Rain or Shine

1

Purpose and Motivations

1.1. Purpose and Motivations

🧭 Overview

🧠 One-sentence thesis

This book serves as a multi-faceted learning resource for introductory soil physics, focusing on how physical principles explain soil properties, processes (especially water and energy balances), and their responses to environmental or management changes.

📌 Key points (3–5)

  • What soil physics studies: uses physics principles to understand soil properties, soil processes, and how environment or management affects them.
  • Book's specific focus: processes in the soil water balance and surface energy balance, and how soil physical properties influence those processes.
  • Target audience: upper undergraduate or master's students, or motivated learners with basic algebra, foundational soil science, and key physics concepts (mass, energy, force, pressure).
  • Common confusion: lacking background in math, soil science, or physics is not a barrier—extra effort and supplementary resources can help.
  • Practical goal: prepares learners for the soil physics portion of the Soil Science Fundamentals Exam (Soil Science Society of America).

📚 What soil physics covers

🔬 The discipline's scope

Soil physics: a scientific discipline focused on using the principles and techniques of physics to understand soil properties, processes in and on soils, and how those properties and processes were, are, or would be affected by changes in environment or management.

  • It is not just describing soil; it is applying physics to explain why and how soil behaves.
  • The excerpt lists three core areas:
    • Properties of the soil
    • Processes which occur in and on soils
    • Effects of environmental or management changes on properties and processes

🌊 Book's particular emphasis

  • The book focuses especially on:
    • Soil water balance processes: how water moves into, through, and out of soil.
    • Surface energy balance processes: how energy (heat, radiation) is exchanged at the soil surface.
  • Both are influenced by soil physical properties (e.g., texture, structure, patterns across spatial scales).

📖 Content and structure

📖 What the book covers

  • Begins with soil patterns, structure, and texture across multiple spatial scales.
  • Then walks through soil water balance and surface energy balance processes step-by-step.
  • Addresses most topics in the Soil Science Fundamentals Exam Performance Objectives (defined by the Soil Science Society of America's Council of Soil Science Examiners).

🎯 Practical outcome

  • Someone who learns this content should be well-prepared for the soil physics portion of the professional examination.
  • Example: a student studying for certification can use this book as a primary resource.

👥 Who this book is for

👥 Primary audience

  • University students: upper undergraduate or master's level, enrolled in an introductory soil physics course.
  • Motivated learners anywhere: curious about basic concepts and applications of soil physics.

🧰 Prerequisites

The book is designed for students who:

  • Are comfortable reading and listening in English.
  • Have basic competence in college-level algebra.
  • Are familiar with foundational principles of soil science.
  • Understand key physics concepts like mass, energy, force, and pressure.

🚧 If you lack prerequisites

  • Don't worry—you can still learn effectively, but may need extra effort in some sections.
  • The excerpt provides supplementary resources:
    • SSSA Glossary of Soil Science Terms
    • Daniels and Haering's "Concepts of Basic Soil Science"
    • Keeney-Kennicutt's "Math Skills Review"
    • Khan Academy video series on Physics

🌍 Practical considerations

🌍 SI units as a learning obstacle

  • The book uses SI (International System of Units).
  • For students in the United States (one of three countries that has not adopted SI), this may be a learning obstacle.
  • The book includes an SI units diagram showing:
    • Seven SI base units
    • Associated derived units: newtons (N), pascals (Pa), joules (J), watts (W), degree Celsius (°C), and siemens (S)
  • Don't confuse: SI units are the standard in science; familiarity with them is essential for soil physics.
2

Content

1.2. Content

🧭 Overview

🧠 One-sentence thesis

This book walks through soil physics step-by-step—from soil structure and texture to water balance and energy balance—preparing learners for both academic study and professional examination.

📌 Key points (3–5)

  • Scope: covers soil patterns, structure, texture, soil water balance, and surface energy balance processes.
  • Alignment with professional standards: addresses most topics in the Soil Science Fundamentals Exam Performance Objectives (Soil Science Society of America).
  • Step-by-step approach: begins with multi-scale spatial properties, then moves through water and energy balance processes sequentially.
  • Preparation outcome: someone who learns this content should be well-prepared for the soil physics portion of the professional examination.

📚 What the book covers

📚 Starting point: soil properties across scales

  • The book begins with soil patterns, structure, and texture examined at multiple spatial scales.
  • This foundation sets up understanding of how physical properties influence processes.

💧 Core processes: water and energy

  • After establishing soil properties, the book walks through:
    • Soil water balance processes (step-by-step)
    • Surface energy balance processes (step-by-step)
  • The excerpt emphasizes that these processes are influenced by soil physical properties covered earlier.

🎯 Alignment with professional standards

🎯 Exam preparation focus

  • The book addresses most topics in the Soil Science Fundamentals Exam Performance Objectives defined by the Soil Science Society of America's Council of Soil Science Examiners.
  • A learner who masters this content should be well-prepared for the soil physics portion of that professional examination.
  • Don't confuse: this is not just an academic textbook—it is explicitly designed to meet professional certification standards.

📋 What "most topics" means

  • The excerpt states the book covers "most" (not all) of the exam objectives.
  • The focus is specifically on the soil physics portion, not the entire exam.
3

1.3. Audience

1.3. Audience

🧭 Overview

🧠 One-sentence thesis

This book targets upper-level undergraduates and master's students in soil physics who have foundational skills in English, algebra, soil science, and basic physics concepts.

📌 Key points (3–5)

  • Primary audience: university students in introductory soil physics courses (upper undergraduate or master's level) and motivated self-learners worldwide.
  • Four prerequisite areas: comfort with English, college-level algebra, foundational soil science principles, and key physics concepts (mass, energy, force, pressure).
  • Common confusion: lacking one prerequisite doesn't disqualify you—you can still learn effectively but may need extra effort in some sections.
  • Learning obstacles: SI units (especially for US students), unit conversions, and Greek letters used for variables.
  • Support resources: the book points to glossaries, foundational reviews, math skills resources, and reference tables to help overcome gaps.

👥 Who this book is for

👥 Primary readers

  • University students: those enrolled in an introductory soil physics course at the upper undergraduate or master's level.
  • Self-learners: motivated learners anywhere in the world who are curious about basic concepts and applications of soil physics.

📋 Expected background

The book assumes four areas of competence:

AreaWhat is expected
LanguageComfortable reading and listening in English
MathBasic competence in college-level algebra
Soil scienceFamiliar with foundational principles of soil science
PhysicsUnderstand key concepts like mass, energy, force, and pressure

🤔 What if you lack a prerequisite?

  • Don't worry—you can still use the book to learn effectively.
  • You may need to invest extra effort to fully understand some sections.
  • The book provides supplementary resources (glossaries, reviews, video series) to help fill gaps.

🚧 Learning obstacles

🌍 SI units challenge

  • Who faces this: students in the United States, one of three countries that has not adopted the International System of Units.
  • What the book uses: SI units throughout.
  • Help provided: an SI units diagram (Fig. 1‑1) showing the seven base units and derived units (newtons, pascals, joules, watts, degree Celsius, siemens).

📏 SI prefixes

To use SI units effectively requires knowing and understanding the SI unit prefixes, which modify the magnitude of the units.

  • The book includes a table (Table 1‑1) of common SI prefixes from peta (10¹⁵) down to femto (10⁻¹⁵).
  • Being familiar with these prefixes helps you correctly understand the size or intensity of physical variables.
  • Example: knowing that "milli" means 10⁻³ helps you interpret millimeters or milligrams.

🔄 Unit conversion

  • Challenge: many students (especially in the US) need to convert from customary US units to SI.
  • Tools available: online unit conversion tools; you can often type "convert 3.7 inches to centimeters" into a web browser.
  • Helpful equivalencies to memorize:
    • 1 inch = 2.54 centimeters (exactly)
    • 1 pound = 0.454 kilograms
    • 1 acre = 0.405 hectares (1 hectare = 10,000 m² or 100 m × 100 m)
  • The book points to a video resource for help with unit conversions.

🔤 Greek letters

  • Why Greek letters are used: the 26 English letters may not be enough to uniquely represent all constants and variables; also reflects the influence of early Greek scholars on Western thought.
  • What the book provides: Table 1‑2 lists the entire Greek alphabet (uppercase, lowercase, and name) for easy reference.
  • Goal: learning these letters makes equations throughout the book easier to understand.
  • Don't confuse: "It's Greek to me" is an English saying meaning "I don't understand"—but here, Greek letters are simply symbols for variables, not a barrier if you learn them.

📚 Support resources

📚 Foundational help

If you lack comfort with math, soil science, or physics, the book recommends:

ResourceWhat it covers
SSSA Glossary of Soil Science TermsSoil science terminology
Daniels and Haering's "Concepts of Basic Soil Science"Foundational soil science principles
"Math Skills Review" by Keeney-KennicuttCollege-level algebra review
Khan Academy video series on PhysicsKey physics concepts

📚 In-book aids

  • SI units diagram (Fig. 1‑1): visual overview of base and derived units.
  • SI prefixes table (Table 1‑1): common prefixes with factors and powers.
  • Greek alphabet table (Table 1‑2): all Greek letters for reference.
  • Unit conversion tips: basic equivalencies and links to video help.
4

Learning Obstacles and Helps

1.4. Learning Obstacles and Helps

🧭 Overview

🧠 One-sentence thesis

Mastering SI unit prefixes, unit conversions, and Greek letter notation removes common barriers to understanding physical variables and equations in scientific study.

📌 Key points (3–5)

  • SI prefixes modify magnitude: prefixes like kilo, milli, and micro scale base units up or down by powers of ten.
  • Unit conversion is a practical skill: knowing basic equivalencies (e.g., 1 inch = 2.54 cm) and using conversion tools helps translate between measurement systems.
  • Greek letters represent variables: science uses Greek alphabet symbols because 26 English letters are insufficient and because of historical Greek influence on Western thought.
  • Common confusion: distinguishing prefix magnitudes—mega (10⁶) vs giga (10⁹) vs tera (10¹²)—requires memorizing the power-of-ten scale.

📏 SI Unit Prefixes

📏 What prefixes do

SI unit prefixes modify the magnitude of base units.

  • They scale units up or down by factors of ten.
  • Understanding prefixes is essential to correctly interpret the size or intensity of physical variables.
  • Example: "kilo" means 1000 times the base unit, so 1 kilometer = 1000 meters.

📊 Common prefix scale

PrefixSymbolFactorPower
petaP1,000,000,000,000,00010¹⁵
teraT1,000,000,000,00010¹²
gigaG1,000,000,00010⁹
megaM1,000,00010⁶
kilok1,00010³
(base)(none)110⁰
millim0.00110⁻³
microμ0.00000110⁻⁶
nanon0.00000000110⁻⁹
picop0.00000000000110⁻¹²

⚠️ Don't confuse prefix order

  • Larger prefixes go upward: kilo < mega < giga < tera.
  • Smaller prefixes go downward: milli > micro > nano > pico.
  • Each step typically represents a factor of 1000 (10³).

🔄 Unit Conversion

🔄 Why conversion matters

  • Students often need to convert between measurement systems (e.g., US customary to SI).
  • Conversion is a "related challenge" that complements understanding prefixes.

🧰 Conversion tools and methods

  • Online conversion tools are widely available.
  • Web search engines can perform conversions directly (e.g., typing "convert 3.7 inches to centimeters").
  • Knowing basic equivalencies by memory is still helpful.

📐 Key equivalencies to remember

The excerpt recommends memorizing:

  • 1 inch = 2.54 centimeters (exactly)
  • 1 pound = 0.454 kilograms
  • 1 acre = 0.405 hectares (1 hectare = 10,000 m² or 100 m × 100 m)

Example: To convert inches to centimeters, multiply by 2.54; to convert acres to hectares, multiply by 0.405.

🔤 Greek Letters in Science

🔤 Why Greek letters are used

  • The 26 English letters may not uniquely represent all constants and variables in a field.
  • Early Greek scholars strongly influenced Western scientific thought.
  • Greek letters provide additional symbols for equations.

📖 The Greek alphabet

The excerpt provides a reference table with uppercase, lowercase, and names:

  • Common examples: α (alpha), β (beta), γ (gamma), δ (delta), π (pi), σ (sigma), ω (omega).
  • Learning these letters makes equations "easier to understand."

💡 Overcoming the "It's Greek to me" barrier

  • The phrase "It's Greek to me" means not understanding a topic.
  • The book does not use Greek language, but does use Greek letters as symbols.
  • Familiarity with the alphabet removes this obstacle to reading scientific notation.
5

2.1. Global Scale

2.1. Global Scale

🧭 Overview

🧠 One-sentence thesis

At the global scale, soil types follow clear spatial patterns strongly influenced by climate variables like temperature and precipitation, and these patterns both reflect and influence Earth's climate and physical processes across thousands of kilometers.

📌 Key points (3–5)

  • What the global scale shows: clear spatial patterns in soil order distribution spanning thousands of kilometers, organized by the USDA soil taxonomy system.
  • What drives the patterns: global climate variables—temperature and precipitation—strongly influence where different soil orders occur.
  • Two-way relationship: soil physical properties are not only influenced by climate but also influence climate itself.
  • Why it matters: these large-scale soil patterns affect countless processes in Earth's coupled human and natural systems, including climate feedback loops.
  • Common confusion: soil patterns don't just passively reflect climate—they actively shape climate through processes like carbon storage and greenhouse gas emissions.

🌍 Soil order patterns and climate drivers

🗺️ What soil orders are

Soil orders: the highest level of the USDA soil taxonomy system, representing major soil types distributed globally.

  • The USDA soil taxonomy is one of several widely-used soil classification systems worldwide.
  • At the global scale, these orders show clear spatial patterns across continents.
  • The patterns span thousands of kilometers.

🌡️ How climate shapes soil distribution

The excerpt identifies three clear climate-driven patterns:

RegionClimate characteristicDominant soil orderKey feature
Above 60°N latitudeFrigidGelisolsSubsurface layer remains frozen year-round
Equatorial South America & AfricaHot and humidOxisolsMost highly weathered soils on Earth
Southern United StatesEast-to-west decreasing precipitationUltisols → Mollisols → AridisolsProgression reflects moisture gradient
  • Temperature example: frigid regions = frozen subsurface layers; hot regions = extreme weathering.
  • Precipitation example: the southern U.S. shows a clear east-to-west gradient—wetter southeast has Ultisols, drier southwest has Aridisols.
  • Don't confuse: these are not random distributions; climate variables create predictable, mappable patterns.

🔄 Two-way soil-climate relationship

🔄 Soil patterns indicate physical properties

  • The soil order patterns spanning thousands of kilometers also indicate large-scale patterns in soil physical properties.
  • These physical properties influence countless processes in Earth's coupled human and natural systems.
  • The patterns are not just descriptive—they have functional significance.

🌡️ How soils influence climate back

Key mechanism: soil physical properties don't just respond to climate; they actively shape it.

  • The excerpt emphasizes: "these global distributions of soil physical properties are not only influenced by climate, but they also influence climate."
  • This creates a feedback loop rather than a one-way cause-and-effect.

❄️ Permafrost carbon feedback example

❄️ The permafrost carbon uncertainty

The excerpt highlights permafrost soils as a significant uncertainty in future climate projections:

  • What's stored: a large amount of organic carbon in the world's permafrost soils.
  • What's happening: as global temperatures rise, these soils are thawing and the organic carbon is decomposing.
  • Result: increased carbon dioxide and methane emissions to the atmosphere.

🔁 The feedback loop mechanism

  1. Rising global temperatures → permafrost thaws
  2. Thawing → organic carbon decomposes
  3. Decomposition → greenhouse gases (CO₂ and methane) released
  4. If not offset by increased plant CO₂ uptake from warmer climate → more greenhouse gases
  5. More greenhouse gases → accelerates climate change → more warming → more thawing

Don't confuse: this is a potential feedback loop, not a certainty—the outcome depends on whether increased plant growth can offset the emissions.

🔬 Why soil physical properties matter here

  • Gas emissions and plant growth in thawing soils both depend strongly on soil temperature and moisture dynamics.
  • These temperature and moisture dynamics are themselves influenced by soil physical properties.
  • This creates a growing need to better understand these processes in thawing soils.
  • Example: soil physical properties determine how quickly permafrost thaws, how much water is available for plants, and how easily gases can escape—all critical to predicting the climate feedback.
6

Basin or Watershed Scale

2.2. Basin or Watershed Scale

🧭 Overview

🧠 One-sentence thesis

At the basin or watershed scale (tens to hundreds of kilometers), large spatial patterns in soil properties—often created by wind and water movement—significantly impact human communities and ecosystems, even though we may fail to notice them in daily life.

📌 Key points (3–5)

  • Scale definition: basin or watershed scale typically spans 10s to 100s of kilometers.
  • Why patterns go unnoticed: we travel only a few kilometers daily or don't take time to observe soil changes, so large patterns are easy to miss.
  • How patterns form: large-scale movement of wind and water generates distinctive spatial patterns (e.g., windblown sands deposited along rivers).
  • Real-world impacts: these patterns explain social and ecological characteristics, from rural economies to native plant distributions.
  • Common confusion: "large" does not mean "obvious"—basin-scale patterns can be so extensive that they escape everyday perception, yet they have significant consequences.

🌊 Scale and visibility

📏 What basin or watershed scale means

Basin or watershed scale: typically 10s to 100s of kilometers in extent.

  • This is the spatial scale at which river basins and watersheds organize the landscape.
  • Soil property patterns at this scale span distances much larger than a single farm or neighborhood.

👁️ Why we fail to notice these patterns

  • Daily travel is limited: most people travel only a few kilometers from home in a typical day.
  • Lack of attention: even when traveling farther, we may not take time to notice changes in soil around us.
  • Paradox of scale: the patterns are so large that they become invisible in everyday experience.
  • Don't confuse: "large spatial pattern" does not mean "easily visible"—the excerpt emphasizes that size can actually hide patterns from casual observation.

🏜️ How basin-scale patterns form

💨 Wind and water movement

  • Large-scale movement of wind and water generates distinctive spatial patterns in soil properties.
  • The excerpt provides a concrete mechanism: windblown sands deposited along rivers.

🗺️ Example: Oklahoma surface soils

The excerpt describes a map of sand content across Oklahoma, USA:

  • Pattern observed: dark brown bands of high sand content trending northwest to southeast in the northwestern part of the state.
  • Location: these sandy areas follow along the north side of rivers (e.g., Cimarron River, North Canadian River).
  • Formation mechanism: windblown sands were likely deposited along the north sides of rivers by prevailing south winds during past severe drought events that stripped the landscape bare.
  • Associated features: these sandy areas overlie alluvial aquifers that provide groundwater for irrigation and municipal use.

Example: During a severe drought, vegetation dies off, exposing bare soil; prevailing winds from the south pick up sand and deposit it along the north banks of rivers, creating a persistent pattern visible across the entire state.

🌾 Why basin-scale patterns matter

🏘️ Impacts on human communities

  • Soil patterns at this scale have significant impacts on human communities.
  • The excerpt links soil patterns to economies of rural communities: sandy soils overlying aquifers support irrigation and municipal water supply, shaping local economic activities.

🌱 Impacts on ecosystems

  • Basin-scale soil patterns help explain the distributions of native plant species on the landscape.
  • Different soil textures (e.g., high sand content vs. other soils) create habitat variation that influences which plants can thrive where.

📊 Social and ecological linkages

AspectHow soil patterns explain it
Rural economiesSandy areas overlie aquifers → groundwater for irrigation and municipal use → economic base for communities
Native plant distributionsSoil texture and moisture patterns → habitat variation → species sorting across the landscape
  • The excerpt emphasizes that these patterns "help explain a variety of social and ecological characteristics," showing that soil is not just a physical substrate but a driver of coupled human and natural systems.
7

2.3. Hillslope Scale

2.3. Hillslope Scale

🧭 Overview

🧠 One-sentence thesis

Erosion and deposition processes, along with topographic position and drainage differences, create organized soil patterns at the hillslope scale (tens to hundreds of meters) that significantly affect land use and management decisions.

📌 Key points (3–5)

  • Scale definition: hillslope patterns span roughly 10s to 100s of meters, one level below the watershed scale.
  • Main processes: erosion and deposition create spatial patterns such as alluvial fans and catenas (toposequences).
  • Catena concept: a series of related but distinct soil types arranged along a hillslope, formed from the same parent material but differing in drainage, slope, and aspect.
  • Common confusion: soils in a catena look different (color, horizon thickness) even though they share the same parent material—the key distinction is topographic position and resulting drainage.
  • Practical impact: hillslope soil patterns drive management practices like subsurface drainage systems, which increase crop production but also reduce wetland habitat and risk water contamination.

🏔️ Erosion and deposition patterns

🏔️ Alluvial fans

Alluvial fans: fan-shaped deposits of water-transported sediment at the base of hills or mountains.

  • These are classic examples of hillslope-scale spatial organization created by erosion and deposition.
  • The excerpt describes an aerial view of a well-developed alluvial fan in the French Pyrenees mountains.
  • Visible features include trails perpendicular to the slope created by grazing livestock, indicating focused zones of soil compaction.
  • Example: sediment washes down from a mountain and spreads out in a fan shape at the base, creating distinct soil patterns.

🔗 Catena (toposequence) concept

🔗 What a catena is

Catena (or toposequence): a series, or chain, of distinct but related soil types arranged along a hillslope.

  • Soils in a catena typically form from the same parent material.
  • They have different physical properties arising from differences in:
    • Slope
    • Aspect
    • Drainage
  • The excerpt emphasizes that these soils are "distinct but related"—they share a common origin but diverge due to position on the hillslope.

🌾 Clarion–Nicollet–Webster catena example

The excerpt describes a common catena found in Iowa and Minnesota farmland:

Soil seriesTopographic positionDrainage classSoil color (subsurface)
ClarionSummitModerately well drainedRed/brown (oxidizing conditions)
NicolletShoulder slopesSomewhat poorly drainedIntermediate
WebsterBackslopesPoorly drainedGray (reducing conditions)
GlencoeClosed depressionsVery poorly drained(More limited extent)
  • All formed in unsorted glacial till (parent material deposited by prior glaciers).
  • Don't confuse: same parent material does not mean identical soils—topographic position and drainage create the differences.
  • Differences in drainage impact soil color and thickness of the A horizon (visible in profile photographs).

🎨 Why soil color varies in a catena

  • Summit soils (Clarion): red/brown subsurface colors indicate oxidizing conditions (better drainage, more oxygen).
  • Footslope soils (Webster): gray subsurface colors indicate reducing conditions (poor drainage, less oxygen, waterlogged).
  • Color is a visual indicator of drainage and oxygen availability.

🚜 Management implications

🚜 Subsurface drainage systems

  • Many farmers in the Iowa/Minnesota region install elaborate subsurface drainage systems to remove water from poorly drained portions of fields.
  • Why: improves crop growth in areas occupied by poorly drained soils (e.g., Webster, Glencoe, Parnell, Flom series).
  • The excerpt describes a system draining approximately 129 hectares (320 acres) in western Minnesota.
  • Drainage pipes are placed in low areas and depressions where poorly drained soils (classified as aquolls in US soil taxonomy) are located.

