Soil moisture

Last updated

Soil moisture is the water content of the soil. It can be expressed in terms of volume or weight. Soil moisture measurement can be based on in situ probes (e.g., capacitance probes, neutron probes) or remote sensing methods. [1] [2]

Contents

Water that enters a field is removed from a field by runoff, drainage, evaporation or transpiration. [3] Runoff is the water that flows on the surface to the edge of the field; drainage is the water that flows through the soil downward or toward the edge of the field underground; evaporative water loss from a field is that part of the water that evaporates into the atmosphere directly from the field's surface; transpiration is the loss of water from the field by its evaporation from the plant itself.

Water affects soil formation, structure, stability and erosion but is of primary concern with respect to plant growth. [4] Water is essential to plants for four reasons:

  1. It constitutes 80–95% of the plant's protoplasm.
  2. It is essential for photosynthesis.
  3. It is the solvent in which nutrients are carried to, into and throughout the plant.
  4. It provides the turgidity by which the plant keeps itself in proper position. [5]

In addition, water alters the soil profile by dissolving and re-depositing mineral and organic solutes and colloids, often at lower levels, a process called leaching. In a loam soil, solids constitute half the volume, gas one-quarter of the volume, and water one-quarter of the volume of which only half will be available to most plants, with a strong variation according to matric potential. [6]

Water moves in soil under the influence of gravity, osmosis and capillarity. [7] When water enters the soil, it displaces air from interconnected macropores by buoyancy, and breaks aggregates into which air is entrapped, a process called slaking. [8] The rate at which a soil can absorb water depends on the soil and its other conditions. As a plant grows, its roots remove water from the largest pores (macropores) first. Soon the larger pores hold only air, and the remaining water is found only in the intermediate- and smallest-sized pores (micropores). The water in the smallest pores is so strongly held to particle surfaces that plant roots cannot pull it away. Consequently, not all soil water is available to plants, with a strong dependence on texture. [9] When saturated, the soil may lose nutrients as the water drains. [10] Water moves in a draining field under the influence of pressure where the soil is locally saturated and by capillarity pull to drier parts of the soil. [11] Most plant water needs are supplied from the suction caused by evaporation from plant leaves (transpiration) and a lower fraction is supplied by suction created by osmotic pressure differences between the plant interior and the soil solution. [12] [13] Plant roots must seek out water and grow preferentially in moister soil microsites, [14] but some parts of the root system are also able to remoisten dry parts of the soil. [15] Insufficient water will damage the yield of a crop. [16] Most of the available water is used in transpiration to pull nutrients into the plant. [17]

Soil water is also important for climate modeling and numerical weather prediction. The Global Climate Observing System specified soil water as one of the 50 Essential Climate Variables (ECVs). [18] Soil water can be measured in situ with soil moisture sensors or can be estimated at various scales and resolution: from local or wifi measures via sensors in the soil to satellite imagery that combines data capture and hydrological models. Each method exhibits pros and cons, and hence, the integration of different techniques may decrease the drawbacks of a single given method. [19]

Moisture level concepts

ECMWF soil moisture forecast for the East Asia region, showing the key moisture levels and intermediate measurements ECMWF soil moisture forecast, East Asia, 2022-04-12.png
ECMWF soil moisture forecast for the East Asia region, showing the key moisture levels and intermediate measurements
Field capacity
A flooded field will drain the gravitational water under the influence of gravity until water's adhesive and cohesive forces resist further drainage at which point it is said to have reached field capacity. [20] At that point, plants must apply suction to draw water from a soil. By convention it is defined at 0.33 bar suction. [20] [21]
Available water and unavailable water
The water that plants may draw from the soil is called the available water. [20] [22] Once the available water is used up the remaining moisture is called unavailable water as the plant cannot produce sufficient suction to draw that water in.
Wilting point
The wilting point is the minimum amount of water plants need to not wilt and approximates the boundary between available and unavailable water. By convention it is defined as 15 bar suction. At this point, seeds will not germinate, [23] [20] [24] plants begin to wilt and then die unless they are able to recover after water replenishment thanks to species-specific adaptations. [25]

Water retention

Water is retained in a soil when the adhesive force of attraction that water's hydrogen atoms have for the oxygen of soil particles is stronger than the cohesive forces that water's hydrogen feels for water oxygen atoms. [26] When a field is flooded, the soil pore space is completely filled by water. The field will drain under the force of gravity until it reaches what is called field capacity, at which point the smallest pores are filled with water and the largest with water and gases. [27] The total amount of water held when field capacity is reached is a function of the specific surface area of the soil particles. [28] As a result, high clay and high organic soils have higher field capacities. [29] The potential energy of water per unit volume relative to pure water in reference conditions is called water potential. Total water potential is a sum of matric potential which results from capillary action, osmotic potential for saline soil, and gravitational potential when dealing with downward water movement. Water potential in soil usually has negative values, and therefore it is also expressed in suction, which is defined as the minus of water potential. Suction has a positive value and can be regarded as the total force required to pull or push water out of soil. Water potential or suction is expressed in units of kPa (103 pascal), bar (100 kPa), or cm H2O (approximately 0.098 kPa). Common logarithm of suction in cm H2O is called pF. [30] Therefore, pF 3 = 1000 cm = 98 kPa = 0.98 bar.

