. Above that level, any addition of fertilizer will not increase
yield further and will remain the same.
Applying a fertilizer rate that exceeds the critical level results in yield loss due to salinity
or specific toxicity of one of the applied elements. Other adverse effects of over-
fertilization are fertilizer waste, accumulation of salinity in the soil and potential
pollution of water sources. Pollution is caused mainly by nitrogen and phosphorus
leaching or runoff.

141
Yield response curves can also be drawn for individual nutrients. When drawing such
curves, it is assumed that other factors are not limiting, including other nutrients.
In actual field conditions, there are many limiting factors that might restrict the potential
yield. Therefore, different fields, or even different sections of the same field, may have
a local yield response curve.

The law of minimum


According to Liebig’s law of minimum, the least available resource is the one that limits
the yield, even if other resources are at their optimal level. The same is true for
resources that are in excess. Liebig illustrated the concept using a metaphorical barrel,
in which each stave represents a different nutrient. The shortest stave limits the ability
of the barrel to hold water (yield).

Therefore, a deficiency of a single nutrient can restrict the yield of the crop. Up to a
certain threshold level, excess of the nutrient will neither increase nor decrease the
yield. Above that threshold, an excess of the nutrient will decrease the yield, as a result
of a specific toxicity it may cause, or due to imbalance between nutrients.

Boron can be used as a good example to illustrate this concept. The range between
deficiency and toxicity of boron is very narrow. Generally, boron soil level of less than
0.5 ppm is considered deficient, while a level of 2 ppm and above may be considered
toxic.

142
Yield decrease, resulting from excess of nutrients such as nitrogen, phosphorus,
potassium, calcium, and magnesium is usually related to salinity or nutrient
imbalances. For example, excess of calcium might restrict the uptake of potassium;
excess of nitrogen might stimulate excessive vegetative growth, on the expense of
reproductive growth and increase the susceptibility of the crop to disease.

Economic considerations
Achieving maximum yield is not necessarily the most profitable path. The cost of
fertilizer added must be compared with the economic returns expected as a result of
the yield increase. The net return is the difference between the additional income,
resulted from applying the fertilizer, and the fertilizer cost. The optimum fertilizer rate
is the rate at which the net return is maximum.

A practical approach for fertilizer management would be to base fertilizer


recommendations on soil, plant and water analysis, and customize your fertilizer
program according to the specific conditions in your field.

143
Calculating Fertilizer Application Rates
Calculating fertilizer application rates starts with knowing the following:

• The required nutrient application rate


• The fertilizer grade / composition
• The field’s area

Required nutrient application rate - this is the rate of nutrient to apply, or the nutrient
recommendation. For soil applications it is usually expressed in units of lbs/acre or
kg/ha.
The nutrient recommendation should take into consideration the crop nutrient uptake,
soil analysis, the mineral content of the irrigation water and other parameters.
The guaranteed analysis of the fertilizers refers to the guaranteed nutrient content of
the fertilizer and is expressed as a percentage by weight.
It is important to note that both nutrient requirements and fertilizer grade can be
expressed either in the oxide form or in the elemental form.
Nutrients that are usually expressed in the oxide form are:

Nutrient Elemental form Oxide form


Phosphorus P P2O5
Potassium K K2O
Calcium Ca CaO
Magnesium Mg MgO

Total nitrogen, sulfur and micronutrients are usually expressed in their elemental form.
The grade of a fertilizer usually refers to the percentage of the three major nutrients –
nitrogen, phosphorus and potassium. The grade appears as three big numbers on the
fertilizer label.
The first number is the total nitrogen concentration in the fertilizer, the second is the
phosphorus concentration and the third refers to potassium.
In most countries, the content of phosphorus and potassium is indicated in the oxide
form, not in the elemental form, so the grade refers to N-P2O5-K2O and not N-P-K.
For example, a fertilizer with a grade of 13-10-27 contains 13% nitrogen, 10% of
phosphorus as P2O5 and 27% of potassium as K2O.
A fertilizer with a grade of 46-0-0 contains 46% nitrogen only.

144
However, the grade alone does not always provide enough information about the
fertilizer content.
Fertilizer may contain additional nutrients that are not indicated in the grade and,
therefore, it is extremely important to review the fertilizer label.
The information on the fertilizer label includes the complete guaranteed analysis.
For example, the grade of ammonium sulfate is 21-0-0. Looking at the grade alone,
one might think that this fertilizer contains only nitrogen. However, as the fertilizer’s
name suggests, ammonium sulfate also contains 24% sulfur.
The type of information is provided on the label, as part of the guaranteed analysis.

Worth Noting
The nitrogen percentage in the fertilizer grade always refers to total nitrogen and does
not give an indication of the form or forms of nitrogen in the fertilizer. Fertilizers usually
contain nitrogen as ammonium (NH4), Nitrate (NO3), urea (CO(NH2)2), or a
combination of those. The guaranteed analysis should detail the nitrogen forms
available in the fertilizer.

Calculating the fertilizer rate


For solid fertilizers, the following formula can be used:
Fertilizer rate = required nutrient application rate X 100 / % nutrient in the fertilizer
Let’s see a few examples:

145
Example 1
A grower wants to apply 50 pounds of nitrogen to his 6 acres field, using Urea 46-0-0
fertilizer. How many pounds of urea he needs to apply?
Fertilizer application rate = 50 x 100 / 46 = 108.7 lbs

Example 2
A grower applied 80 kg of Mono Ammonium Phosphate (MAP 12-61-0) to his 5-
hectares field. The grade represents the complete guaranteed analysis of the
fertilizers.
Which nutrients did he apply? At what rate per hectare?
The first number of the fertilizer grade is the percentage of nitrogen in the fertilizer
(12%).
The second number refers to the percentage of phosphorus as P 2O5 (61%)
Calculating the nitrogen application rate:
80 = N x 100 / 12
N = 80 x 12 / 100 = 9.6 kg per 5 hectares or 9.6/5 = 1.92 kg/ha nitrogen
Calculating the phosphorus application rate:
80 = P2O5 x 100 / 61
P2O5 = 48.8 kg per 5 hectares or 9.76 kg/ha

For liquid fertilizers, the weight of the fertilizer must be accounted for, because the
fertilizer nutrient content is given as percentage by weight.
For example, Urea Ammonium Nitrate (UAN), a liquid fertilizer with a grade of 30-0-0
contains 30% nitrogen by weight. The weight of UAN is 10.86 lbs. per gallon.
A grower wants to apply 15 lbs. of nitrogen. How much UAN 30-0-0 should he use?
As a first step, we can use the previous formula:
Fertilizer rate = Required nutrient application rate X 100 / % Nutrient in the fertilizer
Fertilizer rate = 15 x 100 / 30 = 50 lbs.
Since 1 gallon of the fertilizer weights 10.86 lbs:
50 lbs. / (10.86 lbs./gallon) = 4.6 gallons.
The formula for calculating liquid fertilizer rates becomes, therefore:

146
Fertilizer rate = (Required nutrient application rate X 100) / (% nutrient in the fertilizer
x W)
Where W is the fertilizer weight.

147
Timing of Fertilizer Application
Fertilizer recommendations consist of rate and timing of fertilizer application. Although
applying the right fertilizer rate is important, the timing of the application is equally
important. Applying the right fertilizer at the right time can significantly affect crop yield,
reduce nutrient losses, and avoid damage to the environment.

Nutrient uptake pattern


The uptake rate of different nutrients varies with the growth stages of the crop. Each
crop requires different nutrient ratios at different times throughout its growth cycle. For
example, N:K ratio is usually higher at the vegetative growth stages, in which the crop
takes up more nitrogen than potassium, and lower at the reproductive stages.

The timing of fertilizer application may affect the availability of nutrients. Therefore, for
optimum growth, nutrients must be available when the crop needs them. Nutrients
applied too early in the growing season might be lost by leaching or runoff, while an
application that is too late might not meet the nutrient requirements of the crop.

DAP

Example: nutrient uptake pattern of potato

Salt tolerance of crops


High fertilization levels directly affect the salinity of the soil solution. Fertilizers differ in
their effect on soil salinity, which can be expressed as Salt Index. Fertilizers with a
high salt index are more likely to increase soil salinity, cause injury to germinating
seeds and seedlings and damage roots. Splitting fertilizer application to smaller
applications reduce the risk of salt injury.

148
Therefore, factors such as the salt tolerance of the crop, irrigation water quality and
soil salinity must be considered when making fertilizer recommendations.

Effect of soil properties on timing and frequency of


fertilizer application
Soil mineral composition and texture significantly affect the ability of the soil to retain
nutrients and, therefore, dictate the required timing and frequency of fertilizer
application.

Soil cation exchange capacity (CEC) – Soil CEC is an index of the soil capacity to hold
nutrients, such as potassium, calcium, magnesium, and ammonium nitrogen. Soils
with a higher CEC can hold and store more nutrients than soils with a lower CEC. In
order to avoid loss of nutrients, soils with a low CEC, such as sandy soils, require
smaller and more frequent fertilizer applications, than soils with a high CEC.

Soil texture – Soil texture and CEC are highly correlated. Coarse-textured soils usually
have a lower CEC than fine-textured soils. However, soil texture also determines the
irrigation schedule. Coarse-textured soils require more frequent irrigations, as they
hold less water than fine-textured soils. A higher irrigation frequency results in a
stronger leaching of nutrients and, therefore, splitting the fertilizer application may be
required in order to minimize losses.

Timing of nitrogen application


Of all nutrients, nitrogen is the most susceptible to losses through leaching,
denitrification, and volatilization. Therefore, timing of nitrogen application is crucial. As
mentioned above, leaching is more prevalent in coarse-textured soils, in which nitrate
losses might reach even up to 50% of the applied nitrogen. Therefore, preplant
application of nitrogen, especially in sandy soils and where rainfall is high, increases
the risk of leaching.

Splitting nitrogen application to better match nitrogen supply to crop requirements


enhances nitrogen use efficiency.

Timing of phosphorus application


Phosphorus is not mobile in soil and does not readily leach or move towards the roots.
Because phosphorus is fixed to soil particles, it is often applied as preplant application.
Phosphorus losses occur mainly through runoff and erosion. Therefore, phosphorus
application should be avoided during periods of high rainfall intensity.

149
Applying phosphorus in smaller rates, through fertigation, increases phosphorus
uptake efficiency because the application of dissolved phosphorus allows the plant to
absorb it before fixation occurs.

150
Pre-plant fertilizer application
Depending on the soil type, fertilizers can be applied to crops before planting, right
after planting or throughout the growth of the crop.
Applying fertilizers prior to seeding or planting is referred to as Basal Fertilization. This
practice can help bring soil nutrients to an adequate level and ensure that the crop has
a sufficient pool of nutrients to use at the time the nutrients are required.
However, over-application of fertilizers too soon, before the crop is planted, might
result in losses in nutrients, waste and even damage the crop due to salt damage.

Fertilizer rate to apply as a pre-plant application


The amount of fertilizer to be applied as a pre-plant application depend on the soil
nutritional status, soil texture and the properties of the nutrients to be applied.
Soil analysis - the decision of which nutrients to apply in the pre-plant application and
at what rates, should be based on soil test results.
Soil texture – light, coarse soils are usually able to hold less nutrients than heavy-
textured soils. Therefore, pre-plant application in sandy soils might result in nutrient
losses by leaching and, as a result the crop might suffer nutrient deficiencies and the
applied fertilizer might be wasted. On the other hand, silt and clay soils can retain
higher amount of nutrients and higher pre-plant application rates can be applied.
Pre-plant application of nitrogen – nitrogen in the nitrate form (NO3-), is not retained
by soil particles and, therefore, might quickly leach to below the root zone of the crop.
Therefore, the recommended pre-plant application of nitrate is up to 30% of the total
nitrogen requirement of the crop, depending on the soil texture. In coarse-textured
soils, pre-plant application might not be effective.
Pre-plant application of phosphorus – phosphorus is generally immobile in soils as
a result of interactions with other elements in the soil, such as calcium, aluminum and
iron. Therefore, pre-plant applications of phosphorus are common. The common
phosphorus rate applied as a pre-plant application is within the range of 50-100% of
the total phosphorus requirement, where 100% is more often applied in silt and clay
soils.
Pre-plant application of potassium – potassium is cation, i.e., carries a positive
charge. Therefore, it can be adsorbed to soil particles. The mobility of potassium in
soil is intermediate. Therefore, the rate of pre-plant potassium application lays within
the range of 20-50% of the potassium requirements.

151
Risks involved in pre-plant fertilizer application
Although pre-plant application is convenient and easier to apply than in-season split
applications, it also involves some risks.
1. Nutrient leaching – as mentioned above, nutrients that are applied too early in
the season, might leach below the root zone, especially in light-textured soils.
High levels of precipitation or irrigation amount that is too high, might result in
nutrients leaching to below the root zone and to nutrient deficiencies.

2. Runoff – phosphorus is not mobile in soil and remains mainly in the topsoil.
Therefore, it might be lost by runoff. Phosphorus lost by runoff reaches lakes
and river and cause an environmental problem.

3. Volatilization of nitrogen as ammonia gas – occurs when urea or ammonium-


based nitrogen fertilizers are applied close to the surface of warm, moist and
high pH soils. Under such conditions, urea and ammonium are converted into
ammonia gas (NH3) and volatilize. Pre-plant application of ammonium and urea
nitrogen sources may lead, therefore, to major nitrogen losses if the nitrogen
fertilizer is not incorporated properly into the soil.

4. Salinity – the application of high rates of fertilizer prior to planting raises the
salinity of the soil. This might damage young seedlings or avoid germination.
The salt index of the fertilizer, which is the measure of the potential of the
fertilizer to damage the crop, should be considered.

5. Toxicity – ammonia applied as anhydrous ammonia or formed as a result of


urea application and ammonium at high concentrations are toxic to seeds and
seedlings. Therefore, the conditions under which urea or ammonia-based
fertilizers are applied and the placement method should be carefully
considered.

Fertilizer placements methods


The placement method can significantly affect the efficiency of the Pre-Plant
fertilization. There are various methods of fertilizer placement which are commonly
used for Pre-Plant fertilization:

Broadcast– Fertilizers are applied uniformly to the soil surface. Broadcasting can be
followed by incorporation into the soil.
Banding– Fertilizer is applied in bands (strips), prior to planting/seeding, near to
where the seeds or seedlings are about to be placed.

152
The Ammonium:Nitrate Ratio
Nitrogen is a unique nutrient, in the sense that plants can absorb it both as nitrate
(NO3-) and as ammonium (NH4+). The total amount of nitrogen absorbed by plants is
usually a combination of these two forms.

Ammonium and nitrate have different assimilation and metabolism pathways. The
application ratio between them affects plant growth, uptake of other nutrients and the
pH of the rhizosphere.

Most crops will benefit if they are supplied with both nitrate and ammonium nitrogen.
However, the ratio between them, that will result in an optimal growth, may vary.

The importance of this application ratio is greater in fertigation of intensive crops,


where fertilizers are applied daily. In less intensive crops, where the application of
nitrogen is done just a few times during the growing season, applied ammonium will
often convert to nitrate and will be used by the crop in this form.

A ratio of 1:1 (50% ammonium and 50% nitrate) is considered high. In order to know
the optimal ammonium:nitrate application ratio, it is important to understand the paths
of nitrogen metabolism in plants and the effect of specific field conditions.

Nitrogen metabolism in plants


The metabolism of ammonium and nitrate take place in two distinct organs of the plant
- while the metabolism of ammonium takes place in the roots, the metabolism of nitrate
occurs in the leaves.

When ammonium is absorbed by roots, it must immediately react with carbohydrates


to produce amino acids. Otherwise, it converts to ammonia (NH3), which is toxic to the
plant and might kill root cells within microseconds. As a result, root tips die, and leaves
may become chlorotic.

Nitrate, in contrast, may be either immediately reduced to ammonium and then


converted to amino acids, or it may be transported to the leaves through the xylem,
where it is reduced to ammonium and reacts with carbohydrates.

Carbohydrates are produced in the leaves, in the photosynthesis process. In order to


become available for ammonium metabolism, they must be first transported from the
leaves to the roots, where the reaction takes place.

The chemical equations describing the production of the amino acid Glutamate are:

153
1 1
From ammonium: NH4+ + HCO3- + 13 C6H12O6 + 3 2O2= C5H9O4N + 4CO2 + 6H2O
1 1
From nitrate: NO3- +H+ + 13 C6H12O6 + 1 2O2 = C5H9O4N + 3CO2 + 4H2O

It is clear, from the above equations, that the metabolism of ammonium requires more
oxygen than the metabolism of nitrate.

Factors affecting the adequate ammonium:nitrate


ratio
The optimal ammonium:nitrate application ratio is crop specific and depends on field
conditions, such as temperature and soil pH.

Temperature
Cell respiration process uses both carbohydrates and oxygen. At high temperatures,
the respiration rate of the plant cell increases, and in addition, oxygen becomes less
soluble in water and its availability decreases.

Therefore, at high temperatures, carbohydrates and oxygen become less available for
ammonium metabolism in the roots. Under such conditions, application of ammonium
nitrogen might result in ammonia toxicity, as explained above.

At lower temperatures, ammonium may become a preferred source of nitrogen. While


carbohydrates and oxygen are available for ammonium metabolism in the roots, the
transport of nitrate to the leaves may become restricted.

Crop type and growth stage


Factors that affect the consumption of carbohydrates will also affect the metabolism
of nitrogen. For example, in leafy vegetables such as lettuces, where most of the
growth is in the leaves, carbohydrates are quickly consumed there and are less
available for ammonium assimilation. Therefore, this type of crop is more sensitive to
high ammonium levels.

During the fruit development of crops, fruits consume more carbohydrates and limit
the amount of carbohydrates that can be transported to roots for ammonium
assimilation. Therefore, the efficiency of ammonium metabolism decreases and a
lower ammonium:nitrate application ratio is required.

Effect of ammonium:nitrate ratio on rhizosphere pH


In order to maintain an electrical neutrality within root cells, plants release hydrogen
ions (H+) into the soil solution in exchange for cations and decrease the pH in the

154
immediate surroundings of the roots. The uptake of ammonium, which carries a
positive charge, results, therefore, in a decrease of the pH around the roots.

In a similar way, the uptake of nitrate, which carries a negative charge, results in an
increase of the pH around the roots, as it triggers the release of hydroxide (OH -) or
bicarbonate (HCO3-) ions.

In soilless culture, where the volume of the roots is relatively large compared with the
volume of the substrate or the nutrient solution, the applied ammonium:nitrate ratio
can significantly affect the pH of the growing medium and of the nutrient solution.

In addition, unlike soil, the buffering capacity of the growing medium is very low and
even negligible. Therefore, in soilless culture it is extremely important to maintain an
adequate ammonium:nitrate ratio.

Effect of ammonium/nitrate ratio on nutrient uptake


Competition with other nutrients: Excess ammonium might reduce the availability
of other nutrients, as it competes for uptake with other cations, such as calcium and
magnesium. This may result in nutrient deficiencies. For example, excess ammonium
at the fruit set growth stage of tomato may cause blossom end rot as a result of calcium
deficiency.

Availability of nutrients: a decrease in the rhizosphere pH, to an adequate level,


may increase the availability of micronutrients.

155
Types of Fertilizers
Fertilizers are compounds that are used in agriculture to supply nutrients for
plants. They can be classified in a variety of ways. A specific fertilizer can belong to
multiple fertilizer types.

Classification according to the nature of the fertilizer


Fertilizers can be either organic or inorganic (mineral).