⚖️ Trade-offs of artificial drainage

BenefitCost
Greatly increases crop production potentialDecreases wetland habitat
Allows farming on poorly drained soilsIncreases risk of contaminating receiving water bodies
  • The excerpt notes that improved drainage water management practices exist to mitigate these risks.
  • Example: a farmer installs drainage pipes in a low-lying Webster soil area to prevent waterlogging, enabling corn production, but the drainage water may carry nutrients or contaminants downstream.
8

Soil Profile Scale

2.4. Soil Profile Scale

🧭 Overview

🧠 One-sentence thesis

Soil profiles display vertical organization through distinct horizons spanning tens to hundreds of centimeters, and the specific combination and sequence of these layers differentiate soil types and influence the movement of water, energy, and organisms.

📌 Key points (3–5)

  • What defines this scale: vertical spatial patterns spanning 10s to 100s of cm, organized into layers called horizons.
  • What horizons are: layers that differ in physical and/or chemical properties; their specific combination and sequence define different soil types.
  • How horizons interact with the environment: they influence and are influenced by water, energy, and living organisms passing through the profile.
  • Visual contrasts: adjacent horizons can show dramatic differences in color, texture, and composition due to processes like organic matter accumulation or leaching.
  • Common confusion: horizons are not just color bands—they represent functional differences in properties that affect soil behavior.

🏗️ What soil horizons are and why they matter

🏗️ Definition and basic concept

Horizons: layers within a soil profile that differ from each other in physical and/or chemical properties.

  • Together, all horizons make up the complete soil profile.
  • The specific combination and sequence of horizons differentiate one soil type from another.
  • Adjacent horizons can have very different characteristics even though they are right next to each other vertically.

🔄 How horizons interact with their environment

  • Horizons both influence and are influenced by the presence and passage of:
    • Water
    • Energy
    • Living organisms
  • This is a two-way relationship: the horizons shape what moves through them, and those movements shape the horizons over time.

🎨 Example: The Tetonka soil series

🎨 Dramatic visual contrasts between horizons

The excerpt describes a grassland soil profile in South Dakota (Tetonka series) that shows striking differentiation:

HorizonDepthAppearanceWhy it looks that way
A horizonUppermost, >30 cm thickDark gray to blackHigh soil organic matter content
E horizonJust below ALight gray to white (strong contrast)Downward leaching (eluviation) of clays, iron oxides, and aluminum oxides
Bt horizonBelow EAccumulation zoneClays leached from above accumulate here; shows prismatic structure

🔍 What these contrasts reveal

  • Eluviation: the process of materials being leached or washed downward from upper horizons.
  • The E horizon loses material (becomes lighter), while the Bt horizon gains material (accumulates clays).
  • Even within a single horizon (the Bt), there can be additional levels of spatial organization, such as prismatic soil structure.
  • Example: If you dug a pit and saw a very dark layer on top, a white layer in the middle, and a clay-rich layer below, you would be seeing the result of long-term processes moving materials vertically through the profile.

⚠️ Don't confuse

  • Horizon color is not arbitrary—it reflects real differences in composition and processes.
  • A light-colored horizon is not "empty"; it has been altered by leaching, which is itself an important process.
  • The sequence matters: the same horizons in a different order would represent a different soil type with different behavior.
9

Learning Obstacles and Helps

2.5. Soil Aggregate Scale

🧭 Overview

🧠 One-sentence thesis

Mastering SI unit prefixes, unit conversions, and Greek letters used in scientific notation removes common barriers to understanding physical variables and equations in soil science.

📌 Key points (3–5)

  • SI prefixes modify unit magnitude: from femto (10^-15) to peta (10^15), these prefixes scale base units up or down.
  • Unit conversion is a practical skill: knowing basic equivalencies (e.g., 1 inch = 2.54 cm exactly) and using conversion tools helps translate between measurement systems.
  • Greek letters represent constants and variables: the 26 English letters are insufficient for all scientific symbols, so Greek letters fill the gap.
  • Common confusion: students often struggle converting between US customary units and SI units—memorizing a few key equivalencies helps.

📏 SI Unit Prefixes

📏 What SI prefixes do

SI unit prefixes: modifiers that change the magnitude of base units by powers of ten.

  • The excerpt emphasizes that effective use of SI units requires knowing these prefixes.
  • They scale units from extremely small (femto = 10^-15) to extremely large (peta = 10^15).
  • Example: "kilo" means multiply by 1000 (10^3), so 1 kilometer = 1000 meters.

📊 Common prefix ranges

PrefixSymbolFactorPower
petaP1,000,000,000,000,00010^15
teraT1,000,000,000,00010^12
gigaG1,000,000,00010^9
megaM1,000,00010^6
kilok1,00010^3
(none)(none)110^0
centic0.0110^-2
millim0.00110^-3
microμ0.00000110^-6
nanon0.00000000110^-9
picop0.00000000000110^-12
femtof0.00000000000000110^-15

🎯 Why familiarity matters

  • Understanding prefix magnitude helps you correctly grasp the size or intensity of physical variables throughout the book.
  • Without knowing prefixes, you cannot tell whether a measurement is tiny or enormous.

🔄 Unit Conversion Challenges

🔄 The conversion problem

  • The excerpt identifies unit conversion as a related challenge for many students.
  • US students especially need to convert from customary US units to SI units.
  • Don't confuse: conversion is not just about calculation—it requires knowing which equivalencies to apply.

🧮 Practical conversion strategies

  • Online tools: many conversion tools are available; you can type phrases like "convert 3.7 inches to centimeters" into a web browser.
  • Memory aids: the excerpt recommends knowing some basic equivalencies by heart:
    • 1 inch = 2.54 centimeters (exactly)
    • 1 pound = 0.454 kilograms
    • 1 acre = 0.405 hectares (1 hectare = 10,000 m² or 100 m × 100 m)
  • Example: if you know 1 inch = 2.54 cm, you can quickly estimate that 4 inches ≈ 10 cm without a calculator.

📹 Additional help

  • The excerpt mentions a video resource for help with unit conversions (link provided in original text).

🔤 Greek Letters in Science

🔤 Why Greek letters appear

  • The excerpt notes a common English saying: "It's Greek to me" (meaning complete lack of understanding).
  • This book does not use Greek language, but does use Greek letters to represent constants and variables.
  • Two reasons for using Greek letters:
    1. The 26 English letters may not be enough to uniquely represent all different constants and variables.
    2. Early Greek scholars strongly influenced Western scientific thought.

📖 The Greek alphabet

  • The excerpt provides a complete Greek alphabet table (uppercase, lowercase, and name for each letter).
  • Selected examples from the table:
    • α (alpha), β (beta), γ (gamma), δ (delta)
    • θ (theta), λ (lambda), μ (mu), π (pi)
    • σ (sigma), τ (tau), φ (phi), ω (omega)

🎓 Learning benefit

  • Learning these Greek letters makes equations easier to understand throughout the book.
  • Don't confuse: Greek letters are symbols, not words—they stand for specific quantities in formulas.
  • Example: if an equation uses μ (mu) for a coefficient, recognizing the letter helps you follow the logic without confusion.
10

Scale of Primary Soil Particles

2.6. Scale of Primary Soil Particles

🧭 Overview

🧠 One-sentence thesis

The excerpt does not contain substantive content about the scale of primary soil particles; it presents only auxiliary material on SI units, Greek letters, problem sets, and an introduction to soil water content.

📌 Key points (3–5)

  • The excerpt includes reference tables for SI prefixes and the Greek alphabet to help students understand scientific notation and symbols.
  • Unit conversion is identified as a common challenge, with basic equivalencies provided (e.g., 1 inch = 2.54 cm).
  • Greek letters are used in science because the English alphabet may not provide enough unique symbols for all variables and constants.
  • The excerpt transitions to a new chapter on soil water content but does not discuss primary soil particle scales.

📏 Supporting reference material

📊 SI unit prefixes

The excerpt provides a table of common SI prefixes to help understand magnitude:

PrefixSymbolFactorPower
petaP1,000,000,000,000,00010^15
teraT1,000,000,000,00010^12
gigaG1,000,000,00010^9
megaM1,000,00010^6
kilok1,00010^3
millim0.00110^-3
microμ0.00000110^-6
nanon0.00000000110^-9
  • Understanding these prefixes helps correctly interpret the size or intensity of physical variables.
  • The table spans from very large (peta) to very small (femto) scales.

🔤 Greek alphabet reference

The excerpt explains why Greek letters appear in scientific texts:

  • The 26 English letters may not uniquely represent all constants and variables in a field of study.
  • Early Greek scholars strongly influenced Western scientific thought.
  • A complete Greek alphabet table is provided (alpha through omega, uppercase and lowercase).

🔄 Unit conversion guidance

🔧 Common conversion challenges

The excerpt identifies unit conversion as a learning obstacle, especially for US students converting to SI units.

Basic equivalencies to remember:

  • 1 inch = 2.54 centimeters (exactly)
  • 1 pound = 0.454 kilograms
  • 1 acre = 0.405 hectares (1 hectare = 10,000 square meters or 100 m × 100 m)

💻 Modern conversion tools

  • Online unit conversion tools are now widely available.
  • Web browsers can perform conversions directly (e.g., typing "convert 3.7 inches to centimeters" in the search bar).
  • Despite tool availability, knowing basic equivalencies by memory remains helpful.

🌊 Chapter transition note

🚀 Soil water content introduction

The excerpt transitions to a new chapter focusing on soil water rather than particle scale:

  • Soil liquid phase is not pure water but a complex solution containing organic and inorganic solutes, microorganisms, and colloids.
  • NASA's Phoenix Mars Lander included a special sensor to detect water in Martian soil, developed in partnership with soil physicists.
  • Water is described as "the life-blood of the soil"—almost all terrestrial life and soil processes are influenced by soil water content.

Note: This section does not address the scale of primary soil particles as indicated by the title.

11

Learning Obstacles and Helps

2.7. Terminology

🧭 Overview

🧠 One-sentence thesis

Understanding SI unit prefixes, unit conversions, and Greek letters used in scientific notation removes major barriers to comprehending physical variables and equations in soil science.

📌 Key points (3–5)

  • SI prefixes modify magnitude: prefixes like kilo, milli, and micro scale base units up or down by powers of 10.
  • Unit conversion is essential: students must convert between measurement systems (e.g., US customary to SI) using known equivalencies.
  • Greek letters represent variables: science uses Greek alphabet symbols because 26 English letters are insufficient for all constants and variables.
  • Common confusion: prefix magnitude—memorizing the power-of-10 scale (e.g., kilo = 10³, milli = 10⁻³) prevents errors in calculations.

📏 SI unit prefixes

📏 What prefixes do

SI unit prefixes modify the magnitude of base units.

  • Prefixes scale a unit up or down by multiplying by a power of 10.
  • They allow expressing very large or very small quantities without writing many zeros.
  • Example: 1 kilometer = 1000 meters (kilo = 10³); 1 millimeter = 0.001 meters (milli = 10⁻³).

🔢 Common prefix scale

The excerpt provides a table of prefixes from peta (10¹⁵) down to femto (10⁻¹⁵). Key prefixes include:

PrefixSymbolFactorPower
petaP1,000,000,000,000,00010¹⁵
teraT1,000,000,000,00010¹²
gigaG1,000,000,00010⁹
megaM1,000,00010⁶
kilok1,00010³
(none)(none)110⁰
millim0.00110⁻³
microμ0.00000110⁻⁶
nanon0.00000000110⁻⁹
picop0.00000000000110⁻¹²
femtof0.00000000000000110⁻¹⁵

🎯 Why familiarity matters

  • The excerpt states that being familiar with these prefixes helps you "correctly understand the size or intensity of the physical variables" studied in the book.
  • Without knowing prefixes, you cannot interpret measurements correctly.
  • Example: confusing milligrams (10⁻³ g) with micrograms (10⁻⁶ g) leads to thousand-fold errors.

🔄 Unit conversion

🔄 The conversion challenge

  • The excerpt identifies unit conversion as "a related challenge for many students."
  • Students in the US often need to convert from US customary units to SI units.
  • Conversion requires knowing equivalencies between systems.

🧮 Basic equivalencies to memorize

The excerpt recommends knowing these by memory:

  • 1 inch = 2.54 centimeters (exactly)
  • 1 pound = 0.454 kilograms
  • 1 acre = 0.405 hectares (1 hectare = 10,000 m² or 100 m × 100 m)

🛠️ Conversion tools and methods

  • Many online unit conversion tools are available.
  • You can type phrases like "convert 3.7 inches to centimeters" into a web browser search bar.
  • The excerpt mentions a video resource for help with unit conversions.
  • Don't confuse: having tools available does not eliminate the need to understand basic equivalencies and conversion logic.

🔤 Greek letters in science

🔤 Why Greek letters are used

The excerpt explains two reasons:

  1. Not enough English letters: The 26 letters of the English alphabet may not uniquely represent all constants and variables in a particular area of study.
  2. Historical influence: Early Greek scholars strongly influenced the development of Western thought.

📖 The Greek alphabet

The excerpt provides a complete table of uppercase and lowercase Greek letters with their names:

UppercaseLowercaseName
Aαalpha
Bβbeta
Γγgamma
Δδdelta
Eεepsilon
Θθtheta
Λλlambda
Ππpi
Σσsigma
Φφphi
Ωωomega

(The table includes all 24 letters; only a sample is shown here.)

🎓 Learning Greek letters helps comprehension

  • The excerpt states that "learning these Greek letters should make the equations presented throughout this book easier to understand."
  • Greek letters represent important constants and variables in the book's equations.
  • Example: if you see θ in an equation and don't recognize it as "theta," you cannot follow the mathematical reasoning.

💬 "It's Greek to me"

  • The excerpt references the English saying "It's Greek to me," meaning "we do not understand anything about a given topic."
  • The book does not use Greek language, but it does use Greek letters.
  • Don't confuse: Greek letters are symbols, not words; you only need to recognize their shapes and names, not learn the Greek language.
12

Soil Water Content

3.1. Soil Water Content

🧭 Overview

🧠 One-sentence thesis

Soil water content—the amount of water in soil—is critical to nearly all terrestrial life and soil processes, and modern measurement techniques range from simple oven-drying to satellite-based remote sensing.

📌 Key points (3–5)

  • Why water content matters: water is the "life-blood" of soil; almost all biological, chemical, and physical soil processes depend on it.
  • Two ways to express content: gravimetric (mass of water per mass of dry soil) vs volumetric (volume of water per volume of soil).
  • Measurement methods span a huge range: from the simplest gravimetric oven-drying to cutting-edge satellite, cosmic-ray neutron, and fiber-optic technologies.
  • Common confusion: gravimetric vs volumetric water content—they measure different ratios (mass/mass vs volume/volume) and can be converted using bulk density.
  • Growing data availability: improved sensors and remote sensing mean soil water content data are increasingly available worldwide for applications like drought monitoring and streamflow forecasting.

💧 What soil water content means

💧 The soil liquid phase is more than pure water

  • The excerpt emphasizes that "soil liquid phase is not simply water, but rather a complex solution containing organic and inorganic solutes," plus micro-organisms and colloids.
  • Analogy: soil water is like a river transporting life-giving (and sometimes life-threatening) substances.
  • This chapter focuses on the amount and condition of soil water, leaving solutes and organisms for later chapters.

🌍 Why water content is so important

  • NASA's Phoenix Mars Lander included a special sensor (built with help from soil physicists) to test for water in Martian soil.
  • Finding water on Mars was key to answering questions about the possibility of life.
  • On Earth: "Water is the life-blood of the soil. Almost all terrestrial life and almost all soil biological, chemical, and physical processes are influenced by the water content of the soil."
  • Example: farmers, horticulturalists, turf managers, and government agencies all need accurate soil water content measurements to manage resources.

📏 How to quantify soil water content

📏 Gravimetric water content (θ_g)

Gravimetric water content: the mass of water per unit mass of oven-dry soil particles.

  • Dimensions: mass over mass.
  • Written as a unitless decimal, percentage, or with units like kg/kg or g/g (this book uses g/g).
  • Range for mineral soils: approximately 0 to 0.60 g/g; organic soils may be higher.

📦 Volumetric water content (θ)

Volumetric water content: the volume of water per unit volume of soil.

  • Dimensions: volume over volume.
  • Written as a unitless decimal, percentage, or with units like m³/m³ or cm³/cm³ (this book uses cm³/cm³).
  • Range for mineral soils: approximately 0 to 0.60 cm³/cm³; organic soils may be higher.
  • Conversion: if you know gravimetric water content, you can calculate volumetric water content using soil bulk density (ρ_b) and the density of water (ρ_w).
    • Formula (Eq. 3-1): θ = θ_g × (ρ_b / ρ_w)

📐 Soil water storage (S)

Soil water storage: the equivalent depth of soil water contained in a soil layer of a specified thickness (Δz).

  • Dimensions: length (cm or mm).
  • Calculation: S = θ × Δz (volumetric water content times layer thickness).
  • This tells you the total depth of water if you "squeezed out" all the water from a soil layer.

🔍 Don't confuse gravimetric and volumetric

  • Gravimetric measures mass ratio; volumetric measures volume ratio.
  • They are related but not the same; conversion requires knowing bulk density.
  • Example: a soil with 0.20 g/g gravimetric content might have a different volumetric content depending on how compacted (dense) the soil is.

🔬 Measurement techniques

🔬 The gravimetric method (simplest and oldest)

  • Procedure:
    1. Weigh a soil sample.
    2. Oven-dry at 105°C to constant weight (or for 24 hours for convenience).
    3. Record final dry weight.
  • Why 105°C? Above water's boiling point but below the temperature where organic matter would be lost.
  • What you can measure:
    • If sample volume is unknown: only gravimetric water content.
    • If sample volume is known: also bulk density and volumetric water content.
  • Requirements: an accurate balance and an oven that can hold constant temperature unattended for 24 hours.

🛰️ Satellite remote sensing

  • Satellites map electromagnetic radiation (microwave wavelength) emitted from the land surface.
  • These data estimate volumetric water content of the 0–5 cm soil layer globally with ~9 km spatial resolution.
  • Researchers also extract soil water information from GPS satellite signals reflected off the land surface.

☢️ Cosmic-ray neutron method

  • Neutrons generated by cosmic rays are detected at the land surface (stationary or mobile counters).
  • Used to estimate volumetric water content of surface soil to depths up to 40 cm with ~400 m spatial resolution.

🌡️ Fiber-optic distributed temperature sensing

  • Systems measure soil temperature at 1-m resolution and 1-min intervals along paths ≥1 km.
  • Soil temperature data can be used to infer volumetric water content.
  • Some systems include active heating capabilities.

📡 Electromagnetic and thermal sensors

  • Long-term, automated monitoring of volumetric water content at a single point.
  • Development began with time domain reflectometry (TDR) by Topp et al.
  • Steady progression of new sensors since around the end of World War II.

🧪 The MOISST field experiment

  • Marena, Oklahoma, In Situ Sensor Testbed (MOISST): a site where many established and emerging soil moisture measurement technologies are inter-compared.
  • Equipment includes cosmic-ray neutron sensors, GPS reflectometry sensors, and numerous below-ground soil water content sensors.

🌐 Why better measurements matter

🌐 Increasing data availability

  • Soil water content data are becoming increasingly available around the world.
  • Development of improved measurement methods has been a major focus in soil physics and related disciplines.

🚀 Societal applications

  • Drought monitoring: better data improve early warning and response.
  • Streamflow forecasting: more accurate predictions of river flow.
  • Wildfire preparedness: understanding soil moisture helps predict fire risk.
  • Other potential applications: the excerpt notes "a host of other potential applications."
  • Opportunity for researchers: find creative ways to use these data for societal benefit.

🔗 Interdisciplinary importance

  • Researchers in soil science, hydrology, ecology, agronomy, meteorology, and civil engineering all need accurate soil water content measurements.
  • Practitioners include farmers, horticulturalists, turf managers, and government agencies managing natural resources.
Measurement methodSpatial scaleDepthKey technology
Gravimetric (oven-drying)Single sampleDepends on sampleBalance + oven
Satellite remote sensing~9 km resolution0–5 cmMicrowave radiation
GPS reflectometryVariesSurfaceReflected GPS signals
Cosmic-ray neutron~400 m resolution0–40 cmNeutron counters
Fiber-optic temperature1-m resolution, ≥1 km pathVariesTemperature sensing
Electromagnetic/thermal sensorsSingle pointVariesTDR and newer sensors
13

Soil Water Potential

3.2. Soil Water Potential

🧭 Overview

🧠 One-sentence thesis

Soil water potential measures the potential energy of soil water relative to a reference state, and differences in this potential drive water movement from higher to lower potential regions.

📌 Key points (3–5)

  • What soil water potential measures: the potential energy of the soil solution relative to pure water at atmospheric pressure, same temperature, and a specified elevation.
  • How water moves: water flows from regions of higher water potential to regions of lower water potential; when potential is equal throughout, the system reaches equilibrium (no flow).
  • Three main components: gravitational potential (position in gravity field), osmotic potential (effect of solutes), and pressure potential (capillarity, adsorption, submergence, air pressure).
  • Common confusion—units: soil water potential can be expressed three ways (energy per mass, per volume, or per weight), each with different units and convenience for different problems.
  • Measurement range: matric potential varies over several orders of magnitude, from 0 kPa (saturated pores) to far below -1500 kPa (dry surface soil).

🔋 What soil water potential is

🔋 Definition and reference state

Soil water potential (ψ): the potential energy of the soil solution relative to that of water in a standard reference state.

The reference state is defined by four conditions:

  • Pure water
  • At atmospheric pressure
  • At the same temperature as the soil solution
  • At a specified, constant elevation

🌊 Potential energy analogy

  • Potential energy is energy a substance has by virtue of its location or internal condition, always defined relative to a reference.
  • Example: A book on a table has less potential energy than the same book held above the table; when released, the book spontaneously falls from higher to lower potential energy due to gravity.
  • The same principle applies to soil water: it moves spontaneously from higher to lower potential energy.

⚖️ Equilibrium concept

Equilibrium: a condition where water potential is equal throughout a region of soil, resulting in no water flow.

  • Equilibrium is important for theory and lab experiments.
  • Rarely, if ever, occurs in the field.

🧩 Components of soil water potential

🧩 The additive model

Most scientists consider total soil water potential to be the simple sum of component potentials from several contributing factors.

The three main components are:

ComponentSymbolWhat it measures
Gravitational potentialψ_gPotential energy due to position in a gravitational field, measured relative to an arbitrary reference elevation
Osmotic potentialψ_oDecrease in water potential due to the presence of solutes
Pressure potentialψ_pWater potential attributable to factors besides gravity and solutes (capillarity, adsorption, submergence, air pressure)

🌍 Gravitational potential

  • Depends on the position of the substance or object in a gravitational field.
  • Measured with respect to an arbitrary reference elevation.

🧂 Osmotic potential

  • Represents the decrease in water potential caused by solutes.
  • Affects soil water flow whenever there is a gradient in solute concentration within the soil.
  • Such gradients are especially persistent where there is a membrane (e.g., plant cell wall) or barrier (e.g., soil surface) that restricts movement of water or solutes.

💧 Pressure potential and matric potential

  • Pressure potential refers to soil water potential from any factors besides gravity and solutes.
  • Factors include capillarity, adsorption, submergence, and air pressure.

Matric potential (ψ_m): an important subcategory of pressure potential; the decrease in water potential due to the interaction of water with the soil's solid matrix, primarily through capillarity and adsorption.

📏 Units and expression

📏 Three ways to express soil water potential

The excerpt identifies three common ways, each with advantages and potential confusion:

ExpressionUnitsNotes
Energy per unit massJ kg⁻¹Most fundamental and theoretically sound, but not as widely used
Energy per unit volumeJ m⁻³ or kPaSame units as pressure; convenient for pressure potential and osmotic potential; some uncertainty arises because water density (and thus mass-to-volume relation) is slightly temperature dependent
Energy per unit weightJ N⁻¹ or m or cmSometimes called hydraulic head; when water flow velocity is negligibly small, indicated by the height of free water surface above or below a reference elevation; convenient for solving soil water flow problems

📏 This book's convention

  • The excerpt states that energy per unit volume (kPa) and energy per unit weight (m or cm) will be primarily used.

