Você está na página 1de 20

1

Module No. 1

Soil-Water-Plant Relationship

Crop production depends on the use of soil and water. The soil provides the mechanical
and nutrient support necessary for plant growth and development. In your Hydrometeorology
class, you have learned how water evolves and becomes available to plants through the
hydrologic cycle. Effective management of both soil and water resources for crop production
requires in-depth understanding of their interrelationships. This is our topic in this module.

After thorough and diligent study of this module, you should be able to:

a. Enumerate the components of a soil system and describe its compositions;

b. Describe why water is essential to plants;

c. Compile a list of at least 10 crops commonly grown in Ilocos Sur showing their rooting
depth and corresponding growth stage most susceptible to drought stress;

d. Enumerate the different classifications of soil moisture or soil water and describe their
relative quantities in clayey and sandy soils;

e. Define plant available water (PAW) mathematically;

f. Describe the different factors that determine when, where, and how much water a plant
needs; and

g. Synthesize the relationship of soil, water and plants.


2

I. Soil
Soil as a three-dimensional, dynamic, natural body occurring on the surface of the Earth
that is a medium for plant growth and whose characteristics have resulted from the forces of
climate and living organisms acting upon parent materials over a period of time (Luther, 2016).

The soil system is composed of three major components (Figure 1): air, water, and solids.
The soil solid is a mixture of minerals and organic matters. The mineral matter consists of small soil
particles such as sand, silt, and clay. Organic matter is made up of decaying plant and animal
substances and is distributed in and among the mineral particles. The pore spaces that occur
around the soil particles are important because they store air and water in the soil (Evans, Cassel
& Sneed, 1996).

Source: Brown & Wherrett (2017)

Figure 1. Definition sketch of soil components

A typical soil is made up of about 47 percent mineral particles, 3 percent organic matter,
and 50 percent pores (Figure 2). The amount of water and air present in the pore spaces varies
over time in an inverse relationship (Rogers & Sothers, 1996), that is, more water means less air
and vice versa. Note that air is as important as water to promote normal plant growth and
development.
3

Source: Rogers & Sothers (1996)

Figure 2. Typical soil composition

Root respiration needs air and hence soil aeration is an important consideration in
irrigation and drainage. Hasan reiterated that an active root system requires a delicate balance
between air and water stored in the pore spaces because it regulates root activity and plant
growth processes (2017). Thus, depending on the crop type and soil properties, this air-water
interface has to be considered during irrigation. Application rate should be just enough for the
water holding capacity of a particular soil so that excess water may not be detrimental to
normal plant growth and development.

II. Water
Plants, as in any other living organisms, need water in order to grow and develop.
According to Haman & Izuno, water is transported throughout the plant system almost
continuously from the soil to the roots and from the roots into the various parts of the plant, then
into the leaves where it is released into the atmosphere as water vapor through the stomata
(2003). Based on the cohesion-tension water transport model of Joly and Dixon (Figure 3), water
moves across the ground into the xylem and enters the root hairs by osmosis. The dissolved
minerals in water, on the other hand, move into the roots by diffusion. From the roots, water
moves upward throughout the plant system by root pressure and by tension in the water
molecules created by transpiration (leavingcertbiology.net, 2018).

Water is essential to plants for several reasons (Haman & Izuno, 2003):

 It is the principal medium for the chemical and biochemical processes that support plant
metabolism
 Under pressure within plant cells, water provides physical support for plants.
 It acts as a solvent for dissolved sugars and minerals transported throughout the plant
system through the xylem.
 Evaporation within intercellular spaces provides the cooling mechanism that allows
plants to maintain the favorable temperatures necessary for metabolic processes.
4

Well-watered plants maintain their shape due to the internal pressure in plant cells known
as turgor pressure. This pressure is necessary for plant cell expansion and consequently for plant
growth. Loss of turgor pressure due to insufficient water supply can be noticed as plant wilting
(Haman & Izuno, 2003). When too little water is available in the root zone, the plant will reduce
the amount of water lost through transpiration by partial or total stomatal closure. This results in
decreased photosynthesis because the CO2 required for transpiration that enters the plant
through the stomata is restricted. Decreased photosynthesis reduces biomass production and
results in decreased yields (Ahuja, 2006).