Mineral fertilizers – also referred to as chemical fertilizers or synthetic fertilizers. This


type of fertilizers is produced from minerals and gasses, using a chemical process.
However, many of them are naturally occurring minerals.

For example, urea is produced by reacting ammonia with carbon dioxide; mono
potassium phosphate is produced by reacting phosphoric acid with potassium
hydroxide or with potash.

Potassium chloride (Muriate of Potash – MOP), a mineral fertilizer by itself, is mined


directly from mineral deposits.

Mineral fertilizers have a high content of nutrients that are readily available for plants.

Organic fertilizers – Organic fertilizers are derived from plant and animal sources.
Nutrient concentration in organic fertilizers is lower than in mineral fertilizers. Since
plants cannot absorb organic nutrients directly, nutrients must be mineralized first. The
mineralization process occurs naturally in soil and is driven by soil microorganisms.

Compost used as an organic fertilizer

156
Granular versus liquid fertilizers
Fertilizers can be either dry or liquid. Dry, granular fertilizers can be easily stored, are
cheaper than liquid fertilizers and usually have higher nutrient content.

Liquid fertilizers can provide a more uniform application and their nutrient availability
is usually higher than in granular fertilizers.

Note that when calculating the application of liquid fertilizers, the weight of the fertilizer
should be taken into consideration, as nutrient content is usually provided in
percentage by weight in all types of fertilizers.

Classification according to the nutrient release rate


Slow-release fertilizers supply nutrients slowly and steadily. This allows the nutrients
to become available to the crop over an extended period of time and minimizes nutrient
losses. The fertilizer granules are coated with a polymer that protects the soluble
fertilizer and controls the nutrient release rate.

Some distinguish between slow-release fertilizers (SRF) and controlled release


fertilizers (CRF). Controlled release fertilizers use a different coating technology. The
claim is that while nutrient release from slow-release fertilizers is affected by many
factors, including soil moisture, temperature, and pH, in control release fertilizers the
nutrient release is affected only by temperature and lasts longer.

Soluble versus insoluble fertilizers


Granular fertilizers can be either soluble or insoluble. Highly soluble fertilizers are
easily dissolved in water and can be used in fertigation. Solubility varies among
fertilizers and with temperature.

Insoluble fertilizers last longer in the soil and are less prone to leaching.

Fertilizer types according to their nutrient content


Fertilizers can be classified according to the major nutrients that they contain. For
example: Granular urea contains 46% nitrogen; therefore, it is considered a nitrogen
fertilizer. Ammonium sulfate is also considered a nitrogen fertilizer, as nitrogen is the
main nutrient it provides.

Micronutrient fertilizers contain one or more of the essential micronutrients: boron,


iron, manganese, zinc, copper, and molybdenum.

157
Micronutrient fertilizers can be further classified as chelated or not chelated. A chelate
is a compound that surrounds the metallic ion and make it available for plants.
Chelated micronutrients are available to plants at a wide range of pH, while non-
chelated micronutrients will be only available in soils with pH <7.0.

Non-chelated micronutrients usually come in a form of sulfate micronutrients. For


example, iron sulfate and copper sulfate.

Label of a chelated iron fertilizer

Compound versus straight fertilizers


Compound fertilizers are a type of fertilizers that contain more than two of the major
nutrients. There are many possible N-P-K combinations that can form compound
fertilizers. For example, 17-10-27 fertilizer contains 17% nitrogen, 10% phosphorus as
P2O5 and 27% potassium as K2O.

Simple fertilizers contain only one or two major elements. MAP, urea, potash, and
ammonium sulphate are examples of straight fertilizers.

158
Urea
Urea fertilizers are widely used in agriculture. They are considered an economic
nitrogen source.

The chemical formula of urea is CO(NH2)2 and, in nature, urea is excreted in the urine
of mammals. Commercial urea fertilizers are produced by reacting ammonia with
carbon dioxide.

In its solid form, urea is provided as either prills or granules. Granules are slightly
larger than prills and are denser. Both prilled and granular urea fertilizers contain 46%
N.

Nitrogen leaching and volatilization rates are usually higher when using the prilled
form. Therefore, granular urea fertilizers are 15-20% more efficient than prilled.

Urea fertilizers are highly soluble (solubility of 1079 g/L at 20ºC). Therefore, in addition
to soil applications, urea fertilizers can be also used in fertigation or as a foliar
application. However, urea fertilizers should not be used in soilless culture, as urea
will immediately leach out of the container.

The NPK grade of a solid urea fertilizer is 46-0-0. Another fertilizer containing high
concentration of urea is Urea Ammonium Nitrate (UAN). UAN is a liquid fertilizer
containing between 28 and 32% nitrogen. 50% of the nitrogen is urea, 25% ammonium
nitrogen and 25% nitrate nitrogen.

Reactions of urea in the soil


Plants cannot absorb urea nitrogen. For the plant to absorb nitrogen applied as urea,
nitrogen must be converted into ammonium (NH4+) and nitrate (NO3-), which are the
nitrogen forms that plants can use.

159
Once applied, the urea fertilizer reacts with water in the soil and with urease, an
enzyme that exists abundantly in soils, and goes through a hydrolysis process, in
which urea is converted into ammonium carbonate.

Ammonium carbonate is then converted into ammonium or to ammonia gas (NH 3),
depending on conditions such as pH, temperature, and soil moisture.

Ammonia gas readily volatilizes from the soil and, as a result, significant losses of
nitrogen may occur if conditions favor formation of ammonia rather than ammonium.

High soil pH and temperature result in greater losses of nitrogen.

Soil pH – High soil pH increases the rate of volatilization, as more ammonium is


converted into ammonia gas.

Soil temperature – high soil temperatures increase the rate of urea hydrolysis, as it
increases the activity of urease. At 70ºC urease becomes inactive due to denaturation.
The hydrolysis reaction is as follows:

(NH2)2CO + 2H2O –> (NH4)2CO3


(NH4)2CO3 + H2O –> 2NH3 + 2H2O + CO2
NH3 + H2 –> NH4+ +OH– (ammonification)
2NH4+ + 4O2 –> 2NO3– +4H++2H2O (nitrification)

As can be noted from the equation, the application of urea fertilizers results in an initial
increase in the soil pH around the applied fertilizer, as hydrogen ions are consumed.
However, the nitrification of ammonium to nitrate, which is carried out by soil bacteria,
results in an a slightly acidifying net effect.

Urea mobility in soil


Since urea molecule is not electrically charged, it readily moves in soil. Leaching of
nitrogen as urea will depend, therefore, on soil humidity and the time until urea
hydrolysis is complete. Once urea is converted to ammonium, leaching reduces,
because ammonium, which carries a positive charge, attracts to the negatively
charged particles of the soil and, therefore, is relatively immobile.

Urea is more mobile than ammonium but a bit less mobile than nitrate.

Application of urea fertilizers


160
Urea fertilizers should be applied carefully. If not applied correctly, nitrogen losses due
to volatilization may occur and, in some cases, urea might cause damage to
germinating seeds.

Urea should be incorporated into the soil by irrigation or rainfall soon after its
application. Application of urea fertilizers to the soil surface without incorporating them
into the soil results in greater losses of nitrogen. Losses are greater in soils of high pH.

Urea fertilizers should be applied when temperature is not too low or too high. Soil
temperatures of 15- 20°C (70°F) are considered adequate.

Using urea fertilizers with urease inhibitors – urease inhibitors reduce the rate of
hydrolysis and, therefore, of ammonia production and volatilization. This allows
additional one or two weeks for incorporating the urea fertilizer into the soil by rain,
irrigation or other means.

Urea fertilizers containing biuret – biuret is a chemical compound with the formula
[H2NC(O)]2NH, which is formed in the manufacturing process of urea fertilizers. In high
concentrations biuret might be toxic to crops.

Most urea fertilizers contain 1.0 to 1.3% biuret, which is considered safe to use.
However, some crops are more sensitive. For foliar applications on sensitive crops,
low-biuret fertilizers (contain approximately 0.25% biuret) should be used. Biuret might
also damage seedlings if the urea fertilizer is placed too close to germinating seeds.

161
Compost: Benefits and Quality Parameters
Compost is a decomposed organic material used as a soil amendment. It is created in
a process called “composting”, which is an aerobic process in which microorganisms
break-down organic waste, either plant material or animal wastes.

Adding compost to soil increases the organic matter content of the soil, including
humus, which is the stable form of organic matter that cannot be further decomposed
by microorganisms.

Using compost has multiple benefits:

Improves soil structure


Improves soil water retention
Increases soil CEC
Contains plant nutrients
Improves the availability of micronutrients
Enhances microbial activity in the soil

Soil structure. The decomposed organic matter binds soil particles together, creating
soil aggregates. In clay soil, humus added with the compost surrounds the clay
particles and cause them to aggregate into larger clumps. Therefore, it improves soil
aeration and water infiltration, enhances root system growth and reduces soil crusting,
which is a result of high sodium levels in the soil.

Water retention. Decomposed organic matter can hold a large amount of water,
relative to its weight. Therefore, adding compost to soil improves the water holding
capacity of the soil.

Plant nutrients. Compost contains a certain amount of essential nutrients. The actual
nutrient content of the compost varies between compost samples, even if they are of
the same type. Most composts contain essential plant nutrients, such as nitrogen,
phosphorus, potassium, iron, copper, zinc, and boron. However, compost cannot
provide sufficient amount of nutrients for commercial crops.

It is important to take into consideration that, in addition to the relatively low


concentration of nutrients, compost releases nutrients over a long period of time –
months or years. Therefore, only a small amount of nutrients is readily available after
compost application.

162
Typical optimum range (% of dry
Nutrient weight)

Nitrogen (N) 1-2

Phosphorus (P) 0.3-0.9

Potassium (K) 0.5-1.5

Calcium (Ca) 1.5-3.5

Magnesium (Mg) 0.25-0.7

Sulfur (S) 0.25-0.8

Availability of micronutrients. The humic substances in compost form chelate


complexes with micronutrients in soil, keeping them in the soil solution and hence
making them available for plants.

Compost quality parameters


Moisture content. An ideal compost would have a moisture content of 40% to 60%.
Lower moisture levels result in limited microbial activity and slower decomposition.
High moisture levels may produce anaerobic conditions and the compost will develop
a bad smell.

Organic matter. A stable compost contains 25% to 65% organic matter, expressed
as dry weight. Less than 25% may indicate that sand or soil were mixed into the
compost. Organic matter content above 65% may indicate that the decomposition was
not completed.

C:N ratio. The ratio of total carbon and total nitrogen in the compost, on a dry weight
basis. For example, C:N ratio of 30:1 means the compost contains 30 times as much
carbon as nitrogen.

163
The initial C:N ratio of the compost is one of the most important properties required for
proper composting.

Microorganisms are using carbon as an energy source and nitrogen as a building block
for proteins. They require more carbon than nitrogen, where the ideal C:N ratio is
between 25:1 to 30:1.

A C:N ratio higher than 40:1 may result in slow decomposition and longer composting
time, as there may be not enough nitrogen for microorganisms to grow. When a
compost with high C:N ratio is added to soil, soil microorganisms will have to find
additional nitrogen sources in order to consume the composted material. This may
result in nitrogen immobilization; less nitrogen will be available for the crop and
increased rates of nitrogen fertilizer may be required.

A C:N ratio lower than 20 means the compost contains more nitrogen, in relation to
carbon. This excess nitrogen may be lost to the atmosphere as ammonia gas, which
also gives the compost a bad smell. Adding a compost with a low C:N ratio to the soil
supply available nitrogen to plants.

Composts with large amounts of green material will have a low C:N ratio, while
composts with more dry material will have a higher C:N ratio.

Salinity. Compost materials contain salts in the form of mineral ions. Due to the high
variability in crop sensitivity to salts, application rate standards based on the salt
content of the compost are not defined. However, compost with EC of 0-2 ds/m can
be safely applied at any application rate.

Compost with EC of 2-4 ds/m can be applied at moderate application rates and if the
EC is greater than 4 ds/m, then application rate should be considered carefully.

164
Slow-release and controlled-release
fertilizers
Slow-release (SRF) and controlled-release fertilizers (CRF) are compounds designed
to supply the crop with nutrients, in a rate that meets its nutrient demand. Unlike quick
release fertilizers (QRF) that quickly dissolve in the soil and provide nutrients for a
relatively short period of time, slow-release, and controlled-release fertilizers release
nutrients over a longer period of time.
The advantages of SRF and CRF include:

• Enhanced nutrient-use efficiency


• Decreased nutrient leaching and nitrogen losses by volatilization
• Decreased environmental pollution risk
• A smaller number of applications is required

Under certain soil and weather conditions, nitrogen losses may occur due to leaching,
volatilization, runoff, and denitrification. Such losses not only affect the crop, but they
also pose a major environmental concern. Nitrate contamination of water resources
and emissions of nitrous oxide to the atmosphere are considered a serious
environmental risk.
Crops take up nutrients at a certain rate throughout their growing cycle. Applying a
large dose of fertilizer at the beginning of the crop cycle may result in nutrient losses
before the crop can use them. Therefore, in order to avoid losses and meet the nutrient
requirements of the crop, quick release fertilizers must be applied in several split
applications.
Slow and controlled-release fertilizers can be an alternative to the split applications.
They improve the nutrient use efficiency (NUE) by releasing nutrients at a slow rate,
extending the availability of the nutrient, and minimizing potential losses. Ideally, the
release rate should match the nutrient uptake rate of the crop, so that nutrients become
available as the crop needs them.
From an economical perspective, although SRF and CRF are more expensive than
conventional fertilizers, reducing the fertilizer application frequency saves labor and
energy costs.

What is the difference between slow-release


fertilizers and controlled-release fertilizers?
Slow-release fertilizers contain nutrients in a form that is not immediately available
to plants. They may contain macro nutrients (nitrogen, potassium phosphorus) as well
as micronutrients (iron, manganese, zinc, copper etc.) . The most common slow-

165
release fertilizers include nitrogen fertilizers in which urea is combined with an
aldehyde. Urea formaldehyde (UF), Nitroform (UF derivative), methylene urea (MU)
are examples of such fertilizers.
To release available nitrogen, SRF must be broken down by microorganisms.
Therefore, the rate of release depends on the activity of microorganisms in soil and,
hence, on soil moisture and temperature.
Other SRF, such as IBDU (Isobutylidendiurea) are not dependent on microbial activity
and decompose by hydrolysis. The slow-release effect is achieved due to their low
solubility.
Plant manures, animal manures and compost are considered as natural sources of
slow-release fertilizers. However, their nutrient release rate is usually very slow and
highly depends on microbial activity.
Controlled-release fertilizers refer to coated fertilizers or to matrices. Coated
fertilizers are coated either with a polymer or with an inorganic materials, such as
sulfur. Using the matrices technique, the matrix is dispersed within the fertilizer and
slows down its dissolution. The materials used for the matrix include rubber,
polyolefins, or gel-forming polymers.
The nutrient release pattern of slow-release fertilizers depends on soil conditions and
on the activity of microorganisms and, therefore, is difficult to predict. The nutrient
release pattern of controlled-release fertilizers can be better predicted, as it is not
significantly affected by soil conditions such as pH, microbial activity, salinity etc., but
rather by soil temperature and the properties of the coating materials.
Sulfur-coated urea is classified as a ‘slow-release fertilizer’ because the nitrogen
release rate might be inconsistent. Cracks in the coating may cause a release of over
30% of the nitrogen immediately.

The nutrient release pattern


SRF and CRF can be further classified according to their nutrient release pattern,
which can be either linear or sigmoidal (S-shaped pattern). The sigmoidal pattern
consists of an initial period in which only a small amount of nutrient is released, a linear
release stage and a final release stage, in which the release rate decreases.
In CRF, the release duration and pattern can be controlled by the selection of the
coating materials, the specific ratio between the polymers, and the thickness of the
coating and particle size. The coating acts as a semi-permeable membrane. The
controlled-release effect is achieved when water diffuses through the coating and
slowly dissolves the nutrients. Diffusion rate and solubility are affected by soil
temperature, where higher temperatures result in a higher release rate.

166
Chelated Micronutrients
The availability of essential micronutrients such as iron, zinc, copper, and manganese
greatly depends on soil pH. At soil pH higher than 7.0, these micronutrients tend to
precipitate as insoluble minerals or oxidize and become unavailable for plants. For
example, ferrous iron (Fe2+), which is the plant-available iron form, readily oxidizes to
the unavailable form of ferric iron (Fe3+). The application of micronutrients in the sulfate
form, directly to a soil that has a neutral or alkaline pH level, is not efficient, as they
readily react with hydroxide ions )OH-) and form insoluble compounds.

Chelates are complexes that consist of a large organic molecule, also called a ligand,
and a metallic ion. The organic compound binds and surrounds the metallic ion and
protects it from reacting with other ions. This way, the micronutrients remain available
for plants even at a high soil pH.

There are different types of ligands. They differ in their stability and cost. The most
common ligands used in agriculture are EDTA (Ethylenediamine tetraacetic acid),
DTPA (Diethytenetriamine pentaacetic acid) and EDDHA (Ethylenediamine-di(o-
hydroxyphenylacetic acid)).

FE(III) EDDHA FE(III) EDTA FE(III) DTPA

Stability of chelates
The stability of a chelate refers to its ability to hold the micronutrient under different
conditions. A ligand that forms a stable chelate with one metal at a given soil pH, might
not be effective with another metal or at different soil conditions.

167
The stability constant is a measure of the affinity of the ligand for a particular metallic
ion. It is expressed as:

[Chelate]
K=
[Metal][Ligand]

Chelates can be compared by comparing their log10(K) value. The higher the log10(K)
value the stronger the affinity. For example, log10(K) of Fe3+ EDTA is 25.0 and log10(K)
of Fe3+ EDDHA is 33.9. Therefore, Fe3+-EDDHA is considered more stable than Fe3+-
EDTA.

Other metals in the solution can displace the chelated metal. For example, at high
concentration, calcium might replace the metallic ion in the chelate. The likelihood that
such replacement would happen can be estimated by comparing the stability
constants of the chelates.

The stability constant equation given above is a general one and describes only the
affinity between the ligand and the metal. Specific stability constants are pH-
depended. Specific stability constants for EDTA and DTPA are given in the tables
below.

EDTA
Log K
pH 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0
Fe3+ 8.2 11.5 13.9 14.7 14.8 14.6 14.1 13.7 13.6
Cu2+ 3.4 6.1 8.3 10.2 12.2 14.0 15.4 16.3 16.6
Zn2+ 1.1 3.8 6.0 7.9 9.9 11.7 13.1 14.2 14.9
Mn2+ -- 1.4 3.6 5.5 7.4 9.2 10.6 11.7 12.6

DTPA
Log K
pH 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0
Fe3+ 6.4 10.1 12.7 14.0 14.5 14.5 14.5 14.5 14.2
Cu2+ 1.6 5.3 8.2 10.3 12.1 13.8 15.8 17.6 18.6
Zn2+ -- 3.2 6.0 8.2 9.8 11.2 13.1 15.1 16.5
Mn2+ -- -- 2.1 4.3 6.3 8.2 12.2 12.2 13.8

168
logK of Fe, Cu, Zn and Mn EDTA vs. pH

As can be seen from the figures above, the stability of the chelate is lower at low pH
levels. An increased concentration of hydrogen ions (lower pH) results in less ionized
ligands and hydrogen ions compete with the metal. In very high pH levels, stability
decreases again.