📏 Typical range

  • Soil water potentials vary over several orders of magnitude.
  • A soil with pore spaces completely filled with water: matric potential of 0 kPa.
  • A surface soil exposed to dry atmospheric conditions for extended periods: matric potential far below -1500 kPa.

🔬 Measurement techniques

🔬 Why measure soil water potential

  • Researchers use measurements to understand water movement in the soil-plant-atmosphere continuum.
  • Crop and turf managers use measurements to guide irrigation management.

🧪 Tensiometer

Tensiometer: the oldest type of measuring device for soil water potential, consisting of a ceramic cup connected to a water reservoir and a pressure gauge.

How it works:

  • The cup is embedded in the soil.
  • Water flows into or out of the reservoir through the ceramic until the water potential inside the cup equals the water potential outside the cup.
  • The movement of water causes the pressure at the gauge to increase or decrease relative to atmospheric pressure.
  • The gauge pressure, corrected for the height of the water column between the cup and the gauge, indicates the pressure potential or matric potential at the location of the cup.

Limitations:

  • Tensiometers do not indicate osmotic potential because solutes can move freely through the ceramic.
  • Typically limited to measuring pressure potentials greater than approximately -80 kPa; below this, air passes through the ceramic and invalidates the pressure gauge reading.

🌡️ Heat dissipation sensor

A more recent technique that can measure matric potential at potentials far below the tensiometer range.

Structure:

  • Small temperature sensor and heating element encased in the center of a ceramic plug.

How it works:

  • When embedded in soil, water moves into or out of the ceramic until the matric potential inside equilibrates with the matric potential in the surrounding soil.
  • A pulse of electric current through the heater generates a temperature rise, recorded by the temperature sensor.
  • The magnitude of the temperature rise is related to the absolute value of the matric potential:
    • Relatively dry soil and ceramic → large temperature rise → large negative matric potential
    • Relatively wet soil and ceramic → small temperature rise → small negative matric potential

Limitations:

  • Heat dissipation sensors lose sensitivity when the ceramic is saturated.
  • Upper measurement limit of approximately -10 kPa.

💧 Dewpoint potentiometer

A laboratory instrument for measuring soil water potential on soil samples.

How it works:

  • A small soil sample is placed in a sealed chamber containing a small mirror.
  • The water potential of the air above the sample equilibrates with the soil water potential.
  • The instrument lowers the temperature of the mirror until finding the precise temperature at which dew forms on the mirror.
  • Based on this dewpoint temperature, the water potential of the soil sample is determined.

Range and limitations:

  • Can measure water potentials over a large range.
  • Even less sensitive in the wet range than heat dissipation sensors.
  • Current dewpoint potentiometers have an upper limit of approximately -100 kPa.

🔍 Don't confuse: measurement ranges

Each technique has a different effective range:

  • Tensiometer: greater than approximately -80 kPa (wet end)
  • Heat dissipation sensor: approximately -10 kPa to much drier (intermediate to dry)
  • Dewpoint potentiometer: approximately -100 kPa to very dry (dry end)
14

Soil Water Retention

3.3. Soil Water Retention

🧭 Overview

🧠 One-sentence thesis

Soil water retention curves describe how soil water content relates to matric potential, and this relationship is fundamentally shaped by soil texture, pore size distribution, and whether the soil is wetting or drying.

📌 Key points (3–5)

  • Core relationship: soil water content decreases as matric potential becomes more negative; coarse-textured soils retain less water than fine-textured soils at the same matric potential.
  • Key features of the curve: saturated water content at 0 matric potential, air-entry potential where air first enters pores, an inflection point with practical significance, and a flat tail at very negative potentials.
  • Hysteresis phenomenon: the equilibrium water content at a given matric potential differs between wetting and drying, caused by air entrapment, contact angle hysteresis, and the "ink bottle" effect.
  • Common confusion: wetting vs. drying curves—wetting curves show lower water content than drying curves at the same matric potential; initial conditions also matter.
  • Measurement and modeling: retention curves are determined empirically in the lab using devices like hanging water columns and pressure plates; mathematical functions (Brooks-Corey, Campbell, van Genuchten) fit measured data for practical use.

🌊 The fundamental water retention relationship

🌊 Positive relationship between content and potential

Soil water retention curve: the relationship between soil water content and matric potential.

  • Soil water content is positively related to soil matric potential.
  • As water content decreases, matric potential also decreases (becomes more negative).
  • This is not intuitive at first: "more negative" means stronger suction or lower energy state.

💧 Saturated water content

  • When all pores are filled with water, the soil is at saturated water content (θ_s) and matric potential is 0.
  • Example: for Rothamsted loam, θ_s ≈ 0.51 cm³ cm⁻³.
  • This is the starting point of the retention curve on the left axis.

📉 Plotting convention: absolute value and log scale

  • The absolute value of matric potential (sometimes called suction) is plotted on the x-axis.
  • A logarithmic scale is used because matric potential ranges over many orders of magnitude while water content varies less.
  • Moving right on the curve means moving toward more negative (lower) matric potentials.

🏞️ Key features of the retention curve

🚪 Air-entry potential

Air-entry potential (ψ_e): the highest matric potential at which air has displaced water in some pores of a previously saturated soil.

  • The curve is flat near saturation (between 1 cm and ~100 cm for Rothamsted loam), then bends downward.
  • For Rothamsted loam, ψ_e ≈ -128 cm of water.
  • This marks the transition from fully saturated to partially saturated conditions.

🔄 Inflection point

  • Halfway down the descending limb, the curve changes from concave to convex.
  • Practical significance:
    • Water content at this point may be optimum for tillage (produces the most small aggregates).
    • The slope at this point may indicate soil quality.

🏜️ Tail of the curve

  • At very negative matric potentials (dry conditions), the curve flattens.
  • Large decreases in matric potential produce only small decreases in water content.
  • This represents tightly bound water that is hard to remove.

🧱 Soil properties that shape retention

🧱 Texture: coarse vs. fine

Soil textureSaturated water contentAir-entry potentialWater retention
Sand (L-soil)0.18 cm³ cm⁻³-32 cm (less negative)Much lower
Sandy loam (Royal)IntermediateIntermediateIntermediate
Loam (Rothamsted)0.51 cm³ cm⁻³-128 cm (more negative)Much higher
  • Coarse-textured soils (sand) have larger pores, lower saturated water content, and less negative air-entry potential.
  • Fine-textured soils (loam, clay) have smaller pores, higher saturated water content, and more negative air-entry potential.
  • Example: walking on a sandy beach near water's edge, the ground is dry and firm; near a pond with fine-textured soil, the ground is wet and muddy—this difference reflects retention capability.

🏋️ Bulk density and compaction

  • Compacted soil (higher bulk density) has:
    • Lower porosity.
    • Lower saturated water content.
    • Lower (less negative) air-entry potential.
  • Sufficiently compacted soils can hold more water at matric potentials below the air-entry potential than uncompacted soil.
  • Example: comparing samples of silt loam at bulk densities from 1.01 to 1.34 g cm⁻³ shows clear differences in retention curves.

🌱 Organic matter: unclear evidence

  • Advocates claim increasing soil organic matter improves water retention.
  • Scientific evidence is mixed: about half of studies find a positive effect, half find no effect.
  • One hypothesis: organic matter may decrease bulk density, indirectly improving retention (similar to the compaction effect).

🔁 Hysteresis: wetting vs. drying

🔁 What is hysteresis?

Hysteresis: the phenomenon where the soil water retention curve differs between wetting (sorption) and drying (desorption).

  • For any given matric potential, the equilibrium water content is lower during wetting than during drying.
  • The initial water content also matters: drying from full saturation differs from drying that starts at a lower water content.
  • Example: in silty clay loam, the drying curve from saturation is clearly different from the drying curve starting at point "B."

🫧 Cause 1: Air entrapment

  • When a partially drained soil is rewetted, small pockets of air become trapped in interior pore spaces.
  • This trapped air cannot easily be removed, even if the soil is submerged.
  • Result: higher water contents occur along the primary drainage curve than during subsequent rewetting.
  • The maximum water content during rewetting is called the satiated water content (lower than true saturated water content).
  • X-ray computed tomography images show entrapped air (pink) in the macropore network of a satiated soil column.

💧 Cause 2: Contact angle hysteresis

Contact angle: the angle at which a liquid-gas interface meets a solid surface.

  • Hydrophilic soils (mineral soils): contact angle < 90°, affinity for water.
  • Hydrophobic soils (organic soils or mineral soils with organic coatings): contact angle > 90°, repel water.
  • Thought experiment: adding a small volume of liquid increases the contact angle slightly without moving the drop edge; removing liquid decreases it.
  • Contact angles are different for wetting vs. drying, leading to higher (less negative) pressure potentials during wetting for the same water content.

🍶 Cause 3: Ink bottle effect

Ink bottle effect: drainage from a large cavity is restricted if fluid must drain through a narrow opening.

  • Related to capillary rise: the rise of liquid against gravity due to attraction of liquid molecules to a solid surface and to each other.
  • Capillary rise height (h) depends on surface tension, contact angle, fluid density, gravity, and capillary radius: smaller radius → greater rise.
  • Example: two capillary tubes inserted into water—uniformly narrow tube has greater capillary rise than a tube with an enlarged section.
  • During wetting, capillary rise can only reach the bottom of the enlarged section.
  • During drying from an initially filled state, the enlarged section remains water-filled, so both tubes have equal water height.
  • Non-uniform pore radii cause hysteresis in the retention curve.

🌍 Practical effects of hysteresis

  • Increases water storage near the soil surface after infiltration and drainage.
  • Slows solute leaching under natural rainfall, with greater effects in coarse-textured than fine-textured soils.

🔬 Measuring retention curves

🔬 Why empirical methods?

  • We cannot yet theoretically predict retention curves from first principles (progress is being made).
  • Measurements are performed primarily by empirical methods: based on measurements and experience rather than theory.

🧪 Laboratory devices for different matric potential ranges

Matric potential rangeDeviceSample typePrinciple
0 to -10 kPaHanging water column (tension table)IntactPrecisely control matric potential; measure water content by mass change
-10 to -100 kPaTempe cellsIntactPressurized chamber with porous ceramic plate; air pressure equals desired matric potential; measure water outflow or mass change
-100 to -1500 kPaPressure plate extractorsHomogenized, smallerMultiple samples on pressure plates; similar principle to Tempe cells
Below -100 kPaDewpoint potentiometersAlternative approach; pressure plate data may be unreliable below -100 kPa
  • Near saturation: intact samples are needed because soil structure and inter-aggregate pores strongly influence retention.
  • Below -15 kPa: soil structure effects are negligible; smaller homogenized samples are used.
  • Pressure plates at low potentials: true equilibrium may take many weeks or never be reached; growing evidence suggests unreliability below -100 kPa.

📐 Mathematical functions for retention

📐 Why use functions?

  • After measuring retention at several matric potentials, we fit a mathematical function to allow calculation of water content for all other possible values of matric potential.

📐 Three widely-used functions

Brooks and Corey function:

  • Defined by residual water content (θ_r), air-entry potential (ψ_e), and pore size distribution parameter (λ).
  • Residual water content: conceptually, the water content below which liquid water flow is no longer possible.
  • Larger λ indicates more uniformly sized pores; small λ indicates wide pore size distribution.
  • Predicts a sharp drop in water content at the air-entry potential.

Campbell function:

  • Simpler, more convenient for hand calculations.
  • Uses saturated water content (θ_s), air-entry potential (ψ_e), and pore size distribution parameter (b).
  • Does not include a residual water content.
  • Also predicts a sharp drop at air-entry potential.

van Genuchten function:

  • More flexible and more widely used.
  • Parameters: residual water content (θ_r), saturated water content (θ_s), α (inversely related to air-entry potential), n (pore size distribution index), and m (often defined as m = 1 - 1/n).
  • Provides a smoother curve without a sharp drop at air-entry potential.
  • Generally has the lowest root mean square error (RMSE) when fit to measured data.

📐 Fitting and accuracy

  • The most accurate way to estimate parameters: obtain measurements across a broad range of matric potentials and adjust parameters for best agreement.
  • Example: for Tifton loamy sand and Waukegan silt loam, all three functions fit reasonably well, but van Genuchten has the lowest RMSE.

📐 Pedotransfer functions

Pedotransfer function: a statistical tool for estimating unknown soil properties from known soil properties.

  • If you lack measurements, you can estimate retention curve parameters from soil textural class alone using tables of average parameters.
  • These estimates are suitable for education and general approximations, not for research or design.
  • More accurate pedotransfer functions use additional properties: percent sand/silt/clay, bulk density, or one or more retention measurements.

🎯 Practical significance

🎯 Why retention matters

  • Understanding the retention curve is crucial to understanding:
    • Soil water storage.
    • Water flow.
    • Plant water uptake.
  • Differences in retention between soil textures dramatically influence water movement, plant growth, and related processes in managed and natural ecosystems.

🎯 Don't confuse

  • Saturated vs. satiated: saturated means all pores filled with water (matric potential = 0); satiated means maximum water content during rewetting after air entrapment (lower than saturated).
  • Wetting vs. drying curves: at the same matric potential, wetting curves show lower water content than drying curves due to hysteresis.
  • Residual water content: a model parameter representing the water content below which liquid flow is no longer possible; it is not always included in all retention functions (e.g., Campbell function omits it).
15

Pores and Pore Networks

4.1. Pores and Pore Networks

🧭 Overview

🧠 One-sentence thesis

Soil pore networks—especially macropores—create complex, tortuous flow paths that profoundly influence water movement, yet our fundamental theories largely ignore them because they were developed using homogenized soil rather than intact field samples.

📌 Key points (3–5)

  • What pores do: provide vital flow paths for water, oxygen, carbon dioxide, and nutrients; also serve as habitat for soil organisms.
  • Macropores matter most: pores larger than approximately 0.3 mm diameter are major contributors to water flow and solute transport.
  • Network structure, not just size: pore connectivity and organization influence flow—large dead-end pores may contribute less than smaller continuous pores.
  • Common confusion: our core theories were developed on homogenized soil/sand/glass beads in labs, so they largely ignore macropore effects that are critical in real field conditions.
  • Grand challenge: developing new concepts and models that incorporate macropores and their effects on water flow.

🕳️ What soil pores are and why they matter

🕳️ The role of pore spaces

Soil pores: the spaces between soil solid particles that provide flow paths and habitat.

  • Pores are the "complementary spatial patterns" to the solid phase—wherever there is no solid, there is pore space.
  • Without pores, water, oxygen, carbon dioxide, and nutrients could not move through soil, making life impossible.
  • Pores also provide habitat for living organisms in the soil.

🌐 Complexity of pore networks

  • Soil pore networks are "incredibly complex."
  • Key features:
    • Connectivity: different degrees of connection between pores.
    • Tortuosity: flow paths are winding and indirect, not straight.
    • Size variation: pore sizes vary by orders of magnitude even within the same soil.
  • Understanding pores and pore networks is a prerequisite for understanding water flow in soil.

🔍 Macropores: the largest and most influential pores

🔍 What macropores are

Macropores: the largest soil pores, defined (with respect to water flow and solute transport) as those with equivalent diameters greater than approximately 0.3 mm.

  • The size threshold is "somewhat subjective," but empirical evidence supports the 0.3 mm criterion for flow and transport.
  • Macropores are frequently major contributors to water flow and solute transport in soils.

🐛 How macropores form

Macropores can be created by:

  • Biological activity: soil-dwelling animals, plant roots.
  • Human activity: tillage practices.
  • Physical processes: soil shrinkage (resulting in cracks) and internal erosion (resulting in natural soil pipes).

🌱 Biological and chemical significance

  • Macropores are often "hot-spots" where:
    • Plant roots grow preferentially.
    • Soil organisms move more easily than in the bulk soil.
  • Enrichment of macropore walls: organic matter and clay content can be higher on macropore walls relative to the surrounding soil matrix.
  • Effects of enrichment:
    • Greater sorption and more rapid mineralization of organic contaminants.
    • Greater sorption of trace metals.
    • Greater microbial activity.
    • More rapid nitrogen cycling.

Example: A plant root creates a macropore as it grows; after the root dies, the pore remains and its walls are enriched with organic matter, making it a hotspot for microbial activity and nutrient cycling.

🕸️ Pore network organization and connectivity

🕸️ Why network structure matters

  • It is not only the size but also the organization of pores into networks that influences soil water flow.
  • Dead-end pores: large pores that do not connect to other pores may contribute little to water movement.
  • Continuous pores: smaller pores that provide a continuous flow path can be more important for flow than larger dead-end pores.

Don't confuse: A large pore is not automatically more important for flow—connectivity and continuity matter as much as or more than size.

📸 Visualizing pore networks

  • Technologies from medicine (e.g., X-ray computed tomography) now allow researchers to "see through" intact soil and visualize 3-D pore networks.
  • These images reveal networks that differ dramatically in shape and connectivity.
  • The images challenge current mental and mathematical models of soil water flow and storage.

Example: The excerpt references Fig. 4-2, which shows four soil samples with pore networks visualized using X-ray CT; the networks differ dramatically in their structure.

🚧 The gap between theory and reality

🚧 How our theories were developed

  • Fundamental concepts and theories for soil water storage and flow were mostly developed in the first half of the 20th century.
  • Developers: engineers and physicists.
  • Laboratory conditions: often worked with homogenized soil, sand, or even glass beads, instead of intact soil in the field.
  • Result: the soil water fundamentals that form the core of our knowledge largely ignore the existence and effects of macropores on soil water processes.

🧩 The grand challenge

One of the current grand challenges for soil physicists and hydrologists is to develop new, unifying concepts and quantitative models that effectively incorporate the now substantial body of information on soil macropores and their effects on water flow.

  • There is now a "substantial body of information" on macropores.
  • Existing concepts have limitations because they were developed without considering macropores.
  • As we learn established concepts of soil water flow, we need to keep in mind this grand challenge and the limitations of our existing concepts.

Don't confuse: The established theories are fundamental and still useful, but they are incomplete—they work well for homogenized soil but may not capture the full complexity of intact field soil with macropores.

🧠 A historical perspective

  • Leonardo Da Vinci is quoted: "We know more about the movement of celestial bodies than about the soil underfoot."
  • This is "probably true" with regards to our knowledge of macropores' effects on soil water flow.
  • The excerpt emphasizes that our understanding of soil pores and their effects is still in its infancy.
16

Soil Water Potential for Systems at Equilibrium

4.2. Soil Water Potential for Systems at Equilibrium

🧭 Overview

🧠 One-sentence thesis

Differences in soil water potential—not simply wetness or gravity—drive water flow in soil, and understanding equilibrium conditions is essential before solving flow problems.

📌 Key points (3–5)

  • What drives flow: Water flows in response to differences in total soil water potential, not necessarily downward or from wetter to drier soil.
  • Hydraulic equilibrium: When total water potential is uniform throughout the soil, no flow occurs, even though component potentials (gravitational, pressure) may vary.
  • Component offset: At equilibrium, gravitational and pressure potentials vary in ways that perfectly cancel each other out, keeping total potential constant.
  • Common confusion: Intuition suggests water always flows downward or from wet to dry, but water can flow upward, horizontally, or even from drier to wetter soil under certain potential gradients.
  • Foundation skill: Correctly determining potentials in equilibrium systems is a prerequisite for solving flow problems.

💧 The fundamental driver of soil water movement

💧 Potential differences, not intuition

  • The excerpt emphasizes that differences in soil water potential drive soil water flow.
  • Intuitive assumptions often mislead:
    • Water does not always flow downward.
    • Water does not always flow from wetter soil to drier soil.
  • Reality: water can flow upward, horizontally, or from drier to wetter soil depending on potential gradients.
  • Example: If a drier soil layer has higher total water potential than a wetter layer below it, water will flow from the drier layer downward.

🧮 Why understanding potential is essential

  • To predict water movement, you must supplement intuition with a solid grasp of soil water potential.
  • The excerpt stresses this as "the most fundamental concept for understanding soil water flow."

⚖️ Hydraulic equilibrium and component potentials

⚖️ What is hydraulic equilibrium?

Hydraulic equilibrium: the condition in which soil water potential is the same throughout the soil, so no water flow occurs.

  • When total potential is uniform, water is at rest.
  • This does not mean all component potentials are uniform.

🔀 Component potentials vary but cancel out

  • Even at equilibrium, gravitational potential and pressure potential often vary throughout the soil.
  • These variations perfectly offset one another so that total potential remains constant.
  • The excerpt notes: "if we neglect any differences in osmotic potential, then for soil at hydraulic equilibrium, the variations in gravitational potential and pressure potential will perfectly offset one another."
  • Example: At a higher elevation in the soil column, gravitational potential is higher, but pressure potential is correspondingly lower, keeping total potential the same.

📊 How to analyze equilibrium systems

  • The excerpt directs readers to use diagrams and tables to determine soil water potentials for systems in equilibrium.
  • A video is referenced that explains how to create and fill in these tables step-by-step.
  • Don't confuse: equilibrium means no flow, but it does not mean all potentials are zero or equal—only that their sum is uniform.

🎯 Practical skill: solving equilibrium problems

🎯 Why this skill matters

  • Before you can solve soil water flow problems, you must be able to correctly determine potentials in equilibrium systems.
  • The excerpt emphasizes this as a foundational skill: "Before you can solve soil water flow problems, you need to be able to correctly determine soil water potentials for systems in equilibrium."

🛠️ Tools and methods

  • Use sketches, tables, and systematic calculation of gravitational head, pressure head, and total head at different points.
  • The accompanying video provides worked examples.
  • Practice is essential: the excerpt encourages readers to "get a pencil and paper and watch the video now, taking time to create your own tables and fill in the blanks as you go."
17

Poiseuille's Law

4.3. Poiseuille’s Law

🧭 Overview

🧠 One-sentence thesis

Poiseuille's Law for flow through tubes reveals three key principles—flow is proportional to the hydraulic gradient, strongly dependent on pore size, and influenced by fluid viscosity—that help us understand soil water flow even though the law itself cannot be directly applied to soil.

📌 Key points (3–5)

  • Flow direction and rate: water flows from higher to lower potential; Poiseuille's Law relates flow rate to the pressure difference (hydraulic gradient).
  • Pore size matters enormously: flux depends on the square of the tube radius, so a 10× larger radius means 100× larger flux for the same gradient.
  • Temperature and viscosity: fluid viscosity increases as temperature decreases, reducing flow rates in colder conditions.
  • Common confusion: Poiseuille's Law provides insights but cannot be directly applied to soil because soil is not a smooth straight tube and lacks a meaningful single "radius."
  • Why it matters: these principles set the stage for understanding Darcy's Law, which does work for soil water flow problems.

🌊 Flow driven by potential differences

🌊 Direction of flow

  • Water flows from regions of higher potential to regions of lower potential, unless blocked by an impermeable layer.
  • Knowing soil water potentials at different locations is enough to determine flow direction.
  • To estimate the rate of flow, we need the relationship between potential differences and flow rate.

📐 Poiseuille's Law for tubes

Poiseuille's Law: Q = (π r⁴ Δp) / (8 η L), where Q is volume flow per unit time, r is tube radius, Δp is pressure difference, η is dynamic viscosity, and L is tube length.