Source: leavingcertbiology.net (2018)

Figure 3. Cohesion-tension water transport model in plants

III. Soil-Water Relationship


Soil properties directly affect the availability of water and nutrients to plants. Soil water
affects plant growth directly through its controlling effect on plant water status and indirectly
through its effect on aeration, temperature, nutrient uptake and transport, and transformation
(Ali, 2010).
5

A. Types of soil water

As mentioned earlier, the relative amounts of air and water stored in the pores changes
as water is added to or lost from the soil. The pore volume is actually a reservoir for holding soil
water. This is commonly known as water holding capacity of the soil. Water in the soil reservoir
comes in the form of gravitational water, capillary water and hygroscopic water (Figures 4 & 5).
Not all of the water in the soil reservoir is available for plant use (Evans, Cassel & Sneed, 1996).

Images from StudyLib (2018)

Figure 4. Definition sketch of types of soil water

Source: Juhász & Pregun (2013)

Figure 5. Types of soil water in the soil reservoir


6

1. Gravitational water

Gravitational water is that portion of the soil water that drains freely from the soil because
of insufficient tension [< 0.1 bar (Shreeja, n.d)] that holds the water to remain in the soil reservoir
against the force of gravity. As it percolates deeper, gravitational water will eventually be
deposited and forms part of the groundwater in the aquifer. The soil pore spaces vacated by
gravitational water are replaced by air, which provide aeration in the root zone.

2. Capillary water

When all the gravitational water has drained out, what remains in the soil is called
capillary water. According to Lajos, capillary water occurs as a film around soil particles and in
between the pores. Since it can move through the soil in all directions even upwards for up to
two meters in response to suction, capillary water is the main source of plant moisture. As this
water is withdrawn, the capillary water in between the pores drains first before the film around
the soil particles. As the withdrawn of water continues, the film becomes thinner and harder to
detach from the soil particles (2008).

3. Hygroscopic water

When all the capillary water has been extracted by plant roots, what remains in the soil is
referred to as hygroscopic water. This water is held very tightly on the surface of soil colloidal
particles (Figure 5). Hygroscopic water is not available for plant use because, according to
Shreeja (n.d), the soil water tension is 31 bars which is beyond the 15 bars suction pressure of
plant roots. Unlike capillary water which evaporates easily at atmospheric temperature,
hygroscopic water cannot be separated from the soil unless it is heated.

B. Soil moisture content

The soil holds water in two ways: (1) as a film coating on soil particles, and (2) as water
stored in the pore spaces between soil particles (Klocke & Hergert, 1990). The moisture content
(MC) of the soil indicates the amount of water stored in it (Brouwer, Goffeau & Heibloem, 1985).
It is described as saturation, field capacity, and permanent wilting point (Figure 6).

A definition sketch of these soil MC is reflected in Figure 7. The actual volume of water
present in the soil is a function of soil texture (Evans, Cassel & Sneed, 1996), that is, the heavier
the soil texture, the more soil moisture it can hold as shown in Figure 8.

To determine soil MC, an oven-dry soil sample is primarily used as reference point.
7

(a) Saturation (b) Field capacity (c) Permanent wilting point


Source: Lajos (2008)

Figure 6. Schematic representation of soil moisture content

Figure 7. Definition sketch of soil moisture content


8

Source: Anonymous (n.d)

Figure 8. Relationship of soil moisture content and soil texture

1. Saturation

During basin irrigation or prolonged heavy rainfall, the soil becomes saturated, that is,
when all pore spaces are filled with water (Figure 9). When the soil is at or near saturation, some
of the water freely drains due to gravity (Evans, Cassel & Sneed, 1996). This is the water we
referred to earlier as gravitational water.

Source: University of California at Santa Cruz (n.d)

Figure 9. Soil at saturated condition


9

2. Field capacity

After one to three days of free drainage, the soil is said to be at field capacity (Figure 10).
The water remaining in the soil at field capacity is what we previously called capillary water.
Under field capacity condition, the soil contains the greatest amount of water that is potentially
available to plants. The actual volume of water present when the soil is at field capacity
depends on the soil texture (Evans, Cassel & Sneed, 1996), that is, the heavier the soil texture, the
more soil moisture it can hold as shown in Figure 8. In general, the soil water tension at field
capacity is often estimated at 0.1 to 0.3 bar (Shreeja, n.d), which is low enough that plants can
very easily extract water from the soil.