Plants can absorb chelates from soil or through their leaves. For correction of
micronutrient deficiencies, foliar application is generally more efficient. The ligand
penetrates the mesophyll and releases the micronutrient.

169
Foliar Fertilization
Foliar fertilization, also referred to as foliar feeding, is the application of plant nutrients
directly to the foliage of the plant by spraying the crop with a liquid fertilizer solution.
While it can be used on a large variety of crops, it is more commonly applied in
horticultural crops.
Generally, plants can obtain all their essential nutrients from soil. The nutrient
quantities required by plants are large and cannot be supplied through the foliage.
However, under some conditions and circumstances, the crop may benefit from foliar
fertilization of small quantities of nutrients, mostly of micronutrients. The purpose of
foliar fertilization is, therefore, to supplement soil applications, not to replace them.
Nutrients applied by foliar fertilization are absorbed quicker than nutrients applied
through soil. They penetrate the plant through hydrophilic pores in the cuticle and, to
some extent, also through the stomata.

Why foliar feed?


Foliar feeding can be effective in the following circumstances:
Correcting nutrient deficiencies – Nutrients applied by foliar sprays are absorbed
faster and more efficiently than soil applications. Therefore, foliar nutrient sprays can
provide a quick correction of nutrient deficiencies. However, the correction is
temporary, and a long-term correction should consist of both soil and foliar
applications. The most common nutrients deficiencies that can be corrected with foliar
sprays are of potassium, zinc, boron, manganese and iron.
Inadequate soil conditions – Unfavorable soil conditions restrict nutrient uptake from
soil and, therefore, applying nutrients directly to the foliage can be beneficial.
High soil pH may restrict the uptake of micronutrients and induce deficiencies. Soil
application of chelates may help, but foliar applications can provide a quicker, more
efficient solution.
Nutrient imbalances in soil may cause competition for uptake and restrict the uptake
of some nutrients. For example, excess calcium can interfere with potassium uptake.
Foliar application can bypass the competition and unwanted interactions in the soil.
Waterlogging conditions and excess moisture in soil may cause nutrient deficiencies.
Under such conditions of oxygen stress, roots cannot function efficiently.
Unfavorable environmental conditions – Foliar nutrient applications can improve
crop growth when environmental conditions are unfavorable. Conditions such as
cloudy weather, temperatures that are too high or too low and high humidity impede
nutrient uptake from soil.

170
Damaged or diseased root system – Any conditions that reduce or damage root
activity also reduce nutrient uptake from soil. Damaged roots, or roots affected by
diseases are less effective in nutrient uptake. Foliar feeding can alleviate the negative
effect.
Supplemental application to boost growth – Nutrient requirements vary throughout
the growth cycle of the crop. Supplemental foliar applications of nutrients, at the time
when they are needed, can positively influence growth and yield.

Best practices of foliar feeding


Weather conditions – Temperature, humidity and wind speed affect the efficiency of
the foliar application. Foliar feeding is most efficient at temperatures below 25˚C, high
humidity and when weather is calm. Such conditions often occur in the late evening
hours or early morning.

Under these conditions, plant leaves can better absorb the applied nutrient. When
applied at high temperatures or low humidity, the spray droplets dry quickly and, in
addition, pores on leaf surfaces close. As a result, the absorption rate reduces.

The pH of the spray solution – Acidic pH of the foliar nutrient solution improves the
absorption of nutrients by leaves. A solution pH of around 5.0 is suitable for most foliar
applications. However, the optimal pH varies between nutrients. For example, the
adequate pH range for phosphorus application is 3.0-3.7, ideal pH for zinc is between
4.1 and 4.9 etc.
The pH of the spray solution affects the solubility of the applied fertilizers and the
penetration of nutrients through the cuticle:
1. Solubility - Nutrients must be present in their soluble form in the spray solution
in order to be absorbed by plants. The solubility of many nutrients is higher in
slightly acidic pH and, therefore, at this pH range they remain in the solution
and do not precipitate.

2. Cuticle charge - Plant cuticles contain fixed ionic charges that are pH
dependent and a negative net charge at pH >3.0. Therefore, the pH of the spray
solution can affect cuticle charge and nutrient penetration through it.

Spray droplet size and pressure – Large drops cover a smaller leaf area than small
drops and tend to fall to the ground. Therefore, smaller drops are preferred. However,
droplets that are too small may result in nutrient losses as a result of drift.

171
The droplet size is determined by the type of spray nozzle used and the spray
pressure. Selecting a proper sprayer nozzle and applying the right pressure are,
therefore, important factor in foliar application. However, no specific recommendations
are available for foliar feeding.
Using surfactants – Due to surface tension, spray droplets tend to have a minimum
contact with the leaf and may fall off. Surfactants are materials that reduce the surface
tension and increase the contact surface area of the droplets with the leaves. As a
result, adding a surfactant to the spray solution increases the efficiency of the foliar
feeding.

Without surfactant With surfactant


Photos by Imri Sela

Constraints of foliar feeding


Although foliar feeding is considered an effective way of applying nutrients to plants in
a variety of situations, it has some limitations.

Potential leaf damage – As the spray solution dries on leaf surfaces, the
concentration of salts (the applied fertilizers) increases and may damage leaves.
Leaves may become chlorotic and necrotic if the concentration of the spray solution is
too high or if the spray was applied when conditions are unfavorable or when the plant
is stressed.

Short timeframe for absorption – Once the spray droplets dry on leaf surfaces,
nutrients become unavailable and absorption rate decreases dramatically.

172
Low application rates – Due to phytotoxicity, potential leaf damage and limited
absorption rate through the leaves, foliar application rates must be relatively low and,
therefore, do not meet the nutrient requirements of the crop, especially the
macronutrient requirements. Therefore, foliar nutrition is mostly supplementary.

Cost/benefit considerations – The cost of foliar feeding is often higher than the cost
of soil applications. Furthermore, because of the low application rates, several
applications may be required, which make the costs even higher.

173
Chapter 5

Fertigation and
Soilless Culture

174
Fertigation
Fertigation is the application of fertilizers through the irrigation water. It can be applied
in almost any irrigated crop, in both open field and greenhouses. Fertigation allows
for a more efficient use of both water and fertilizers.

In irrigated crops, water must be applied in a relatively high frequency, depending on


the soil type and water requirements of the crop. Such crops usually have an irrigation
system in place, such as drip irrigation, pivot, or sprinklers.

The use of an irrigation system provides an opportunity for a more efficient fertilizer
application.

Some of the advantages of fertigation are:

• Allows to apply nutrients when the crop needs them


• Increases the availability of nutrients to plants
• Allows for uniform application of nutrients
• Minimizes leaching nutrients, mainly nitrogen in the nitrate form
• Saves labor

Timing of nutrient application in fertigation


In fertigation, plant nutrients can be applied at the time they are needed by the plant.
In other application methods, such as broadcast or band application, split fertilizer
applications are more complex and costly, making them impractical.

Plants absorb nutrients at different rates throughout their growth. Applying nutrients
too early or too late might significantly affect the yield. For example, nitrogen in its
nitrate form, is not retained by soil particles.

Therefore, it tends to easily leach out the soil. If applied too early, it might be lost by
leaching, or volatilization and, as a result, much less nitrogen will be available to the
crop when it actually needs it, later on in its growth cycle.

It was demonstrated in many trials that split application of nitrogen, using fertigation,
results in higher efficiency and higher yields. This is mostly due to minimization of
nitrogen losses.

Phosphorus might be lost by runoff. However, in many crops and soil types, it is
common to apply at least 50% of the phosphorus at pre-plant.

175
An additional advantage of applying smaller fertilizer doses via fertigation, according
to the growth stage of the crop, is that a lower soil salinity can be maintained.

Nutrient distribution and availability


Because in fertigation the fertilizers are delivered in a dissolved form, with the irrigation
water, the distribution of the nutrients in the soil is almost the same as the distribution
of water. This way, nutrients can be applied directly to the root zone. This is particularly
true for nutrients that are not adsorbed to soil particles, such as nitrates (NO 3-) and
sulfates (SO42-).

Other nutrients, such as phosphorus and potassium do not readily move in soil and
will tend to remain in the topsoil. Phosphorus reacts with calcium, aluminum, and other
elements, depending on the soil pH and potassium binds to soil clay particles, as it
carries a positive charge, while soil clay particles have a negative charge.

Because nutrients in fertigation are applied a dissolved form, the availability and
mobility of potassium and phosphorus increase. Best efficiency is achieved when
nutrients are applied uniformly throughout the irrigation duration, as it may avoid salt
accumulation or nutrient leaching.

Drip irrigation systems allow for a more efficient and precise distribution of nutrients,
with minimum waste, as only the root zone of the plant is irrigated. Fertilizer is not
applied in between rows and waste is avoided.

Fertigation methods
There are two main fertigation methods: quantitative and proportional.

Quantitative fertigation is the simplest method. Fertilizers are delivered to the field
through the water, but fertilizer injection is not proportional to irrigation flow rate. As a
result, in most cases, the application is not uniform.

The fertilizer can be dissolved in advance in a tank and injected to the irrigation water
using a simple injector, such as a venturi injector. Or, if a bypass fertilization tank is
used, a solid fertilizer is placed in the tank and dissolves as irrigation water flow
through the tank.

The bypass tank is a pressurized tank, connected as a bypass to the irrigation line.
Part of the water flow is diverted to flow through the tank and dissolve the fertilizer in
it. The concentration of fertilizer in the water is, therefore, high at the beginning of the
irrigation and declines as fertigation proceeds.

176
In both cases, the solubility of the fertilizer must be accounted for. As a rule of thumb,
a water volume of 5 times the tank volume must flow through the tank in order to
dissolve the fertilizer.

If fertilizer solution is prepared and injected into the irrigation water, the amount to
dissolve in the tank must not exceed the solubility limit of the fertilizer.

In quantitative fertigation, application rates are expressed in units of mass or volume,


e.g., kilograms, liters, pounds etc.

Fertilizer bypass tank. Photo by Guy Sela.

Proportional fertigation is a more sophisticated fertigation method. The application of


fertilizers is proportional to the water flow rate and, as a result, the concentration of
fertilizer in the irrigation water is uniform throughout the irrigation. This allows for a
more efficient distribution of fertilizers in the root zone.

This method is suitable for any soil type, but it is most commonly used in sandy soils
and soilless culture. Injectors that can inject a proportional rate of fertilizers, such as
positive displacement pumps, or venturi injectors with solenoid control, must be used.

In proportional fertigation, fertilizer rates are expressed as concentration in water in


units such as mg/L, ppm, mmol/L, and meq/L. Concentration can be converted to an
amount of applied nutrient, using the following equations:

Nutrient amount (kg) = ppm x m3 /1,000

And in imperial units:

Nutrient amount (lbs) = ppm x acre inch / 4.422

177
Example:

How many kilograms of MAP (Mono Ammonium Phosphate) were applied to a 2


hectares field if the concentration of MAP in the irrigation water was 60 ppm and the
irrigation amount applied was 40 m3/ha?

60 ppm = 60 mg/L = 60 g/m3

60 g/m3 x 40 m3/ha = 2,400 g/ha = 2.4 kg/ha

Therefore, the application of MAP on 2 hectares was 2.4 x 2 = 4.8 kg

Note that in proportional fertigation, increasing the water application rate will increase
the amount of nutrient applied, even if the concentration of the nutrient in the water
remained unchanged.

How much MAP is applied per hectare, if the irrigation rate in the previous example
was changed to 50 m3/ha?

60 g/m3 x 50 m3/ha = 3,000 g/ha = 3 kg/ha.

Proportional fertigation systems


There are several types of proportional fertigation systems. Some are mechanical,
without automation or control, and others are fully automated and controlled.
Most modern fertigation systems use fertilizer injectors, which can be adjusted to
apply specific rates of fertilizers.

Controlled systems usually include EC (Electrical Conductivity) and pH sensors to


adjust the injection of fertilizers according to the grower’s preferences.

There are fertigation systems enable application of specific rates of fertilizers to


different sections of the field. The fertigation system and the fertilizer recipe can be
designed in such a way that different fertilizer rates are applied to different sections of
the field, customized to the type of crop, growth stage, soil variations etc. The
customization is done by injecting different rates of fertilizer solutions to the different
sections of the field.

For example, If the fertigation system includes 3 fertilizer stock solutions, applied
nutrient rates can be adjusted by setting different combinations of injection rates. Note
that the composition of the stock solutions remains the same.

178
An automated fertigation system Fertigation stock tanks

Dosatron injectors (proportional, not controlled)

Challenges with using fertigation


Fertilizer solubility - In order to be able to deliver fertilizers through the irrigation
water, only soluble fertilizers can be used.

Note that fertilizers differ in their solubility and that the solubility varies with
temperature. Usually, higher water temperature will result in higher solubility (i.e., a
larger amount of fertilizer can be dissolved in the same volume of water).

Fertilizer compatibility - Furthermore, in fertigation, certain fertilizers are not


compatible or have a limited compatibility and, therefore, must be dissolved in
separate tanks, as reactions between different fertilizers might occur and result in a
precipitation of minerals in the tank.

For example, fertilizers containing phosphorus or sulfur should not be mixed in the
same tank with fertilizers containing calcium.

179
Examples of soluble fertilizers that are commonly used in fertigation

Fertilizer Formula Solubility at 20°C (g/liter)

Potassium nitrate KNO3 209

Potassium sulfate K2SO4 120

Calcium nitrate Ca(NO3)2 1290

Magnesium nitrate Mg(NO3)2 2560

Magnesium sulfate MgSO4 710

Ammonium sulfate (NH4)2SO4 750

Mono ammonium phosphate (MAP) NH4H2PO4 374

Mono potassium phosphate (MKP) KH2PO4 230

Urea NH2 1200

Potassium chloride KCl 264

The distance from the injection point – Many fertigation systems are designed in
such a way that the injection point is at the irrigation head. The distance between the
injection point and the field might be long and it would take longer for the fertilizer to
reach the field. The water that remains in the pipe from the previous irrigation gets to
the field first.
Therefore, If the volume of the irrigation pipe, from the injection point to the field, is
significant, compared to the volume of the irrigation, it may be difficult to precisely
apply the required rates of fertilizer to the field.
Maintaining a uniform concentration of fertilizer in the irrigation water can help
overcome this issue.

180
Hydroponics
Hydroponics is a growing method in which plants are grown in a mineral nutrient
solution, with or without the use of a growing medium.
This method is gaining popularity, as it has many advantages:

• Uses less space


• Saves water, as water is recirculated or can be reused
• There are no weeds
• Plant environment is controlled
• There are no constrains related to soil, such as pH and salinity
• Reduced plant diseases

On the other hand, growing in hydroponics also has some disadvantages:


• It requires deeper knowledge
• There is less room for errors
• Once occurred, diseases can spread quickly
• The initial costs are high

Materials that are commonly used as growing media include perlite, peat, coco peat,
volcanic ash, montmorillonite, rockwool etc.
The main purpose for using a growing medium is to provide support for plant roots.
Growing media do not retain much nutrients and many are inert materials and do not
retain nutrients at all. The nutrient solution is, therefore, the main source of nutrients
for plants grown in hydroponics. It consists of water and mineral nutrient, which are
added with dissolved fertilizers.
In several hydroponic system types, the nutrient solution can be recirculated and re-
used.

Hydroponic systems
Hydroponics systems can have different designs and use different methods. The main
system types are:
• Nutrient Film Technique (NFT)
• Deep water culture
• Ebb and flow
• Wick systems
• Hydroponic drip systems
• Aeroponics

181
Nutrient film technique (NFT) – in this method, plant roots are exposed to a continuous
flow of a nutrient solution film. Plants are placed in planting holes along channels,
where the channels are placed at an angle that allows water to flow by gravity towards
a drain pipe. The thin film of nutrient solution is continuously recirculated.

NFT system

Deep water culture – Plant roots are placed directly in a nutrient solution. Oxygen is
injected to the water.

Deep water culture

182
Ebb and flow – Potted plants are placed in a growing tray. The tray is flooded with a
nutrient solution, until it reaches a certain level. Then, the water drains and can be
reused.

Wick system – A passive hydroponic system in which one side of fibrous nylon ropes
are placed around the plants, in the growing medium, while their other side is placed
in the nutrient solution reservoir. The nutrient solution goes up the wicks to the plants
by capillary action.

Drip system – In these systems, the nutrient solution is delivered to the plants by a
drip system. The plants are placed containers such as bags, pots, slabs, beds etc. The
drainage coming out from the bottom of the container can be recirculated.

Drip-irrigated tomatoes

Aeroponics – Plant roots are suspended in air, while nozzles spray the nutrient solution
directly on the roots at certain intervals.

Aeroponics

183
Criteria for a Balanced Nutrient Solution
A balanced nutrient solution is a solution that meets the following criteria:
1. It provides the crop with all the essential nutrients.
2. The nutrients are at adequate concentrations in the solution.
3. The ratios between nutrients and nutrient forms are adequate.
4. The concentration of elements that might be toxic is below the toxicity threshold
for the crop.
5. The pH of the solution is within the adequate range.
6. The EC of the solution is within the adequate range for the crop.

Providing the crop with all the essential nutrients


In hydroponics, the nutrient solution is often the only source of nutrients for the plant.
Therefore, it must contain all essential nutrients, regardless of whether the nutrients
are added with fertilizers or were present naturally in the source water.

The nutrients are at adequate concentrations in the solution


Each crop has specific nutrient requirements at different growth stages. Environmental
conditions, such as temperature, light intensity and relative humidity also affect the
required nutrient concentration, as they affect the transpiration rate, photosynthesis,
and potential yield.
Nutrient concentrations in the solution must be in an adequate range for each crop
and growth stage, considering the environmental conditions.
Specific data is not readily available for each crop and conditions and, therefore, the
use of standard nutrient solution compositions is common. An example of two such
standard solutions is given in the table below. Different strengths of the solutions is
used for different crops, where the maximum strength is used for crops such as
tomato, pepper and cucumber and a diluted solution (1/2 or 1/4 of the original
concentration) for crops with a lower nutrient demand, such as lettuce, parsley,
strawberries etc.
Nutrient Arnon & Hoagland, 1940 Steiner, 1961
(ppm) (ppm)
N 210 168
P 31 30
K 235 260
Ca 200 185
Mg 48 48

184
S 64 111
B 0.5 0.5
Fe 1.5 1.35
Mn 0.5 0.6
Zn 0.05 0.11
Cu 0.02 0.127
Mo 0.01 0.2

Ratios between nutrients and nutrient forms


A balanced nutrient solution will have adequate ratios between nutrients and between
nutrient forms. The optimal ratios vary with the growth stage of the crop. For example,
for many crops, such as tomatoes, cucumber and pepper, the optimal N:K ratio in the
vegetative stage is 1:1, while in the fruit set stage, the required ratio increases to 1:2
– 1:2.5, as more potassium is required for fruit development.
In leafy crops, the ratios may remain 1:1 throughout the growth cycle.
The ratios between elements that might compete for uptake must be taken into
consideration. Excess of one element can result in the deficiency of another. This can
occur with ions that have similar charge and size.
For example, Ca:Mg, K:Mg, K:NH4. Excess of calcium, magnesium can cause
potassium deficiency and vice versa.
The ammonium to nitrate ratio (N-NH4:N-NO3) is unique, as it is a ratio between
different forms of the same nutrient. Excess ammonium might be harmful for many
crops, including tomato and strawberries, as it can cause toxicity. It may also reduce
the pH in the root system. The optimum ammonium:nitrate ratio for most crops is
approximately 1:10. Acid-loving crops, such as blueberries and azaleas, will benefit
from a higher ammonium:nitrate ratio.