The forces with which water is held in soils determine its availability to plants. Forces of adhesion hold water strongly to mineral and humus surfaces and less strongly to itself by cohesive forces. A plant's root may penetrate a very small volume of water that is adhering to soil and be initially able to draw in water that is only lightly held by the cohesive forces. But as the droplet is drawn down, the forces of adhesion of the water for the soil particles produce increasingly higher suction, finally up to 1500 kPa (pF = 4.2). [31] At 1500 kPa suction, the soil water amount is called wilting point. At that suction the plant cannot sustain its water needs as water is still being lost from the plant by transpiration, the plant's turgidity is lost, and it wilts, although stomatal closure may decrease transpiration and thus may retard wilting below the wilting point, in particular under adaptation or acclimatization to drought. [32] The next level, called air-dry, occurs at 100,000 kPa suction (pF = 6). Finally the oven dry condition is reached at 1,000,000 kPa suction (pF = 7). All water below wilting point is called unavailable water. [33]

When the soil moisture content is optimal for plant growth, the water in the large and intermediate size pores can move about in the soil and be easily used by plants. [9] The amount of water remaining in a soil drained to field capacity and the amount that is available are functions of the soil type. Sandy soil will retain very little water, while clay will hold the maximum amount. [29] The available water for the silt loam might be 20% whereas for the sand it might be only 6% by volume, as shown in this table.

Wilting point, field capacity, and available water of various soil textures (unit: % by volume) [34]
Soil TextureWilting PointField CapacityAvailable water
Sand3.39.15.8
Sandy loam9.520.711.2
Loam11.727.015.3
Silt loam13.333.019.7
Clay loam19.731.812.1
Clay27.239.612.4

The above are average values for the soil textures.

Water flow

Water moves through soil due to the force of gravity, osmosis and capillarity. At 0 to 33 kPa suction (field capacity), water is pushed through soil from the point of its application under the force of gravity and the pressure gradient created by differences in the pressure of water; this is called saturated flow. At higher suction, water movement is pulled by capillarity from wetter toward drier soil. This is caused by water's adhesion to soil solids, and is called unsaturated flow. [35] [36]

Water infiltration and movement in soil are controlled by six factors:

  1. Soil texture
  2. Soil structure. Fine-textured soils with granular structure are most favourable to infiltration of water.
  3. The amount of organic matter. Coarse matter is best and if on the surface helps prevent the destruction of soil structure and the creation of soil crusts.
  4. Depth of soil to impervious layers such as hardpans or bedrock
  5. The amount of water already in the soil
  6. Soil temperature. Warm soils take in water faster while frozen soils such as permafrost may not be able to absorb depending on the type of freezing. [37]

Water infiltration rates range from 0.25 cm per hour for high clay soils to 2.5 cm per hour for sand and well stabilized and aggregated soil structures. [38] Water flows through the ground unevenly, in the form of so-called gravity fingers, because of the surface tension between water particles. [39] [40]

Tree roots, whether living or dead, create preferential channels for rainwater flow through soil, [41] magnifying infiltration rates of water up to 27 times. [42]

Flooding temporarily increases soil permeability in river beds, helping to recharge aquifers. [43]

Water applied to a soil is pushed by pressure gradients from the point of its application where it is saturated locally, to less saturated areas, such as the vadose zone. [44] [45] Once soil is completely wetted, any more water will move downward, or percolate out of the range of plant roots, carrying with it clay, humus, nutrients, primarily cations, and various contaminants, including pesticides, pollutants, viruses and bacteria, potentially causing groundwater contamination. [46] [47] In order of decreasing solubility, the leached nutrients are:

In the United States percolation water due to rainfall ranges from almost zero centimeters just east of the Rocky Mountains to fifty or more centimeters per day in the Appalachian Mountains and the north coast of the Gulf of Mexico. [49]

Water is pulled by capillary action due to the adhesion force of water to the soil solids, producing a suction gradient from wet towards drier soil [50] and from macropores to micropores. [51] The so-called Richards equation allows calculation of the time rate of change of moisture content in soils due to the movement of water in unsaturated soils. [52] Interestingly, this equation attributed to Richards was originally published by Richardson in 1922. [53] The soil moisture velocity equation, [54] which can be solved using the finite water-content vadose zone flow method, [55] [56] describes the velocity of flowing water through an unsaturated soil in the vertical direction. The numerical solution of the Richardson/Richards equation allows calculation of unsaturated water flow and solute transport using software such as Hydrus, [57] by giving soil hydraulic parameters of hydraulic functions (water retention function and unsaturated hydraulic conductivity function) and initial and boundary conditions. Preferential flow occurs along interconnected macropores, crevices, root and worm channels, which drain water under gravity. [58] [59] Many models based on soil physics now allow for some representation of preferential flow as a dual continuum, dual porosity or dual permeability options, but these have generally been "bolted on" to the Richards solution without any rigorous physical underpinning. [60]