  • The term Δp is analogous to the difference in water potential between two points in soil.
  • The term Δp / L defines the hydraulic gradient: the ratio of pressure difference to distance.
  • Key insight: flow rate through a tube is proportional to the hydraulic gradient—if the gradient decreases, flow rate decreases.
  • Example: a steeper pressure drop over the same distance → higher hydraulic gradient → faster flow.

🔍 Three key insights from Poiseuille's Law

🔍 Insight 1: Proportionality to hydraulic gradient

  • Flow rate is proportional to the hydraulic gradient (Δp / L).
  • This principle also holds true for soil water flow, even though soil is not a tube.

📏 Insight 2: Strong dependence on pore size

  • Flux (q) is volumetric flow rate per unit area.
  • By inspecting Poiseuille's Law and noting that tube cross-sectional area is π r², flux is given by: q = (r² Δp) / (8 η L).
  • The term shows that flux depends strongly on tube radius.
  • If radius is larger by a factor of 10, flux is larger by a factor of 100.
  • Don't confuse: this is not a linear relationship; doubling the radius quadruples the flux for the same gradient.
  • Example: larger soil pores will transmit water much more rapidly than smaller pores under the same hydraulic gradient.

🌡️ Insight 3: Fluid properties influence flow

  • The term η⁻¹ (inverse of viscosity) shows that as fluid viscosity increases, flux decreases.
  • Water and air viscosities both increase as temperature decreases.
  • Therefore, flow rates through soil will decrease as temperature decreases.
  • Example: colder soil water flows more slowly than warmer soil water under the same gradient.

🚧 Limitations for soil applications

🚧 Why Poiseuille's Law cannot be directly applied to soil

  • Published in 1841, Poiseuille's Law provides helpful insights but has little practical value for solving soil water flow problems.
  • Soil is not a smooth straight tube, nor is it a bundle of smooth straight tubes.
  • We usually cannot identify any meaningful "radius" for the soil pore network that would allow direct application of Poiseuille's Law.
  • We need another approach → this sets up the need for Darcy's Law (covered in the next section).
18

Darcy's Law

4.4. Darcy’s Law

🧭 Overview

🧠 One-sentence thesis

Darcy's Law provides a practical, flexible relationship for predicting water flow through saturated porous media by relating flow rate to the hydraulic gradient and a single empirically determined constant, the hydraulic conductivity.

📌 Key points (3–5)

  • What Darcy's Law states: water flow through saturated porous media is proportional to the hydraulic gradient, with hydraulic conductivity as the proportionality constant.
  • Why it's more practical than Poiseuille's Law: Darcy's Law requires only an empirical constant (hydraulic conductivity) rather than specific geometric details like tube radius, making it applicable to virtually any porous media.
  • Key trade-off: flexibility comes at a cost—hydraulic conductivity varies enormously (several orders of magnitude) between soil types and even within the same soil type.
  • Common confusion: saturated vs unsaturated hydraulic conductivity—these differ dramatically, and Darcy's original work focused only on saturated conditions.
  • What controls hydraulic conductivity: soil texture, macropore presence and connectivity, soil structure, and chemical properties all strongly influence flow rates.

📐 The law itself

📐 Darcy's discovery and formulation

  • In 1856, French engineer Henry Darcy proved that water flow through sand beds was proportional to the hydraulic gradient—the same relationship Poiseuille had found for laminar flow through tubes.
  • The proportionality constant is called the hydraulic conductivity (K).

Darcy's Law: q = -K_s × (Δψ_t / L)

where Δψ_t is the difference in total water potential between two points in saturated porous media separated by distance L.

  • The symbol K_s clarifies we are referring to saturated hydraulic conductivity, which differs dramatically from unsaturated soil conductivity.

🔬 What hydraulic conductivity represents

  • Hydraulic conductivity is a measure of a material's ability to transmit water.
  • Darcy noted it depends on the permeability of the porous media.
  • By comparing with Poiseuille's Law, we can see it also depends on fluid viscosity, which varies with temperature.
  • Although experiments used only sand, the relationship accurately describes flow in a wide array of soil types and other porous media.
  • It has become one of the most important relationships in soil physics, hydrology, and hydrogeology.

🆚 Darcy's Law vs Poiseuille's Law

🆚 Key advantage: flexibility

  • Poiseuille's Law requires specific geometric information (the radius of the tube).
  • Darcy's Law requires only an empirically determined constant (hydraulic conductivity).
  • This gives Darcy's Law the flexibility to apply to virtually any porous media, as long as flow is laminar.
  • Example: Soil is not a smooth straight tube, nor a bundle of smooth straight tubes—we usually cannot identify any meaningful "radius" for the soil pore network to directly apply Poiseuille's Law.

⚖️ Key trade-off: variability

  • The flexibility comes at a cost: hydraulic conductivity can differ by several orders of magnitude from one soil type to another.
  • It can also differ by an order of magnitude or more from one sample to another within the same soil type.
  • This enormous variability requires understanding the primary factors that influence hydraulic conductivity.

🌍 Factors affecting saturated hydraulic conductivity

🏖️ Soil texture

  • Soil texture strongly influences saturated hydraulic conductivity.
  • Sand-dominated soils: relatively large pore spaces → large K_s values.
  • Clay-dominated soils: relatively small pore spaces → small K_s values.
  • Example from Reynolds et al. measurements using intact soil cores in Canada:
    • Sand: 29 cm per hour
    • Loam: 4.1 cm per hour
    • Clay loam: 0.091 cm per hour
    • The sand's K_s was more than 300 times larger than the clay loam's.

🕳️ Macropores

  • The presence, size, and continuity of macropores can strongly influence saturated hydraulic conductivity.
  • Poiseuille's Law indicates flux through a tube increases with the square of the radius; water flux through soil is similarly sensitive to large pores, even if few in number.
  • Many other pore features matter: connectivity, internal roughness, and tortuosity (a measure of the extent of twists and turns taken by the pores).
  • Don't confuse texture with structure: clayey soils with large numbers of well-connected macropores generated by living organisms (e.g., earthworms) can have K_s values greater than coarse-textured soils that lack macropores.

🧱 Soil structure

  • The type and degree of soil structure affect saturated hydraulic conductivity.
  • Strongly developed, fine blocky structure: contributes to high K_s values.
  • Massive, featureless structure: often indicates compaction and low K_s values.
  • By carefully measuring structural features (length, width, number of inter-aggregate pores) and applying an appropriate model, one can reasonably predict K_s.
  • Practical application: some US states use visual inspection of soil profiles by trained soil scientists as a primary factor in determining suitability for septic system drain fields.

🧪 Chemical properties

  • Chemical properties of the soil and the flowing solution can impact saturated hydraulic conductivity.
  • These effects arise when soil and solution characteristics promote swelling and chemical dispersion of clay.

Chemical dispersion: the process in which soil particles, previously held together in close contact within soil aggregates, respond to a changed chemical environment by expanding and separating from one another, breaking down the soil aggregates.

  • Swelling and dispersion can reduce saturated hydraulic conductivity by a factor of 100 or more.
  • Conditions that promote swelling and dispersion:
    • Irrigation with sodic water
    • High content of 2:1 clays, particularly montmorillonite
    • Low electrical conductivity of the flowing solution
    • High exchangeable sodium percentage (ESP) in the soil
19

Factors Affecting Saturated Hydraulic Conductivity

4.5. Factors Affecting Saturated Hydraulic Conductivity

🧭 Overview

🧠 One-sentence thesis

Saturated hydraulic conductivity varies by orders of magnitude among soils because it is controlled by texture, macropore structure, and chemical conditions that cause clay swelling or dispersion.

📌 Key points (3–5)

  • Texture dominates: sand-dominated soils have large pores and high conductivity; clay-dominated soils have small pores and low conductivity (differences can exceed 300×).
  • Macropores and structure matter: even clayey soils can have high conductivity if they contain well-connected macropores (e.g., from earthworms) or strong blocky structure.
  • Chemical effects: swelling and dispersion of clay—triggered by sodic water, low electrical conductivity, or high exchangeable sodium—can reduce conductivity by 100× or more.
  • Flocculation reverses dispersion: adding polyvalent cations (Ca²⁺, Mg²⁺) through gypsum or acid can restore conductivity by bringing dispersed particles back together.
  • Common confusion: don't assume texture alone determines conductivity—macropore connectivity and chemical environment can override texture effects.

🏗️ Physical factors

🏗️ Soil texture

Soil texture strongly influences saturated hydraulic conductivity.

  • Sand-dominated soils: large particles → large pore spaces → high conductivity.
  • Clay-dominated soils: small particles → small pore spaces → low conductivity.
  • Example: measured values were 29 cm/h for sand, 4.1 cm/h for loam, and 0.091 cm/h for clay loam—the sand was more than 300 times higher than the clay loam.
  • This is the most obvious factor, but not the only one.

🕳️ Macropores and connectivity

  • Poiseuille's Law insight: flux through a tube increases with the square of the radius, so even a few large pores can dominate flow.
  • Other pore features matter:
    • Connectivity (are pores linked?)
    • Internal roughness
    • Tortuosity (twists and turns in the pore path)
  • Clayey soils can have high conductivity if they contain large numbers of well-connected macropores created by organisms like earthworms—sometimes exceeding coarse-textured soils that lack macropores.
  • Don't confuse: texture alone does not determine conductivity; macropore structure can override texture.

🧱 Soil structure

  • Strongly developed, fine blocky structure → high conductivity (many inter-aggregate pores).
  • Massive, featureless structure → often indicates compaction → low conductivity.
  • The excerpt describes measuring structural features (length, width, number of inter-aggregate pores) and using models to predict conductivity.
  • Practical application: some US states use visual inspection of soil profiles by trained scientists to estimate flow rates and determine suitability for septic system drain fields.

🧪 Chemical factors

🧪 Chemical dispersion

Chemical dispersion: the process in which soil particles, previously held together in close contact within soil aggregates, respond to a changed chemical environment by expanding and separating from one another, breaking down the soil aggregates.

  • What happens: clay swells and aggregates break apart, reducing pore space and conductivity.
  • Magnitude: swelling and dispersion can reduce saturated hydraulic conductivity by a factor of 100 or more.
  • Conditions that promote dispersion:
    • Irrigation with sodic water
    • High content of 2:1 clays, especially montmorillonite
    • Low electrical conductivity of the flowing solution
    • High exchangeable sodium percentage (ESP) in the soil
  • ESP definition: amount of exchangeable sodium divided by the sum of exchangeable calcium, magnesium, potassium, and sodium.
  • Example: leaching a sandy loam column (initial ESP 10%) with distilled water caused a 90% reduction in conductivity.

🔄 Flocculation (reversing dispersion)

Flocculation: the process in which dispersed soil particles come together, often due to a change in the chemical environment.

  • What promotes flocculation: high proportion of polyvalent cations (Ca²⁺, Mg²⁺, Al³⁺).
  • What promotes dispersion: high proportion of monovalent cations, especially Na⁺.
  • Remediation strategies:
AmendmentMechanismBest for
Gypsum (CaSO₄)Provides Ca²⁺ to displace Na⁺ on cation exchange sitesSodic and saline-sodic soils
Sulfuric acid (H₂SO₄)Dissolves calcium carbonate (CaCO₃) in soil, releasing Ca²⁺ to displace Na⁺Calcareous sodic soils
  • Don't confuse: flocculation is not just "drying out"—it is a chemical process driven by cation composition.

🔍 Why variability is so large

🔍 Multiple interacting factors

  • The excerpt emphasizes that hydraulic conductivity "can differ by several orders of magnitude from one soil type to another" and "by an order of magnitude or more from one sample to another within the same soil type."
  • Reasons for enormous variability:
    • Texture sets a baseline, but macropores and structure can shift it dramatically.
    • Chemical conditions can reduce conductivity by 100× even in the same soil.
    • All these factors interact: a clayey soil with good structure and low sodium can have higher conductivity than a sandy soil with poor structure and high sodium.
  • This variability is the "cost" of the flexibility mentioned at the start: Darcy's Law applies broadly, but the conductivity parameter itself is highly variable.
20

Darcy's Law for Layered Soil

4.6. Darcy’s Law for Layered Soil

🧭 Overview

🧠 One-sentence thesis

Darcy's Law can be adapted to calculate saturated water flow through soil profiles with multiple layers by summing the hydraulic resistances of each layer.

📌 Key points (3–5)

  • What the adaptation does: extends Darcy's Law to handle soils with distinct horizons that have different hydraulic conductivities.
  • How it works: hydraulic resistance of each layer (thickness divided by conductivity) is added to calculate total flow.
  • Key limitation: this form applies only to saturated flow, not to unsaturated conditions.
  • Common confusion: saturated vs unsaturated flow—many field situations (infiltration, redistribution, drainage) involve unsaturated soil, which requires a different approach.
  • Why it matters: allows estimation of water flow rates through layered soil profiles commonly encountered in the field.

🧱 The layered soil problem

🧱 Why layers matter

  • Saturated hydraulic conductivity can vary by orders of magnitude between different soil horizons.
  • Soil profiles consist of horizons with differing properties, so a single-layer approach is insufficient.
  • The excerpt asks: "what does that mean for saturated flow through soil profiles?"

🔧 The solution approach

  • Darcy's Law is rearranged into a convenient form for solving saturated water flow problems when multiple layers are present.
  • Each layer contributes a hydraulic resistance term to the overall flow calculation.

🧮 Hydraulic resistance concept

🧮 Definition and formula

Hydraulic resistance of a soil layer: the thickness of the layer (L) divided by the hydraulic conductivity of the layer, R_h = L / K.

  • This is the key building block for the layered-soil version of Darcy's Law.
  • Resistance increases with layer thickness and decreases with higher conductivity.

📐 Two-layer equation

  • For a soil with two distinct layers, Darcy's Law can be written as an equation (Eq. 4-4) that includes R_h1 and R_h2 (hydraulic resistances for layers 1 and 2).
  • The equation structure: flow is driven by the total hydraulic head difference, divided by the sum of the resistances.

🔢 Extending to more layers

  • For soils with more than two layers, additional hydraulic resistances are added to the denominator for each layer.
  • This modular approach allows the method to scale to any number of distinct horizons.

⚠️ Scope and limitations

⚠️ Saturated flow only

  • Both the standard form and the layered-soil form of Darcy's Law apply only to saturated water flow.
  • The excerpt emphasizes this limitation explicitly.

🌱 Field conditions often unsaturated

  • Many field situations require understanding or estimating water flow rates when the soil is unsaturated.
  • Examples of unsaturated processes:
    • Infiltration into soil profiles
    • Redistribution through soil profiles
    • Drainage from soil profiles
  • Don't confuse: layered Darcy's Law is useful for saturated profiles, but a different relationship is needed for unsaturated conditions.

⏳ Historical context

  • It took 50 years after Darcy's landmark paper before someone developed a relationship to predict water flow in unsaturated soil.
  • This sets up the transition to the next section (Buckingham-Darcy Law), which addresses unsaturated flow.

🎯 Practical application

🎯 What the method enables

  • Allows estimation of water flow rates for layered soils likely to be encountered in the field.
  • The excerpt references a video example showing how to apply Darcy's Law for layered soil to calculate flow through a soil profile with two distinctly different layers.

📝 Problem-solving approach

  • The method involves:
    • Identifying each distinct layer
    • Calculating the hydraulic resistance for each layer (thickness / conductivity)
    • Summing the resistances
    • Using the modified Darcy's Law equation to find flow rate
  • Example scenario: A soil profile with a sandy upper layer and a clayey lower layer—each has a different thickness and conductivity, so each contributes a different resistance term to the total flow calculation.
21

Buckingham-Darcy Law

4.7. Buckingham-Darcy Law

🧭 Overview

🧠 One-sentence thesis

The Buckingham-Darcy Law generalizes Darcy's Law to describe water flow in both saturated and unsaturated soil by accounting for how hydraulic conductivity and the driving gradient both depend on soil water content.

📌 Key points (3–5)

  • Historical breakthrough: Edgar Buckingham (1902) recognized that matric potential gradients drive unsaturated flow and that hydraulic conductivity depends on water content.
  • The generalized equation: Buckingham-Darcy Law includes hydraulic conductivity as a function of water content, K(θ), and uses total water potential gradient (pressure + gravitational) as the driving force.
  • Three key differences from Darcy's Law: (1) conductivity varies strongly with water content; (2) the gradient itself depends on water content through matric potential; (3) the law must be applied as a differential equation for infinitesimal layers, not directly to finite layers.
  • Common confusion: Unlike Darcy's Law, which can be applied directly to a saturated layer, Buckingham-Darcy Law cannot use single values for a varying unsaturated layer—integration or simplifying assumptions are needed.
  • Why it matters: This law enables prediction of water flow in unsaturated conditions (infiltration, redistribution, drainage), which are common in field soils.

🔬 Buckingham's conceptual breakthrough

🧪 The historical context

  • In 1902, physicist Edgar Buckingham was hired by the US Department of Agriculture's Bureau of Soils.
  • He worked there only four years, but his studies led to a conceptual breakthrough that changed soil physics and hydrology.
  • Building on earlier work by John Maxwell and Lyman Briggs, Buckingham developed the foundation for unsaturated flow theory.

💡 What Buckingham discovered

Buckingham made several key insights:

  • Capillary potential (matric potential): Water in unsaturated soil is attracted to and held by soil solid surfaces through what he called "capillary potential" (now known as matric potential).
  • Spatial gradients drive flow: Gradients in this potential act as a force to drive soil water flow; the resulting flow is proportional to and in the opposite direction of that gradient.
  • Conductivity depends on water content: In unsaturated soil, the "capillary conductance" (hydraulic conductivity) depends largely on soil water content.
  • First retention curves: Buckingham measured equilibrium soil water content variation with height above a water table in laboratory columns, producing the first published soil water retention curves relating water content and capillary potential.

Example: Water flows from areas of higher matric potential to areas of lower matric potential, just as water flows downhill in response to gravitational potential gradients.

🧮 The resulting equation

Buckingham-Darcy Law: A single equation describing both saturated and unsaturated water flow in soil and other porous media, written as a differential equation where K(θ) indicates hydraulic conductivity as a function of volumetric water content and d(ψ_p + ψ_g)/dz indicates the gradient of total water potential in the z direction.

  • The equation can be viewed as a generalization of Darcy's Law.
  • Osmotic potential can also be included with the other potentials if necessary.
  • The notation d(ψ_p + ψ_g)/dz means the gradient is defined for an infinitesimal layer thickness.

🔑 Three key differences from Darcy's Law

🔑 Difference 1: Hydraulic conductivity depends strongly on water content

How conductivity varies:

  • In unsaturated soil, hydraulic conductivity depends strongly on soil water content.
  • Water content can vary in space and time.
  • The sensitivity is dramatic: conductivity can drop by more than one order of magnitude with small water content changes.

Example from Grenoble sand:

  • At water content 0.30 cm³/cm³: hydraulic conductivity ≈ 15 cm/h
  • At water content 0.20 cm³/cm³: hydraulic conductivity ≈ 1 cm/h
  • A drop of only 0.10 cm³/cm³ in water content causes conductivity to fall by more than a factor of 10.

Why this happens (the physical mechanism):

  1. From the soil water retention curve: soil water content is positively related to soil matric potential.
  2. When water content decreases (e.g., from 0.30 to 0.20 cm³/cm³), matric potential also decreases.
  3. From the capillary rise equation: as matric potential decreases, water is held in capillary spaces with smaller radii, r.
  4. From Poiseuille's Law: flow through a capillary depends on r².
  5. If capillary radius is reduced by a factor of 4, flow is reduced by a factor of 16.

Don't confuse: Soils are not capillary tubes, but this reasoning helps explain why hydraulic conductivity depends so strongly on soil water content.

🔑 Difference 2: The gradient driving flow depends on water content

How the gradient varies:

  • The gradient term in Buckingham-Darcy Law includes ψ_p, the pressure potential.
  • For unsaturated soil, the pressure potential term includes matric potential, ψ_m.
  • Matric potential is related to soil water content by the water retention curve.

What this means:

  • Within an otherwise homogeneous soil, gradients in water content indicate gradients in matric potential.
  • These matric potential gradients drive soil water flow from areas of higher water content to areas of lower water content.

Example: If one part of a soil layer has higher water content than another part, water will flow from the wetter area to the drier area, even if both are at the same elevation.

🔑 Difference 3: Cannot be applied directly to finite layers with varying water content

The application problem:

  • Darcy's Law can be applied directly to determine flux through a saturated soil layer of any desired thickness, so long as hydraulic conductivity and hydraulic gradient are defined for that layer.
  • Buckingham-Darcy Law as written cannot be applied directly to determine flux through an unsaturated layer when water content varies throughout the layer.

Why this limitation exists:

  • In varying unsaturated layers, hydraulic conductivity and hydraulic gradient can vary by orders of magnitude within the layer.
  • No single value for those variables can represent the entire layer.
  • That is why Buckingham-Darcy Law is written as a differential equation with d(ψ_p + ψ_g)/dz indicating the gradient is defined for a layer of infinitesimal thickness.
  • The corresponding hydraulic conductivity must also be evaluated for the water content of that infinitesimal layer.
  • The product of gradient and conductivity gives the flux for that vanishingly thin layer.

🧮 How to apply the law in practice

🧮 Integration for finite layers

To calculate flux across a real soil layer of finite thickness, the Buckingham-Darcy Law must be integrated across that layer.

Limitations:

  • Such integration is possible only for some relatively simple and specific flow problems.
  • Example application mentioned: evaporation from an underground water table (covered later in the source material).

🧮 Simplifying assumptions

In other cases, it is possible to apply simplifying assumptions to the Buckingham-Darcy Law:

  • Unit hydraulic gradient assumption: Assuming a unit hydraulic gradient results in an algebraic expression to calculate flux at a particular depth in the soil.
  • Example application mentioned: soil water redistribution and deep drainage (covered later in the source material).

Don't confuse: The need for integration or simplifying assumptions is a consequence of the first two key differences—both conductivity and gradient vary with water content in ways that cannot be captured by single representative values.

22

4.8. Models for Soil Hydraulic Conductivity

4.8. Models for Soil Hydraulic Conductivity

🧭 Overview

🧠 One-sentence thesis

Mathematical functions for hydraulic conductivity allow calculation of conductivity at any water content, corresponding to each soil water retention model and using pore-size distribution parameters.

📌 Key points (3–5)

  • Why models are needed: measurements exist only at saturation and a few points below, but calculations require conductivity values for all water contents.
  • Three main models: Brooks and Corey, Campbell, and van Genuchten functions each correspond to a water retention model.
  • Pore-size distribution parameter: each model uses a parameter (λ, b, or n) that describes whether pores are uniform or varied in size.
  • Common confusion: larger λ means more uniform pores (narrow distribution), while small λ means wide distribution of pore sizes—don't reverse the interpretation.
  • Most common choice: the van Genuchten function is the most commonly used.

🔬 Why hydraulic conductivity models exist

🔬 The measurement gap

  • In practice, we measure hydraulic conductivity at saturation (K_s) and perhaps at one or two water contents below saturation.
  • However, soil water flow calculations require knowing conductivity at all water contents, not just a few measured points.
  • Mathematical functions fill this gap by allowing calculation of K(θ) for any volumetric water content θ.

🔗 Link to retention models

  • Each hydraulic conductivity function corresponds to a soil water retention function from Chapter 3.
  • The same parameters that describe how water is retained also describe how easily water flows.
  • This pairing ensures consistency between the retention curve and the conductivity function.