Source: University of California at Santa Cruz (n.d)

Figure 10. Soil at field capacity

3. Temporary and permanent wilting points

According to Evans, Cassel & Sneed, as water is depleted from the soil, the soil water
tension becomes high [up to < 10 bars (Shreeja, n.d)] and the plant hardly takes up water. This is
called temporary wilting point (Figure 11). At temporary wilting point, wilting is evident during
daytime but the plant recovers during nighttime when evapotranspiration is relatively less. When
temporary wilting is manifested, irrigation water must be applied within 3 to 5 days (1996).
10

Source: University of California at Santa Cruz (n.d)

Figure 11. Soil at permanent wilting point

When temporary wilting condition persists and still no water is added to the soil, the soil
water tension becomes even higher [15 bars (Shreeja, n.d)] until the plant is unable to take up
anymore of the remaining water and permanent wilting results. When the plant has removed all
available water, the soil moisture content has reached the permanent wilting point as shown in
Figure 7 (Evans, Cassel & Sneed, 1996). The effect of permanent wilting point is irreversible, that is,
the plant will eventually die even when water is added to the soil.

Most field crops will recover overnight from temporary wilting if less than 50 percent of
the total available moisture (TAM) has been depleted. TAM is the volume of water stored in the
soil reservoir that can be used by plants. It is the difference between the volume of water stored
when the soil is at field capacity (FC) and the volume still remaining when the soil reaches
permanent wilting point (PWP) (Evans, Cassel & Sneed, 1996). Mathematically,

TAM = FC – PWP (Eq. 1)

TAM, sometimes called plant available water or available water capacity of the soil,
stored in the soil reservoir is commonly expressed as the depth of water per unit depth of soil.
Typical units are inches of TAM per inch or foot of soil depth (in/in or in/ft). Different types of soil
have different TAMs (Table 1). Note that the values in Table 1 are used for irrigation scheduling,
which is beyond the scope of this course.
11

Table 1. Total available moisture capacity based on soil texture


Soil Texture Total Available Moisture Capacity (in/ft)
Coarse sand1/ 0.25 – 0.75
Fine sand1/ 0.75 – 1.00
Loam sand1/ 1.10 – 1.20
Sandy loam1/ 1.25 – 1.40
Fine sandy loam1/ 1.50 – 2.00
Silt loam1/ 2.00 – 2.50
Silty clay loam1/ 1.80 – 2.00
Silty clay1/ 1.50 – 1.70
Clay1/ 1.20 – 1.50
Peat and muck2/ 2.00 – 3.00
Sources: 1/ Plant and Soil Sciences eLibrary (2018)
2/ Anonymous (n.d)

The amount of TAM remaining in the soil decreases as plants extract water from the soil.
The amount of TAM removed since the last irrigation or rainfall is called depletion volume. Crop
yield and quality are not affected for as long as the amount of water used by the crop does not
exceed the allowable depletion volume. This is indicated in Figure 7 as readily available moisture
(RAM). The allowable depletion of TAM depends on the soil and the crop (Evans, Cassel &
Sneed, 1996).

In irrigation management particularly irrigation scheduling, the allowable depletion


volume is generally set at 50 percent. However, according to Abubaker, the RAM may range
from 40 percent or less in sandy soils to greater than 60 percent in clayey soils. The allowable
depletion is also dependent on the type of crop, its stage of development, and its sensitivity to
drought stress. For example, the allowable depletion recommended for some drought-sensitive
crops particularly vegetable crops is only 20 percent during critical stages of development. The
allowable depletion may approach 70 percent during non-critical periods for drought-tolerant
crops such as soybeans or cotton (2009).

4. Oven-dry condition

An oven-dry soil does not exist naturally under Earth’s condition. It can be attained only
under controlled laboratory environment. This is when all water has been removed from the soil.
According to Shreeja, the soil water tension of an oven-dried soil sample is 10,000 bars (n.d). if
we are to determine the soil MC, then soil samples must be taken from the field and oven-dried
in the laboratory. The oven-dry weight of the sample is then used as an input to the soil MC
equation.
12

IV. Plant Factors


Aside from climatic factors such as air temperature, humidity and wind speed, plant
factors such as growth stage, rooting depth and sensitivity to drought stress including soil and
water quality determine when, where, and how much water a plant will use.

A. Growth stage

A plant needs water differently at different growth stages. A young plant requires less
water than when it is in the vegetative stage. When the plant approaches maturity, its water
need declines. In general, according to Rogers & Sothers (1996), this water use and crop growth
stage relationship resembles that of the curve in Figure 12.

Source: Rogers & Sothers (1996)

Figure 12. Crop water use at different growth stages

B. Potential and effective root depths

Potential rooting depth of crops is different from their effective rooting depth, which is
always less than the former. The potential rooting depth is the maximum rooting depth of a crop
when grown in a moist soil with no barriers or restrictions that inhibit root elongation. For most
agricultural field crops, the potential rooting depth ranges from about 2 to 5 feet (Evans, Cassel
& Sneed, 1996).