Water uptake by plants

Of equal importance to the storage and movement of water in soil is the means by which plants acquire it and their nutrients. Most soil water is taken up by plants as passive absorption caused by the pulling force of water evaporating (transpiring) from the long column of water (xylem sap flow) that leads from the plant's roots to its leaves, according to the cohesion-tension theory. [61] The upward movement of water and solutes (hydraulic lift) is regulated in the roots by the endodermis [62] and in the plant foliage by stomatal conductance, [63] and can be interrupted in root and shoot xylem vessels by cavitation, also called xylem embolism. [64] In addition, the high concentration of salts within plant roots creates an osmotic pressure gradient that pushes soil water into the roots. [65] Osmotic absorption becomes more important during times of low water transpiration caused by lower temperatures (for example at night) or high humidity, and the reverse occurs under high temperature or low humidity. It is these processes that cause guttation and wilting, respectively. [66] [67]

Root extension is vital for plant survival. A study of a single winter rye plant grown for four months in one cubic foot (0.0283 cubic meters) of loam soil showed that the plant developed 13,800,000 roots, a total of 620 km in length with 237 square meters in surface area; and 14 billion root hairs of 10,620 km total length and 400 square meters total area; for a total surface area of 638 square meters. The total surface area of the loam soil was estimated to be 52,000 square meters. [68] In other words, the roots were in contact with only 1.2% of the soil volume. However, root extension should be viewed as a dynamic process, allowing new roots to explore a new volume of soil each day, increasing dramatically the total volume of soil explored over a given growth period, and thus the volume of water taken up by the root system over this period. [69] Root architecture, i.e. the spatial configuration of the root system, plays a prominent role in the adaptation of plants to soil water and nutrient availability, and thus in plant productivity. [70]

Roots must seek out water as the unsaturated flow of water in soil can move only at a rate of up to 2.5 cm per day; as a result they are constantly dying and growing as they seek out high concentrations of soil moisture. [71] Insufficient soil moisture, to the point of causing wilting, will cause permanent damage and crop yields will suffer. When grain sorghum was exposed to soil suction as low as 1300 kPa during the seed head emergence through bloom and seed set stages of growth, its production was reduced by 34%. [72]

Consumptive use and water use efficiency

Only a small fraction (0.1% to 1%) of the water used by a plant is held within the plant. The majority is ultimately lost via transpiration, while evaporation from the soil surface is also substantial, the transpiration:evaporation ratio (T/ET) varying according to vegetation type and climate, peaking in tropical rainforests and dipping in steppes and deserts. [73] Transpiration plus evaporative soil moisture loss is called evapotranspiration. Evapotranspiration plus water held in the plant totals to consumptive use, which is nearly identical to evapotranspiration. [72] [74]

The total water used in an agricultural field includes surface runoff, drainage and consumptive use. The use of loose mulches will reduce evaporative losses for a period after a field is irrigated, but in the end the total evaporative loss (plant plus soil) will approach that of an uncovered soil, while more water is immediately available for plant growth. [75] Water use efficiency is measured by the transpiration ratio, which is the ratio of the total water transpired by a plant to the dry weight of the harvested plant. Transpiration ratios for crops range from 300 to 700. For example, alfalfa may have a transpiration ratio of 500 and as a result 500 kilograms of water will produce one kilogram of dry alfalfa. [76]

Related Research Articles

<span class="mw-page-title-main">Soil</span> Mixture of organic matter, minerals, gases, liquids, and organisms that together support life

Soil, also commonly referred to as earth or dirt, is a mixture of organic matter, minerals, gases, liquids, and organisms that together support the life of plants and soil organisms. Some scientific definitions distinguish dirt from soil by restricting the former term specifically to displaced soil.

<span class="mw-page-title-main">Xylem</span> Water transport tissue in vascular plants

Xylem is one of the two types of transport tissue in vascular plants, the other being phloem. The basic function of the xylem is to transport water from roots to stems and leaves, but it also transports nutrients. The word xylem is derived from the Ancient Greek word ξύλον (xylon), meaning "wood"; the best-known xylem tissue is wood, though it is found throughout a plant. The term was introduced by Carl Nägeli in 1858.

<span class="mw-page-title-main">Evapotranspiration</span> Process by when water moves into the air from plants and soil.

Evapotranspiration (ET) is the combined processes which move water from the Earth's surface into the atmosphere. It covers both water evaporation and transpiration. Evapotranspiration is an important part of the local water cycle and climate, and measurement of it plays a key role in agricultural irrigation and water resource management.

Water potential is the potential energy of water per unit volume relative to pure water in reference conditions. Water potential quantifies the tendency of water to move from one area to another due to osmosis, gravity, mechanical pressure and matrix effects such as capillary action. The concept of water potential has proved useful in understanding and computing water movement within plants, animals, and soil. Water potential is typically expressed in potential energy per unit volume and very often is represented by the Greek letter ψ.