The concentration of toxic elements


Elements such as chlorides, sodium, boron manganese, zinc, iron, fluorides can cause
phytotoxicity if their concentration exceeds the threshold tolerated by the crop.
Groundwater may contain high levels of chlorides, sodium, boron, and fluorides. In
certain circumstances the water must be treated prior to using it for the nutrient
solution. Such treatment is usually costly.

185
The EC of the nutrient solution
The EC of the nutrient solution is used to monitor the total salt concentration in the
solution. Since nutrients are applied in the form of salts, the EC of the nutrient solution
can give an indication on the total concentration of nutrients. However, it does not
provide information about specific nutrients and elements that are not required by
plants, such as sodium, that also contribute to the EC level. The EC of the nutrient
solution must be within a range that does not exceed the salinity threshold of the crop.
A balanced nutrient solution will contain all essential nutrients at adequate levels and
its EC level will be within the recommended EC level for the specific crop. The EC of
the nutrient solution must be monitored frequently. In closed hydroponic systems,
where the nutrient solution is recirculated, an increase in the EC of the solution may
indicate accumulation of undesired elements and that the nutrient solution has to be
refreshed.

The pH of the nutrient solution


A balanced nutrient solution will have its pH within the optimal range for the crop. In
hydroponics, the pH of the nutrient solution can fluctuate as a result of root activity.
Therefore, it is important to monitor it frequently. The stability of the pH depends also
on the alkalinity of the water. Nutrient solutions that have a low bicarbonate
concentration will be more prone to fluctuations.
Examples for optimum ranges of EC and pH values for different crops are given in the
table below.

Crop Optimum EC range Optimum pH range


(ds/m)
Basil 1.0 – 1.6 5.5 – 6.5
Broccoli 2.8 – 3.5 6.0 – 6.5
Cabbage 2.5 – 3.0 6.5 – 6.8
Celery 1.8 – 2.4 6.3 – 6.7
Chives 1.8 – 2.4 6.0 – 6.5
Cucumber 1.7 – 2.2 5.5 – 6.0
Eggplant 2.5 – 3.5 5.5 – 6.5
Lettuce 0.8 – 1.5 5.5 – 6.5
Melon 1.8 – 2.5 5.5 – 6.0
Parsley 1.2 – 1.8 5.5 – 6.5
Pepper 2.0 – 3.5 5.5 – 6.5
Spinach 1.8 – 2.3 5.5 – 6.5
Strawberries 1.0 – 1.4 5.5 – 6.5
Tomato 2.0 – 4.0 5.5 – 6.5

186
Calculating Nutrient Solution Formulas
A nutrient solution is the mix of irrigation water with added nutrients. It is mainly used
in soilless culture and hydroponics, where the nutrient solution is the only source of
nutrients to the plants.

Calculating and balancing the nutrient solution formula might be confusing. Following
the proper steps is the best way to avoid mistakes.

Step 1 – Test your source water


The source water may contain essential nutrients, which must be accounted for.
Therefore, it is important to test the irrigation water before calculating the fertilizers to
be added to the nutrient solution.

Water source Recommended tests


Groundwater EC, pH, K, Ca, Mg, SO4, B, HCO3, Na,
Cl, Fe
Surface water EC, pH, Ca, Mg, HCO3, Na, Cl
Desalinated water EC, pH, Ca, Mg, SO4, B, HCO3, Na, Cl

Note that the lab may provide the test results in different units, such as ppm, mg/l,
meq/L, mmol/L.

To simplify the calculation, it is recommended to convert all values to ppm. See


appendix I.

Step 2 – Obtain the nutrient requirements of the crop


Determine the nutrient rates your crop requires. You can use recommendations from
the literature, or your own experience.

If nutrient requirements are provided in units other than ppm or mg/l, it is


recommended to convert the values to ppm (1ppm = 1 mg/l).

Step 3 – Calculate the nutrient rates to be added with


fertilizers
To do so, simply deduct the source water test results from the nutrient requirements
of the crop. Repeat that for each nutrient.

187
For example, if magnesium requirement is 60 ppm and the source water contain 40
ppm, then 60-40=20 ppm of magnesium must be added with fertilizers. This means
20 milligrams of magnesium must be added to each liter of nutrient solution.

A negative value means your source water contains more of the nutrient than the crop
needs and, therefore, you should not use fertilizers that contain that nutrient.

Step 4 – Make a list of your available fertilizers


The fertilizer you select must contain all nutrients that need to be added, as per the
calculation in step 3.

Check which fertilizers you have available that contain those nutrients.

Step 5 – Calculate fertilizer rates


Begin with the fertilizer that contains a unique nutrient, that other fertilizers do not
contain. For example, if the only available source of calcium you have is calcium
nitrate, start the calculation with this fertilizer.

For solid fertilizers:

FR = 100 x NA / %N

Where:

FR is the fertilizer rate to apply in ppm (or mg/l).


NA is the required nutrient concentration to be added by this fertilizer in ppm (1 ppm
= mg/l = 1 g/m3).
N is the concentration of the nutrient in the fertilizer.

For liquid fertilizers:

FR = NA / (% N x D x 10)

Where D is the density of the fertilizer in kg/L and FR is the fertilizer rate in of ml/L or
L/m3.
And when the density of the fertilizer (D) is given in lbs/gallon, the formula becomes:

FR = NA / (%N x D x 11.98)

Where FR is the required application rate in gal/1000gal.

188
The percent nutrients that a fertilizer contains is usually given as a percent weight to
weight and, therefore, dividing by the density of the fertilizer is required.

Example:

Calculate the rate of magnesium sulfate (a solid fertilizer), required to apply 20 ppm of
magnesium (Mg).

Magnesium sulfate contains 9.1 % Mg (Magnesium) and 14% S (sulfur). Therefore:

NA = 20 mg/l

N = 9.1

FR = 100 x 20 / 9.1 = 220 mg/l

Therefore, adding 220 mg/l of magnesium sulfate to the nutrient solution would add
20 ppm of magnesium.

Using the same formula, the rate of sulfur added by this fertilizer can be calculated:

220 = 100 x NA / 14

NA = 220 X 14 / 100 =30.8 ppm = 30.8 mg/L sulfur

Step 6: Repeat the calculation to balance all nutrients


Now, the right combination of fertilizers that balances all the required nutrients, must
be found.
Since for some nutrients one may have to use multiple fertilizer sources, it is
recommended to prepare an Excel table that includes the fertilizers to be used and the
rate of nutrients each of them contributes.

189
Important to remember:

1. The concentration of the nutrient in the fertilizer may be given in the oxide form.
Pay attention for which form you do the calculation. For example, MAP 0-52-34
contains 52% K2O and 34% P2O5. To calculate the rate of the fertilizer required to
apply certain concentrations of K and P, you should first convert % K 2O and P2O5
to % of K and P.

2. 52% P2O5 = 22.7% P


38% K2O = 31.6% K

3. Nitrogen may be applied in two different forms – nitrate (NO3-) and ammonium
(NH4+). The ratio between these two forms is of a great importance in nutrient
solutions. Therefore, you may want to break down the nitrogen calculation to these
two forms. For example, MAP (mono ammonium phosphate) contains 12% N-NH4,
while potassium nitrate contains 13% N-NO3. This means the nitrogen in MAP is
in its ammonium form, while in potassium nitrate it is in the nitrate form. Some
fertilizers may contain both forms.

3. Often, the pH of the nutrient solution has to be lowered. This is done by adding
acid to the nutrient solution. The three most common acids are nitric acid,
phosphoric acid and sulfuric acid. As can be suggested from their names, those
acids contribute nutrients to the nutrient solution. Nitric acid contributes
nitrogen, phosphoric acid – phosphorus and sulfuric acid – sulfur. These
nutrients must be accounted for when doing the calculation.

190
Closed Hydroponic Systems
In open hydroponic systems, excess nutrient solution is disposed to the environment.
This results in wastage of nutrients and water. Furthermore, nutrients, mainly nitrogen
and phosphorus, might end up in groundwater or surface water, causing serious
environmental problems.
On the other hand, in closed hydroponic systems the nutrient solution is recirculated
continuously. Closed systems include several types, such as NFT, raft/pond systems,
reusing drainage collected from containers and ebb and flow. This type of systems
allows for substantial savings in water and fertilizers, while avoiding environmental
pollution. Recirculating the nutrient solution is becoming legally required in numerous
countries.
However, recirculating the nutrient solution creates several challenges. Unlike open
systems, where a fresh nutrient solution is applied with each irrigation cycle, the
composition of the recirculated nutrient solution changes over time. Nutrients and
elements can accumulate in the nutrient solution or be depleted from it.
This might result in salinity build up, accumulation of harmful ions and nutrient
imbalances. Therefore, mismanagement of the system might cause yield reduction or
negatively affect crop quality.

A raft system

191
Accumulation of harmful ions and salinity build up - If present in the source water,
ions such as sodium and chloride may accumulate in the nutrient solution. Sodium is
considered harmful to plants when present in concentrations higher than 50 ppm.
Chloride is required by some crops in very small quantities, but its concentration in
water may be significant and chloride accumulation can be toxic to plants.
Accumulation of these ions and of other ions that the crop does not take up may result
in an increase in the electrical conductivity (EC) of the nutrient solution to a level that
exceeds the salinity threshold of the crop.
Nutrient imbalances – The crop takes up nutrients at a certain ratio. Often, the initial
nutrient solution does not contain nutrients at the exact ratios as required by the crop.
As a result, the nutrient solution might become unbalanced – the concentration of
certain nutrients might increase in the solution, while the concentration of others may
decrease. For example, the uptake rate of nitrogen, phosphorus and potassium is
usually higher than the uptake rate of nutrients such as calcium, magnesium and
sulfur. This causes the balance between nutrients in the nutrient solution to change as
the nutrient solution is recirculated.

Nutrient management strategies


To maintain an adequate nutrient balance and avoid buildup of harmful elements,
growing in closed hydroponic systems requires applying several strategies.
Analysis of the source water supply – Before deciding on a recipe for the nutrient
solution, the composition of the water supply must be known. Some elements that may
be present in the water, such as calcium, magnesium, sulfur, and boron are
considered essential plant nutrients and must be accounted for. Furthermore, the
grower must be aware of the presence of elements, such as sodium and chlorides,
that might accumulate in the nutrient solution.
Disposal of a fraction of the nutrient solution – In this practice, a small fraction of
the nutrient solution is disposed on a continuous basis, in order to avoid accumulation
of unwanted ions. The specific fraction that must be disposed depends on the quality
of the source water supply and should be determined by periodic lab analysis of the
nutrient solution. In some cases, the entire nutrient solution must be replenished.
Provided that water quality is good enough, disposal may not be required for crops
with a short growing cycle.
Monitoring EC and pH levels – EC and pH of the nutrient solution must be routinely
monitored, and corrections must be made, in order to maintain the required EC and
pH levels of the solution. Corrections can be made by adjusting the EC and pH of the
nutrient solution, with or without automation.
The pH level of the nutrient solution also changes over time due to ion exchange
between plant roots and the nutrient solution. Changes might be rapid so close

192
attention must be paid to fluctuations in pH. Ammonium/nitrate ratio is one of the main
causes for pH fluctuations.
Testing the composition of the nutrient solution –EC measurements do not
provide detailed enough information on the concentration of individual ions in the
solution. Therefore, a periodic lab analysis of the nutrient solution is required. The
analysis can tell which nutrients are accumulated and which are depleted. Nutrient
ratios and disposal fraction can be adjusted based on the test results.

193
Fertilizer Solubility and Compatibility
Soluble fertilizers are fertilizers that can be dissolved in water. The solubility is the
maximum amount of fertilizer that can be fully dissolved in a defined volume of distilled
water and at a given temperature. For example, the solubility of Mono Potassium
Phosphate (MKP 0-52-34) is about 230 grams per liter at 20˚C (31 oz/gal at 68˚F).
The term ‘solubility’ is used mainly when fertilizers are applied through the irrigation
water, in fertigation or hydroponics. Fertilizers are dissolved in either stock solutions
or directly in the nutrient solution. Fertilizer stock solutions are concentrated fertilizer
solutions that are injected into the irrigation water. Precipitation of undissolved
fertilizer, or solid compounds, that are formed as a result of adding fertilizers to the
water, may result in reduced nutrient availability and in clogging of emitters, fertilizer
injectors and other parts of the irrigation system.
Any addition of fertilizer beyond its solubility limit, will result in a formation of a solid
precipitate, as the additional amount will not dissolve.
Not all fertilizers are considered soluble. Fertilizers such as superphosphate, slow-
release fertilizers, most organic fertilizers and amendments such as gypsum and lime,
are considered insoluble. In fact, all mineral fertilizers are soluble to a certain extent,
as in order to become available to plants, nutrients must be dissolved in the soil
solution. However, ‘insoluble’ fertilizers cannot be used in fertigation.
Some fertilizers may also contain insoluble residues or impurities, such as clay, silt
and sand particles.
Water temperature greatly affects the solubility of fertilizers and the dissolution rate.
The solubility of most fertilizers increases with water temperature. This means that at
a higher water temperature, a larger amount of fertilizer can be dissolved in the same
amount of water.
When dissolved in water, the dissolution reaction may result in either decrease or
increase of the water temperature. The reaction is called ‘endothermic’ when heat is
absorbed and water temperature decreases, or exothermic, when water temperature
increases. This change in water temperature, resulted from the dissolution of the
fertilizer, may in turn affect solubility of the fertilizer and its dissolution rate.

194
Solubility g/l
Fertilizer / Temperature (C˚) 5 10 20 25 30 40

Potassium nitrate 133 170 209 316 370 458

Ammonium nitrate 1183 1510 1920 . . .

Ammonium sulfate 710 730 750 . . .

Calcium nitrate 1020 1130 1290 . . .

Magnesium Nitrate 680 690 710 720 . .

MAP (Mono Ammonium Phosphate) 250 295 374 410 464 567

MKP (Mono Potassium Phosphate) 110 180 230 250 300 340

Potassium chloride 229 238 255 264 275 .

Potassium sulfate 80 90 111 120 . .

Urea 780 850 1060 1200 . .

The solubility product constant


The solubility product constant (Ksp) describes the maximum concentration of fertilizer
that can be dissolved. The higher the Ksp, the more soluble the fertilizer is.
Consider the following dissolution reaction:
aA (s) →bB (aq) + cC (aq)
Where A is an undissolved solid (fertilizer) and B and C are dissociated ions which
constitute the solid fertilizer.
a, b and c are the number of moles of each specie.
*Reminder: Mole is unit of amount, referring to the number of particles of a substance,
and equals 6.022 x 1023.
The Ksp for this reaction is defined as:
Ksp = [B]b[C]c
Where B and C are the concentrations of the ions in a saturated solution in units of
mole/liter (Molar). Therefore, adding any amount of one of the ions, B or C, will result
in a precipitation of A.
195
Example:
What is the solubility of potassium sulfate at 25˚C if the solubility product is 1.32?
K2SO4 (s) → 2K+ (aq) + SO42-(aq)
Ksp = [K+]2 [SO42-]
One mole of K2SO4 dissociates to two moles of K+ and one mole of SO42-.
Since there are twice as many moles of K+ than of SO42-:
1.32 = [2X]2 [X]
1.32 =4X3
X = 0.691
Where X is the number of moles.
The molecular weight of K2SO4 is 174 grams/mole
Therefore, the solubility of potassium sulfate at the given temperature is:
0.691 mole/liter x 174 grams/mole = 120.2 grams/liter

The common ion effect


Solubility data of fertilizers is given for fertilizers dissolved in distilled water. The
presence of ions in the water may affect the amount of fertilizer that can be dissolved.
If the water already contains the same ion as the ion added with the fertilizer, the
solubility of the fertilizer will decrease.

Consider following dissociation reaction of potassium sulfate:

K2SO4 (s) → 2K+(aq) + SO42- (aq)


The reported solubility of potassium sulfate in distilled water at 25˚C is 120 grams/liter.
However, if the water already contains potassium or sulfate ions, a precipitation will
occur before this concentration is reached.

Since the solubility equilibrium must be maintained, dissolving potassium sulfate in


water that already contains either sulfate or potassium will shift the equilibrium to the
left, towards precipitation of potassium sulfate.

This is also true for dissolving two or more fertilizers in the same stock solution. For
example, mixing magnesium nitrate and magnesium will reduce the solubility of both
fertilizers.

196
Fertilizer compatibility
Two fertilizers are considered compatible if they can be dissolved in the same stock
solution without causing precipitation, as long as their solubility is not exceeded.
Fertilizers that may form insoluble compounds when mixed together, or cause an
unwanted chemical reaction, are considered incompatible.

For example, preparing a concentrated solution that contains magnesium sulfate and
calcium nitrate will most likely result in a fast precipitation of calcium sulfate. Therefore,
these two fertilizers are not compatible and must be mixed in different tanks or applied
in separate irrigation cycles.

The compatibility of different fertilizers is given in the following chart:

Common practices
The interactions between fertilizers and the irrigation water, as well as between the
one fertilizer and another may be complex and difficult to evaluate or calculate.
Therefore, it is recommended to perform a “jar test”. In this easy to perform test,
fertilizers are mixed in a small volume of water (using a jar, beaker ,or a bucket), using
the same dilution rate as in the stock solution or in the nutrient solution.

Watch if any precipitates form on the bottom of the jar, or if the water appears milky or
cloudy. If no precipitates occur within an hour, the mixture is safe to use. If turbidity
occurs, it is likely that emitters or other parts of the irrigation system may become
clogged using that particular fertilizer mixture.

197
A “jar test”. Photo by Guy Sela.

For example, consider a 500 liters tank in which you would like to dissolve 30 kg
potassium sulfate and 50 kg potassium nitrate.
How should the jar test be performed if you have a 1-liter jar?
Potassium sulfate – 30 kg / 500 liter = 30,000 g / 500 L = 60 g/L
Potassium nitrate – 50 kg / 500 liter = 50,000 g / 500 L = 100 g / L

Therefore, dissolve, 60 grams of potassium sulfate and 100 grams of potassium nitrate
in the jar, mix well and wait 1-2 hours. If a formation of a precipitate occurs – the
solution must be diluted.

198
Fertilizer Stock Solutions
Fertilizer stock solutions are concentrated fertilizer solutions that are prepared by
dissolving fertilizers in water and are injected into irrigation water.
To avoid precipitation on the bottom of the stock tank, fertilizers must be fully dissolved
and remain in the solution. Otherwise, deposits might clog parts of the fertigation
system. Furthermore, the precipitated nutrients are no longer available for plants.
At the same time, the fertilizer solution must be concentrated enough (‘strong’ enough)
so that the injected amount would meet the nutrient requirement of the crop.
Several considerations must be taken into account when preparing stock solutions:

• Interactions between fertilizers


• Solubility limit of the different fertilizers
• Water quality
• Required fertilizer rates
The above considerations should also be part of the design phase of the fertigation
system, as they are also used to determine design parameters, such as the number
of tanks required and injector flow rates.