📐 The three main conductivity models

📐 Brooks and Corey function

K(θ) is the hydraulic conductivity as a function of volumetric water content, K_s is the saturated hydraulic conductivity, and λ is the pore-size distribution index.

  • Uses the same λ parameter as the Brooks and Corey water retention curve.
  • The function is given by Equation 4-6 (specific formula not reproduced here).

📐 Campbell function

  • Corresponds to the Campbell water retention model.
  • Uses parameter b, which is related to pore size distribution.
  • The function is given by Equation 4-7.

📐 van Genuchten function

  • Corresponds to the van Genuchten water retention model.
  • The excerpt states this is the most commonly-used hydraulic conductivity function.
  • Uses two parameters:
    • n: a pore-size distribution index similar to λ
    • m: defined as m = 1 – 1/n
  • The function is given by Equation 4-8.

🧱 Understanding pore-size distribution parameters

🧱 What the parameter tells you

Parameter valueWhat it meansPhysical interpretation
Larger λ (or n)More uniformly-sized poresNarrow distribution of pore sizes
Small λ (or n)Wide distribution of pore sizesPores vary greatly in size

🧱 Why pore distribution matters

  • The pore-size distribution controls both water retention and flow.
  • Uniform pores (large λ) behave more predictably; varied pores (small λ) create more complex flow patterns.
  • The same parameter appears in both the retention curve and the conductivity function, linking the two behaviors.

⚠️ Don't confuse

  • "Larger λ" does NOT mean "more pores" or "larger pores on average."
  • It means the pores are more uniform in size—a narrower range of sizes.
  • Small λ means a wide range of pore sizes, from very small to very large.
23

Soil Water Flow: Conclusion and Integration

4.9. Conclusion

🧭 Overview

🧠 One-sentence thesis

This chapter integrates the fundamental concepts of soil pore networks, water potential, and three foundational flow laws (Poiseuille's, Darcy's, and Buckingham-Darcy) to prepare for studying the dynamic processes of the soil water balance that sustain life on Earth.

📌 Key points (3–5)

  • What was covered: complexity of soil pores and pore networks, how to determine gravitational/pressure/total water potentials at equilibrium and during steady saturated flow, and three foundational laws for water flow in soil.
  • Hydraulic conductivity's role: determines soil water flow rates; exhibits enormous variability controlled by soil and water factors.
  • Common confusion: distinguishing among the three laws—Poiseuille's Law (single tube), Darcy's Law (saturated bulk soil), and Buckingham-Darcy Law (unsaturated soil with variable conductivity).
  • Why it matters: these fundamental concepts are the foundation for understanding each process in the soil water balance, starting with water inputs before water even reaches the soil.

🌊 The three foundational flow laws

🔬 Poiseuille's Law

  • Applies to flow through a single tube of known diameter.
  • Used to calculate volumetric flow rate through individual pores.
  • Example: when a 1.0-mm diameter tube is pushed through a soil column, water flows through that tube according to Poiseuille's Law while the surrounding soil follows Darcy's Law.

🏞️ Darcy's Law

  • Applies to saturated bulk soil (the entire porous medium).
  • Calculates flux based on saturated hydraulic conductivity and hydraulic gradient.
  • Used for steady, saturated flow through homogeneous or layered soil.

💧 Buckingham-Darcy Law

  • Applies to unsaturated soil where water content varies with depth.
  • Includes a differential equation term d(ψ_p + ψ_g)/dz indicating the gradient is defined for an infinitesimally thin layer.
  • Hydraulic conductivity must be evaluated for the water content of each infinitesimal layer.
  • The product of gradient and conductivity gives flux for that vanishingly thin layer.

Don't confuse: To calculate flux across a real soil layer of finite thickness, the Buckingham-Darcy Law must be integrated across that layer; this is only possible for relatively simple flow problems or by applying simplifying assumptions (e.g., unit hydraulic gradient).

📊 Water potential tables and equilibrium

📐 How to construct water potential tables

The excerpt describes making "a simple table to determine the gravitational, pressure, and total water potentials for soil-water systems."

  • At equilibrium: total head is constant throughout the profile; gravitational and pressure heads vary with depth.
  • During steady, saturated flow: hydraulic gradient drives flow; differences in total head across layers determine flux.

🔄 Practical applications

  • Determining pressure head when soil reaches hydraulic equilibrium with a hanging water column.
  • Calculating hydraulic head differences across soil columns with ponded water.
  • Analyzing layered soil systems with different saturated hydraulic conductivities.

🧮 Hydraulic conductivity models

📈 Why mathematical functions are needed

The excerpt states: "We sometimes have measurements of soil hydraulic conductivity at saturation and perhaps at one or two water contents below saturation, but we often need a mathematical function to allow calculation of hydraulic conductivity for all other values of water content."

  • Measurements are limited to a few water contents.
  • Mathematical functions allow calculation for all water content values.

🔢 Three main models

The excerpt presents three hydraulic conductivity functions corresponding to water retention models:

ModelKey parameterWhat it represents
Brooks and Coreyλ (lambda)Pore-size distribution index; larger values = more uniformly-sized pores; small values = wide distribution
CampbellbParameter related to pore size distribution
van Genuchtenn and mn is pore-size distribution index similar to λ; m = 1 - 1/n

Note: The van Genuchten model is described as "the most commonly-used hydraulic conductivity function."

🔀 Variability in hydraulic conductivity

📉 Enormous variability occurs

The excerpt emphasizes "the enormous variability which occurs in hydraulic conductivity."

  • Hydraulic conductivity is not a single fixed value for a soil.
  • It varies with water content (much higher at saturation than in unsaturated conditions).
  • Multiple soil and water factors contribute to this variability.

🧩 Factors controlling variability

The excerpt mentions "the soil and water factors which contribute to that variability" but does not detail them in this conclusion section. Earlier sections (referenced but not included in this excerpt) covered:

  • Pore network complexity and connectivity.
  • Water content (saturated vs. unsaturated).
  • Soil structure and pore size distribution.

🌍 Transition to soil water balance processes

🚀 What comes next

The excerpt states: "Equipped with understanding of these fundamental concepts, we are now ready to begin studying one-by-one the processes of the soil water balance, that dynamic ebb and flow which sustains life on Earth."

The study sequence will be:

  1. Water inputs (precipitation and irrigation) — "our study must begin before the water ever reaches the soil"
  2. Interception by plant canopies and residue
  3. Impact of water drops on soil surface
  4. Infiltration into soil
  5. Runoff and erosion by water
  6. Redistribution within soil profile
  7. Drainage from bottom of profile
  8. Groundwater pollution and soil salinization
  9. Evaporation from bare soil and wind erosion
  10. Root water uptake and transpiration
  11. Land surface energy balance processes

🔗 Integration of concepts

The conclusion emphasizes that all the concepts covered—pore networks, water potential, flow laws, and hydraulic conductivity—are foundational and necessary for understanding each subsequent process in the soil water balance.

24

Precipitation

5.1. Precipitation

🧭 Overview

🧠 One-sentence thesis

Precipitation patterns—shaped by amount, intensity, and raindrop size—are fundamental drivers of Earth's ecosystems and human settlement, with extreme rainfall events capable of delivering erosive energy far exceeding the soil's capacity to absorb water.

📌 Key points (3–5)

  • Global precipitation patterns: annual amounts range from <100 mm in deserts to >2000 mm (even >5000 mm in some locations) in rainforests, strongly influencing natural ecosystems and human population distribution.
  • Precipitation intensity extremes: world-record intensities (e.g., 305 mm in 42 minutes) far exceed soil infiltration capacity and require horizontal transport of water vapor into storms, since local atmospheric water content is typically <60 mm.
  • Raindrop size and energy: raindrops typically range 1–4 mm in diameter, with larger drops forming at higher intensities; intense rainfall delivers substantial kinetic energy (approaching 10,000 J m⁻² h⁻¹) that can degrade soil structure.
  • Common confusion: extreme rainfall totals vs. atmospheric water capacity—record events depend on continuous vapor transport, not just the water already present above one location.
  • Historical foundation: Ewald Wollny's 1874 experiments were among the first to quantify how vegetative canopies intercept 12–55% of rainfall and protect soil from raindrop impact, establishing early soil physics and conservation science.

🌍 Global precipitation patterns and their impacts

🌍 Amount and distribution

Average annual precipitation: ranges from <100 mm in the world's great deserts to >2000 mm in equatorial rainforests, with a few locations exceeding 5000 mm.

  • The excerpt emphasizes that this global pattern is "arguably one of the most important influences on the Earth's coupled human and natural systems."
  • Precipitation patterns are clearly reflected in both natural ecosystem distribution and human population patterns.
  • Example: The majority of cities with >1,000,000 people are located in regions with 500–2000 mm average annual rainfall.

🏔️ Orographic lift and rain shadows

  • Orographic lift: mountains lift warm, moist ocean winds, cooling the air and causing precipitation.
  • This process produces:
    • Relatively high precipitation on the upwind side of mountains
    • Relatively low precipitation areas (rain shadows) on the downwind side
  • Example: Eastern India receives >5000 mm annually due to orographic lift, while rain shadows of the Himalayas and Andes receive near 0 mm of precipitable water.

🚰 Human responses to precipitation patterns

  • If climate change substantially alters precipitation patterns, the excerpt predicts shifts in human population or increased water transfers (or both).
  • Large-scale water transfer projects have already begun.
  • Example: China's "South-to-North Water Diversion" project transfers 25 billion cubic meters of water per year from the Yangtze River basin to Beijing and northern China across >1,000 km.

⚡ Precipitation intensity and extreme events

⚡ Record-breaking intensities

The excerpt provides a detailed example: Holt, Missouri, June 22, 1947:

  • A rainstorm dropped 305 mm (12 inches) in only 42 minutes
  • This set a world record for that duration, equivalent to 43.6 cm h⁻¹
  • This rate is "far greater than the rates at which water can enter, i.e. infiltrate, most soils"
  • The storm filled homes with up to 60 cm of water and mud

📊 World record precipitation intensities

DurationAmountLocationDate
1 minute38 mmBarot, Guadeloupe26 Nov 1970
1 hour401 mmShangdi, Inner Mongolia, China3 Jul 1975
1 day1,825 mmFoc Foc, La Réunion7-8 Jan 1966
1 month9,300 mmCherrapunji, IndiaJul 1861
1 year26,461 mmCherrapunji, IndiaAug 1860–Jul 1861
  • The excerpt notes these intensities are "equally incredible" and "far greater than the capacity of the soil to transmit that water."

🌊 Atmospheric water vapor transport

Total column water vapor (or total precipitable water): the depth of water that would result if all atmospheric water vapor above a given land area were condensed and deposited on the surface.

  • Annual average precipitable water ranges from near 60 mm over equatorial oceans and Amazonian rainforest to near 0 mm in mountain rain shadows.
  • Don't confuse: local atmospheric capacity with extreme rainfall totals—record events for durations greater than a few minutes "far exceed the amount of water that the atmosphere can hold in one location."
  • This proves that extreme rainfall depends on "strong horizontal transport of water vapor into the rainstorm from the surrounding atmosphere."
  • Example: Understanding that the atmosphere typically holds <60 mm of precipitable water "makes the record precipitation intensities all the more incredible."

💥 Kinetic energy of rainfall

  • For rainfall intensities of 10 cm h⁻¹, the kinetic energy of falling rain can approach 10,000 J m⁻² h⁻¹.
  • This substantial energy delivery validates Wollny's recognition of "the critical role of the vegetative canopy and plant residues in protecting the soil from the erosive energy of rainfall."

💧 Raindrop characteristics and behavior

💧 Typical size range

Raindrops typically have cross-sectional diameters between 1 and 4 mm.

  • The minimum diameter is approximately 0.5 mm because smaller droplets are generally kept aloft by air currents.
  • The size distribution shifts toward larger drops as rainfall intensity increases.
  • Example: In measurements at Ottawa, Canada, rainfall intensities of 1.0, 2.8, 6.3, and 23.0 mm h⁻¹ showed progressively larger drop size distributions.

🎈 Large drop instability and fragmentation

  • Although raindrops with equivalent diameters up to 9 mm have been observed, such large drops are generally unstable.
  • The process of disintegration:
    1. Drag forces and pressures flatten the drop as it falls
    2. The drop deforms into a short-lived, parachute-like shape
    3. The drop completely disintegrates
  • This fragmentation process may generate the entire distribution of raindrop sizes.
  • The excerpt notes that high-speed video footage has captured this striking process.

🌧️ Relationship to rainfall interception

  • Small raindrops are more effective than large drops in wetting vegetation or residue surfaces.
  • Small drops have less kinetic energy than large drops, increasing the likelihood of retention by vegetation or residue.
  • As a result, interception losses tend to increase as raindrop size decreases.
  • Don't confuse: smaller drops with less impact—while they have lower individual energy, they are more easily intercepted, affecting how much precipitation reaches the soil.

🌿 Historical foundation: Wollny's pioneering work

🌿 Early soil physics experiments

  • Ewald Wollny, a German scientist, became "perhaps the first scientist whose published work focused squarely on soil physics."
  • His experiments as early as 1874 were the first to quantitatively describe:
    • How raindrops falling on bare soil degraded soil structure
    • How vegetative canopy provided vital protection against raindrop impact

📏 Quantifying canopy interception

  • Wollny's subsequent experiments revealed that vegetative canopies of common crops could intercept 12–55% of total rainfall.
  • These discoveries, along with his studies on runoff and erosion, make Wollny "one of the pioneers of soil physics and hydrology and an early leader on the important issue of soil and water conservation."

🏛️ Context in soil physics history

  • The excerpt places Wollny shortly after Henry Darcy, the French engineer whose 1856 work on water filters and groundwater aquifers laid "one of the earliest cornerstones of soil physics."
  • The study of soil water balance processes begins "where Wollny, and arguably soil physics itself, began—by considering precipitation and the interactions between raindrops, vegetative canopies, and the soil."

🍃 Rainfall interception losses

🍃 Event size and interception efficiency

The excerpt describes a study under live oak mottes on the Edwards Plateau in central Texas:

  • Events totals >15 mm were required to ensure at least half of the precipitation reached the soil.
  • Even for the largest events (~80 mm), only about 80% of rainfall reached the soil.
  • The remainder was intercepted by plant canopy or leaf litter.

🌦️ Implications for small rainfall events

  • Regions receiving a relatively large portion of precipitation through small rainfall events will be susceptible to relatively large losses due to rainfall interception.
  • Don't confuse: total annual precipitation with effective precipitation reaching the soil—interception can substantially reduce the water available to the soil, especially where small, frequent events dominate.
25

Rainfall Interception

5.2. Rainfall Interception

🧭 Overview

🧠 One-sentence thesis

Rainfall interception by plant canopies and residue can prevent a substantial portion of precipitation from reaching the soil—sometimes more than half—thereby strongly affecting soil water availability in both natural and managed ecosystems.

📌 Key points (3–5)

  • What interception is: rainfall prevented from reaching soil because plant canopies or residue capture it and it subsequently evaporates; related processes include throughfall, stemflow, and canopy drip.
  • Vegetation characteristics matter: interception storage capacity varies by species due to differences in leaf shape, hairiness, growth habit, biomass, and litter accumulation.
  • Per-unit-mass vs per-unit-area: a species may intercept more rainfall per unit mass but less per unit area if its total biomass is lower—a common source of confusion when comparing species.
  • Rainfall event size is critical: small events can be completely intercepted, while larger events allow a higher percentage to reach the soil; regions with many small events lose more water to interception.
  • Magnitude across ecosystems: interception can account for 10–45% of annual rainfall in croplands and rangelands, and up to 46% in forests, making it a major component of the water balance.

🌿 What interception is and related processes

🌿 Rainfall interception definition

Rainfall interception: the process by which rainfall is prevented from reaching the soil because of capture by plant canopies or plant residue and subsequent evaporation.

  • Water is held on leaves, stems, branches, or surface litter and evaporates before it can infiltrate the soil.
  • This process reduces the amount of water available for soil moisture replenishment.
  • Similar interception occurs with sprinkler irrigation systems.

💧 Related water pathways

ProcessWhat happens
ThroughfallRain passes through the canopy and reaches the soil
StemflowRain striking the canopy flows down leaves and branches to the main stem, then down the stem to the soil
Canopy dripWater from rain or dew drips off the outer edges of the plant canopy
  • These processes redistribute water spatially rather than preventing it from reaching the soil.
  • Example: stemflow can deliver concentrated water near the tree trunk, potentially giving trees a competitive advantage.

🌾 How vegetation characteristics control interception

🌾 Interception storage capacity

Interception storage capacity: the maximum volume of rainfall per unit area that a particular plant canopy can hold.

  • This capacity determines how much rain must fall before water begins to reach the soil.
  • Different species have very different capacities due to their physical structure.

🌱 Case study: Edwards Plateau grasses

The excerpt compares two grass species in central Texas rangelands:

SpeciesPer-unit-mass interceptionPer-unit-area interceptionStorage capacityAnnual interception
CurlymesquiteHigher (flat, hairy leaves + stolons)Lower (less total biomass)1.0 mm10.8%
Sideoats gramaLowerHigher (more standing biomass)1.8 mm18.1%
  • Don't confuse: a plant that is more effective per unit mass may still intercept less rainfall overall if the site has lower total plant biomass.
  • The excerpt emphasizes that both the intrinsic effectiveness (per mass) and the total amount of vegetation (per area) matter.

🌳 Live oak mottes: extreme interception

Live oak clumps on the Edwards Plateau show much higher interception than grasses:

  • Dense canopy plus thick leaf litter on the soil surface.
  • Leaf litter holds approximately twice as much rainfall per unit mass as grasses.
  • Litter interception storage capacity: 8.7 mm.
  • Combined effect: canopy intercepts 25.4% of annual rainfall, litter intercepts another 20.7%, so only 53.9% of precipitation reaches the soil.

Example: In a year with 523 mm of rainfall, 133 mm is intercepted by the canopy, 108 mm by the litter, and only 282 mm reaches the mineral soil.

🌊 Stemflow and competitive advantage

  • Stemflow accounted for only 3.3% of annual precipitation in live oak mottes.
  • However, it delivered a volume equivalent to 222% of annual precipitation to the soil within a 10-cm radius of each tree trunk.
  • This concentrated water delivery may help trees establish a competitive advantage over other plant species.
  • Similar mechanisms have been suggested for juniper encroachment in rangelands.

🌍 Interception across ecosystems

The excerpt provides interception percentages from multiple studies:

Croplands:

  • Winter wheat: 33%
  • Maize: 30%
  • Soybean: 35%
  • Switchgrass: 25–31%
  • Forage sorghum: 27–45%
  • Center pivot irrigation on maize: ~8% of applied water

Natural ecosystems:

  • Eastern redcedar: 36%

  • Tallgrass prairie: 44%

  • Amazonian rainforest: 9%

  • Deciduous forest: 20–29%

  • These values show that interception is a major water balance component across diverse environments.

🌧️ How rainfall characteristics affect interception

🌧️ Rainfall event size is the strongest factor

The excerpt states that rainfall amount per event exerts the strongest influence on interception among the three characteristics (amount, intensity, duration).

Live oak motte example (Edwards Plateau):

  • Events <5 mm: completely intercepted by canopy and litter; no water reaches soil.
  • Events >15 mm: required to ensure at least half of precipitation reaches the soil.
  • Even the largest events (~80 mm): only about 80% of rainfall reaches the soil.

Implication for regional water balance:

  • Regions that receive a large portion of their precipitation through small rainfall events will lose relatively more water to interception.
  • Example: A climate with many 3–4 mm events will have higher interception losses than one with fewer, larger storms totaling the same annual rainfall.

💧 Visual pattern from the excerpt

The excerpt describes Figure 5-7, showing percent of precipitation intercepted versus event size:

  • Small events: nearly 100% intercepted.
  • As event size increases, the percentage intercepted decreases.
  • Even for large events, some interception still occurs (only ~80% reaches soil at 80 mm events).

🌂 Raindrop size distribution

  • Small raindrops are more effective than large drops at wetting vegetation or residue surfaces.
  • Small drops have less kinetic energy, so they are more likely to be retained by vegetation or residue rather than bouncing off or dripping through.
  • Result: interception losses tend to increase as raindrop size decreases.

🔬 Historical context and significance

🔬 Early soil physics research

  • Ewald Wollny, a German scientist, conducted experiments as early as 1874 that were the first to quantitatively describe how raindrops degrade bare soil structure and how vegetative canopy protects against raindrop impact.
  • His subsequent experiments revealed that vegetative canopies of common crops could intercept 12–55% of total rainfall.
  • These discoveries, along with his studies on runoff and erosion, make Wollny a pioneer of soil physics and hydrology and an early leader on soil and water conservation.

🌍 Why interception matters for water balance

  • Interception can substantially reduce soil water availability in rangelands, forests, and croplands.
  • It affects both natural and managed ecosystems.
  • Understanding interception is essential for predicting water availability, especially in water-limited environments where soil water is the limiting factor for vegetative productivity.
  • The magnitude of interception (10–46% of annual rainfall across various ecosystems) means it is a major component of the water balance that cannot be ignored.
26

Raindrop Impact

5.3. Raindrop Impact

🧭 Overview

🧠 One-sentence thesis

Raindrop impact degrades soil structure and delivers substantial kinetic energy to the land surface, making vegetative canopy protection vital for soil conservation.

📌 Key points (3–5)

  • Historical foundation: Ewald Wollny's 1874 experiments were the first to quantitatively show how raindrops degrade bare soil structure and how vegetation protects against raindrop impact.
  • Kinetic energy delivery: intense rainfall (e.g., 10 cm/h) can deliver nearly 10,000 J/m²/h of kinetic energy to the land surface.
  • Raindrop size matters: typical raindrops are 1–4 mm in diameter; larger drops carry more kinetic energy but are unstable and fragment.
  • Common confusion: raindrop size vs. rainfall intensity—as intensity increases, the size distribution shifts toward larger drops, which increases erosive energy.
  • Why it matters: understanding raindrop impact explains why vegetative canopies and plant residues are critical for protecting soil from erosion.

🌱 Historical discovery and significance

🌱 Wollny's pioneering work (1874)

  • Ewald Wollny, a German scientist, became perhaps the first researcher whose published work focused squarely on soil physics.
  • His experiments were the first to quantitatively describe two key processes:
    • How raindrops falling on bare soil degraded the soil structure.
    • How the vegetative canopy provided vital protection against raindrop impact.
  • Subsequent experiments revealed that vegetative canopies of common crops could intercept 12–55% of total rainfall.
  • These discoveries, along with his studies on runoff and erosion, make Wollny a pioneer of soil physics and hydrology and an early leader on soil and water conservation.

🏛️ Foundation of soil physics

  • Wollny's work began as early as 1874, shortly after Henry Darcy's 1856 work on water filters and groundwater aquifers.
  • The study of soil water balance processes begins where Wollny—and arguably soil physics itself—began: by considering precipitation and the interactions between raindrops, vegetative canopies, and the soil.

⚡ Kinetic energy of rainfall

⚡ Energy delivery to the land surface

  • Atmospheric processes can generate precipitation intensities far greater than the capacity of soil to transmit water.
  • The kinetic energy delivered to the land surface by intense rainfall events can be substantial.
  • Quantitative example: For rainfall intensities of 10 cm/h, the kinetic energy of falling rain can approach 10,000 J/m²/h.