<span class="mw-page-title-main">Ecohydrology</span>

Ecohydrology is an interdisciplinary scientific field studying the interactions between water and ecological systems. It is considered a sub discipline of hydrology, with an ecological focus. These interactions may take place within water bodies, such as rivers and lakes, or on land, in forests, deserts, and other terrestrial ecosystems. Areas of research in ecohydrology include transpiration and plant water use, adaption of organisms to their water environment, influence of vegetation and benthic plants on stream flow and function, and feedbacks between ecological processes, the soil carbon sponge and the hydrological cycle.

<span class="mw-page-title-main">Water content</span> Quantity of water contained in a material

Water content or moisture content is the quantity of water contained in a material, such as soil, rock, ceramics, crops, or wood. Water content is used in a wide range of scientific and technical areas, and is expressed as a ratio, which can range from 0 to the value of the materials' porosity at saturation. It can be given on a volumetric or mass (gravimetric) basis.

<span class="mw-page-title-main">Infiltration (hydrology)</span> Process by which water on the ground surface enters the soil

Infiltration is the process by which water on the ground surface enters the soil. It is commonly used in both hydrology and soil sciences. The infiltration capacity is defined as the maximum rate of infiltration. It is most often measured in meters per day but can also be measured in other units of distance over time if necessary. The infiltration capacity decreases as the soil moisture content of soils surface layers increases. If the precipitation rate exceeds the infiltration rate, runoff will usually occur unless there is some physical barrier.

Field capacity is the amount of soil moisture or water content held in the soil after excess water has drained away and the rate of downward movement has decreased. This usually takes place 2–3 days after rain or irrigation in pervious soils of uniform structure and texture. The physical definition of field capacity is the bulk water content retained in soil at −33 kPa of hydraulic head or suction pressure. The term originated from Israelsen and West and Frank Veihmeyer and Arthur Hendrickson.

<span class="mw-page-title-main">Permanent wilting point</span>

Permanent wilting point (PWP) or wilting point (WP) is defined as the minimum amount of water in the soil that the plant requires not to wilt. If the soil water content decreases to this or any lower point a plant wilts and can no longer recover its turgidity when placed in a saturated atmosphere for 12 hours. The physical definition of the wilting point, symbolically expressed as θpwp or θwp, is said by convention as the water content at −1,500 kPa (−15 bar) of suction pressure, or negative hydraulic head.

<span class="mw-page-title-main">Water retention curve</span>

Water retention curve is the relationship between the water content, θ, and the soil water potential, ψ. This curve is characteristic for different types of soil, and is also called the soil moisture characteristic.

Pore water pressure refers to the pressure of groundwater held within a soil or rock, in gaps between particles (pores). Pore water pressures below the phreatic level of the groundwater are measured with piezometers. The vertical pore water pressure distribution in aquifers can generally be assumed to be close to hydrostatic.

Moisture stress is a form of abiotic stress that occurs when the moisture of plant tissues is reduced to suboptimal levels. Water stress occurs in response to atmospheric and soil water availability when the transpiration rate exceeds the rate of water uptake by the roots and cells lose turgor pressure. Moisture stress is described by two main metrics, water potential and water content.

In soil, macropores are defined as cavities that are larger than 75 μm. Functionally, pores of this size host preferential soil solution flow and rapid transport of solutes and colloids. Macropores increase the hydraulic conductivity of soil, allowing water to infiltrate and drain quickly, and shallow groundwater to move relatively rapidly via lateral flow. In soil, macropores are created by plant roots, soil cracks, soil fauna, and by aggregation of soil particles into peds. Macropores can also be found in soil between larger individual mineral particles such as sand or gravel.

<span class="mw-page-title-main">Agricultural soil science</span> Branch of soil science

Agricultural soil science is a branch of soil science that deals with the study of edaphic conditions as they relate to the production of food and fiber. In this context, it is also a constituent of the field of agronomy and is thus also described as soil agronomy.

Vegetation and slope stability are interrelated by the ability of the plant life growing on slopes to both promote and hinder the stability of the slope. The relationship is a complex combination of the type of soil, the rainfall regime, the plant species present, the slope aspect, and the steepness of the slope. Knowledge of the underlying slope stability as a function of the soil type, its age, horizon development, compaction, and other impacts is a major underlying aspect of understanding how vegetation can alter the stability of the slope. There are four major ways in which vegetation influences slope stability: wind throwing, the removal of water, mass of vegetation (surcharge), and mechanical reinforcement of roots.

A xerophyte is a species of plant that has adaptations to survive in an environment with little liquid water. Examples of xerophytes include cacti, pineapple and some gymnosperm plants. The morphology and physiology of xerophytes are adapted to conserve water during dry periods. Some species called resurrection plants can survive long periods of extreme dryness or desiccation of their tissues, during which their metabolic activity may effectively shut down. Plants with such morphological and physiological adaptations are said to be xeromorphic. Xerophytes such as cacti are capable of withstanding extended periods of dry conditions as they have deep-spreading roots and capacity to store water. Their waxy, thorny leaves prevent loss of moisture.