Water uptake by specific crop is closely related to its root development and distribution
in the soil. As shown in Figure 13, about 70 percent of plant roots are found in the upper half of
the crop’s maximum rooting depth. When adequate moisture is present, water uptake by the
crop is about the same as its root distribution. Thus, about 70 percent of the water used by the
crop comes from the upper half of the root zone. This zone is the effective root depth
(onlineworksuite.com, 2016).
13

Source: Rogers & Sothers (1996) Source: Anonymous (n.d)

Figure 13. Moisture extraction at varying root depths

The effective root depth is the depth that should be used to compute the volume of TAM
in the soil reservoir (Anonymous, n.d). For a mature crop, the effective rooting depth is estimated
to be one-half the potential rooting depth reported in literatures. For example, if a given crop
has a maximum rooting depth of 2 feet, then the maximum effective root depth is estimated to
be 1 foot. According to Evans, Cassel & Sneed, effective root depth is influenced by the stage of
crop development. Effective root depth for most agricultural field crops increases as top growth
increases until the reproductive stage is reached. After this time, it remains fairly constant (1996).

C. Crop sensitivity to drought stress

The reduction in crop yield or quality resulting from drought stress depends on the stage
of crop development. For most agricultural field crops, the most critical irrigation period typically
begins just before the reproductive stage and lasts about 30 to 40 days to the end of the fruit
enlargement or grain development stage. For example, corn is most susceptible to drought
stress at tasseling and silking stages (Figure 14), which is about 65 to 75 days after seeding. For a
given level of stress, the corn yield reduction would be four times greater at these stages than at
knee-high stage. From the yield standpoint, applying irrigation water at tasseling and silking
stages would be worth four times more than if the same amount of water was applied during the
knee-high stage. Knowledge of this relationship is most useful when the irrigation capacity or
water supply is limited. When water is in short supply, irrigation should be delayed or cancelled
during the least susceptible crop growth stages. This water can then be reserved for use during
more sensitive growth stages (Evans, Cassel & Sneed, 1996).
14

Source: Anonymous (n.d)

Figure 14. Crop water use at different growth stages

D. Soil and water quality

Another factor on the amount of soil water available to the plant, according to Evans,
Cassel & Sneed, is the soil and water quality. For optimum plant growth and development, the
soil must have adequate room for water and air movement, and for root growth. The soil
structure can be altered by certain soil management practices. For example, excessive tillage
can break apart aggregated soil and excessive traffic can cause compaction. Both of these
practices reduce the amount of pore space in the soil, and thus reduce the availability of water
and air, and reduce the room for root development (1996).

They further mentioned that the quality of the water is also important to plant growth and
development. Irrigation water with a high content of soluble salt is not as available to the plant,
so a higher soil MC must be maintained in order to have water available to the plant. Increasing
salt content of the water reduces the potential to move water from the soil to the roots. Some
additional water would also be needed to leach the salt below the crop root zone to prevent
build-up in the soil. Poor quality water can affect soil structure.
15

Passing Score: 75 points


Due Date: ASAP but not later than ____________________________________.
Penalty for Late Submission: 5 points deduction per day of delay

If space is not enough, continue at the left-side directly opposite the item being answered.

1. What are the components of a soil system? Describe the composition of each. (15 pts)

2. Explain briefly why water is essential to plants? (15 pts)


16

3. List down at least 10 crops commonly grown in Ilocos Sur with their corresponding rooting
depth and growth stage most susceptible to drought stress. Cite the source(s) of said
information. (15 pts)

Crop Rooting Depth (cm) Growth Stage Most Source/Reference


Susceptible to
Drought Stress
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.

4. What are the different classifications of soil moisture? Describe their relative amounts in
clayey and sandy soils. (15 pts)
17

5. Express total available moisture (TAM) mathematically. (10 pts)

6. In your own words, describe the factors that determine when, where, and how much water
a plant needs. (15 pts)
18

7. Synthesize in not more than 150 words the relationship of soil, water and plants. (15 pts)
19

Abubaker, J. (2009). Irrigation scheduling for efficient water use in dry climates. Swedish University
of Agricultural Sciences {MS thesis]. Retrieved from
https://stud.epsilon.slu.se/619/1/abubaker_j_091116.pdf

Ahuja, M. (Ed.). (2006). Life Science-I. Delhi, India: Isha Books. Retrieved from
https://books.google.com.ph/books?isbn=8182053196

Ali, M.H. (2010). Fundamentals of irrigation and on-farm water management. New York: Springer.
Retrieved from https://books.google.com.ph/books?isbn=1441963359

Anonymous. (n.d). Pemanasan global dan perubahan iklim. Retrieved from


http://tep.fateta.unand.ac.id/images/MATERI_KULIAH/Bahan_Ajar/HUBUNGAN_TANAH_
AIR.pdf.