Number of tanks required


The minimum number of tanks required is determined by the compatibility of the
fertilizers used. Incompatible fertilizers must be separated and dissolved in separate
tanks, as mixing them in the same tank will result in a precipitate.
For example, fertilizers containing calcium are incompatible with fertilizers containing
phosphorus or sulfur and mixing them together may result in a precipitation of calcium
sulfate or calcium phosphate. Therefore, if calcium nitrate and potassium sulfate are
included in the fertilizer recipe, then at least two tanks are required.
It is recommended to make a list of the fertilizers included in the fertilizer recipe and
divide them into groups, according to their compatibility restrictions.
Often, acids are used to lower the pH of the irrigation water. Using acid in one of the
nutrient tanks is possible, subject to compatibility and fertilizer stability restrictions.
However, it is recommended to use a separate tank for the acid, as it allows for more
control over the pH and better monitoring.
Example:
How many tanks are required if the fertilizer recipe consists of potassium nitrate,
calcium nitrate, mono ammonium phosphate (MAP) and magnesium sulfate?

199
Compatibility check:
Fertilizer Compatible with Incompatible with
Potassium nitrate Calcium nitrate, mono
ammonium phosphate,
magnesium sulfate
Calcium nitrate Potassium nitrate Mono ammonium
phosphate,
magnesium sulfate
Mono ammonium Potassium nitrate Calcium nitrate,
phosphate (MAP) magnesium sulfate
Magnesium sulfate Potassium nitrate Calcium nitrate, mono
ammonium phosphate

Therefore, the possible distributions are:


Tank 1 Tank 2 Tank 3
Potassium nitrate,
Option 1 Calcium nitrate Magnesium sulfate
MAP
Potassium nitrate,
Option 2 MAP Magnesium sulfate
calcium nitrate
Potassium nitrate,
Option 3 Calcium nitrate MAP
magnesium sulfate

As it can be noted from the tables above, at least three tanks are required.

Fertilizer amounts in the tanks


The amounts of fertilizers to dissolve in each tank depend on the:
1. The required fertilizer rate
2. The solubility of the fertilizers
3. Tank volume
4. Injector flow rate

For proportional fertigation, the amount of each fertilizer in the tanks can be calculated
using the following equation:
Ci x V
Fi =
IR x 1,000

200
Where
Fi – The amount of fertilizer i the tank (kg)
Ci – The desired concentration of fertilizer i in the nutrient solution (g/m 3)
V – Tank volume (L)
IR – Injection ratio (L/m3)

The injection ratio is defined as the ratio between the flow rate of the injector and the
irrigation flow rate. It can be expressed as a ratio (e.g., 1:100, 1:200), or in units such
as L/m3, gals/100 gals and percent (%).
For example, if the irrigation flow rate is 40 m3/h and the injector flow of 100 L/h, then
the injection ratio is:
100 (L⁄hr.) L
In L/m3: = 2.5
40 m3 ⁄ℎ𝑟. m3

100 L⁄hr.
In %: x100 = 0.25%
40,000 L⁄ℎ𝑟.

100 𝐿
As a ratio: −→ 1: 400
40,000 𝐿

Injection ratio can be converted to irrigation duration, using the following equation:
Injection duration (min.) = (I X D X IR) / f

Where:
I = Irrigation flow rate (m3/h)
D = Irrigation duration (min.)
IR = Injection ratio (L/m3)
f = Injector Flow Rate (L/h)

Given the above injection ratio, a tank volume of 500 L and a required fertilizer
concentration of 400 g/m3, what is the amount of fertilizer that should be dissolved in
the tank?
CxV 400 x 500
F= = = 80 𝑘𝑔
IR x 1,000 2.5 x 1,000

201
The solubility limit
The fertilizer concentration in the stock tank must not exceed the solubility limit of the
fertilizer. Therefore, for each fertilizer:

Fi x 1,000 / V must be < The solubility of fertilizer i

The solubility of the fertilizer from the previous example, must therefore be greater
than 80 x 1,000 / 500 = 160 g/L.
If the concentration of one of the fertilizers exceeds its solubility limit, then the solution
must be further diluted, and the injection ratio must be increased (the more diluted the
stock solution is, the higher the injection ratio should be, in order to reach the same
nutrient concentration in the nutrient solution).
Remember that the solubility of a fertilizer depends on water temperature and on other
fertilizers dissolved in the same tank. Dissolving fertilizers that contain a common
nutrient in the same tank will lower the solubility of both fertilizers. This is known as
the common ion effect.

A precipitate formed as a result of mixing incompatible fertilizers or exceeding solubility limits. Photo by: Guy Sela

Some rules of thumb:

• If the solubility of a fertilizer is unknown – consider a solubility of 200 g/L.


• If fertilizers containing a common nutrient are dissolved in the same tank - dilute
the stock solution by 20%.
• Always separate calcium fertilizers from sulfur and phosphorus.
• Avoid mixing magnesium with fertilizers containing ammonium phosphate.
• Different nitrate sources can be mixed; however, the common ion effect should
be considered.
202
• Always add acid to water, not water to acid!
• The stock solution should be stirred well.

203
Fertilizer Injectors
Fertilizer injectors are used in fertigation systems to inject fertilizer solutions and acids
into the irrigation water. The injectors draw a small amounts of fertilizer solution from
the fertilizer stock tanks and inject them into the irrigation line at a certain ratio.
The ratio of stock solution injected to irrigation water is referred to as the “injection
rate” or “injection ratio”.

Because both the injected stock solution and irrigation water are liquids, the injection
ratio is expressed as either percent or as a volume:volume ratio. For example, an
injection rate of 1:100 means that 1 part of stock solution is injected into 100 parts of
irrigation water.

Generally, the two main types of injectors used in most fertigation systems are venturi
injectors and positive displacement injectors.

Venturi injectors
Venturi injectors use differential pressure to create suction. Water enters the venturi
and flows through a tapered constriction. The pressure in the outlet is lower than in
the inlet. This pressure difference creates a vacuum at the suction inlet.

Venturi injectors do not need an external energy source. They are easy to use and
have a relatively low cost. However, the injection rate is not constant and might change
with fluctuations in pressure.

This type of injectors is used mainly for quantitative fertigation. However, automated
fertigation systems also use venturi injectors. EC, pH and pressure control allow for
an accurate injection of fertilizers. Most automatic control systems use venturi injectors
that are controlled by solenoid valves. Often, the injectors are connected to a
rotameter that measures the flow rate of the injector.

204
An automated fertigation system

It is important to make sure that the injector has a capacity to inject the required rate
of stock solution. This depends primarily on the irrigation flow rate:

Required injector flow rate = Injection rate x Irrigation flow rate

For example, if the required injection rate is 5 L/m 3 (0.5 %) and irrigation flow rate is
20 m3/h (88 GPM), then the injector flow rate must be at least 5 L/m 3 x 20 m3/h = 100
L/h. And in imperial units: 0.5/100 x 88 GPM = 0.44 GPM.

In case the capacity of injector does not meet the required injection rate, either or both
of the following measures can be considered:

• Decreasing the irrigation flow rate. For example, by reducing areas that are
irrigated simultaneously.
• Increasing the concentration of the stock solution, if possible, considering
solubility limitations.

Positive displacement injectors


Positive displacement fertilizer injectors provide precise and controlled injection rates.
The injector draws the required percentage of stock solution and mixes it with the
water. Injection is proportional to the volume of water entering the injector and the
same proportion is kept over a wide range of pressures and water flow rates. Examples
of this type of injectors include Dosatron, Gewa and Dosmatic (all brand names). In
these injectors water pressure is used as a power source.

205
There is a large selection of positive displacement injectors that can provide different
injection rate ranges and operate under a wide range of pressures and flow rates.

Examples of positive displacement injectors and their properties

Injector Injection range Operating Water flow range


pressure
Dosatron D25 RE5 1 – 5% 0.3 – 6.0 bar 10 L/h – 2.5 m3/h
(0.04 – 11 GPM)
Dosatron D25 RE 2 0.2 – 2% 0.3 – 6.0 bar 10 L/h – 2.5 m3/h
(0.04 – 11 GPM)
Dosatron D20 S 0.2 – 2% 0.12 – 10 bar 1 m3/h – 20 m3/h
(4.4 – 88 GPM)
Dosmatic MiniDos 12-1% 0.2 – 1.0% 0.4 – 10 bar 30 L/h – 2.7 m3/h
(0.13 – 12 GPM)
Dosmatic SuperDos 20 – 2.5% 0.2 – 2.5% 0.3 – 35 bar 10 L/h – 4.5 m3/h
(0.04 – 20 GPM)

206
Calibration of Fertilizer Injectors
In order to inject the right amount of stock solution, fertilizer injectors must be properly
calibrated. Several methods can be used for calibration, depending on the injector type
and system setup.

Calibrating positive displacement injectors


Positive displacement injectors allow for adjusting the injection rate to the required
level, where the injection ratio is expressed either as percentage or as a ratio (e.g.,
1:200 or 0.5%).

Dosatron injectors. Photo by Guy Sela.

The following method can be used for calibrating positive displacement injectors.

Equipment needed:
• Graduated beaker
• 5-20 liters (1.3 – 5 gallons) bucket

Procedure:
1. Fill the graduated beaker with water or stock solution. Use about 0.5 – 1 liters
(15-35 ounces).
2. Place the suction tube of the injector inside the beaker. If the tube has a filter,
make sure that it is fully covered with water/stock solution.
3. Remove air bubbles by running water through the irrigation system, allowing
the injector to operate.
4. Once air bubbles are removed, fill up the beaker up to the required volume.
Record the volume of solution in the beaker: V1beaker .
5. Turn on water again. Collect the water in the bucket.

207
6. Close the water once the bucket is full or when a sufficient amount of solution
was taken up by the injector.
Record the volume of solution in the beaker again: V2beaker
Record the volume of the water collected in the bucket: V bucket
7. Calculate the injection ratio using the following formula:

Vbucket 𝑥 𝐾
Injection ratio =
(V1beaker − V2beaker )

Where
Vbucket is the volume of the water collected in the bucket
V1beaker is the volume of solution in the beaker before the injection - step 4
V2beaker is the volume of the solution in the beaker after step 6
K equals 1,000 when Vbucket is in liters and Vbeaker in milliliters
K equals 128 when Vbucket is in gallons and Vbeaker in ounces

Example:
Vbucket = 4 Liters
V1beaker = 500 mL
V2beaker = 420 mL

Injection ratio = (4 x 1,000) / (500-420) = 50

Therefore, the injection ratio is 1:50 or 1/50 x 100 = 2% = 20 L/m3

8. If the calculated injection ratio equals to the injection ratio set on the injector,
then the injector is calibrated.
9. If not: check that water pressure and flow rate are within the suitable range for
the injector. Increase or decrease the injection ratio according to the calibration
results.

Calibration of automated control fertigation systems


In automatic control systems, injectors are usually adjusted to a specific flow rate. The
same flow rate must be defined in the controller settings. This type of systems also
consist of a water meter that measures the irrigation flow rate and EC and pH
electrodes that measure the EC and pH of the water going to the field.

With both irrigation and injector flow rates given, the controller can calculate the
number of injection pulses required in order to meet a defined injection ratio. This is
further adjusted by EC and pH control, where the number of injection pulses is
increased or decreased, in order to maintain the target EC and pH set in the controller.

208
Some systems allow setting a fertilization program by defining injection ratio, target
EC and target pH. Others enable only EC and pH control, and do not include an option
to set an injection ratio.

To calibrate an automatic control system of the first type, follow the below steps:

Equipment needed:
• EC and pH calibration solutions
• EC and pH meters
• Graduated beaker
• 5-20 liters (1.3 – 5 gallons) bucket

Calibration process:

1. Calibrate the EC and pH electrodes, according to the instructions provided by the


manufacturer.

Calibration of EC and pH electrodes. Photo by: Guy Sela

2. In systems with roatmeters, operate the system and adjust the injector to the
required flow rate, using the flow control valve on the rotameter.

209
3. Make sure that the injector flow rate, defined on the controller settings, matches
the actual flow rate of the injector.

EC and pH electrodes

Flow control valve

Suction tube

Photo by Guy Sela

4. On the controller settings, define a fertilization program (injection ratio, target EC


and pH). Use just one of the dosing channels if a single injector is being tested.
5. Perform a jar test to make sure that the defined injection ratio results in the target
EC and/or pH levels.
6. Fill the graduated beaker with water or stock solution. Use about 0.5 – 1 liters (15-
35 ounces).
7. Place the suction tube of the injector inside the beaker. If the tube has a filter, make
sure that it is fully covered with water/stock solution.
8. Remove air bubbles by running water through the irrigation system, allowing the
injector to operate.
9. Once air bubbles are removed, register the volume of solution in the beaker.
10. Operate the irrigation for about 5-10 minutes, or until a sufficient amount of solution
was taken up by the injector.
11. Register the volume of water applied, as it appears on the controller.
12. Calculate the actual injection ratio, using the same formula as above, where V bucket
is replaced with the applied water volume taken from the controller (step 10).

Since automation compensates for deviations in the injector flow rate and in the
correlation between injection rate, EC and pH, steps 1-5 are the actual calibration of
the system, while steps 6-12 are used to validate it.

210
Controlling the Irrigation Water pH
In fertigation, adjusting the pH of the irrigation water to be within an optimal range is
of a great importance. Often, the water pH has to be lowered, but there are cases, like
when using desalinated water, in which measures to avoid the pH from dropping must
be taken.

Except of reducing the risk of emitter clogging, maintaining the pH within the range of
between 5.5-6.7 is important in order to improve the availability of nutrients, such as
phosphorus, iron, zinc and other micronutrients.

Plants can take up nutrients only from the soil solution or from the nutrient solution. A
nutrient that has precipitated out of the solution is no longer available to plant uptake.

A solid precipitate forms when the solubility of a mineral salt is exceeded. The pH of
the irrigation water affects both the solubility of minerals as well as the formation of
less soluble forms compounds. For example, at a pH level higher than 7.0, calcium
and magnesium carbonates, iron minerals and other compounds tend to precipitate
and become unavailable for the crop.

Adjusting the pH of the irrigation water in fertigation is not likely to change the pH of
the soil, especially if the soil has a high buffering capacity (e.g., clayey or calcareous
soil).Therefore, adjusting the irrigation water pH to increase nutrient availability is
more efficient in soilless culture.

Precipitated minerals form deposits which might accumulate in emitters (e.g., drip
system) and clog them. As mentioned above, the pH of the nutrient solution will
determine whether precipitation will occur, depending also on the composition of the
nutrient solution and other elements in the irrigation water.

Acidifying the irrigation water


The amount of acid to add to the irrigation water depends on:
1. The type of acid used
2. The concentration of the acid
3. The alkalinity of the water
4. The initial pH of the water

Upon adding acid to the water, it will first react with carbonates in the water and the
pH will decrease gradually. Once most of the bicarbonates are neutralized, the pH will
decline sharply. Therefore, it is recommended to add an acid amount that will still keep

211
a certain concentration of bicarbonates in the water, in order to avoid a sharp drop of
the pH. A level of 60 ppm bicarbonate is often recommended.

Application of acid through fertigation system with pH control – Automated


fertigation systems, that have pH control, allow the user to define a target pH. The
controller then keeps the irrigation water pH at the desired level. In this type of
systems, an injector is connected to a tank containing acid. A pH sensor is installed
on the irrigation line or on a bypass. The controller adjusts the number of pulses
applied by the injector, based on the actual water pH and the target pH defined for the
controller.

EC and pH sensors on a fertigation system. Photo by Guy Sela.

Most systems have settings for stopping the irrigation if the measured pH is below or
above a certain threshold. This is particularly important in soilless culture.

The efficiency of the pH adjustment by the control system is greatly affected by the
concentration of the acid in the tank and the injector flow rate. For example, if the acid
concentration or the injector flow rate are too low, the controller may not be able to
lower the pH and vice versa. If either the concentration of the acid or the injector flow
rate are too high, the pH might not be stable and may fluctuate.

212
In fertigation systems without a controller, both the injector flow rate and acid
concentration have to be adjusted carefully, as the adjustment is done manually.

It is recommended to perform a jar test prior to injecting acid into the irrigation water.

Acid injection patterns


1 2 3
pH

Irrigation time

In the above figure, irrigation water pH is plotted against the irrigation time for three
different acid injection patterns. Although the average pH is the same for all three,
pattern 3 represents a more uniform and stable pH throughout the irrigation.

Example:

Jar test results showed that in order to lower the pH of his irrigation water to the desired
level, a grower must inject 300 ml of acid per 1 cubic meter of irrigation water (3.84
oz/100gal).

Flow rate of the acid injector: 50 L/h (26.4 gal/h)


Irrigation flow rate: 40 m3/h (176.1 GPM).
Average irrigation duration: 30 minutes
Acid tank volume: 100 L

What would be a proper dilution rate of the acid in the tank?

Irrigation amount in 30 minutes = 0.5 hr. x 40 m3/h = 20 m3

Amount of acid required: 300 ml/m3 x 20 m3 = 6000 ml = 6 L

In 30 minutes, the injector will inject 25 liters of solution into the irrigation water. These
25 liters must contain 6 liters of acid, in order for the pH to be lowered to the desired
level.

Therefore:
100 L (tank volume) / 25 L = 4
4 x 6 liters of acid = 24 L

For a uniform acid application, 24 liters of acid must be diluted in a 100 L tank.

213
Acids as fertilizers
In agriculture, the most commonly acids used to lower the water pH are nitric, sulfuric
and phosphoric acid. These acids contribute nitrogen, sulfur and phosphorus,
respectively. Nutrients added with the acid must be taken into consideration in the
nutrition plan of the crop, as the amount added may be considerable.

For example, 100 ml of 85% phosphoric acid contribute 45 grams of phosphorus.


Acidifying the irrigation water using phosphoric acid at a rate of 50 ml/m 3 will result in
22.5 ppm phosphorus in the irrigation water. This phosphorus concentration is more
than half of the common concentration used in soilless culture.

To convert this concentration to amount of phosphorus applied, multiply it by the


irrigation amount. For example, if 20 m3 of water are applied, then 20 m3 x 22.5 gr/m3
= 450 grams of phosphorus are applied with each irrigation.

214
Growing Media and Their Properties
Growing in greenhouses enables producing high quality, out of season vegetables,
fruits, and ornamentals.
Many greenhouse growers have shifted from growing in soil to growing in soilless
culture in containers. Growing in soilless culture enables growers to successfully grow
crops where soil conditions are inadequate, it allows for better management of water
and nutrient availability and better control of plant pathogens.
On the other hand, growing plants in confined containers restricts root growth and,
therefore, requires more intensive and careful care.
While growing hydroponically, in a nutrient solution, is suitable for leafy green
vegetables and low crops, such as strawberries, crops such as tomatoes, peppers,
cut flowers etc. require support for their root system.
Growing media are materials used for growing plants in containers. They are often
also referred to as ‘substrates’. Growing media consist of a mixture of components
and their physical structure should provide plants with water, air and support. Certain
types of growing media also serve as a reservoir for nutrients.
These properties are determined by the components used and the ratios between
them.
There are many types and mixtures of growing media. Using the correct media mix
and handling it properly can greatly affect plant performance.

Criteria for choosing growing media


When choosing a growing media, it is important to consider the following properties:
Bulk density – The bulk density is the weight of a given volume of the medium and
measured in g/cm3, kg/m3 or lbs./ft3. The bulk density of the media mix should be high
enough, to allow the medium to physically support the plant, and low enough, so that
the medium can be transported and handled easily. Bulk density of growing media is
much lower than the bulk density of soil.
The bulk density range of various substrate materials is given in the table below.
Material Bulk density (g/cm3)
Light peat 0.06-0.1
Dark peat 0.1-0.15
Perlite 0.08-0.12
Rockwool 0.08-0.09
Vermiculite 0.09-0.15

215
Porosity and pore size distribution –The porosity of growing media determines the
available space for water, air, and root growth. The growing medium must be able to
store a sufficient amount of water to meet plant needs and, at the same time, have
enough pores filled with air.
Gas exchange is required for root respiration and plant growth. Inadequate oxygen
supply to roots can cause nutrient disorders, such as ammonium toxicity, manganese
toxicity and iron deficiency.
Pores are the spaces between the solid components of the growing medium. They
can be filled either by air or by water. The total porosity is defined as:
Total porosity = Air porosity (%) + Water Holding Porosity (%)
The air porosity refers to the volume of pores that are occupied by air, after the growing
medium was allowed to drain, and the water holding capacity is the volume of pores
that remain filled with water after gravitational water drained.
An ideal growing medium should have a total porosity of 60-80% and air porosity of
15-25%.