<span class="mw-page-title-main">Transpiration</span> Process of water moving through a plant parts

Transpiration is the process of water movement through a plant and its evaporation from aerial parts, such as leaves, stems and flowers. It is a passive process that requires no energy expense by the plant. Transpiration also cools plants, changes osmotic pressure of cells, and enables mass flow of mineral nutrients. When water uptake by the roots is less than the water lost to the atmosphere by evaporation plants close small pores called stomata to decrease water loss, which slows down nutrient uptake and decreases CO2 absorption from the atmosphere limiting metabolic processes, photosynthesis, and growth.

Hydraulic redistribution is a passive mechanism where water is transported from moist to dry soils via subterranean networks. It occurs in vascular plants that commonly have roots in both wet and dry soils, especially plants with both taproots that grow vertically down to the water table, and lateral roots that sit close to the surface. In the late 1980s, there was a movement to understand the full extent of these subterranean networks. Since then it was found that vascular plants are assisted by fungal networks which grow on the root system to promote water redistribution.

DPHM-RS is a semi-distributed hydrologic model developed at University of Alberta, Canada.

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

References

  1. Zhang, Lijie; Zeng, Yijian; Zhuang, Ruodan; Szabó, Brigitta; Manfreda, Salvatore; Han, Qianqian; Su, Zhongbo (2021-12-02). "In Situ Observation-Constrained Global Surface Soil Moisture Using Random Forest Model". Remote Sensing. 13 (23): 4893. Bibcode:2021RemS...13.4893Z. doi: 10.3390/rs13234893 . ISSN   2072-4292.
  2. Albergel, Clement; de Rosnay, Patricia; Gruhier, Claire; Muñoz-Sabater, Joaquin; Hasenauer, Stefan; Isaksen, Lars; Kerr, Yann; Wagner, Wolfgang (March 2012). "Evaluation of remotely sensed and modelled soil moisture products using global ground-based in situ observations". Remote Sensing of Environment. 118: 215–226. Bibcode:2012RSEnv.118..215A. doi:10.1016/j.rse.2011.11.017.
  3. Wallace, James S.; Batchelor, Charles H. (1997). "Managing water resources for crop production". Philosophical Transactions of the Royal Society B: Biological Sciences . 352 (1356): 937–47. doi:10.1098/rstb.1997.0073. PMC   1691982 . Retrieved 14 August 2022.
  4. Veihmeyer, Frank J.; Hendrickson, Arthur H. (1927). "Soil-moisture conditions in relation to plant growth". Plant Physiology . 2 (1): 71–82. doi: 10.1104/pp.2.1.71 . PMC   439946 . PMID   16652508.
  5. Donahue, Miller & Shickluna 1977, p. 72.
  6. Ratliff, Larry F.; Ritchie, Jerry T.; Cassel, D. Keith (1983). "Field-measured limits of soil water availability as related to laboratory-measured properties". Soil Science Society of America Journal . 47 (4): 770–75. Bibcode:1983SSASJ..47..770R. doi:10.2136/sssaj1983.03615995004700040032x . Retrieved 14 August 2022.
  7. "Water movement in soils". Oklahoma State University, Department of Plant and Soil Sciences. Stillwater, Oklahoma. Retrieved 14 August 2022.
  8. Le Bissonnais, Yves (2016). "Aggregate stability and assessment of soil crustability and erodibility. I. Theory and methodology". European Journal of Soil Science. 67 (1): 11–21. doi:10.1111/ejss.4_12311. S2CID   247704630 . Retrieved 14 August 2022.
  9. 1 2 Easton, Zachary M.; Bock, Emily (22 March 2016). "Soil and soil water relationships" (PDF). Virginia Tech . hdl:10919/75545 . Retrieved 14 August 2022.
  10. Sims, J. Thomas; Simard, Régis R.; Joern, Brad Christopher (1998). "Phosphorus loss in agricultural drainage: historical perspective and current research". Journal of Environmental Quality . 27 (2): 277–93. doi:10.2134/jeq1998.00472425002700020006x . Retrieved 14 August 2022.
  11. Brooks, R.H.; Corey, Arthur T. (1964). Hydraulic properties of porous media (PDF). Fort Collins, Colorado: Colorado State University . Retrieved 14 August 2022.
  12. McElrone, Andrew J.; Choat, Brendan; Gambetta, Greg A.; Brodersen, Craig R. "Water uptake and transport in vascular plants" (PDF). Retrieved 14 August 2022.