Brouwer, C., Goffeau, A. & Heibloem, M. (1985). Irrigation water management: Training Manual
No. 1 - Introduction to irrigation. Rome, Italy: Food and Agriculture Organization.
Retrieved from http://www.fao.org/docrep/r4082e/r4082e03.htm

Brown, K & Wherrett, A. (2017). Bulk density measurement. Soilquality.org.au Factsheet. Retrieved
from http://soilquality.org.au/factsheets/bulk-density-measurement

Clipart Library. (n.d). Agricultural and biosystems engineering [Logo]. Retrieved from
http://clipa.cash/agricultural-biosystems-engineering-logo.html

Evans, R., Cassel, D.K. & Sneed, R.E. 1996. Soil, water and crop characteristics important to
irrigation scheduling. Publication No. AG 452-1. North Carolina Cooperative Extension
Service. Retrieved from https://content.ces.ncsu.edu/soil-water-and-crop-characteristics-
important-to-irrigation-scheduling

Haman, D.Z. & Izuno, F.T. (2003). Soil plant water relationships. CIR1085. University of Florida.
Retrieved from http://ufdc.ufl.edu/IR00004480/00001

International Commission on Irrigation and Drainage. (n.d). ICID [Logo]. Retrieved from
https://icid2019.com/international-commision-on-irrigation-and-drainage/

Juhász, C. & Pregun, C. (2013). Water management. Retrieved from


https://www.tankonyvtar.hu/en/tartalom/tamop412A/2011_0009_Juhasz_Csaba_Pregun
_Csaba-Water_Management/ch05s03.html

Klocke, N.L. & Hergert, G.W. (1990). How soil holds water. Field Crops G-21, Cropping Practices.
University of Nebraska Cooperative Extension. Retrieved from
digitalcommons.unl.edu/cgi/viewcontent.cgi?article=1720&context=extensionhist

Lajos, B. (2008). Soil science. Retrieved from


https://www.tankonyvtar.hu/en/tartalom/tamop425/0032_talajtan/ch07s02.html
20

leavingcertbiology.net. (2018). Chapter 25: Nutrition in the flowering plant. Retrieved from
http://www.leavingcertbiology.net/chapter-25-nutrition-in-the-flowering-plant.html

Luther, G. (2016). Pasture: Soil-plant-water to animals inter-relationships. Retrieved from


http://godwinluther.blogspot.com/2016/03/pasture.html

National Academies of Sciences, Engineering and Medicine. (n.d). NASEM [Logo]. Retrieved
from http://sites.nationalacademies.org/PGA/biso/SS/index.htm

onlineworksuite.com. (2016). Crop productivity index CPI soil root shoot crop yield. Retrieved
from https://vdocuments.mx/crop-productivity-index-cpi-soil-root-shoot-crop-yield-
56ce6e94d4984.html

Plant and Soil Sciences eLibrary. (2018). Irrigation management. Retrieved from
http://croptechnology.unl.edu/pages/informationmodule.php?idinformationmodule=11
30447123&topicorder=3&maxto=13&minto=1

Rogers, D.H. & Sothers, W.M. (1996). Irrigation management: Soil, water and plant relationships.
Cooperative Extension Service. Kansas State University, Manhattan, Kansas. Retrieved
from https://www.ksre.k-state.edu/irrigate/reports/L904.pdf

Sabia Landscaping and Tree Service. (n.d). SLTS [Logos]. Retrieved from
http://www.sabialandscaping.com/landscaping/seasonal-yard-garden-services/ and
http://www.sabialandscaping.com/irrigation-drainage/

Shreeja, D. (n.d). Soil water: Importance, concepts and classification. Retrieved from
http://www.soilmanagementindia.com/soil-water/soil-water-importance-concepts-and-
classification/1790

StudyLib. (2018). Air, tanah and tanaman. Retrieved from


https://studylib.net/doc/9408664/hub.tanah--air-dan-tanaman.pp-t

University of California at Santa Cruz. (n.d). Irrigation principles and practices. Retrieved from
http://casfs.ucsc.edu/.../PDF-downloads/1.5-irrigation.pdf

Você também pode gostar