Material Porosity (%)


Light peat 90-95
Dark peat 85-90
Perlite 85-90
Rockwool 94-97
Vermiculite 90-95

The size distribution of the pores is one of the most important properties of the growing
medium. It determines the proportion between air porosity and water-holding porosity.
Due to capillary forces, smaller pores (micropores) retain water stronger than larger
pores (macropores). Capillary forces involve two types of attraction – cohesion and
adhesion, where cohesion is the attraction between water molecules and adhesion is
the attraction between water and the solid components of the medium. In small pores,
the solid particles around the pore are closer one to each other and capillary forces
are higher. This enables micropores to hold water against gravity.
In macropores, capillary forces are weaker than the gravitational forces and, therefore,
they drain more easily. Because water is held in pores against gravity, the amount of
work required to extract it from the pores is called ‘tension’.
A good growing medium will contain both micropores, for water holding capacity, and
macropores, for drainage and aeration.

216
Classification of pore sizes

Pore type Radius


Macropores >75 micrometers
Mesopores 30-75 micrometers
Micropores 3-30 micrometers
Ultra-micropores <3 micrometers

Not all the water in the growing medium is available for plants. Water that is held
tightly in very small pores cannot be taken up by plant roots. Therefore, water
that is held at tensions between 1 and 5 kPa (1.0 - 50 cm) is considered readily
available water. Water held at tensions between 5 and 10 kPa (50-100 cm) is the
water buffering capacity and is not readily available to plants. Water held at
tensions above 100 cm are considered unavailable.

Gravitational forces in soil are the main driver of drainage. However, in containers,
which are evidently lower in depth/height than soil, the gravitational forces are less
dominant in driving drainage to provide proper aeration. Therefore, the physical
properties of the growing medium must be such that allow for proper drainage.
The tension at the bottom of the container equals to zero and increases along the
container height. The tension is zero at the bottom of the container because the
drainage holes at its bottom expose the medium to the atmospheric pressure (taken
as zero).
Due to the relatively high number of larger pores, that can drain more easily, the water
content of the growing medium varies along the height of the container, where the
medium at the top of the container is drier than the medium at the bottom of it. Larger
pores at the top of the container will drain, while pores with the same size, at the
bottom of the container may remain filled with water.
Water retention curves describe the relationship between the volumetric water content
of the growing medium and the tension.

217
Water retention curves of two growing
100
media
Volumetric water content (%)

80

60

40

20

0
0.1 1 10
Tension (kpa)
Medium A (Pine Bark) Medium B (Peat)

It can be seen from the curves above, that medium A (pine bark), drains faster than
medium B (peat). For example, at a tension of 1 kPa (10 cm) above the bottom of the
container), medium A holds 30% water, while medium B holds 60%. Most probably,
medium B contains a higher portion of small pores than medium A. The total volumetric
water content of the medium in the container, at water holding capacity, is the integral
below the curve, from 0 tension to a tension equal to the height of the container.
The container shape and size determine the actual amount of water that the growing
medium can hold.
This is demonstrated in the image below. Both container A and container B have the
same volume and are filled with the same growing medium. However, because water
in both containers will rise to the same height, the amount of water they can hold is
different, and container A will hold more water than container B.
Container shape and height, therefore, significantly affect both air/water ratio in
the substrate and irrigation scheduling.

218
Air

Container A Container B

Hydraulic conductivity – The hydraulic conductivity of the growing medium is its


ability to transmit water. It is one of the most limiting factors for water uptake when
growing in substrates. If the transpiration rate of the plant exceeds the hydraulic
conductivity of the growing medium, water uptake will be restricted.
The hydraulic conductivity of substrates used in soilless culture is very high when the
substrate is wet. However, it declines sharply as the substrate dries out because larger
pores empty first and the continuity of water in the medium is interrupted. As a result,
water will tend to flow in ‘preferential paths’, through the larger pores, and leach out of
the container.
Therefore, when the substrate is too dry, it is a better practice to apply small amounts
of water at a high frequency, rather than applying one large amount.

219
In-house Nutrient Monitoring in Container
Plants
The root system of plants grown in containers is restricted to the container size and
occupy a large portion of the growing medium. As a result, the root volume to substrate
volume ratio is much higher than in soil.

In addition to the high root:substrate ratio, the buffering capacity of the growing
medium is significantly lower than of soil. Therefore, changes in the rhizosphere are
rapid and frequent monitoring is required.

In order to avoid salt accumulation in the substrate, excess water is applied. The
excess water comes out of the container through holes at the bottom of the container.
Depending on the quality of the water and the type of substrate, an excess of 15-30%
is usually required.

Testing the drainage water coming out of the container is a useful tool for both irrigation
and fertilization management. Decisions about fertilization levels and irrigation rates
can be made based on simple on-site tests.

A routine, weekly analysis of both the drip water and drainage is considered a best
practice. Parameters such as EC, pH and nitrate levels can be easily tested in the
greenhouse. Portable meters and simple testing kits can be used for that purpose,
without having to send samples to the laboratory. It is recommended to test at least
the levels of EC and pH of both drip and drainage water.

A complete analysis should be performed in the following cases:


• Periodically
• When a diagnosis of a problem is required.
• For new crops whose nutritional requirements are not known to the grower.

EC and nutrient monitoring


The concentration of salts (including dissolved nutrients) in the drainage solution
varies throughout the irrigation. The first drops coming out from the bottom of the
container will have a different EC than the drops at the end of the irrigation.

The reason for that is that salts keep leaching out of the container as the irrigation
proceeds. Assuming uniform wetting of the growing media, the EC of drainage will
eventually be equal to the EC of the drip solution.

220
Therefore, the amount or the percentage of drainage, out of the volume of the drip
solution, must be accounted for when measuring the samples.

In order to learn about trends in the growing medium, it is essential to compare


samples that were obtained at the same percentage of drainage. Comparing the
amount of salts or individual nutrients in both drip solution and drainage solution can
give an indication of accumulation of salts or depletion of nutrients.

Let’s see a few examples:

Example 1:

Assume a container with a single dripper.


Drip solution EC: 1.0 ds/m
Drainage solution EC: 1.3 ds/m
Drip solution volume (full irrigation sample): 600 ml.
Drainage solution volume: 100 ml (100/600 x 100 = 16% drainage)
Using the general rule of thumb that 1 ds/m = 640 mg/L TDS, the concentration of
salts in the drip solution is 640 mg/L.

Therefore, 800 ml drip solution contain 0.8 x 640 = 512 mg of salts.


Doing the same for the drainage solution:
1.3 x 640 = 832 mg/L
832 mg/L x 0.1 L = 83.2 mg

From this example we can see that although the drainage EC is higher than the drip
EC, there is no indication of accumulation of salts, as 512 mg were applied but only
83.2 mg leached out. In fact, these results may indicate a potential nutrient deficiency.
Now, let’s assume that the same amount of salts that came out of the container was
obtained in 300 ml of drainage:
83.2 mg / 0.3 = 277.33 mg/L
And the estimated EC is 277.33/640 = 0.43 ds/m

In terms of drainage percentage, the first sample was obtained at 100 ml/ 800 ml =
12.5% and the second at 300 ml / 800 ml = 37.5%.

Example 2
For the previous example, what would the drainage EC that indicates zero
accumulation or depletion be, if the drip solution volume is 800 ml and the drainage
volume is 300 ml?

Zero accumulation or depletion means that the amount of salts that came out of the
container is equal to the amount of salts entered. Therefore:
221
ECdrip x Vdrip = ECdrainage x Vdrainage

1.0 x 0.8 = ECdrainage x 0.3

ECdrainage = 2.66 ds/m

For simplicity, the following general rules of thumb can be used:

Results % Drainage Interpretation


ECdrainage ≈ ECnutrient solution + 0.5 ds/m 20-50 Desirable
ECdrainage > ECnutrient solution + 0.8 ds/m 20-50 Salt accumulation and/or
excess fertilization.
ECdrainage < ECnutrient solution -- Indication for a potential
nutrient deficiency
ECdrainage > ECnutrient solution +0.5 <20 Excess fertilization or
insufficient irrigation

Drip water sampling. Photo by: Guy Sela

pH Monitoring
pH level in the growing medium is dynamic, and greatly depends on the nutrient
balance in the nutrient solution and on the buffering capacity of the water and the
growing medium. For example, ammonium to nitrate ratio is one of the main factors
affecting the pH in the rhizosphere, where a high ratio may result in a decrease in pH.

222
Therefore, the pH of the nutrient solution is rarely the actual pH in the root zone.

The pH of the drainage water can provide an indication of these dynamics. If it drops
to values that are too low (<5.0 for most crops), then the ratios between nutrients in
the nutrient solution should be assessed.

In some cases, increasing the buffering capacity of the nutrient solution may be
necessary. For example, when using desalinated water. This can be achieved by
adding potassium bicarbonate to the fertilizer recipe.

223
Chapter 6

Irrigation

224
Water Requirements of Crops
Plants absorb water in the transpiration process. Water is transferred to the
atmosphere, from the surface area of its leaves, where larger leaves surface results
in higher rate of water absorbed and transferred to the atmosphere.
The water requirement of a crop is the amount of water the crop requires in order to
meet both the transpiration demand and the water lost to the atmosphere from the soil
and leaf surfaces (evaporation). The combined process of transpiration and
evaporation is called evapotranspiration (ET).
Weather conditions, which include temperature, relative humidity and radiation,
provide the driving force for evapotranspiration.

Reference evapotranspiration (ET0)


Reference evapotranspiration is defined as the evapotranspiration from a hypothetical
grass reference crop with an assumed height of 0.12m, surface resistance of 70 s m -
1 and albedo of 0.23.

The surface resistance is related to water vapor diffusion through the stomata. Albedo
is a measure of reflectivity, the higher the albedo the more solar radiation is reflected
back from the surface.
There are several methods used to determine the reference evapotranspiration. The
most common methods are the evaporation pan method and the Penman-Monteith
equation. In the evaporation pan method, a standard pan, with specific dimensions is
used. There are several types of pans. For example, the “Class A” pan is cylindrical
and has a diameter of 120.7 cm and a depth of 25 cm. The Sunken Colorado pan is
square and has dimensions of 0.92 x 0.92 x 0.46 m.

The pan is filled with a known amount of water. The daily amount of water evaporated
from the pan is then measured in units of millimeters or inches.
Environmental conditions at the specific location of the pan, such as wind speed and
humidity, also influence evaporation.
The rate of water evaporation from the pan is multiplied by a coefficient, K pan, in
order to account for the variability in pan types and environmental conditions.
The reference evapotranspiration is therefore calculated as:

225
ET0 = K pan x E pan
Where
Epan – pan evaporation (mm/day)
K pan - pan coefficient

An evaporation pan

The Penman-Monteith equation is a model used to approximate ET0 based on energy


balance and aerodynamics, using also surface resistance.

G = soil heat flux density (MJ m-2 day-1)


Rn = net radiation at the crop surface (MJ m-2 day-1)
ρa = air density (kg m-3)
Cp = specific heat of dry air (MJ kg-1 ˚C-1)
es0= mean saturated vapor pressure (kPa)
ea = mean daily ambient vapor pressure (kPa)
rav = bulk surface aerodynamic resistance for water vapor (s m-1)
rs = the canopy surface resistance (s m-1)
ˠ = psychrometric constant (kPa °C-1)
Δ = slope vapor pressure curve [kPa °C-1]

226
The FAO (Food and Agriculture Organization) recommended the following
simplified equation:

In which:
T = mean air temperature (°C)
u2 = wind speed (m s-1) at 2 m above the ground

Crop evapotranspiration (ETc).


Since the actual crop evapotranspiration is different than the reference
evapotranspiration, ET0 is multiplied by a crop coefficient, Kc, to obtain the crop
evapotranspiration, ETc.
Kc is determined by the crop type, its growth stages and climate conditions. Crop
properties that are integrated within the crop coefficient include stomata density and
position, crop height etc. Kc values for a specific crop vary throughout the growth of
the crop, taking into account leaves area, ground cover and water requirements at the
specific growth stage.
Therefore, the ETc for a specific crop usually increases as the crop grows. Tall crops
with high planting density may have Kc values can larger than 1.
Kc values range between 0.3 and 1.25.
For example, the Kc of a certain crop at its initial growth stage is 0.3, and ET0 is 7mm.
The water requirement of the crop on that specific day is, therefore:
ETc = Kc*ET0 = 0.3 X 7 = 2.1 mm

227
Irrigation Scheduling Using Soil Water
Budget Approach
Irrigation scheduling is defined as the frequency and amount of water applied to the
crop through irrigation. There are several methods that can be used for irrigation
scheduling, including the soil water budget approach, measuring water tension in the
soil and using soil moisture sensors.

The key to optimal irrigation scheduling is applying water at the right time and in the
right amount, while avoiding water stress or excess irrigation.

There are several methods for irrigation scheduling. The method presented here
involves doing the balance between the estimated soil moisture content and the
amount of water lost from the soil by transpiration from plants’ leaves and by
evaporation from the soil surface.
It requires knowledge of the soil type, its water holding capacity and the actual
evapotranspiration, considering the development stage of the crop.
The soil provides the medium in which water is stored. Soil voids are filled either with
air or water. Since each soil has a different distribution of particle sizes, the size of the
pores, and as a result, the amount of water that they can hold varies. For example,
sandy soils can hold less water than clay soils.
However, not all the water in the soil is available to plants. A certain amount of water
is held strongly inside tiny pores and cannot be extracted by the plant. Therefore, for
each soil, the ‘available water’ should be determined. The available water is defined
as the fraction of the soil volume between field capacity and permanent wilting point.
For example, a value of 0.2, means that when the soil is saturated, the available water
amount equals to 20% of the soil volume.
Multiplying the available water fraction by the relevant soil volume gives the Total
Available Water. Soil volume = area X active root system depth.
For example, if the active root system reaches 30 cm, and available water fraction is
0.2, then:
TAW (Total Available Water) = 0.2 X 10,000 m2/ha X 0.3 m = 600 m3
When the soil is sufficiently wet, the water supply to the crop is high enough meet its
water demand, and water uptake is equal to ETc.
As the water content of the soil decreases, the soil water becomes more difficult to
extract.

228
When the soil water content falls below a certain threshold value, the soil water can
no longer be transported quickly enough towards the roots to meet the
evapotranspiration rate. The fraction of TAW that a crop can extract from the root zone
without suffering water stress water is the Readily Available Water.
Although this fraction is crop specific and depends also on crop evapotranspiration,
ETc, a value of 0.35 can be used as a rule of thumb.
Therefore, Readily Available Water (RAW) = 0.35 X TAW.
Since the RAW provides the available water in the soil, it also determines the
frequency of the irrigation applications, or the gap between irrigations.
In order to know when to irrigate, we now need to understand how much water is
being depleted, or what is the water depletion rate.
The depletion depends on the amount of water that is absorbed by the plant and the
amount of water that is evaporated from soil which is not covered by plants.
The process in which water is transferred to the atmosphere from leaves and from soil
surface is called Evapotranspiration (ET) – evaporation + transpiration.
Plants absorb water in the transpiration process. Water is transferred to the
atmosphere, from the surface area of its leaves, where larger leaves surface results
in higher rate of water absorbed and transferred to the atmosphere.
Weather conditions, which include temperature, relative humidity and radiation,
provide the driving force for evapotranspiration.

Putting it all together


Once the water holding capacity and the Readily Available Water (RAW) in the soil
are known, an adequate irrigation schedule can be determined.
Here are the steps:
1. Record the daily ET0
2. Multiply ET0 by the Kc that corresponds with the crop growth stage, to obtain
ETc.
The daily water requirement in m3/ha equals 10 times the ETc.
3. The RAW (in m3) is the required water amount per irrigation. Divide RAW (in
m3) by the daily water requirement, to obtain the irrigation interval.
4. Add 15-20% to the required water amount, in order to account for the leaching
requirement (“extra” water applied in order to leach excess of salts from the root
zone).

Let’s see a numerical example, for a 0.4ha plot.

229
Assume the following:
Average weekly ET0 = 5 mm/day.
Kc of the crop = 0.4.
RAW = 60 m3/ha.

ETc = 0.4x5 = 2 mm/day = 20 m3/ha/day = 0.078 inch/day.


Irrigation interval = RAW/daily water requirement = 60/20 = 3 days.
Required water amount per irrigation = RAW + 0.15RAW = 60 + 9 = 69 m 3/ha.
Therefore:
Apply 69 m3/ha every 3 days.
Since the plot’s area is 0.4ha, apply to that plot 27.6 m 3 every 3 days.

Note that by using this method, the water application amount does not change, only
the water application interval.
This way, we always bring the soil to the same moisture status with each irrigation.
Another way to manage it, is by subtracting the daily water requirement from the RAW
and irrigate when it becomes zero.
For example, if RAW is 200 m3 and we have the following readings of ET0:
Day 1 – 4mm, Day 2 – 5mm, Day 3 – 6mm, Day 4 – 4mm, Day 5 – 7mm, Day 6 – 4.
At day 1, the balance is: 200 – 40 = 160
At day 2: 160 – 50 = 110
At day 3: 110 – 60 = 50
Remaining available water on day 4: 50 – 40 = 10 → This is the moment to irrigate.

230
Irrigation Scheduling Using Soil Moisture
Sensing
Two main methods are used in agriculture to measure water in soil - soil moisture
sensors and tensiometers. Soil moisture sensors measure the volumetric water
content of the soil and tensiometers measure the soil water tension in the soil.

In order to understand the difference between the two methods, let us examine what
determines the ability of plants to absorb water from the soil.

The water in soil is held in soil pores, which are the spaces between soil particles.
This is very similar to the way water is held inside a sponge. Smaller pores can retain
water more tightly than bigger ones, due to stronger capillary forces.

In order to absorb water, plants must overcome the force in which water is held within
the pores.

Tensiometers measure the actual force that has to be applied in order to move water
in the soil, while sensors that measure soil moisture measure the volumetric water
content (θ) in the soil, i.e. the percentage of the soil volume that is occupied by water.

These are two related parameters. Under the same tension, a soil with larger pores,
such as sandy soils, will hold less water than a soil with smaller pores (e.g., clay soils).
This means that the volumetric water moisture of a clay soil may be relatively high, but
a large portion of the water might not be available to the plant, because it is held too
tight in small soil pores.

Therefore, tensiometers installed in two different soils may measure the same value
of tension, while the moisture content of the two soils might be different.

Sandy soils require more frequent, small irrigations, while clay soils require less
frequent irrigations, but a higher water amount in each irrigation event.

Soil moisture status


As previously mentioned, three important terms are used associated with the water
status in the soil: Field capacity, Wilting point and Available water.

Field capacity is the water content of the soil after excess water has stopped draining
after irrigation or rain. It is estimated to be the water content in the soil at a matric
potential (Ψ) of -1/3 bars.