  13. Steudle, Ernst (2000). "Water uptake by plant roots: an integration of views". Plant and Soil . 226 (1): 45–56. doi:10.1023/A:1026439226716. S2CID   3338727 . Retrieved 14 August 2022.
  14. Wilcox, Carolyn S.; Ferguson, Joseph W.; Fernandez, George C.J.; Nowak, Robert S. (2004). "Fine root growth dynamics of four Mojave Desert shrubs as related to soil moisture and microsite". Journal of Arid Environments . 56 (1): 129–48. Bibcode:2004JArEn..56..129W. doi:10.1016/S0140-1963(02)00324-5.
  15. Hunter, Albert S.; Kelley, Omer J. (1946). "The extension of plant roots into dry soil". Plant Physiology . 21 (4): 445–51. doi: 10.1104/pp.21.4.445 . PMC   437296 . PMID   16654059.
  16. Zhang, Yongqiang; Kendy, Eloise; Qiang, Yu; Liu, Changming; Shen, Yanjun; Sun, Hongyong (2004). "Effect of soil water deficit on evapotranspiration, crop yield, and water use efficiency in the North China Plain". Agricultural Water Management. 64 (2): 107–22. doi:10.1016/S0378-3774(03)00201-4 . Retrieved 14 August 2022.
  17. Oyewole, Olusegun Ayodeji; Inselsbacher, Erich; Näsholm, Torgny (2014). "Direct estimation of mass flow and diffusion of nitrogen compounds in solution and soil". New Phytologist . 201 (3): 1056–64. doi:10.1111/nph.12553. PMID   24134319.
  18. "Essential Climate Variables". Global Climate Observing System. 2013. Retrieved 14 August 2022.
  19. Brocca, Luca; Hasenauer, Stefan; Lacava, Teodosio; Moramarco, Tommaso; Wagner, Wolfgang; Dorigo, Wouter; Matgen, Patrick; Martínez-Fernández, José; Llorens, Pilar; Latron, Jérôme; Martin, Claude; Bittelli, Marco (2011). "Soil moisture estimation through ASCAT and AMSR-E sensors: an intercomparison and validation study across Europe". Remote Sensing of Environment . 115 (12): 3390–3408. Bibcode:2011RSEnv.115.3390B. doi:10.1016/j.rse.2011.08.003 . Retrieved 14 August 2022.
  20. 1 2 3 4 Wadleigh 1957, p. 48.
  21. Richards & Richards 1957, p. 50.
  22. Richards & Richards 1957, p. 56.
  23. Wadleigh 1957, p. 39.
  24. Richards & Richards 1957, p. 52.
  25. Snyman, Henny A.; Venter, W.D.; Van Rensburg, W.L.J.; Opperman, D.P.J. (1987). "Ranking of grass species according to visible wilting order and rate of recovery in the Central Orange Free State". Journal of the Grassland Society of Southern Africa . 4 (2): 78–81. doi:10.1080/02566702.1987.9648075 . Retrieved 14 August 2022.
  26. Donahue, Miller & Shickluna 1977, pp. 72–74.
  27. "Soil and water". Food and Agriculture Organization of the United Nations . Retrieved 21 August 2022.
  28. Petersen, Lis Wollesen; Møldrup, Per; Jacobsen, Ole H.; Rolston, Dennis E. (1996). "Relations between specific surface area and soil physical and chemical properties". Soil Science. 161 (1): 9–21. Bibcode:1996SoilS.161....9P. doi:10.1097/00010694-199601000-00003 . Retrieved 21 August 2022.
  29. 1 2 Gupta, Satish C.; Larson, William E. (1979). "Estimating soil water retention characteristics from particle size distribution, organic matter percent, and bulk density". Water Resources Research . 15 (6): 1633–35. Bibcode:1979WRR....15.1633G. CiteSeerX   10.1.1.475.497 . doi:10.1029/WR015i006p01633 . Retrieved 21 August 2022.
  30. "Soil water potential". AgriInfo.in. Archived from the original on 17 August 2017. Retrieved 15 March 2019.
  31. Savage, Michael J.; Ritchie, Joe T.; Bland, William L.; Dugas, William A. (1996). "Lower limit of soil water availability". Agronomy Journal . 88 (4): 644–51. doi:10.2134/agronj1996.00021962008800040024x . Retrieved 21 August 2022.
  32. Al-Ani, Tariq; Bierhuizen, Johan Frederik (1971). "Stomatal resistance, transpiration, and relative water content as influenced by soil moisture stress" (PDF). Acta Botanica Neerlandica . 20 (3): 318–26. doi:10.1111/j.1438-8677.1971.tb00715.x . Retrieved 21 August 2022.
  33. Donahue, Miller & Shickluna 1977, pp. 75–76.
  34. Rawls, Walter J.; Brakensiek, Donald L.; Saxtonn, Keith E. (1982). "Estimation of soil water properties". Transactions of the ASAE . 25 (5): 1316–1320. doi:10.13031/2013.33720 . Retrieved 28 August 2022.
  35. Donahue, Miller & Shickluna 1977, p. 85.