🛡️ Why vegetation protection is critical

  • Wollny was correct in recognizing the critical role of the vegetative canopy and plant residues in protecting the soil from the erosive energy of rainfall.
  • Without this protection, the kinetic energy from raindrops directly impacts bare soil, degrading its structure and increasing erosion risk.
  • Don't confuse: interception (water captured by vegetation) vs. impact protection (energy absorption by vegetation)—both processes reduce soil degradation but work differently.

💧 Raindrop characteristics

💧 Typical raindrop sizes

Raindrops typically have cross-sectional diameters between 1 and 4 mm.

  • The minimum diameter of raindrops is approximately 0.5 mm because smaller droplets are generally kept aloft by air currents.
  • Although raindrops with equivalent diameters up to 9 mm have been observed, such large drops are generally unstable.

🔄 How large drops fragment

  • The drag forces and pressures acting on a large drop as it falls result in a sequence of deformations:
    1. Flattening of the drop.
    2. Deformation into a short-lived, parachute-like shape.
    3. Complete disintegration of the drop.
  • This process of fragmentation of large drops may generate the entire distribution of raindrop sizes.
  • Example: A 9 mm drop falling through the atmosphere will not remain intact; it will break apart into multiple smaller droplets.

📊 Size distribution and rainfall intensity

  • The raindrop size distribution shifts toward larger drops as the rainfall intensity increases.
  • Data from Ottawa, Canada, showed four different rainfall intensities (1.0, 2.8, 6.3, and 23.0 mm/h) with progressively larger average drop sizes.
Rainfall intensityDrop size trendImplication
Low (1.0 mm/h)Smaller drops dominateLower kinetic energy per drop
High (23.0 mm/h)Larger drops dominateHigher kinetic energy per drop
  • Don't confuse: more intense rainfall doesn't just mean more drops—it also means larger drops, which carry significantly more kinetic energy and erosive potential.

🔗 Connection to interception

🔗 Small vs. large drops and interception

  • Small raindrops are more effective than large drops in wetting the surfaces of vegetation or residue.
  • Small drops have less kinetic energy than large drops, which increases the likelihood of small drops being retained by the vegetation or residue.
  • As a result, interception losses tend to increase as the raindrop size decreases.

🧩 Interplay between size and impact

  • Smaller drops: easier to intercept, less erosive energy if they reach the soil.
  • Larger drops: harder to intercept, more erosive energy if they reach the soil.
  • This creates a trade-off: high-intensity storms with large drops are both harder to intercept and more damaging to exposed soil.
27

Crust Formation

5.4. Crust Formation

🧭 Overview

🧠 One-sentence thesis

The Horton infiltration model attributes the decline in infiltration capacity over time to flow-restricting changes in a thin surface layer caused by raindrop impact, soil swelling, and pore plugging.

📌 Key points (3–5)

  • What infiltration capacity means: the maximum rate at which a given soil can absorb rainfall under specified conditions.
  • Why infiltration rates decrease: Horton identified three surface-layer changes—packing by raindrop impact, soil swelling, and plugging of pores with fine materials.
  • The Horton equation: a mathematical model that describes infiltration capacity declining exponentially from an initial high rate to a constant minimum rate.
  • Model limitations: parameters are not clearly tied to measurable soil properties, and the conceptual framework misses the most fundamental reason for declining infiltration rates (which had been discovered earlier but was not widely accepted).
  • Common confusion: the Horton model is widely used and successful, but this may be due to its flexible mathematical form rather than the correctness of its underlying physical assumptions.

🌧️ Infiltration basics

💧 What infiltration is

Infiltration: the process by which water enters the soil profile.

  • It is the next step in the soil water balance after precipitation and interception.
  • The infiltration rate is the volume of water flowing into the soil per unit surface area per unit time.

📉 How infiltration rates change over time

  • Infiltration rates typically decrease over time during an infiltration event.
  • They approach a relatively constant, low rate if the event is prolonged.
  • Example: In a silty clay loam soil in Hawaii, infiltration rates initially declined sharply, then approached a fairly constant level during ponded infiltration.
  • Understanding why this decline happens was a major scientific mystery in early soil physics and hydrology.

🔬 Horton's conceptual framework

🧱 What Horton defined

Infiltration capacity: the maximum rate at which a given soil can absorb rainfall when the soil is in a specified condition.

  • Horton was a prominent early American hydrologist and engineer.
  • He studied the mystery of declining infiltration rates in detail.

🛠️ Three surface-layer changes

Horton deduced from field observations that the decrease in infiltration capacity was due to flow-restricting changes in a thin layer at the soil surface:

  1. Packing of the soil surface by raindrop impact: physical compaction from raindrops hitting the surface.
  2. Swelling of the soil: expansion of soil particles when wetted.
  3. Plugging of surface pores with fine materials: small particles clog the openings through which water enters.
  • All three mechanisms reduce the ability of water to enter the soil.
  • Don't confuse: these are surface phenomena, not changes throughout the entire soil profile.

📐 The Horton infiltration equation

🧮 Mathematical form

The Horton infiltration equation is:

  • f = infiltration capacity at time t
  • f_c = constant minimum value of infiltration capacity (the asymptotic lower limit)
  • f_0 = initial infiltration capacity at the start of the infiltration event
  • K_f = a constant controlling the rate at which infiltration capacity decreases over time
  • t = time

📊 What the equation describes

  • Infiltration capacity f starts at a high initial value f_0.
  • It decreases exponentially over time.
  • It approaches a constant minimum value f_c asymptotically (gets closer and closer but never quite reaches it).
  • The rate of decline is controlled by the constant K_f.

Example: If a soil has high initial infiltration capacity, raindrop impact and surface changes cause the capacity to drop quickly at first, then level off to a steady, lower rate.

⚠️ Strengths and weaknesses of the Horton model

✅ Why the model has been widely used

  • The Horton infiltration equation has been widely used in hydrology and related disciplines.
  • Its success may be due to:
    • The correctness of its underlying conceptual framework (the three surface-layer changes), or
    • The flexibility of its mathematical form (the exponential decay function fits many datasets well), or
    • Some combination of both.

❌ Main deficiencies

DeficiencyExplanation
Parameters not tied to measurable propertiesThe constants (f_c, f_0, K_f) are not clearly related to physical properties of the soil that can be measured independently.
Missing the fundamental reasonThe conceptual framework fails to include the most fundamental reason that infiltration rates decrease over time.
  • The fundamental discovery explaining declining infiltration had been made decades before Horton's work.
  • However, it apparently was not widely accepted in Horton's time.
  • Don't confuse: the Horton model is useful and widely applied, but it is not the most physically complete explanation of infiltration decline.
28

6.1. Horton Infiltration Model

6.1. Horton Infiltration Model

🧭 Overview

🧠 One-sentence thesis

Horton's infiltration model attributes the decline in infiltration capacity over time to surface-layer changes (packing, swelling, pore plugging), but it lacks connection to measurable soil properties and misses a more fundamental mechanism discovered earlier.

📌 Key points (3–5)

  • What Horton defined: infiltration capacity as the maximum rate at which soil can absorb rainfall under specified conditions.
  • Why capacity decreases: Horton attributed it to three surface changes—raindrop packing, soil swelling, and pore plugging with fine materials.
  • The Horton equation: describes infiltration capacity declining from an initial value toward a constant minimum, controlled by a decay constant.
  • Model strengths and weaknesses: widely used due to its conceptual framework or flexible math, but parameters are not tied to measurable soil properties and the model omits the most fundamental reason for infiltration decline.
  • Common confusion: Horton's surface-layer explanation was influential, yet a more fundamental discovery had been made decades earlier but was not widely accepted in his time.

🔬 Infiltration capacity concept

🔬 What infiltration capacity means

Infiltration capacity: "the maximum rate at which a given soil can absorb rainfall when the soil is in a specified condition."

  • This is not the actual infiltration rate at any moment, but the maximum possible rate the soil can handle.
  • Horton observed that this capacity decreases over time during an infiltration event.
  • The excerpt emphasizes that infiltration rates "typically decrease over time during the course of an infiltration event, approaching a relatively constant and low rate."

📉 Why capacity declines (Horton's view)

Horton deduced from field observations that the decrease was due to flow-restricting changes in a thin surface layer:

  1. Packing of the soil surface by raindrop impact – physical compaction reduces pore space.
  2. Swelling of the soil – expansion of soil particles narrows pathways.
  3. Plugging of surface pores with fine materials – small particles clog openings.
  • All three mechanisms occur at or near the soil surface.
  • Don't confuse: these are surface-layer effects, not changes deeper in the profile.

🧮 The Horton infiltration equation

🧮 Mathematical form

The Horton infiltration equation (Eq. 6-1) describes how infiltration capacity f changes over time t:

  • f: infiltration capacity at time t
  • f₀: initial infiltration capacity at the start of the event
  • f_c: constant minimum value that f approaches asymptotically (the long-term steady rate)
  • K_f: a constant controlling the rate at which capacity decreases

In words: infiltration capacity starts at an initial high value, then decays exponentially toward a constant minimum.

📊 Why the model succeeded

ReasonWhat the excerpt says
Conceptual frameworkMay be due to correctness of the surface-layer explanation
Mathematical flexibilityMay be due to the flexible form of the equation
CombinationLikely some combination of both
  • The model "has been widely used in hydrology and related disciplines."
  • Example: if a soil starts with high capacity but surface pores clog quickly, the equation can fit the observed decline by adjusting K_f.

⚠️ Deficiencies of the Horton model

⚠️ Parameters not measurable

  • Main deficiency 1: "the parameters are not clearly related to measurable physical properties of the soil."
  • This means you cannot easily predict f₀, f_c, or K_f from standard soil measurements (texture, structure, etc.).
  • The model requires calibration from observed infiltration data rather than prediction from soil properties.

⚠️ Missing the fundamental mechanism

  • Main deficiency 2: "the model's conceptual framework fails to include the most fundamental reason that infiltration rates decrease over time."
  • The excerpt notes that "that fundamental discovery had, in fact, been made decades before but apparently was not widely accepted in Horton's time."
  • Don't confuse: Horton's surface-layer changes (packing, swelling, plugging) are real effects, but they are not the most fundamental cause of infiltration decline.
  • The excerpt does not reveal what that fundamental reason is in this section, only that Horton's model omits it.

🔍 Context and broader infiltration concepts

🔍 What infiltration rate means

Infiltration rate: the volume of water flowing into the soil per unit of surface area per unit time.

  • This is the actual rate at any moment, not the maximum capacity.
  • The video and Figure 6-1 show that infiltration rates "initially declined sharply then approached a fairly constant level during ponded infiltration."
  • Example: in a silty clay loam in Hawaii, the rate dropped quickly at first, then leveled off.

🕵️ The early mystery

  • "The fundamental cause of this decrease in infiltration rates over time was one of the main scientific mysteries in the early days of soil physics and hydrology."
  • Robert Horton, "one of the most prominent early hydrologists," studied this mystery in detail.
  • His model was an influential attempt to explain the phenomenon, even though it was incomplete.
29

6.2. Green-Ampt Infiltration Model

6.2. Green-Ampt Infiltration Model

🧭 Overview

🧠 One-sentence thesis

The Green-Ampt model explains why infiltration rates decrease over time by showing that the hydraulic gradient driving infiltration diminishes as the wetting front moves deeper into the soil.

📌 Key points (3–5)

  • Core mechanism: infiltration rate depends on hydraulic conductivity and the hydraulic gradient between the soil surface and the wetting front.
  • Why infiltration slows: as the wetting front moves farther from the surface, the hydraulic gradient decreases, reducing infiltration rate.
  • Key assumptions: homogeneous soil, uniform initial water content, constant pressure head at the wetting front, and a distinct wetting front.
  • Common confusion: the model works best for coarse-textured soils with sharp wetting fronts; it is less accurate for fine-textured soils or when air-entrapment, crusting, or swelling occur.
  • Fundamental tendency: early in infiltration, wetting front position and cumulative infiltration are approximately proportional to the square root of time.

🏗️ Model foundation and history

📜 Origins

  • Developed in the early 1900s by Heber Green and G.A. Ampt in Australia while studying drainage and soil water flow.
  • Built on early soil physics literature, including Edgar Buckingham's work on unsaturated soil flow (section 4.7).
  • Published in 1911 with physically-based equations for downward, upward, and horizontal infiltration.

🎯 Purpose

  • Provides a simplistic but elegant approximation for infiltration when water is ponded on the soil surface.
  • Represents a specialized application of the Buckingham-Darcy law.

🧮 Model equations and structure

📐 Infiltration rate equation

The Green-Ampt model for downward infiltration: infiltration rate (i) equals hydraulic conductivity (K) multiplied by the ratio of (pressure head at soil surface minus pressure head at wetting front plus distance to wetting front) divided by (distance to wetting front).

  • i = infiltration rate
  • K = hydraulic conductivity of the soil
  • L_f = distance (or length) from soil surface to wetting front
  • H_0 = pressure head at soil surface
  • H_f = pressure head at wetting front
  • The ratio on the right side represents the hydraulic gradient.

📊 Cumulative infiltration equation

Cumulative infiltration (I) equals the change in water content (Δθ) multiplied by the distance to the wetting front (L_f).

  • I = cumulative infiltration
  • Δθ = difference between final water content (θ_f) and initial water content (θ_i)
  • L_f = distance to wetting front

⏱️ Time-dependent equation

  • A third equation relates wetting front position to time, but it must be solved by trial-and-error (implicitly) rather than directly.
  • This makes practical application more challenging.

🔧 Model assumptions and limitations

✅ Required assumptions

The model assumes:

  • Homogeneous soil with uniform initial water content (θ_i)
  • Constant pressure head at the wetting front (H_f)
  • Uniform and constant water content and hydraulic conductivity (K) in the wetted region
  • Constant ponding depth (H_0)

🎯 When the model works best

ConditionWhy it matters
Coarse-textured soilsMore likely to have a distinct, sharp wetting front
Initially dry soilSharp wetting front more likely than in initially wet soil
Homogeneous soil textureUniform properties throughout wetted region
Minimal complicationsAir-entrapment, surface crusting, and soil swelling do not substantially influence infiltration

⚠️ When the model is less accurate

  • Fine-textured soils tend to have diffuse rather than sharp wetting fronts.
  • Initially wet soil may not develop a distinct wetting front.
  • Non-homogeneous soil texture violates uniformity assumptions.
  • Air-entrapment, surface crusting, or soil swelling can alter the infiltration process in ways the model does not capture.

Don't confuse: the model's simplicity is both its strength (easy to understand fundamental mechanisms) and its limitation (less accurate under complex real-world conditions).

💡 Key insight: why infiltration slows

🔽 The hydraulic gradient mechanism

  • The power of the Green-Ampt model lies in succinctly describing the fundamental reason infiltration rates decrease over time.
  • Since ponding depth (H_0) and pressure potential at the wetting front (H_f) are constants, the hydraulic gradient gets smaller as L_f gets larger.
  • In plain language: as the wetting front moves farther from the soil surface, the "pull" driving water downward weakens.

🧪 Experimental evidence

Laboratory experiments with fine sand columns confirm the model's predictions:

  • The wetting front reached 10 cm depth in about 0.05 hours (3 minutes).
  • It took about four times as long (0.2 hours or 12 minutes) to reach 20 cm depth.
  • This demonstrates that the rate of advance of the wetting front decreases over time.

📏 Square root relationship

During the early part of an infiltration event, the wetting front position is approximately proportional to the square root of the event duration.

  • Cumulative infiltration follows this same proportionality.
  • The Green-Ampt model captures this behavior.
  • Example: if it takes 3 minutes to infiltrate to 10 cm, it will take roughly 12 minutes (4 times as long) to reach 20 cm (twice as deep), because 4 is the square of 2.

🌟 Practical value

🔬 Understanding and prediction

Despite its assumptions and caveats, the Green-Ampt model has proven extremely useful for:

  • Helping understand the fundamental physics of infiltration
  • Predicting infiltration behavior under ponded conditions
  • Providing a framework for more complex models

🔗 Connection to broader theory

The model can be viewed as a specialized application of the Buckingham-Darcy law, linking it to fundamental principles of water flow in unsaturated soil.

30

6.3. Infiltration for a Constant Rainfall Rate

6.3. Infiltration for a Constant Rainfall Rate

🧭 Overview

🧠 One-sentence thesis

The Mein and Larson modification of the Green-Ampt model predicts infiltration under constant rainfall by distinguishing between an initial supply-controlled phase and a later soil-controlled phase that begins when the soil's infiltration capacity drops below the rainfall rate and ponding starts.

📌 Key points (3–5)

  • Limitation of basic Green-Ampt: the original model only works for ponded infiltration, not for rainfall at rates below the soil's infiltration capacity.
  • Two phases of infiltration: initially all rain infiltrates (supply controlled), then after ponding begins, the soil limits infiltration (soil controlled).
  • Time to ponding: Mein and Larson provide a three-step method to estimate when ponding begins, which is the earliest moment runoff can start.
  • Common confusion: infiltration capacity vs rainfall rate—if rainfall rate is below initial infiltration capacity but above saturated hydraulic conductivity, ponding will eventually occur as infiltration capacity decreases over time.
  • Why it matters: only after ponding begins can runoff occur, so predicting the time to ponding is essential for understanding surface water accumulation.

🌧️ The problem and the breakthrough

🌧️ What the Green-Ampt model could not do

  • The Green-Ampt model (described in the prior section) is valid only for ponded infiltration—when water is already standing on the surface.
  • It does not predict infiltration when water arrives at a rate below the soil's infiltration capacity (e.g., light to moderate rainfall or sprinkler irrigation).
  • This gap left engineers without a way to model realistic rainfall scenarios.

🔬 The Mein and Larson modification (1971)

  • Engineers Russell Mein and Curtis Larson at the University of Minnesota extended the Green-Ampt model to handle constant-rate rainfall.
  • Their key insight: if rain arrives at a constant rate below the soil's initial infiltration capacity but above the soil's saturated hydraulic conductivity, the infiltration process has two distinct stages.

🔄 Two phases of infiltration

🔄 Phase 1: Supply controlled

  • When: from the start of rainfall until ponding begins.
  • What happens: all water reaching the surface infiltrates immediately.
  • Why: the soil's infiltration capacity is still higher than the rainfall rate.
  • Limiting factor: the rate of water delivery to the surface (the rainfall rate itself).

Supply controlled: the rate of water delivery to the surface limits the infiltration rate.

Example: If rain falls at 1.5 cm/h and the soil can initially absorb 3 cm/h, all 1.5 cm/h will infiltrate.

🔄 Phase 2: Soil controlled

  • When: after ponding begins.
  • What happens: water accumulates on the surface; infiltration rate is now determined by the soil's properties, not the rainfall rate.
  • Why: the soil's infiltration capacity has decreased over time and fallen below the rainfall rate.
  • Limiting factor: the rate of water flow in the soil.

Soil controlled: the rate of water flow in the soil limits the infiltration rate.

  • Important: only after this change point is reached can runoff begin.

📊 Visual summary

The excerpt references Fig. 6-4, which shows:

  • Curves of infiltration rate versus time for different constant rainfall rates.
  • Curve 1 (greatest rainfall rate) through Curve 4 (least rainfall rate).
  • Curve 4 represents a rainfall rate below the soil's saturated hydraulic conductivity (no ponding occurs).
  • Each curve has a horizontal segment (supply-controlled phase) followed by a declining segment (soil-controlled phase).
  • The end of the horizontal segment marks the time to ponding (t_p).

⏱️ Estimating time to ponding

⏱️ The three-step method

Mein and Larson provide a simple procedure to estimate when ponding begins:

  1. Step 1: Find the wetting front depth at ponding

    • Ponding begins when the rainfall rate (r) equals the soil's infiltration capacity (i).
    • Substitute r for i in the Green-Ampt ponded infiltration equation (Eq. 6-2), assume ponding depth H_0 = 0, and solve for the wetting front depth L_f.
    • The result is Eq. 6-5, which defines the depth of the wetting front when ponding begins.
  2. Step 2: Calculate cumulative infiltration at ponding

    • Insert the L_f value from Step 1 into Eq. 6-3 to estimate the cumulative infiltration (I) that has occurred before ponding.
  3. Step 3: Calculate time to ponding

    • Divide the cumulative infiltration (I) by the rainfall rate (r) to get the time required for ponding to occur (t_p).
    • Formula: t_p = I / r (Eq. 6-6).

⚠️ Key condition

  • This method applies when the rainfall rate is:
    • Below the soil's initial infiltration capacity (so water initially infiltrates fully), and
    • Above the soil's saturated hydraulic conductivity (so infiltration capacity will eventually drop below the rainfall rate).
  • If the rainfall rate is below the saturated hydraulic conductivity (like Curve 4 in Fig. 6-4), ponding never occurs.

🧪 Application in practice

🧪 Problem set context

The excerpt includes practice problems that apply the Mein and Larson method:

  • Given soil properties (initial water content, wetted water content, wetting front pressure head, hydraulic conductivity) and a constant rainfall rate.
  • Questions ask for:
    • Depth of the wetting front when ponding begins.
    • Cumulative infiltration by that time.
    • Time required before ponding occurs.
  • One problem varies hydraulic conductivity to show how soil properties affect ponding time.

🧪 Why this matters

  • Predicting the time to ponding is essential for:
    • Runoff prediction: runoff cannot begin until ponding starts.
    • Irrigation design: knowing when water will accumulate helps optimize application rates.
    • Flood forecasting: understanding when and where surface water will form during storms.
31

Infiltration Measurements

6.4. Infiltration Measurements

🧭 Overview

🧠 One-sentence thesis

Infiltration can be measured using several types of infiltrometers—each applying water differently (ponded, tension, or sprinkled)—or by measuring runoff under natural rainfall, with each method having distinct advantages and limitations for estimating soil hydraulic properties.

📌 Key points (3–5)

  • What infiltrometers do: devices that apply water to soil under controlled conditions to measure infiltration.
  • Ring infiltrometers: apply water at positive pressure (ponded), either constant head or falling head; double-ring design reduces lateral flow effects compared to single-ring.
  • Tension infiltrometers: apply water at slightly negative pressure (-1 to -10 cm), are portable, require less water and time, but may yield lower hydraulic conductivity estimates.
  • Common confusion: single-ring vs. double-ring—single-ring has lateral flow that must be accounted for; double-ring's outer ring minimizes lateral flow from the inner ring so data can be analyzed as one-dimensional flow.
  • Field-scale measurement: runoff plots measure infiltration under natural rainfall by calculating the difference between precipitation and runoff.

🔬 Ring infiltrometers

🔬 How ring infiltrometers work

Infiltrometers: devices used to measure infiltration under controlled conditions.

  • Simple metal rings are pushed or pounded into the soil surface.
  • Water is applied inside the rings in one of two ways:
    • Constant head: depth held constant or nearly constant by continually refilling as water infiltrates.
    • Falling head: ponded water depth is allowed to decrease over time as water infiltrates.

🔄 Single-ring vs. double-ring design

FeatureSingle-ringDouble-ring
Flow patternWater flows laterally and vertically away from ringOuter ring reduces lateral flow from inner ring
Analysis complexityMust account for lateral flow effectsInner ring data can be analyzed as one-dimensional flow
Hydraulic conductivity estimatesSimilar to double-ring when proper procedures appliedSimilar to single-ring when proper procedures applied
  • Don't confuse: The outer ring in a double-ring setup is not measured—it exists only to control the flow pattern from the inner ring.
  • The excerpt states that both methods "provide similar estimates of hydraulic conductivity" when proper field procedures and analytical techniques are applied.