231
Permanent wilting point is the soil water content below which plants wilt and cannot
recover. It is estimated to be the water content in the soil at a matric potential of -15
bars.

The available water is defined as the amount of water in the soil between wilting point
and field capacity.

The relationship between the moisture content in the soil and the matric potential is
described in the water retention curve. An example for such curves is given in the
following figure.

Example:

For a sandy soil, the water retention of which is given above:

Wilting point: 0.04 (4%)

Field capacity: 0.06 (6%)

Available water: 0.06 – 0.04 = 0.02 (2%).

232
For the silty loam soil:

Wilting point: 0.07 (7%)

Field capacity: 0.35 (35%)

Available water: 0.35 – 0.07 = 0.28 (28%)

And for the clay soil:

Wilting point: 0.18 (18%)

Field capacity: 0.45 (45%)

Available water: 0.45 – 0.18 = 0.27 (27%)

Therefore, in order to properly interpret the readings of a soil sensor that measures
the volumetric water content of the soil, the reading must be calibrated to the soil type
or, even better, specifically to the water retention curve of the soil. The water content
at field capacity and wilting point must be known.

Measuring the tension in the soil is a direct measurement of the matric potential and
can be used for determining the irrigation schedule, while measuring the volumetric
soil moisture is an indirect measurement.

Irrigation scheduling using soil measurements


Proper irrigation scheduling using soil measurements requires measuring the water
status in at least two soil depths – one where the active root system is and the other
below it. Standard depths are 20 cm, 60 cm and 90 cm, depending on the root system
depth of the crop.

If tensiometers are used, then the tensiometer installed at the depth of the active root
system provides an indication of when to irrigate (the “shallow” tensiometer), while the
reading of the deeper tensiometer is used to determine the amount of water to apply.

Readings below 10 cbar (centibar) usually indicate that the soil is wet enough.
Readings above 50 cbar indicate a dry soil. For most irrigation methods, the common
practice is to start irrigation when up to 50% of the available water was depleted. In

233
drip irrigation the allowed depletion is of 20-30%, because soil moisture must be kept
uniform. Therefore, lower tensiometer readings are used as a trigger for irrigation.

Tensiometer trigger levels (rules of thumb)

Sprinklers Drip irrigation

Sandy, Loamy sand 20-40 cbar 10-15 cbar

Sandy loam, Loam, Silt-loam 40-60 cbar 22-35 cbar

Clay-loam, Clay 50-80 cbar 40-50 cbar

Example:

What would be the volumetric water content of the soil at which a grower must irrigate,
if the soil is silty loam soil with water content of 30% at field capacity and 12% at wilting
point? Assume that 50% depletion is allowed.

Answer:

For 50% allowed depletion: Water content = (30 + 12)/2 = 21%

Installation location
The conditions in the field may be highly variable spatially Therefore, the location
where the soil measurement device is installed may significantly affect the readings,
the number of required devices and the decisions made.

The devices must be installed in locations that represent the field and crop. The higher
the variability in the field, the more soil measurement devices are required.

Factors to consider include:

• Spatial variability soil properties – different soil textures, topography.


• Crop types – different crops have different water uptake patterns.
• Different growth stages of the same crop
• Different types of irrigation system

234
In drip and sprinkler irrigation, it is important to locate the device in an adequate
distance from the emitter. The lateral movement of water in soil varies in different
soil types. Therefore, in sandy soils the measuring device should be placed closer
to the emitter.

Coarse-textured soil Fine-textured soil

235
Principles of Irrigation System Design
Designing an irrigation system requires knowledge of the crop, field conditions, soil
and water. Answering the following questions can help in determining the
characteristics of the irrigation system.
• What is the soil texture? How much water can the soil retain?
• What is the size of the field?
• What is the topography of the field?
• Will the field be divided to irrigation sections?
• What is the planting density of the crop?
• What is the depth of the crop’s root system?
• What are the water requirements of the crop in its specific location?
• What is the source of water supply? What is the water quality?
• What is are the costs and labor requirements of the system?

Choosing the irrigation system type


The main three types of irrigation systems are surface irrigation, overhead irrigation
and localized irrigation.
Surface irrigation refers to irrigation systems that deliver water across the field by
gravity. The system is designed in such a way that water flows from a water supply
ditch at the upper end of the field to the lower end of the field and infiltrates into the
soil as it advances.
Surface irrigation has a high labor requirement, and, of all irrigation methods, surface
irrigation has the lowest water use efficiency (around 55%). It is not suitable for sandy
soils due to their high infiltration rate.
Surface irrigation methods include furrow irrigation, border irrigation and flood
irrigation. Furrow irrigation is more suitable for row crops where only part of the field
surface is irrigated. Border irrigation is appropriate for most crops, except crops that
require flooding conditions, such as rice.
Overhead irrigation refers to permanent sprinkler irrigation and pivot irrigation. This
method has higher efficiency compared to surface irrigation. However, the initial
investments costs are high.
Localized irrigation is the most efficient of all methods and has a water use efficiency
of 90% and higher. However, it is most suited for smaller areas, requires good water
quality and has a high initial and maintenance costs.

236
Localized irrigation includes drip irrigation, sub-drip irrigation systems and micro-
sprinklers.
The following table compares irrigation system types:

Surface irrigation Overhead Localized


irrigation irrigation
Soil types All, except sandy All All
soils
Water availability Suited to areas with Suited to areas Suited to areas
very high water with very high with limited water
availability water availability availability
Water quality Can work with water Relatively low salt Low salt
with high sediment concentration is concentration is
content required. required.
Suitable for
wastewater.
Climate Any Areas with strong Any
winds should be
avoided
Automation Less common Very common Very common
Fertilization Not suitable for Suitable for Suitable for
fertigation fertigation. fertigation. High
Fertilizer losses efficiency.
may occur.
Operational Low High High
flexibility

Irrigation flow rate


Refers to the rate of water application in overhead and localized irrigation. The
irrigation flow rate is expressed in units of m3/h or gallons/min.
In order to minimize soil erosion and runoff, the irrigation flow rate should be planned
in such a way that it does not exceed the soil infiltration rate.
The minimum required irrigation flow rate:
𝐴𝑥𝐷𝑥𝐹
𝑄=
𝑡
Where:
A Field area
D Maximum water requirement of the crop (maximum ETc of the crop in mm)
t Time available for irrigation per day (hours)

237
F=10 when An is in hectares, Dn in millimeters and Q in m3/h
F=452.57 when An is in acres, Dn in inches and Q in gal/min.

For example:
The area of the field is 8 hectares (20 acres), the maximum crop evapotranspiration
rate is 7 mm/day (0.27 inches) and the grower can irrigate maximum 14 hours per day.
Thus, the minimum required irrigation flow rate:
Q = 8*7*10/14 = 40 m3/h
or
20*0.27*452.57 /14 = 174.5 gal/min.

Irrigation duration
An important parameter to know and consider at the design phase is the required
irrigation duration. It can be calculated using the soil properties, the efficiency of the
irrigation system and the leaching requirement (which is a function of water quality and
salt tolerance of the crop).

𝑅𝐴𝑊
𝑡=
𝐼(1 − 𝐿𝑅)𝑒
t=RAW /I(1-LR)e
Where
RAW - Readily available water
I- Soil infiltration rate
LR - Leaching Requirement
e- Irrigation efficiency (fraction)

Irrigation uniformity
A uniformity of 100% means that each point within the irrigated area receives the same
amount of irrigation.
Several indices are used to evaluate irrigation uniformity in different irrigation systems:
DU – Distribution uniformity – used for localized irrigation and sprinkler irrigation.
CU - Christiansen uniformity coefficient – used for sprinkler irrigation.

238
𝑄25%
𝐷𝑈 = 100 𝑥
𝑄𝑛
Where
Q25% Average flow rate of 25% of the emitters with the lowest flow rate.
Qn Average flow rate of all emitters.

DU (%) Classification

95-100 Excellent

85-95 Good

75-85 Regular

65-75 Poor

<60 Unacceptable

To calculate the DU, between 40 and 100 emitters, from different sections of the field,
must be randomly sampled.
Measuring emitter flow rate
For each emitter – measure the volume of water coming out of the emitter for 5 minutes
and perform the following calculation:
60 𝑥 𝑉
Actual emitter flow rate (L/h) = 1000 𝑥 𝑡

v – volume (in cc) of water provided by the emitter in t minutes.


t – the test time during which the sample was taken (in minutes).
For example:
A dripper provides 40 ml of water in 5 minutes
60 𝑥 40
Actual dripper flow rate = = 0.48 L/h
1,000 𝑥 5

Christiansen’s Uniformity Coefficient (CU)


CU is commonly used for sprinkler irrigation systems.
̅|
∑𝑛𝑖=1 |𝐷𝑖 − 𝐷
𝐶𝑈 = 100 (1 −
𝑛𝐷̅
239
Where

𝐷𝑖 amount of water in an individual water catch can


̅
𝐷 average amount of water in all catch cans

n number of catch cans

CU for sprinkler irrigation:

CU (%) Classification
<70 Poor
70%-90 Adequate
>90 Excellent

240
Drip irrigation Systems
Drip irrigation consists of frequent application of water at low pressure and low volume,
on or below the soil surface. The devices that control the water outlet are called
emitters or drippers, and water is applied through drip emitters that are placed along
a water supply line.
The advantages of drip irrigation are:

• High water use efficiency


• High efficiency of nutrient application through fertigation
• Energy requirement is lower compared with other systems, due to the focused,
local application.
• Reduced growth of weeds
• No soil erosion
• Reduced disease stress (leaves are kept dry)

The main challenge with drip irrigation is emitter clogging.

Properties of drip emitters


The desired properties of drip emitters are:

• Low flow rate (0.5-8.0 L/h)


• Low vulnerability to clogging
• Low production cost
• High durability

Achieving a low flow rate requires a large dissipation of the pressure. The dripper flow
rate is determined by the physical characteristics of the dripper and its hydraulic
design:

• Pattern and dimensions of the passage of water through the dripper


• Water pressure at the dripper inlet

The smaller the cross section of the passage is, the lower dripper flow rate at a given
pressure will be. However, the narrower the passage, the greater the risk of clogging
by suspended solid particles and chemical precipitates.

241
Classification of drippers according to their hydraulic characteristics

Long path-drippers – Water flows through a small-diameter long microtube. They


operate at a pressure of up to one atmosphere, and flow rate is sensitive to changes
in pressure

Labyrinth - Water flows along a labyrinth in which the flow direction changes sharply.
This results in a turbulent flow, loss energy and low flow rate. The turbulent flow also
helps reduce the clogging risk.

Flow rate is not affected much by changes in pressure.

Pressure compensating and non-pressure compensating drippers


In pressure-compensating drippers (PC), the dripper flow rate is kept uniform at a wide
range of pressures. A diaphragm adjusts the diameter of the water passage,
depending on the pressure in the line.

Pressure compensated drippers usually also consist of self-flushing mechanism that


makes them more resistant to clogging.

The flow rate of non-pressure compensating drippers varies with the pressure.
Therefore, variations in topography along the line will result in varying flow rates of the
drippers. This type of drippers can be used in a landscape that is relatively flat.

Inline and online drippers


Drip emitters can be categorized into inline and online drippers.
Inline drippers are installed inside the pipe, along its length, at fixed distances.
They are suitable for fields where plants are equally spaced. Advantages include
easier and quicker installation and a reduced risk of mechanical damage.

Online drip systems provide more flexibility in determining the location of the
drippers along the line. Drippers can be placed varying distances, where they are
needed. Drippers are punched manually into the pipe, which is supplied without drip
emitters on it. This type of drip systems is suitable for growing in containers in
greenhouses, where irrigation of individual plants is required, in irrigation of trees, or
in fields where spacing between plants is irregular.

Surface and subsurface drip irrigation (SDI)


In surface drip irrigation, irrigation lines are placed on the soil or hung on support
elements.

Subsurface drip irrigation is drip irrigation that is buried below the soil surface.

242
Advantages of SDI:
• Minimizes mechanical damage to the system
• Decreases weeds infestation
• Reduces the presence of pests and diseases
• Eliminates water runoff and evaporation from the soil surface
• Improves the absorption of nutrients by the roots (especially of phosphorus)

Disadvantages of SDI
• Higher installation costs
• Increased risk of root penetration and clogging by soil particles
• More difficult to monitor the operation of the system
• Requires higher maintenance

Water distribution in drip irrigation


In drip irrigation, the surface area of the wetted area is a small fraction of the field’s
area. Water flows through the soil from the emission point to the areas of the soil with
higher water deficit. Unlike sprinkler irrigation, water movement is two-dimensional, or
three-dimensional, in the case of subsurface drip irrigation. Water moves downwards,
laterally and, in subsurface drip irrigation, also upwards. Water movement is governed
by capillary and gravitational forces, where the balance between these two forces
determines the pattern of water distribution.

The water distribution pattern and the volume of wetted soil depend on soil properties,
dripper flow rate, spacing between drip emitters, and on the irrigation rate and amount.
The volume and shape of the soil wetted by a single dripper is referred to as ‘wet bulb’.

Soil properties
In fine-textured soils (e.g., clay soil) the capillary forces are more predominant than
gravity forces. Therefore, the horizontal width of the wet bulb is greater than its depth.

In light-textured soils (e.g., sandy soil), the vertical water movement is more
predominant than the lateral movement, and the wet bulb gets the shape of a carrot.

Form of the wet bulb in different soil types

243
Soil structure factors, such as soil cracking, or changes in texture across the soil
profile, also affect water distribution in the soil.

Dripper flow rate


For the same soil type and amount of water applied, higher dripper flow rates will result
in a wet bulb that is more horizontally distributed, while in lower flow rates, vertical
water movement will be predominate.

QA > QB

Drip emitter spacing


For the same soil type, irrigation rate and duration, a greater distance between the
drip emitters will result in a wider and shallower wet bulb. A smaller distance between
the drippers on the line, will result in a narrower and deeper wetting pattern.

The diameter of the wet bulb will increase only until overlap between wet bulbs occurs.
After that, most of the water flow will be vertical.

The effect of dripper spacing on wetting pattern

244
Irrigation amount
Assuming a constant dripper flow rate, the wet bulb increases in depth with the amount
of water applied. i.e. with the irrigation duration. However, its horizontal distribution is
limited.

T1 T2

Salt accumulation pattern in drip systems


Dissolved salts move with water. Therefore, in drip irrigation, salts accumulate on the
margins of the wet bulb. Salts accumulate particularly on the surface of the soil, due
to the low water content on it. Concentric rings of accumulated salts are visible on the
margins of the wetted area.

Salts also accumulate at a depth that depends on leaching efficiency.

Salt accumulation does not damage the crop while its root system is in a zone of low
salinity. Due to the high irrigation frequency, the salt level in the root zone is usually
similar to its level in the irrigation water.

However, a rain event might wash salts from the soil surface into the root zone. This
could damage seedlings or emerging seeds. To avoid such a damage, it is
recommended to flush the soil, using sprinklers, before planting or transplanting.

245
Irrigation rate - adjustments for drip irrigation
Because in drip irrigation only part of the soil is wetted, some of the parameters that
are used for irrigation planning must be adjusted:

Crop coefficient (ETc) adjustment:


ETc was developed for crops under sprinkler irrigation or flooding. For drip irrigation:

ETCa = ETC x [0.1(GC)0.5]

Where GC is the percentage of soil coverage by crop foliage.

Adjusted Readily Available Water (RAW):

RAW = p x A x ΘAw X Zr x n

Where A is the wetted area and n is the number of beds/rows.

If the wet bulbs overlap: A – area of the bed/row.

If wet bulbs don't overlap: A = πr2

Where r is the radius of the wet bulb.

246
Causes and Prevention of Emitter
Clogging
Micro irrigation systems have many benefits. However, due to the small opening
diameter of drippers and micro sprinklers, they are very sensitive to clogging. The
causes for clogging can be classified into biological, physical, and chemical causes.
The sensitivity of emitters to clogging depends the quality of the irrigation water, the
filtration system in place and the properties of the emitter.
The source of the irrigation water can provide an initial indication of what type of
potential clogging hazard to expect. The risk of physical and biological clogging is
higher when using surface water, while chemical clogging is more likely when the
water source is groundwater.

Physical Causes
Solid particles are the most dominant cause for emitter clogging. Surface water has
a higher clogging potential due to their high content of suspended solids. Organic
physical particles may include algae, snails, plant parts, seeds etc. Inorganic particles
include sand, silt and clay particles.
Removal of suspended solids can be achieved by filtration or settling ponds. Drip
irrigation systems usually require particles larger than 130 microns to be removed.
However, even very small silt and clay particles, that each one of them individually
does not cause a clogging problem, can clog emitters if they flocculate to form larger
aggregates.
A settling pond is a reservoir where water is collected and allowed to remain almost
still. Suspended particles then settle by gravitation and removed. An additional
filtration is usually required after the settling pond.
Types of filtration systems used for treating water before entering the irrigation line
include hydro cyclones, sand media filtration, disc and screen filters.
Hydrocyclones use centrifugal forces to remove particles from water. They are often
used as pre-filtration to remove sand particles from water. Additional filtration is usually
required.
Screen filters remove suspended solids and sand particles. The necessary size of
the filter is based on the maximum particle size allowed by the emitter diameter. A
common rule of thumb is to use 0.1 times the diameter of the emitter opening.

247
The size of screen filters is often expressed in mesh. A 120-mesh screen has 120
openings in one square inch.

Filtration unit conversion

Mesh Micron Inches


40 420 0.0165
100 150 0.0059
120 125 0.0049
150 105 0.0041
200 75 0.0029
270 53 0.0021
400 37 0.0015

Biological Causes
The conditions in drip irrigation provide a favorable environment for bacterial growth.
Bacteria and algae can cause an accumulation of slime in the irrigation system. This
slime can be a direct cause for drip clogging. Bacterial slime is formed by filamentous
bacteria, particularly in the presence of iron. Ferrous iron (Fe2+) is oxidized to ferric
iron (Fe3+) in the presence of dissolved oxygen. The iron bacteria, which use iron for
their metabolic activity, form colonies around the ferric iron, and form a sticky slime.
Concentrations ferrous iron of above 0.2 ppm are considered a potential hazard for
clogging of drip systems.
Other types of slime forming bacteria can form slime even without the presence of
iron. Slime related to iron bacteria is reddish. Sulfur bacteria can also produce slime.
Slime produced by sulfur bacteria has a dark color.
Except of being a direct cause of clogging, slime can make other physical particles
aggregate and
Other than bacterial slime, algae, organic matter particles and plant residues can also
cause clogging.
Biological clogging can be controlled by oxidation, sand media filtration and
chlorination.
Oxidation – Iron can be removed from water by oxidation, followed by filtration.
Ferrous iron is oxidized to ferric iron, which precipitates and can be filtered out of the
water. Oxidation can be performed by aeration, chlorination or potassium
permanganate.

248
Sand media filters contain gravel and graded sand, of various sizes, inside a
pressurized tank. They can remove clay particles as well as organic matter.
Chlorination – Chlorine kills bacteria and algae. Chlorine should be injected to the
water storage tank, in order to allow adequate contact time. A residual chlorine level
of 0.2-0.5 ppm must be kept at the most distant part of the system.