  36. "Soil water movement: saturated and unsaturated flow and vapour movement, soil moisture constants and their importance in irrigation" (PDF). Tamil Nadu Agricultural University . Retrieved 28 August 2022.
  37. Donahue, Miller & Shickluna 1977, p. 86.
  38. Donahue, Miller & Shickluna 1977, p. 88.
  39. Cueto-Felgueroso, Luis; Juanes, Ruben (2008). "Nonlocal interface dynamics and pattern formation in gravity-driven unsaturated flow through porous media". Physical Review Letters . 101 (24): 244504. Bibcode:2008PhRvL.101x4504C. doi:10.1103/PhysRevLett.101.244504. PMID   19113626. S2CID   21874968 . Retrieved 28 August 2022.
  40. "Finger flow in coarse soils". Cornell University . Retrieved 28 August 2022.
  41. Ghestem, Murielle; Sidle, Roy C.; Stokes, Alexia (2011). "The influence of plant root systems on subsurface flow: implications for slope stability". BioScience . 61 (11): 869–79. doi:10.1525/bio.2011.61.11.6.
  42. Bartens, Julia; Day, Susan D.; Harris, J. Roger; Dove, Joseph E.; Wynn, Theresa M. (2008). "Can urban tree roots improve infiltration through compacted subsoils for stormwater management?". Journal of Environmental Quality . 37 (6): 2048–57. doi:10.2134/jeq2008.0117. PMID   18948457 . Retrieved 28 August 2022.
  43. Zhang, Guohua; Feng, Gary; Li, Xinhu; Xie, Congbao; P, Xiaoyu (2017). "Flood effect on groundwater recharge on a typical silt loam soil". Water. 9 (7): 523. doi: 10.3390/w9070523 .
  44. Nielsen, Donald R.; Biggar, James W.; Erh, Koon T. (1973). "Spatial variability of field-measured soil-water properties". Hilgardia . 42 (7): 215–59. doi: 10.3733/hilg.v42n07p215 .
  45. Rimon, Yaara; Dahan, Ofer; Nativ, Ronit; Geyer, Stefan (2007). "Water percolation through the deep vadose zone and groundwater recharge: preliminary results based on a new vadose zone monitoring system". Water Resources Research . 43 (5): W05402. Bibcode:2007WRR....43.5402R. doi: 10.1029/2006WR004855 .
  46. Weiss, Peter T.; LeFevre, Greg; Gulliver, John S. (2008). "Contamination of soil and groundwater due to stormwater infiltration practices: a literature review" (PDF). Saint Anthony Falls Laboratory . CiteSeerX   10.1.1.410.5113 . Retrieved 28 August 2022.
  47. Hagedorn, Charles; Hansen, Debra T.; Simonson, Gerald H. (1978). "Survival and movement of fecal indicator bacteria in soil under conditions of saturated flow". Journal of Environmental Quality . 7 (1): 55–59. doi:10.2134/jeq1978.00472425000700010011x. S2CID   774611 . Retrieved 28 August 2022.
  48. Donahue, Miller & Shickluna 1977, p. 90.
  49. Donahue, Miller & Shickluna 1977, p. 80.
  50. Ng, Charles W.W.; Pang, Wenyan (2000). "Influence of stress state on soil-water characteristics and slope stability". Journal of Geotechnical & Geoenvironmental Engineering . 126 (2): 157–66. doi:10.1061/(ASCE)1090-0241(2000)126:2(157) . Retrieved 4 September 2022.
  51. Germann, Peter Fritz; Beven, Keith (2006). "Water flow in soil macropores. I. An experimental approach". European Journal of Soil Science. 32 (1): 1–13. doi:10.1111/j.1365-2389.1981.tb01681.x . Retrieved 4 September 2022.
  52. Richards, Lorenzo A. (1931). "Capillary conduction of liquids through porous mediums". Physics . 1 (5): 318–33. Bibcode:1931Physi...1..318R. doi:10.1063/1.1745010 . Retrieved 4 September 2022.
  53. Richardson, Lewis Fry (1922). Weather prediction by numerical process. Cambridge, United Kingdom: Cambridge University Press. p. 262. Retrieved 4 September 2022.
  54. Ogden, Fred L.; Allen, Myron B.; Lai, Wencong; Zhu, Julian; Douglas, Craig C.; Seo, Mookwon; Talbot, Cary A. (2017). "The soil moisture velocity equation". Journal of Advances in Modeling Earth Systems. 9 (2): 1473–87. Bibcode:2017JAMES...9.1473O. doi: 10.1002/2017MS000931 .
  55. Talbot, Cary A.; Ogden, Fred L. (2008). "A method for computing infiltration and redistribution in a discretized moisture content domain". Water Resources Research . 44 (8): W08453. Bibcode:2008WRR....44.8453T. doi: 10.1029/2008WR006815 .
  56. Ogden, Fred L.; Lai, Wencong; Steinke, Robert C.; Zhu, Julian; Talbot, Cary A.; Wilson, John L. (2015). "A new general 1-D vadose zone solution method". Water Resources Research . 51 (6): 4282–4300. Bibcode:2015WRR....51.4282O. doi:10.1002/2015WR017126. S2CID   119834716.