💧 Tension infiltrometers

💧 How tension infiltrometers differ

  • Apply water under slightly negative pressures, often between -1 and -10 cm of water.
  • This contrasts with ring infiltrometers, which apply water at positive pressures (ponded conditions).

✅ Advantages of tension infiltrometers

  • Portable: easier to move and set up.
  • Shorter measurement times: require relatively short durations.
  • Smaller water volumes: require less water than single- or double-ring infiltrometers.
  • Less variability: may result in less variable hydraulic conductivity estimates than other methods.

⚠️ Limitation

  • Hydraulic conductivity estimates are "in some cases, lower than those from other methods."
  • Example: If a ring infiltrometer estimates conductivity at 0.65 cm/h, a tension infiltrometer on the same soil might yield a lower value.

🌧️ Sprinkle infiltrometers and runoff plots

🌧️ Sprinkle infiltrometers

  • Apply water to soil as drops from a collection of capillary tubes suspended a few centimeters above the surface.
  • Water delivery rate is controlled by adjusting pressure inside a sealed water supply reservoir.
  • Measurement process:
    1. Water application continues until runoff is generated beneath the infiltrometer.
    2. Runoff is collected and measured.
    3. Infiltration rate = water delivery rate − runoff rate.

📏 Runoff plots for field conditions

Runoff plots: small areas of land surrounded by borders to keep out surface runoff from adjacent areas, with equipment to collect and measure runoff on the downhill side.

  • Used to measure infiltration under natural rainfall conditions.
  • Infiltration = precipitation − runoff.
  • Example from the excerpt: A runoff plot at an open-cut coal mine site in Australia measured infiltration and runoff. When coal mine spoils lacked vegetation, they were susceptible to raindrop impact and crust formation, which resulted in low infiltration and high runoff volumes.
  • Why borders matter: They prevent surface runoff from adjacent areas from entering the plot, ensuring measurements reflect only the plot's own infiltration and runoff.

🧮 Green-Ampt model context

🧮 Historical background

  • Developed in the early 1900s by Heber Green and G.A. Ampt in Australia.
  • Built on early soil physics literature, including Edgar Buckingham's work on flow in unsaturated soil.
  • Published in 1911 with equations for downward, upward, and horizontal infiltration.

🧮 Model characteristics

  • Describes infiltration when water is ponded on the soil surface.
  • The model is a "simplistic but elegant approximation."
  • Core insight: Infiltration rates tend to decrease over time because the hydraulic gradient driving infiltration decreases over time.
  • Can be viewed as a specialized application of the Buckingham-Darcy law.

🧮 Key assumptions

  • Homogeneous soil with uniform initial water content.
  • Pressure head at the wetting front is constant.
  • Water content and hydraulic conductivity are uniform and constant in the wetted region.
  • Ponding depth is constant.

🧮 When the model works best

  • When a relatively sharp or distinct wetting front exists throughout infiltration.
  • More likely in coarse-textured soils than fine-textured soils.
  • More likely in initially dry soil than initially wet soil.
  • When soil texture is homogeneous throughout the wetted region.
  • When air-entrapment, surface crusting, and soil swelling do not substantially influence infiltration.

🧮 Model equations

The excerpt presents three equations:

  1. Infiltration rate equation (Eq. 6.2): Infiltration rate equals hydraulic conductivity times a term involving the distance to the wetting front, pressure head at the surface, and pressure head at the wetting front.

  2. Cumulative infiltration equation (Eq. 6-3): Cumulative infiltration equals the difference between final and initial water content times the distance to the wetting front.

  3. Implicit time equation (Eq. 6-4): An equation that must be solved by trial-and-error to calculate how the wetting front position changes over time.

  • Challenge: There is no direct (explicit) way to calculate how the wetting front position changes over time; it requires trial-and-error (implicit) solution.
32

Overview of Runoff and Water Erosion Processes

7.1. Overview of Runoff and Water Erosion Processes

🧭 Overview

🧠 One-sentence thesis

Runoff and water erosion involve a sequence of processes—excess water generation, surface storage, detachment, transport, and deposition—that can cause harmful soil loss on cropland but may also be harnessed for beneficial water harvesting in water-scarce regions.

📌 Key points (3–5)

  • Two mechanisms for excess water: infiltration-excess (delivery rate exceeds infiltration) vs saturation-excess (soil is saturated and cannot accept more water).
  • Three stages of erosion: detachment (by raindrop impact, aggregate breakdown, or runoff scouring), transport (sheet, rill, or gully erosion), and deposition (when flow velocity decreases).
  • Common confusion—sheet vs rill vs gully erosion: sheet erosion is uniform and often unnoticed; rill erosion forms small shallow channels that can be tilled; gully erosion creates deep channels that cannot be crossed or removed by tillage.
  • Particle behavior paradox: fine sand (0.2–0.5 mm) is most easily detached, but clay particles, once suspended, remain in suspension even at very low flow velocities.
  • Why it matters: runoff reduces crop water availability, erodes fertile topsoil, pollutes surface water, but can also be captured for irrigation; erosion models help predict and manage these impacts.

💧 How excess water and runoff are generated

💧 Two pathways to excess water

The excerpt identifies two conceptual reasons for excess water on the soil surface:

Infiltration-excess overland flow: the soil is unsaturated but the rate of water delivery to the surface exceeds the infiltration rate into the soil.

Saturation-excess overland flow: the soil is saturated (or satiated) and cannot accept any more water.

  • Both pathways lead to water accumulating on the surface, but the underlying cause differs.
  • Don't confuse: infiltration-excess happens even when the soil is not fully saturated; saturation-excess requires the soil to be completely saturated.

🛑 Surface storage capacity delays runoff

  • Even after excess water is generated, runoff does not begin immediately.
  • The soil surface is never perfectly smooth; roughness creates a finite surface storage capacity—the volume of excess water per unit area that can be retained on the surface.
  • Runoff only begins once this storage capacity is exceeded.
  • Example: small depressions and bumps on a field hold water before it starts flowing downhill.

🌊 The three stages of water erosion

🔨 Detachment: how soil particles break free

Detachment: the separation of soil particles from the bulk soil body.

Three mechanisms cause detachment:

  • Raindrop impact: raindrops strike the soil surface and dislodge particles.
  • Aggregate breakdown upon wetting: soil aggregates disintegrate when wetted.
  • Scouring force of surface runoff: flowing water exerts force that detaches particles.

What controls detachment rate:

  • Degree of surface cover (vegetation, plant residues, or other protective covers)
  • Soil strength (stronger soil resists detachment)
  • Rainfall intensity (higher intensity increases raindrop impact)
  • Velocity of surface runoff (faster flow exerts more scouring force)

🚚 Transport: patterns of water and sediment movement

Once detached, soil particles (sediment) are transported across the surface. The excerpt describes three distinct patterns:

Erosion typeCharacteristicsPractical implications
Sheet erosionWater flow and soil erosion distributed relatively uniformly across the surfaceInsidious—can go unnoticed for years; the excerpt describes a case where >30 cm of topsoil was lost, effectively destroying productive capacity
Rill erosionWater and sediment concentrate in small, shallow channelsRills can be removed using tillage and are easily crossed with field equipment
Gully erosionRills deepen and widen to form gulliesGullies cannot be removed by tillage and cannot be easily crossed with equipment
  • Don't confuse rills and gullies: the key distinction is whether tillage can remove them and whether equipment can cross them.

🏔️ Gravity-driven erosion on slopes

  • Other types of water-related erosion are driven by gravity acting on wet soil along hillslopes, gullies, and streambanks.
  • When soil becomes thoroughly wetted:
    • The weight of the soil body increases
    • Soil strength decreases
    • Risk of gradual soil creep, sudden slumps, and potentially devastating landslides increases
  • Example: the excerpt mentions a 7.6 magnitude earthquake in El Salvador on January 13, 2001, that triggered a massive landslide killing ~585 people; wet soil was thought to be a key contributing factor.

📦 Deposition: where sediment settles

Sediment deposition: the final stage of the erosion process, typically initiated by a decrease in flow velocity.

  • Deposition is a major issue affecting streams, reservoirs, and coastal areas.
  • It is also one of the primary drivers of dramatic spatial variability in alluvial soils.
  • Sediment control is an important management concern in agriculture, construction, and engineering.
  • The excerpt notes that deposition can be approximated using Stokes' Law (a formula from earlier in the text).

📊 The Hjulström-Sundborg diagram: particle behavior in flowing water

📊 What the diagram shows

Hjulström-Sundborg diagram: a diagram summarizing the behavior of particles in a stream or river, showing relationships between particle size and the tendency to be eroded, transported, or deposited at different current velocities.

  • Based on work by researchers at the University of Uppsala in Sweden.
  • Both axes have a logarithmic scale.
  • The diagram depicts the three stages of water erosion (detachment/erosion, transport, deposition) as a function of flow velocity for different particle sizes.

🔍 Key insights from the diagram

Sand-sized particles (1 mm):

  • Detachment requires roughly 20 cm/s flow velocity.
  • Once detached, the particle remains in suspension and is transported until flow velocity drops below 10 cm/s.
  • This means transport can continue at lower velocities than required for detachment.

Most erodible particle size:

  • Sand-size particles in the range 0.2–0.5 mm (fine to medium sand) have the lowest detachment velocity and are thus most erodible.
  • Why? The excerpt explains that for smaller particles (silt and clay), attractive and adhesive forces between particles increase as particle size decreases, making them harder to detach.

Clay-sized particles:

  • Once detached and suspended, clay particles will not be deposited but will remain in suspension indefinitely, even at flow velocities as low as 0.1 cm/s.
  • This explains why clay can travel long distances in water and why streams remain turbid long after flow slows.

⚠️ Paradox to remember

  • Don't confuse detachment ease with transport behavior: fine sand is easiest to detach, but clay, once detached, stays suspended longest.
  • Larger particles settle quickly when flow slows; clay does not.

🧮 Predicting runoff: the Curve Number Method

🧮 Core hypothesis

The Curve Number Method is an empirical model developed by the USDA Soil Conservation Service (now NRCS) in the 1950s.

Core hypothesis: The ratio of actual precipitation retained by the landscape during a rainfall event (F) to the potential maximum retention (S) equals the ratio of runoff (Q) to total precipitation minus initial abstraction (P – I_a).

In words: F divided by S equals Q divided by (P minus I_a).

🔢 How the method works

  • F: amount of precipitation retained during the event
  • S: potential maximum retention for that landscape
  • Q: amount of runoff
  • P: total precipitation amount
  • I_a: initial abstraction (water that does not contribute to runoff or retention, typically estimated as 20% of S, i.e., I_a = 0.2S)

The excerpt provides a rearranged formula (Eq. 7-2) that is only valid for P > I_a.

Curve Number (CN):

  • An empirical number between 0 and 100 selected based on the hydrologic characteristics of the landscape.
  • Used to calculate S (in inches) via a formula (Eq. 7-3).
  • Curve numbers for a wide variety of circumstances are available (the excerpt provides a link).
  • Units of S, P, I_a, and Q must be the same; unit conversion can be applied as needed.

📈 Adjustments and refinements

  • Procedures exist for adjusting CN values based on initial soil moisture conditions:
    • Lower values for drier conditions
    • Higher values for wetter conditions
  • Research has indicated that setting the initial abstraction to 5% rather than 20% of S results in more accurate runoff predictions, but this change also requires revising the existing CN values.

🆚 Empirical vs mechanistic models

Model typeCurve Number Method (empirical)Mechanistic models (e.g., WEPP)
ApproachBased on empirical relationships and a simple hypothesisSimulate underlying mechanisms of soil water balance and overland flow
Data requirementsCurve number based on landscape characteristicsDetailed information about weather, rainfall patterns, soil properties, topography, land use
What it predictsRunoff amountRunoff and associated soil erosion; can include spatial information and simulate overland flow convergence
ExampleCurve Number MethodUSDA Water Erosion Prediction Project (WEPP) model
  • The excerpt mentions that the WEPP model development was led by the National Soil Erosion Research Laboratory of the USDA Agricultural Research Service, with over 200 contributors.

🌍 Why runoff and erosion matter: onsite and offsite effects

❌ Harmful effects on cropland

Runoff from cropland is typically undesirable because it results in:

  • Reduced water availability for the crop: water leaves the field instead of infiltrating.
  • Erosion of fertile topsoil: the most productive soil layer is lost.
  • Pollution of surface water bodies: sediment, phosphorus, and other contaminants are carried into streams, lakes, and rivers.

Example: the excerpt describes runoff and water erosion from a corn (maize) field in Iowa, USA, with harmful onsite and offsite effects.

✅ Beneficial uses: water harvesting

In contrast, runoff from uncultivated or impervious areas can sometimes be beneficial when captured and stored for later use.

Water harvesting strategies: capturing and storing runoff for later use.

  • May be key solutions to meeting critical water challenges in:
    • Developing regions such as sub-Saharan Africa
    • Major metropolitan areas in arid or semi-arid regions of developed nations (e.g., Sydney, Australia)
  • Example: the excerpt describes a runoff-harvesting pit in Uganda used to irrigate bananas, cassava, corn, and vegetables.

🌊 Influence on surface water

  • Runoff is one of the main sources of water to surface water bodies.
  • Any soil or water management practices that influence runoff are likely to influence surface water quantity and quality.
  • This connection makes runoff management critical for both agricultural productivity and environmental protection.
33

Runoff and Erosion Models

7.2. Runoff and Erosion Models

🧭 Overview

🧠 One-sentence thesis

Scientists have developed both simple empirical models (like the Curve Number Method) and complex mechanistic models (like WEPP) to predict runoff and erosion, with the choice depending on data availability and the need for detailed process simulation.

📌 Key points (3–5)

  • Two modeling approaches exist: empirical models use observed relationships (Curve Number Method), while mechanistic models simulate underlying physical processes (WEPP model).
  • Curve Number Method core idea: the ratio of actual retained precipitation to potential retention equals the ratio of runoff to effective precipitation.
  • WEPP model complexity: requires detailed weather, soil, topography, and land use data to simulate infiltration and overland flow processes.
  • Common confusion: empirical models are simpler and faster but less detailed; mechanistic models are more accurate but require extensive input data and computational resources.
  • Initial abstraction debate: research suggests 5% (instead of the traditional 20%) of potential retention gives more accurate predictions, but requires revising existing curve numbers.

📊 Empirical approach: Curve Number Method

📐 Core hypothesis

The Curve Number Method is based on the idea that the ratio between actual precipitation retained (F) and potential maximum retention (S) equals the ratio between runoff (Q) and total precipitation minus initial abstraction (P - I_a).

  • Developed by USDA Soil Conservation Service (now NRCS) in the 1950s.
  • The method expresses this as: F/S = Q/(P - I_a).
  • By substituting F = P - I_a - Q and rearranging, runoff can be calculated directly from precipitation and landscape characteristics.

🔢 Key parameters

ParameterMeaningHow it's determined
SPotential maximum retentionCalculated from curve number (CN): S = (1000/CN) - 10 (in inches)
CNCurve number (0-100)Selected based on hydrologic characteristics of the landscape
I_aInitial abstractionTypically estimated as 20% of S (i.e., I_a = 0.2S)
QRunoff amountCalculated only when P > I_a
  • All units (S, P, I_a, Q) must be the same; unit conversion applied as needed.
  • Curve numbers for various circumstances are available in reference tables.

🌧️ How rainfall becomes runoff

  • The relationship between rainfall and runoff varies with CN values.
  • Higher CN values produce more runoff for the same rainfall amount.
  • The method only predicts runoff when precipitation exceeds the initial abstraction (P > I_a).
  • Example: A landscape with CN = 91 (bare soil, poor condition) generates more runoff than one with CN = 81 (small grain, contour farmed) for the same rainfall event.

🔧 Adjustments and refinements

  • Soil moisture adjustments: CN values can be adjusted based on initial soil moisture conditions.
    • Lower values for drier conditions.
    • Higher values for wetter conditions.
  • Initial abstraction revision: Research indicates setting I_a to 5% (rather than 20%) of S results in more accurate runoff predictions, but this change requires revising existing CN values.
  • Don't confuse: changing the initial abstraction percentage affects accuracy but also requires recalibrating all existing curve numbers.

🔬 Mechanistic approach: Process-based models

⚙️ What mechanistic models simulate

  • In contrast to empirical methods, mechanistic models predict runoff by simulating the underlying mechanisms of soil water balance and overland flow processes.
  • These models simulate the dynamics of:
    • Rainfall interception
    • Infiltration
    • Overland flow as runoff from different landscape portions converges toward a stream
  • Runoff is predicted from these simulated processes rather than from empirical relationships.

📋 Data requirements

Mechanistic models typically require detailed information about:

  • Weather and rainfall patterns

  • Soil properties

  • Topography

  • Land use

  • Many models include spatial information to simulate how water moves across the landscape.

  • Example: A mechanistic model tracks how runoff from an upper slope area flows downhill and combines with runoff from adjacent areas before reaching a stream.

🌊 WEPP model specifics

🏗️ Model development and testing

  • WEPP = USDA Water Erosion Prediction Project model
  • Led by the National Soil Erosion Research Laboratory of the USDA Agricultural Research Service.
  • Over 200 people contributed to development.
  • The team conducted rainfall simulation experiments at over 50 sites to generate data for developing and testing the model.
  • One widely used mechanistic model capable of predicting both runoff and associated soil erosion.

💧 How WEPP simulates infiltration

  • The model simulates infiltration using the Green-Ampt approach (similar to methods described earlier in the source material).
  • Soil hydraulic conductivity is a key parameter in the Green-Ampt approach.
  • In WEPP, hydraulic conductivity is influenced by:
    • Tillage
    • Crusting
    • Surface cover
    • Storm precipitation amount
  • Runoff is predicted based on surface water excess (water that cannot infiltrate).

🖥️ Accessibility vs complexity

  • Despite its complexity, the WEPP model can be set up and run relatively easily using online versions.
  • Don't confuse: "complex" refers to the underlying science and data requirements, not necessarily the user interface—online tools make it accessible even though the model itself is sophisticated.
  • Example workflow (from problem set): users can select a watershed, build channel networks, specify climate and land use, run 10-year simulations, and view erosion maps and discharge summaries.

🔀 Comparing the two approaches

AspectEmpirical (Curve Number)Mechanistic (WEPP)
ComplexityRelatively simpleHighly complex
Data needsCurve number, precipitationWeather, soil properties, topography, land use
BasisObserved empirical relationshipsSimulation of physical processes
SpeedFast calculationsMay take several minutes to run
OutputsRunoff amountRunoff and soil erosion
AdjustabilityCN adjustments for soil moistureDynamic adjustments for tillage, crusting, cover, precipitation
  • Both approaches are widely used; choice depends on available data, required detail, and computational resources.
  • Empirical models are suitable when quick estimates are needed with limited data.
  • Mechanistic models are preferred when detailed process understanding and spatial variability are important.
34

8.1. Redistribution Within Partially Wetted Profiles

8.1. Redistribution Within Partially Wetted Profiles

🧭 Overview

🧠 One-sentence thesis

After infiltration ends, water continues to flow from the wetted upper soil layers to drier deeper layers, and this redistribution slows over time due to decreasing matric potential gradients, falling hydraulic conductivity, and hysteresis effects that help retain water near the surface.

📌 Key points (3–5)

  • What redistribution is: water flow from the infiltration-wetted zone (higher water potential) to drier soil below (lower water potential) after infiltration stops.
  • Why redistribution slows over time: three mechanisms work together—matric potential gradient decreases, hydraulic conductivity in the wetted zone drops as it dries, and hysteresis causes the drying soil to retain more water than it would otherwise.
  • Common confusion: the wetted zone loses water while the zone below gains water; a thin transition layer may first gain then lose water.
  • Hysteresis effect: the drying branch of the water retention curve holds more water at the same matric potential than the wetting branch, so the infiltration-wetted soil retains more water during redistribution than if hysteresis did not occur.

💧 What happens during redistribution

💧 The starting condition

  • At the end of infiltration, the soil profile is far from hydraulic equilibrium.
  • The upper portion wetted by infiltration has higher water potential; deeper soil has lower water potential.
  • Water continues to flow from the wetted region downward into the drier soil.

📉 Water content changes in different zones

The excerpt presents measured profiles for a clay soil after a 1.5-inch (3.8 cm) infiltration event:

Depth zoneWhat happensExample from excerpt
Above the wetting front (infiltration-wetted zone)Water content decreasesFrom ~0.44 cm³ cm⁻³ to ~0.32 cm³ cm⁻³ after one day
Below the wetting front (deeper dry soil)Water content increasesAt 12 cm depth, from ~0.04 cm³ cm⁻³ to ~0.28 cm³ cm⁻³ after one day
Thin transition layer (centered on 11 cm)Water content first increases (after one hour) then decreases (after one day)This layer receives water from above but also loses water downward
  • Don't confuse: the wetted zone is not simply "holding" water—it is actively losing water to the soil below.

🌊 The wetting front evolves

  • The water content gradient associated with the wetting front diminishes over time during redistribution.
  • The matric potential gradient associated with the wetting front also diminishes over time.
  • Example: as water moves deeper, the sharp boundary between wet and dry soil becomes more gradual.

🛑 Three mechanisms that slow redistribution

🛑 Decreasing matric potential gradient

  • As water flows from the wetted zone to the drier zone, the difference in matric potential between the two regions gets smaller.
  • A smaller gradient means a weaker driving force for water flow.
  • Result: the rate of redistribution decreases over time.

🛑 Decreasing hydraulic conductivity

  • As the water content decreases in the infiltration-wetted portion of the profile, the hydraulic conductivity of that soil decreases dramatically.
  • Lower hydraulic conductivity means water flows more slowly even if a gradient still exists.
  • Result: this also causes the rate of redistribution to decrease over time.

🛑 Hysteresis in the water retention curve

Hysteresis: the water content for any given matric potential is higher for the drying branch of the water retention curve than for the wetting branch.

  • During redistribution, the infiltration-wetted part of the soil is getting drier (following the drying branch), while the neighboring parts are getting wetter (following the wetting branch).
  • Because the drying branch holds more water at the same matric potential, the infiltration-wetted part retains more water than it would if hysteresis did not occur.
  • Evidence: numerical model results (Fig. 8‑2) show water content profiles calculated with and without hysteresis; the hysteretic profile retains more water near the surface.
  • Don't confuse: hysteresis is not about the soil "refusing" to drain—it is about the physical relationship between matric potential and water content being different for wetting vs. drying.

🔄 Dynamic nature of redistribution

🔄 A two-sided process

  • Redistribution creates a dynamic situation: the infiltration-wetted part is getting drier while the neighboring parts are getting wetter.
  • This is not a static equilibrium but an ongoing adjustment.

🔄 Sensitivity to hysteresis

  • Because one zone is drying and another is wetting, the redistribution process is sensitive to hysteresis in the water retention curve.
  • The excerpt emphasizes that hysteresis affects how much water is retained near the soil surface after infiltration.

🔄 Summary of retention mechanisms

The excerpt lists three mechanisms that "help" the soil retain water near the soil surface after infiltration:

  1. The decreasing matric potential gradient
  2. The decreasing hydraulic conductivity in the infiltration-wetted zone
  3. The effects of hysteresis
  • All three work together to slow the loss of water from the upper soil layers.
  • Example: after a rainfall event, the surface soil does not immediately drain all its water; these three mechanisms keep water available near the surface for plant roots and evaporation.
35

8.2. Drainage from the Soil Profile

8.2. Drainage from the Soil Profile

🧭 Overview

🧠 One-sentence thesis

Drainage from the soil profile can be estimated using the unit-gradient approach when pressure potential gradients become negligible at depth, but drainage rates depend strongly on soil texture and typically do not have a distinct endpoint.