A chlorination and sand filtration system

Chemical Causes
Precipitation of minerals is a common cause of emitter clogging. Precipitation occurs
when the solubility of a mineral is exceeded. Carbonates, calcium, iron and
manganese are the most common elements that can cause clogging. Clogging
potential increases at water pH greater than 7.0.
At high carbonate and calcium levels and at this pH level, precipitates of calcium
carbonates are formed.
In the presence of oxygen, and at pH above 7.0, soluble ferrous iron (Fe 2+) readily
converts to the insoluble ferric form (Fe3+).
Addition of fertilizers might affect emitter clogging as well. For example, phosphorus
can react with calcium to form calcium phosphates.
To avoid chemical clogging, the pH of the irrigation water must be reduced to below
7.0. Sulfuric, phosphoric, and nitric acids are commonly used to regulate the pH in drip
irrigation systems.
Iron can be removed from water by the oxidation-filtration process described above.

249
Clogging potential of drip emitters

Parameter Clogging Hazard


Low Moderate High
pH <7.0 7.0-8.0 >8.0
Iron (Fe) mg/L <0.2 0.2-1.5 >1.5
Manganese (Mn) mg/L <0.1 0.1-1.5 >1.5
Hydrogen Sulfide (H2S) mg/L <0.2 0.2-2.0 >2.0
Total Dissolved Solids (TDS) mg/L <500 500-2000 >2000
Total Suspended Solids (TSS) mg/L <50 50-100 >100
Bacteria Count (#/ml) <10,000 10,000-50,000 >50,000

250
Irrigation Scheduling in Container Plants
Factors such as the crop water requirements, the size of the container and the
properties of the growing medium determine the amount and frequency of the irrigation
of container plants. Due to the restricted volume of containers, plants grown in
containers require frequent irrigations and close monitoring of the moisture in the
growing medium.

Irrigation intervals may vary from once every few days to several times a day,
depending on the above-mentioned factors.

Estimation of the water requirements


Estimating the water uptake of the crop is relatively easy when growing in containers.
Water requirements can be estimated using standard methods, such as
evapotranspiration (ET) and crop coefficient (Kc) data, or they can be measured in the
greenhouse.

Provided that the containers are not too big, the simplest method to measure the water
uptake of the crop is by weighing a few containers.

1. Select a container or containers that represent the plants in the greenhouse.


2. Apply irrigation until drainage comes out of the bottom of the container.
3. Wait until drainage stops.
4. Weigh the container and register the time and the weight of the container.
5. Wait for 24 hours and weigh the container again (provided that the growing
medium can hold the necessary amount of water for 24 hours).
6. The difference between the first measurement and the second, is the daily
water uptake
7. To obtain the water uptake of the entire irrigation section (valve), multiply the
result by the number of containers in the irrigation section.

For example, let us assume that the first measurement gave 2.5 kg and the second
gave 2.2 kg. The difference is 0.3 kg or 300 ml of water.

If 4,000 containers are irrigated under the same irrigation valve, then the total daily
water requirement for that valve is 0.3 x 4,000 = 1,200 liters or 1.2 m3.

For imperial units, measure container weight in pounds and divide the result by 0.85.

The irrigation interval


Once the daily water uptake is known, the irrigation interval can be calculated.

251
𝐶𝑜𝑛𝑡𝑎𝑖𝑛𝑒𝑟 𝑣𝑜𝑙𝑢𝑚𝑒 𝑋 𝑅𝐴𝑊
𝐼𝑟𝑟𝑖𝑔𝑎𝑡𝑖𝑜𝑛 𝑖𝑛𝑡𝑒𝑟𝑣𝑎𝑙 (𝑑𝑎𝑦𝑠) =
𝐷𝑎𝑖𝑙𝑦 𝑤𝑎𝑡𝑒𝑟 𝑢𝑝𝑡𝑎𝑘𝑒

Where RAW is the readily available water of the growing medium, expressed as a
fraction.
For example:
Container volume: 1 gallon
RAW: 18%
Daily water uptake: 10 fl oz

𝑓𝑙 𝑜𝑧
1 𝑔𝑎𝑙 𝑥 0.18 𝑥 128
𝑔𝑎𝑙
𝐼𝑟𝑟𝑖𝑔𝑎𝑡𝑖𝑜𝑛 𝑖𝑛𝑡𝑒𝑟𝑣𝑎𝑙 = = 2.3 𝑑𝑎𝑦𝑠
10 𝑓𝑙 𝑜𝑧

The irrigation interval should, therefore, be two days.

Several observation methods can aid in finding the right moment to irrigate. For
example, you can observe the plants and identify how long after irrigation they start
showing water stress symptoms. Irrigation should be applied before stress symptoms
are visible.

The Irrigation rate


The water amount to be applied in each irrigation can be calculated as follows:
Irrigation amount = (Daily uptake x irrigation interval) x (1+LR)

252
Where LR is the leaching requirement, expressed as a fraction. It is the percentage of
drainage required for flushing salts out. It should usually be between 15 to 30%.

Once the preliminary irrigation schedule is determined, adjustments should be made


based on a routine sampling of the drip and drainage solutions. It is recommended to
take at least one or two representative samples from each irrigation section. Samples
should be taken at least once a week.

The irrigation amount can be adjusted according to the percentage of drainage. If no


water comes out of the container after irrigation, then the amount of water should be
increased.

The percentage of drainage is calculated as:

% drainage = (Vdrainage / Vdrip) x 100

The desired percentage of drainage can be determined based on the salinity tolerance
of the crop and the quality of the irrigation water. A higher percentage of drainage will
be required for sensitive plants, irrigated with water with high concentration of salts
(high EC). Weekly monitoring of the drip and drainage solutions can provide a good
indication of salt accumulation.

A drainage solution EC that is greater than the EC of the drip solution by more than
0.5-0.7 ds/m indicates that you may need to increase the irrigation rate. It should be
noted, though, that an increase of the EC of the drainage solution may also indicate
excess fertilization. In general, a low percentage of drainage accompanied by a
continuous increase of the drainage EC means that the irrigation rate should be
increased.

253
Variable Rate Irrigation
Variable rate irrigation (VRI) involves applying different amounts of water to different
areas of the field. This innovative technology has many advantages. It allows farmers
to:

• Maximize yields
• Save water
• Avoid or minimize nutrient leaching
• Avoid or minimize runoff
• Save energy
Agricultural fields never are uniform. Variability may be related to the soil or to the crop
itself.
Soil is heterogenous and there will always be variability in soil properties throughout
the field.
Spatial variability in topography and soil texture greatly affect the water availability for
the crop.
Soil texture may vary horizontally as well as throughout the soil profile. Soils of different
textures retain different amounts of water and, therefore, will require different irrigation
rates and scheduling. Sandy soils require more attention, as they can store only a
small amount of water. Therefore, they must be irrigated more frequently and with
smaller amounts of water in each irrigation. Too high irrigation depth in a sandy soil
may result in water waste and leaching of nitrogen below the root zone of the crop.
On the other hand, heavy-textured soils can store more water and, therefore, must be
irrigated at larger intervals and with greater amounts of water.
Land topography will affect, in various ways, how efficiently water is distributed within
the field in:

• Slopes and elevation differences may result in pressure variations in the


irrigation system, and therefore in water application uniformity.
• Water flow within the field is determined by its microtopography.
Planting density - In many fields, especially large ones, the farmer may decide to plant
the crop in different planting densities and/or plant different varieties throughout the
field. This results in different irrigation requirements for each zone of the field.
Planting date – in many cases, the field is planted gradually, resulting in zones with
different development stages of the crop.
In variable rate irrigation, the field is divided into several irrigation zones, based on the
above-mentioned factors, where each zone is considered to be homogenous and is
managed individually.

254
In permanent sprinkler and drip irrigation systems, each irrigation zone is controlled
by an individual valve. Typically, one valve is operated at a time.
Most pivot irrigation systems today are used to water the field uniformly. However,
precision variable rate irrigation systems offer new opportunities and are becoming
more common.
Center pivot systems offer high control flexibility. Center pivot companies offer two
types of variable rate control:
The names given by different pivot companies may vary, commonly named “Sector
VRI” and “Zone VRI”.
“Sector VRI” or “Time VRI” – in this method, a prescription for pie-like slices is fed to
the pivot. The field is divided into pie-like slices and the speed of the pivot varies for
each sector, in order to apply the desired irrigation depth. Slow speed increases the
irrigation depth, while higher speed reduces the application rate. The irrigation
prescription is usually given as a colored map or a table listing the irrigation depth or
relative pivot speed for each sector. Sectors are described as a range of angles.

Sector VRI Zone VRI

Zone VRI allows for even higher accuracy: smaller zones can be created by controlling
the sprinkler valve pulse, where the coverage area is further divided to rings. This
way, thousands of individual management zones can be defined.
The prescriptions are usually generated using a prescription software and are
uploaded to the pivot control system.

255
How is the irrigation prescription built?
That’s the 64-thousand-dollar question. It’s great to have a pivot that can apply
accurate water rates for each point in the field, but how can the grower know the
optimal water rate for each zone?

Soil maps can be created using EC (Electrical Conductivity) mapping. Soil maps
provide information on the variability in the field and can be uploaded into a software
that can process the data and provide baseline application rates for each irrigation
zone. The irrigation rate for each zone can then be adjusted based on the actual status
in the field, such as weather conditions, planting densities etc.

Soil moisture sensors are being increasingly adopted by farmers and provide an
opportunity for making variable rate irrigation fully automated and much more precise.
Sensors can be installed in each irrigation zone and provide real-time data about soil
moisture, salinity and other parameters.

256
Irrigation with Desalinated Water
Desalinated water is being used for irrigation purposes in many places around the
world, where the available water source is too saline to be used. Such water sources
may include groundwater, sea water and sometimes surface water from lakes.

Water salinity is a major problem for agriculture. Saline water contains high
concentrations of salts, which crops might not tolerate.

The source of the salts in water is the surrounding rocks, e.g., the mineralogy of the
aquifer. Therefore, surface water is usually less salty than groundwater, which is
surrounded by rocks. Oceans are salty because when water evaporates, the salts
remain and accumulate.

Groundwater may contain high concentrations of ions such as sodium, chlorides,


sulphate, bicarbonates, calcium, magnesium. It may also contain trace elements such
as boron, iron, manganese, and fluoride, which may be present at relatively low
concentrations but might become toxic to the plant if their concentrations exceed
certain thresholds (such thresholds are usually crop-specific).

Desalination with reverse osmosis


There are various methods of desalination, of which, the main method used today is
Reverse Osmosis.

In reverse osmosis, water is pushed through semi-permeable membranes, using


pressure. The salts do not pass through the membrane, while water molecules do.

Although there is some selectivity to specific ions, desalination with reverse osmosis
removes most salts from the water and the amount of salts removed depends mostly
on the pressure and on the ratio between the amount of desalinated water produced
and the amount of brine water rejected and discharged as waste.

Flow rate of feed water = flow rate of permeate + flow rate of brine

257
Higher rejected brine volume results in better quality of the desalinated water and vice
versa.

Desalination removes most ions (salts) from the treated water. Many water sources,
especially groundwater, may contain elements that are considered essential plant
nutrients. Such elements include calcium, magnesium and sulfur, which are required
in relatively large quantities by plants.

Removing these salts, together with the harmful elements, such as sodium and
chlorides, requires adding back those elements with fertilizers. Such fertilizers may
include calcium nitrate, magnesium nitrate and magnesium sulphate.

Desalination also removes carbonate hardness from water. This results in water that
has a lower resistance to changes in pH. Adding even a small amount of acid to the
desalinated water might result in a sharp drop of its pH. Therefore, in such cases, the
water has to be stabilized by adding sources of bicarbonate, such as potassium
bicarbonate.

Under some circumstances the level of harmful salts can be compromised, depending
on the concentration of such salts in the raw water and the tolerance level of the crop.
In such cases, one of the two following options can be applied, depending on the
specific conditions and water quality:

1. The reverse osmosis plant can be operated at higher permeate flow rates, leaving
more salts in the desalinated water, i.e., leaving higher concentration of harmful
salts, but also higher level of essential nutrients.
2. Desalinated water can be mixed with untreated water.

Applying the above methods, can save energy and fertilizer costs.

It is important to run a simulation of the reverse osmosis plant, prior to its final design
and installation. The manufacturer of the plant will usually run such a simulation, but it
is the grower’s responsibility to provide a full laboratory analysis of the raw water and
to define the required water quality and flow rates. Except of salts, does reverse
osmosis remove other substances?

Since dissolved salts, which are removed by reverse osmosis, are much smaller than
viruses, reverse osmosis also removes viruses and bacteria.

258
Which elements should the raw water be tested for?
Common ions in that should be tested include calcium, magnesium, sulphate,
carbonates, sodium, chlorides, potassium, silica, boron, iron, manganese, fluoride,
barium and strontium.

Does reverse osmosis remove bacteria and viruses?


Since bacteria viruses are much larger than dissolved salts, they are removed in the
reverse osmosis process. However, presence of certain bacteria prior to the RO plant
often results in membrane clogging and operation problems.

A reverse osmosis plant

259
Appendix I: Conversion Tables

From To Multiply by

NH4 N-NH4 0.821

NO3 N-NO3 0.226

N NH4 1.285

N NO3 4.427

P2O5 P 0.436

PO4 P 0.326

P P2O5 2.291

P PO4 3.066

K2O K 0.830

K K2O 1.205

CaCO3 Ca 0.400

CaO Ca 0.714

Ca CaCO3 2.497

MgO Mg 0.603

Mg MgO 1.657

260
Element mmol/L meq/L ppm

N-NO3- 1 1 14

N-NH4+ 1 1 14

P-H2PO4- 1 1 31

K+ 1 1 39

Ca2+ 1 2 40

Mg2+ 1 2 24

Na+ 1 1 23

Cl- 1 1 35.5

S-SO42- 1 2 32

Fe2+ 1 3 56

Mn2+ 1 2 55

Zn2+ 1 2 65

Cu2+ 1 2 64

B-B4O72- 1 2 11

261
Appendix II: Nutrient Uptake by Crop
Crop Yield N P2 O 5 K 2O
Almonds 3.9 t/ha 212 kg/ha 80 kg/ha 360 kg/ha
(3,500 lbs/acre) (189 lbs/acre) (71 lbs/acre) (320 lbs/acre)

Apple 60 t/ha 170 kg/ha 74 kg/ha 240 kg/ha


(1200 bushels/acre) (151 lbs/acre) (66 lbs/acre) (214 lbs/acre)
25 t/ha 180 kg/ha 60 kg/ha 350 kg/ha
Avocado (400 bushels/acre) (160 lbs/acre) (53 lbs/acre) (312 lbs/acre)
Broccoli 12 t/ha 210 kg/ha 60 kg/ha 290 kg/ha
(100 cwt/acre) (187 lbs/acre) (53 lbs/acre) (258 lbs/acre)

Carrot 55 t/ha 145 kg/ha 65 kg/ha 325 kg/ha


(25 short tons/acre) (130 lbs/acre) (58 lbs/acre) (290 lbs/acre)

Cabbage 60 t/ha 195 kg/ha 70 kg/ha 320 kg/ha


(27short tons/acre) (174 lbs/acre) (62 lbs/acre) (285 lbs/acre)

Cherry 15 t/ha 125 kg/ha 55 kg/ha 155 kg/ha


(120 cwt/acre) (110 lbs/acre) (50 lbs/acre) (140 lbs/acre)

Corn 10 t/ha 220 kg/ha 70 kg/ha 180 kg/ha


(160 bushels/acre) (196 lbs/acre) (62 lbs/acre) (160 lbs/acre)

Cucumber 120 t/ha 320 kg/ha 115 kg/ha 470 kg/ha


(955 cwt/acre) (285 lbs/acre) (102 lbs/acre) (420 lbs/acre)

Custard apple 25 t/ha 130 kg/ha 20 kg/ha 100 kg/ha


(22,000 lbs/acre) (116 lbs/acre) (18 lbs/acre) (90 lbs/acre)

Eggplant 80 t/ha 550 kg/ha 85 kg/ha 720 kg/ha


(635 cwt/acre) (490 lbs/acre) (75 lbs/acre) (640 lbs/acre)

Grapefruit 80 t/ha 200 kg/ha 60 kg/ha 350 kg/ha


(635 cwt/acre) (178 lbs/acre) (53 lbs/acre) (310 lbs/acre)

Peach 40 t/ha 135 kg/ha 45 kg/ha 155 kg/ha


(80,000 lbs/acre) (120 lbs.ac) (40 lbs.ac) (138 lbs.ac)

Pepper 100 t/ha 355 kg/ha 95 kg/ha 550 kg/ha


(45 short tons/acre) (316 lbs/acre) (85 lbs/acre) (490 lbs/acre)

Plum 25 t/ha 75 kg/ha 38 kg/ha 110 kg/ha


(22,000 lbs/acre) (67 lbs/acre) (34 lbs/acre) (108 lbs/acre)

Pomegranate 25 t/ha 125 kg/ha 75 kg/ha 250 kg/ha


(22,000 lbs/acre) (111 lbs/acre) (67 lbs/acre) (223 lbs/acre)

Potato 55 t/ha 240 kg/ha 65 kg/ha 380 kg/ha


(25 short tons/acre) (215 lbs/acre) (58 lbs/acre) (339 lbs/acre)

Tomato 160 t/ha 440 kg/ha 110 kg/ha 770 kg/ha


(1275 cwt/acre) (140 lbs/acre) (98 lbs/acre) (685 lbs/acre)

Watermelon 100 t/ha 250 kg/ha 110 kg/ha 440 kg/ha


(90,000 lbs/acre) (220 lbs/acre) (98 lbs/acre) (390 lbs/acre)

262
How does it work?
FERTILIZATION AND IRRIGATION
• LEARN AT YOUR OWN ONLINE COURSE
PAC E
Log in with your password and get
Learn the best practices and Upon completing the course, you
24/7 access to the course
theory of fertilizer management, will master the fundamentals of
recordings. Downloadable
materials. fertigation, irrigation scheduling, fertilization and irrigation and will
soil test interpretation and more. be capable of designing fertilization
plans and irrigation schemes
The course “Fertilization and
• START IMMEDIATELY professionally
irrigation – Strategies and Best
You can start learning at any time,
even right now! Practices” consists of 10 online
sessions, 1.0-1.5 hours each, in
which you will learn the best
• BE CONNECTED fertilization and irrigation
The course instructor will be practices.
questions available to answer any
of your questions via email and
Whatsapp

CLICK HERE TO REGISTER

ht t p s: / / c r o p a i a. c om / f ert i l i z at i o n - a n d - i rr i g at i o n - o n l i n e- c o ur s e/
i nf o @ c r o p a i a. c om 263
+ 97 2 .5 2 3. 59 7 .9 6 4 ( p h on e / Wh ats a pp )
This book covers the essentials of plant nutrition and irrigation management. Throughout the
book, the author shares his knowledge on the most important aspects of crop nutrition and
irrigation, including:

• Plant nutrients, their roles in plants and behaviour in soil and water
• Fertilizer management practices
• The soil – fertility, salinity, acidity, cation exchange capacity, sodicity
• Testing soils, how to interpret soil test results
• Irrigation water quality – water quality parameters, testing the water, interpreting water
analysis.
• Fertigation best practices
• How to give fertilizer recommendations
• Irrigation scheduling methods
• Hydroponics and nutrient solutions

And more

Guy Sela spent many years researching and experimenting ways


to improve agricultural production and develop new agricultural
technologies.
He honed his skills on commercial farming around the world and by
working with thousands of growers and agronomists globally.
Apart of his activity in the AgTech industry, Mr. Sela is an
international speaker, and is often invited to share his knowledge
about fertilization and irrigation at international conferences.

264
2021 edition

© All rights reserved

You might also like