  57. Šimůnek, Jiri; Saito, Hirotaka; Sakai, Masaru; Van Genuchten, Martinus Th. (2013). "The HYDRUS-1D software package for simulating the one-dimensional movement of water, heat, and multiple solutes in variably-saturated media" . Retrieved 11 September 2022.
  58. Bouma, Johan (1981). "Soil morphology and preferential flow along macropores". Geoderma. 3 (4): 235–50. doi:10.1016/0378-3774(81)90009-3 . Retrieved 11 September 2022.
  59. Luo, Lifang; Lin, Henry; Halleck, Phil (2008). "Quantifying soil structure and preferential flow in intact soil Using X-ray computed tomography". Soil Science Society of America Journal . 72 (4): 1058–69. Bibcode:2008SSASJ..72.1058L. CiteSeerX   10.1.1.455.2567 . doi:10.2136/sssaj2007.0179 . Retrieved 11 September 2022.
  60. Beven, Keith; Germann, Peter (2013). "Macropores and water flow in soils revisited" (PDF). Water Resources Research . 49 (6): 3071–92. Bibcode:2013WRR....49.3071B. doi:10.1002/wrcr.20156. S2CID   53132908 . Retrieved 11 September 2022.
  61. Aston, Mervyn J.; Lawlor, David W. (1979). "The relationship between transpiration, root water uptake, and leaf water potential". Journal of Experimental Botany . 30 (1): 169–81. doi:10.1093/jxb/30.1.169 . Retrieved 18 September 2022.
  62. Powell, D.B.B. (1978). "Regulation of plant water potential by membranes of the endodermis in young roots". Plant, Cell and Environment . 1 (1): 69–76. doi:10.1111/j.1365-3040.1978.tb00749.x . Retrieved 18 September 2022.
  63. Irvine, James; Perks, Michael P.; Magnani, Federico; Grace, John (1998). "The response of Pinus sylvestris to drought: stomatal control of transpiration and hydraulic conductance". Tree Physiology. 18 (6): 393–402. doi: 10.1093/treephys/18.6.393 . PMID   12651364.
  64. Jackson, Robert B.; Sperry, John S.; Dawson, Todd E. (2000). "Root water uptake and transport: using physiological processes in global predictions" (PDF). Trends in Plant Science . 5 (11): 482–88. doi:10.1016/S1360-1385(00)01766-0. PMID   11077257. S2CID   8311441 . Retrieved 11 November 2022.
  65. Steudle, Ernst (2000). "Water uptake by plant roots: an integration of views". Plant and Soil . 226 (1): 45–56. doi:10.1023/A:1026439226716. S2CID   3338727 . Retrieved 18 September 2022.
  66. Donahue, Miller & Shickluna 1977, p. 92.
  67. Kaufmann, Merrill R.; Eckard, Alan N. (1971). "Evaluation of water stress control with polyethylene glycols by analysis of guttation". Plant Physiology . 47 (4): 453–56. doi:10.1104/pp.47.4.453. PMC   396708 . PMID   16657642 . Retrieved 18 September 2022.
  68. Wadleigh 1957, p. 46.
  69. Kramer, Paul J.; Coile, Theodore S. (1940). "An estimation of the volume of water made available by root extension". Plant Physiology . 15 (4): 743–47. doi:10.1104/pp.15.4.743. PMC   437871 . PMID   16653671 . Retrieved 18 September 2022.
  70. Lynch, Jonathan (1995). "Root architecture and plant productivity". Plant Physiology . 109 (1): 7–13. doi:10.1104/pp.109.1.7. PMC   157559 . PMID   12228579 . Retrieved 18 September 2022.
  71. Comas, Louise H.; Eissenstat, David M.; Lakso, Alan N. (2000). "Assessing root death and root system dynamics in a study of grape canopy pruning". New Phytologist . 147 (1): 171–78. doi: 10.1046/j.1469-8137.2000.00679.x .
  72. 1 2 Donahue, Miller & Shickluna 1977, p. 94.
  73. Schlesinger, William H.; Jasechko, Scott (2014). "Transpiration in the global water cycle". Agricultural and Forest Meteorology . 189/190: 115–17. Bibcode:2014AgFM..189..115S. doi:10.1016/j.agrformet.2014.01.011 . Retrieved 25 September 2022.
  74. Erie, Leonard J.; French, Orrin F.; Harris, Karl (1968). Consumptive use of water by crops in Arizona (PDF). Tucson, Arizona: The University of Arizona . Retrieved 25 September 2022.
  75. Tolk, Judy A.; Howell, Terry A.; Evett, Steve R. (1999). "Effect of mulch, irrigation, and soil type on water use and yield of maize". Soil and Tillage Research. 50 (2): 137–47. doi:10.1016/S0167-1987(99)00011-2 . Retrieved 25 September 2022.
  76. Donahue, Miller & Shickluna 1977, pp. 97–99.

Bibliography