📌 Key points (3–5)

  • Unit-gradient approach: when pressure potential gradients are negligible at depth, drainage rate equals the hydraulic conductivity at that depth's water content.
  • Texture effects: coarse-textured soils (sand) drain faster and more completely than fine-textured soils (loam, clay) due to larger pore sizes and higher hydraulic conductivity.
  • No distinct endpoint: drainage rates decrease over time but do not cease at a clear moment, which creates problems for the "field capacity" concept.
  • Common confusion: drainage is not constant—it slows progressively as water content decreases, and the relationship between water content and hydraulic conductivity is strongly non-linear.
  • Practical application: the unit-gradient method has been used to estimate groundwater recharge using soil moisture monitoring networks.

💧 The unit-gradient approach

💧 When it applies

  • The excerpt explains that strong forcings (rainfall, irrigation, solar radiation) at the soil surface are moderated deeper in the profile.
  • At sufficient depth, the pressure potential gradient becomes negligible.
  • When this happens, drainage can be estimated using a simplified approach.

🧮 How it works

The approach starts with the Buckingham-Darcy Law:

The water flux (drainage rate) equals the hydraulic conductivity times the gradient of total potential (gravitational plus pressure).

When the pressure potential gradient is negligible and vertical direction is positive downward, the equation simplifies to:

Unit-gradient equation: drainage rate equals the hydraulic conductivity value associated with the soil water content at that depth.

  • The gravitational gradient term equals -1 when z is positive downward.
  • This means q = K(θ), where q is drainage rate and K(θ) is hydraulic conductivity as a function of volumetric water content θ.

⚠️ Key requirement

  • Accurate drainage estimates require careful determination of the relationship between hydraulic conductivity and soil water content.
  • The excerpt emphasizes this relationship is strong and non-linear (referencing Fig. 4-7 from earlier material).
  • Don't confuse: the drainage rate is not simply the water content—it depends on the conductivity function of that water content.

🗺️ Real-world application example

🗺️ Oklahoma Mesonet study

The excerpt describes a study using the unit-gradient approach to estimate groundwater recharge:

  • Daily soil moisture data at 60-cm depth from a statewide monitoring network (Oklahoma Mesonet).
  • Used the Mualem-van Genuchten hydraulic conductivity function with soil parameters estimated by a pedotransfer function called Rosetta.
  • Results showed mean annual drainage rates across Oklahoma for 1998-2014.

🌧️ Precipitation gradient effects

  • Drainage estimates reflected the precipitation gradient: approximately 1420 mm annual precipitation in the southeast to 430 mm in the western Oklahoma Panhandle.
  • Annual drainage maps showed large year-to-year fluctuations driven by differences in annual precipitation.
  • Example: Stillwater had 214 mm/yr, Oklahoma City East had 82 mm/yr, Porter had 166 mm/yr, and Marena had 66 mm/yr.

🏜️ Soil texture effects on drainage

🏜️ Coarse vs fine textures

General principle: Fine-textured soils exhibit lower drainage rates and less cumulative drainage than coarse-textured soils, all other factors being equal.

Why this happens:

  • Coarse-textured soils have larger pore sizes.
  • Larger pores lead to larger hydraulic conductivity values at and near saturation.
  • Fine-textured soils have smaller pores and lower conductivity.

📉 Simulation comparison

The excerpt presents numerical simulations for uniform profiles of loam and sand (Fig. 8-4):

Soil typeInitial water contentWater content after 1 dayChange
Sand~0.45 cm³/cm³~0.20 cm³/cm³Large decrease
Loam~0.47 cm³/cm³~0.35 cm³/cm³Much smaller decrease
  • Sand drained dramatically in the first day.
  • Loam exhibited much less drainage during the same interval.
  • Clay soil would drain even less under the same conditions.

⏱️ Drainage as a continuous process

⏱️ Progressive slowdown

The excerpt emphasizes a key feature visible in the simulations:

  • Drainage rates decreased over time.
  • This is indicated by progressively smaller distances between water content curves at different time points.
  • The curves show measurements at different durations (in days).

⏱️ No distinct endpoint

Critical observation: There is no distinct time at which drainage ceased.

  • Even after 6 days of drainage in uniform sand, water content continued to decrease.
  • The drainage process continued without a clear stopping point.
  • Example: if you measure at day 1, day 2, day 3, etc., each day shows some additional drainage, though the amount gets smaller.

⚠️ Problem for "field capacity"

  • The fact that drainage typically does not have a distinct endpoint highlights a major problem.
  • This creates difficulties for the widely-used concept of "field capacity" in soil science, agronomy, and hydrology.
  • Don't confuse: drainage slowing down is not the same as drainage stopping—the process continues indefinitely, just at progressively lower rates.
36

Field Capacity

8.3. Field Capacity

🧭 Overview

🧠 One-sentence thesis

Field capacity is a flawed concept because drainage does not stop at any clearly defined water content, yet it continues to be used in irrigation and modeling by approximating it with water content at a specific matric potential.

📌 Key points (3–5)

  • What field capacity claims to be: the volumetric water content at which drainage effectively ceases.
  • Why the concept is flawed: drainage typically has no clearly defined breakpoint where it stops; water content continues to decrease gradually over time.
  • Common confusion: field capacity is often incorrectly treated as a soil property, when in reality it is a hypothetical value without a universal physical basis.
  • How it is estimated in practice: by assuming field capacity equals water content at a specific matric potential (commonly -33 kPa in the U.S., -10 kPa elsewhere), though research shows these values are poor approximations.
  • Recent findings: less negative matric potential values (-10 or -6 kPa) appear more suitable than -33 kPa for estimating field capacity.

🧩 The field capacity concept and its problems

🧩 What field capacity is supposed to represent

Field capacity: the volumetric water content at which drainage effectively ceases.

  • This concept has been widely used in irrigation planning, crop modeling, and hydrologic modeling.
  • It is often treated as if it were a property of the soil itself.
  • The underlying assumption is that there is a specific water content value where drainage stops.

❌ Why the concept is fundamentally flawed

  • The core problem: there are typically no clearly defined breakpoints at which drainage stops.
  • Drainage is a continuous process that slows down gradually rather than stopping abruptly.
  • Example from the excerpt: In a drainage experiment on silt loam soil in Israel, gravimetric water content at 60–90 cm depth decreased continuously:
    • Start of drainage: 0.29 g/g
    • After 1 day: 0.20 g/g
    • After 2 days: 0.19 g/g
    • After 30 days: 0.16 g/g
    • After 60 days: 0.15 g/g
  • The excerpt asks: "At which of these water contents could we accurately say that drainage has ceased?" The answer is none—drainage continues throughout.

🚫 Common misconception

  • Don't confuse: Field capacity is not an inherent soil property, even though it is often incorrectly treated as one.
  • It is a hypothetical construct, not a physical characteristic that can be measured directly.

🔬 How field capacity is estimated in practice

🔬 The matric potential approach

  • Most common method: Assume field capacity equals the water content retained at a specific matric potential value.
  • This approach is used despite repeated research proving there is no universal matric potential value that represents field capacity.
  • The method persists because of its convenience.

🌍 Regional differences in matric potential values

RegionMatric potential valueNotes
United States-33 kPaMost commonly used historically
Some other countries-10 kPaMore widely used alternative
Recent research recommendations-10 kPa or -6 kPaLess negative values appear more suitable

📉 Problems with the -33 kPa standard

  • Recent research has shown that -33 kPa provides a poor approximation of field capacity in most cases.
  • Less negative values of matric potential such as -10 kPa or even -6 kPa appear to be more suitable.
  • The excerpt emphasizes that the traditional U.S. standard does not work well in practice.

🔄 Alternative approaches

  • Alternative approaches for estimating field capacity have been proposed.
  • These are based on the parameters of the soil water retention curve.
  • The excerpt mentions these alternatives exist but does not detail them.

🤔 Implications for practice

🤔 The tension between theory and practice

  • The paradox: Field capacity is a flawed concept with no clear physical basis, yet it continues to be used in practical applications.
  • Why it persists: Convenience—having a single reference value simplifies irrigation planning and modeling, even if that value is not physically accurate.
  • The practical compromise: Using matric potential values (especially less negative ones like -10 or -6 kPa) provides a workable approximation despite the conceptual problems.

⚠️ What users should understand

  • Field capacity should not be treated as a precise, universal soil property.
  • Drainage is a continuous process, not one that stops at a specific water content.
  • The matric potential value used to estimate field capacity is an approximation chosen for convenience, not a physical law.
  • Different matric potential values may be more appropriate depending on the soil and situation.
37

Artificial Drainage

8.4. Artificial Drainage

🧭 Overview

🧠 One-sentence thesis

Artificial drainage systems substantially increase agricultural productivity on poorly-drained soils by improving oxygen transport to roots, but they also contribute to large-scale environmental problems by removing nutrients and pesticides that degrade water quality in receiving water bodies.

📌 Key points (3–5)

  • What artificial drainage is: subsurface systems (commonly called "tile drainage" in the US) installed underground to increase soil drainage rates beyond natural capability.
  • Why it's used: improves oxygen transport to crop roots, enabling deeper root systems and higher yields on poorly-drained soils.
  • What it removes: not only excess water but also nutrients (nitrogen, phosphorus) and pesticides from the soil.
  • Environmental trade-off: agronomic benefits come with unintended ecological consequences—drainage systems increase nutrient loss, contributing to problems like the Gulf of Mexico hypoxic "dead zone."
  • Common confusion: drainage itself is not inherently problematic; redistribution and drainage are essential to many natural and managed landscapes, but the scale and consequences of artificial systems can create environmental harm.

🚜 What artificial drainage systems are

🔧 Construction and installation

Tile drainage: a widespread type of artificial drainage; subsurface drainage systems commonly called "tile drainage" in the US.

  • The name comes from the clay tiles originally used to build these underground systems.
  • Modern systems use heavy-duty, corrugated, perforated plastic tubing.
  • Installation: tubing is placed approximately 1.2 m below ground using a large tractor and a special implement called a tile plow.
  • Example: A tractor pulls a tile plow through a field, laying perforated plastic pipe at depth to create a drainage network.

🌍 Scale of adoption

  • Artificial drainage has driven broad adoption in areas with poorly-drained soils.
  • In the US, artificial drainage affects more than 10% of total land area in many counties.
  • In some portions of the Mississippi River floodplain and the Midwest, more than 50% of land area is artificially drained cropland.
  • The excerpt emphasizes that people invest "substantial time, money, and resources" because natural drainage is inadequate for intended agricultural or engineering uses.

🌱 Agronomic benefits

🫁 Oxygen transport and root development

  • Increased drainage improves the soil's ability to transport oxygen to crop roots.
  • Oxygen supports cellular respiration in roots.
  • Better oxygen availability allows crops to develop deeper and larger root systems.
  • Deeper, larger roots enable higher yields.
  • Example: On a poorly-drained soil, a crop's roots might be shallow and stunted; after tile drainage installation, roots can grow deeper, accessing more water and nutrients, leading to better yields.

📈 Productivity gains

  • Subsurface drainage systems have "substantially increased the agricultural productivity of poorly-drained soils around the world."
  • These agronomic benefits are the reason for broad adoption of the technology.

🌊 Environmental consequences

💧 What drainage systems remove

  • Subsurface drainage removes not only excess water but also:
    • Nutrients (particularly nitrogen and phosphorus)
    • Pesticides
  • These compounds flow from the soil into drainage water.
  • The drainage water then flows into surface water bodies (rivers, lakes, coastal areas).

🐟 Water quality and aquatic ecosystems

  • Nutrients and pesticides negatively affect water quality and aquatic ecosystems in receiving water bodies.
  • One large-scale example: the hypoxic zone in the Gulf of Mexico where the Mississippi River flows in.

☠️ The Gulf of Mexico hypoxic "dead zone"

Hypoxic zone: the region where the concentration of dissolved oxygen in the water at the sea floor is less than 2 mg per liter.

How hypoxia develops:

  1. Nutrient-rich water (from drainage and other sources) flows into the Gulf.
  2. Nutrients enhance the growth of algae and other organisms near the surface.
  3. When these organisms die, they settle to the sea floor.
  4. Bacteria decompose the dead organisms, consuming oxygen in the process.
  5. Result: hypoxia (very low oxygen levels).

Consequences:

  • At oxygen levels below 2 mg/L, bottom-dwelling fish and shrimp cannot survive.
  • The hypoxic zone is also called the "dead zone."

Link to artificial drainage:

  • Nutrient loss from cropland in the Mississippi River basin is considered a major contributor to the growth in size of the hypoxic zone since the 1980s.
  • Research has shown that artificial drainage systems have increased nutrient loss from cropland.
  • Therefore, artificial drainage is one contributing factor to this large-scale environmental problem.

⚖️ Balancing benefits and harms

🔄 Drainage is not inherently bad

  • The excerpt emphasizes: "drainage itself is not necessarily problematic."
  • Redistribution of water within the soil profile and drainage of water from the soil profile are essential to the functioning of many natural and managed landscapes.
  • These processes help sustain:
    • Plant growth
    • Biogeochemical cycling
    • Groundwater aquifers

🧪 Influence on solute movement

  • Redistribution and drainage strongly influence the movement of solutes in the environment.
  • The excerpt notes that solute transport processes will be considered in more detail in the next chapter.

⚠️ Don't confuse

  • Natural drainage processes (essential, support ecosystems) vs. large-scale artificial drainage systems (can create environmental harm through nutrient and pesticide removal at scale).
  • The problem is not drainage per se, but the unintended ecological and environmental consequences when artificial systems increase nutrient and pesticide loss beyond what natural systems would produce.
38

Solute Transport

9.1. Solute Transport

🧭 Overview

🧠 One-sentence thesis

Solute transport in soil occurs primarily through advection (bulk water flow) and diffusion (concentration-driven movement), with additional processes like sorption and hydrodynamic dispersion modifying how quickly and how far solutes move through the soil profile.

📌 Key points (3–5)

  • Two fundamental mechanisms: advection moves solutes with flowing water; diffusion moves solutes from high to low concentration due to random molecular motion.
  • Pore water velocity vs flux: the same water flux produces higher pore water velocity when water content is lower, affecting how fast solutes travel.
  • Sorption slows transport: solutes that attach to soil particles move more slowly than non-interacting solutes, quantified by a retardation factor.
  • Common confusion: advective flux (mass per area per time) vs pore water velocity (distance per time)—flux tells you mass movement rate, velocity tells you travel speed.
  • Real-world complexity: chemical reactions, biological processes, and preferential flow make accurate prediction extremely challenging.

🌊 Advection: transport by water flow

🌊 What advection is

Advection: the transport of solutes due to the bulk flow of water.

  • When water flows through soil, it carries dissolved solutes along with it.
  • The advective flux (J_a) depends on both the water flux (q) and the solute concentration (C).
  • Units: mass of solute passing through a unit cross-sectional area of soil per unit time (e.g. grams per square meter per year).
  • Example: if water is flowing downward through soil at a certain rate and contains dissolved nitrate, the nitrate moves downward at a rate proportional to both the water flow rate and the nitrate concentration.

🚀 Pore water velocity

Pore water velocity: the average distance in the direction of the bulk flow which is traveled by an individual water molecule in a unit of time.

  • Pore water velocity (v) equals water flux (q) divided by volumetric water content (θ).
  • Key insight: for a given flux, pore water velocity is higher when volumetric water content is lower.
  • Why: the same amount of water flowing through fewer filled pores must move faster.
  • Don't confuse: flux measures how much water passes through an area; velocity measures how fast individual water molecules travel.

⏱️ Estimating travel time

  • Travel time (t) for a solute to pass through a soil layer of thickness L: t = L / v.
  • If water content is uniform, this becomes: t = (L × θ) / q.
  • Important limitation: these estimates only work for solutes that do not interact with the soil in any way.
  • Example: to estimate how long it takes for a non-reactive tracer to move through 1 meter of soil, divide the thickness by the pore water velocity.

🔍 Complex flow at the pore scale

  • At the microscopic pore scale, water velocities form a complex pattern with different directions and speeds.
  • The simulation in Fig. 9-1 shows brightly colored areas (high velocities) and dark blue areas (low velocities).
  • The average pore water velocity (v) is the net outcome of these innumerable complex flow paths.

🔀 Diffusion and dispersion

🔀 What diffusion is

Solute diffusion: the net movement of solutes from a region of higher concentration to a region of lower concentration due to the random thermal motion of the solute and water molecules.

  • Governed by Fick's Law: diffusive flux (J_d) depends on the diffusion coefficient (D) and the concentration gradient (dC/dz).
  • The diffusion coefficient in soil depends on:
    • The diffusion coefficient of that solute in pure water
    • The volumetric water content of the soil
    • The tortuosity (twistiness) of the diffusion paths within the soil pore network
  • Diffusion is hindered as soil water content decreases and tortuosity increases.

🐌 When diffusion matters

  • Diffusion is often a relatively minor contributor to solute transport in soil compared to advection.
  • However, diffusion can be important over small distances.
  • Example: nutrient uptake by plant roots—nutrients must diffuse through the last few millimeters of soil to reach the root surface.

🌀 Hydrodynamic dispersion

Hydrodynamic dispersion: the process in which the solute appears to be diffusing uniformly upstream and downstream from a plane which is moving downstream at a rate equal to the average pore water velocity.

  • This is a composite phenomenon arising from the interaction of advection and diffusion.
  • Discovered by British physicist Sir Geoffrey Taylor in 1953 for flow in capillary tubes.
  • How it works (in a capillary tube):
    1. Flow velocities have a parabolic distribution: zero at the walls, maximum at the center
    2. Solute initially introduced at a plane gets stretched by this velocity distribution
    3. Strong concentration gradients perpendicular to flow drive strong diffusive transport
    4. Perpendicular diffusion consolidates and homogenizes the solute distribution over time
    5. Eventually the distribution becomes a symmetrical, slowly-expanding, normal distribution centered on a plane moving with the average pore water velocity
  • The excerpt notes this is "utterly counter-intuitive."
  • Although soils bear little resemblance to capillary tubes, this representation was widely adopted to describe solute transport in soil.

📐 Combined transport equation

  • Total solute flux combines advection, diffusion, and hydrodynamic dispersion.
  • Diffusion and hydrodynamic dispersion are grouped together because they cannot be readily distinguished in practice.
  • The combined equation uses an effective diffusion-dispersion coefficient (D_e).
  • Important limitation: this model is only intended to represent transport of solutes that do not interact with the porous media.

🧲 Sorption and retardation

🧲 What sorption is

Sorption: a general term for chemical and physical processes by which solutes become attached to soil particles.

  • Examples:
    • Binding of an organic pesticide like glyphosate by soil organic matter
    • Attachment of ammonium (NH₄⁺) on a soil's cation exchange sites
  • Solutes that undergo sorption are transported less readily than those that do not.

🐢 The retardation factor

  • In the simplest cases, sorption can be described by a linear adsorption isotherm.
  • The adsorbed concentration (C_a, mass of chemical per mass of dry soil) is proportional to the solution concentration (C_l).
  • The proportionality constant is the distribution coefficient (K_d, volume per unit mass).
  • When this linear relationship is valid, we can compute a retardation factor (R) to account for sorption effects.
  • The retardation factor depends on K_d, bulk density (ρ_b), and volumetric water content (θ).

⏳ Impact on travel time

  • Transport time for a sorbing solute: t = (R × L × θ) / q.
  • The transport time for a solute undergoing sorption is R times greater than the transport time for a non-sorbing solute.
  • Example: if the retardation factor is 5, a sorbing solute will take 5 times longer to move through the same soil layer compared to a non-sorbing solute.
  • Don't confuse: retardation slows the solute down; it doesn't stop it completely.

⚠️ Complicating factors

⚗️ Chemical and biological processes

  • Chemical reactions can strongly affect solute behavior:
    • Oxidation/reduction
    • Dissolution/precipitation
    • Association/dissociation type reactions
  • Biological processes also frequently alter fate and transport:
    • Microbial degradation
    • Plant uptake
  • These diverse processes can occur simultaneously, making accurate prediction of solute transport extremely challenging.

🌊 Preferential flow

  • The excerpt identifies preferential flow as "perhaps the single greatest challenge to accurate solute transport prediction."
  • Preferential flow: instances of severe heterogeneity in the water flow rate.
  • This challenge arises not from chemical and biological factors, but from the physical flow patterns themselves.
  • Don't confuse: preferential flow is a physical heterogeneity problem, not a chemical reaction problem.

🌍 Groundwater pollution context

  • Solute transport is one of the primary mechanisms causing groundwater pollution around the world.
  • One common contaminant: nitrate (NO₃⁻), which can be readily leached from agricultural soils receiving nitrogen inputs through fertilizers or manure.
  • In the US, observations of nitrate concentrations in groundwater have been used to develop statistical models that estimate groundwater nitrate concentrations across the nation.
39

Preferential Flow

9.2. Preferential Flow

🧭 Overview

🧠 One-sentence thesis

Preferential flow allows water and solutes to move through a small portion of soil much faster and deeper than uniform flow would permit, dramatically affecting solute transport and pollution prediction.

📌 Key points (3–5)

  • What preferential flow is: uneven movement of water and solutes through a small soil volume at high rates, reaching greater depth in shorter time than uniform flow.
  • Three main causes: macropores (from animals, roots, cracks), hydrophobic soil layers, and certain soil layering patterns (fine-over-coarse textures).
  • Why it matters: solutes can reach groundwater faster than expected; sampling becomes harder due to extreme heterogeneity in flow paths.
  • Common confusion: not all preferential flow is the same—macropore flow, hydrophobic-layer breakthrough, and layering-induced funneling/fingering are distinct mechanisms.
  • Transient behavior: some soils are hydrophobic when dry but become hydrophilic when wetted, complicating predictions.

🌊 What preferential flow is and why it matters

🌊 Definition and core behavior

Preferential flow: the uneven movement of water and solutes through a relatively small portion of the soil volume at relatively high flow rates, allowing these substances to reach greater depth in shorter time than would be possible in a uniform flow situation.

  • It is not uniform movement through the entire soil profile.
  • Instead, water and solutes are channeled through a small fraction of the soil, bypassing much of the soil matrix.
  • The excerpt emphasizes "relatively small portion" and "relatively high flow rates."

🚨 Practical implications

  • Faster solute transport: contaminants can reach groundwater more quickly than uniform-flow models predict.
  • Sampling difficulty: because flow paths are extremely heterogeneous, reliably detecting pollution in soil or groundwater becomes harder.
  • Example: if a pollutant moves through narrow channels, a soil sample taken from the bulk soil may miss the contamination entirely.

🕳️ Macropore flow

🕳️ What macropores are and how they form

  • Macropores are large pores in the soil that create preferential flow paths.
  • They are ubiquitous (found everywhere) and can be formed by:
    • Burrowing animals (earthworms, gophers)
    • Growth and decay of plant roots
    • Soil shrinkage cracks

🔵 Example from New York

  • The excerpt describes a well-structured soil near Ithaca, New York, where blue dye was applied with infiltrating water.
  • The dye moved relatively uniformly through the topsoil.
  • Upon reaching the subsoil, the dye was channeled almost exclusively through macropore flow paths.
  • This shows that macropore flow can dominate solute transport in the subsoil, even if the topsoil behaves more uniformly.

🧩 Macropore flow is common but not the only type

  • The excerpt states that macropore flow is "one of the most common types of preferential flow, but is certainly not the only type."
  • Other mechanisms (hydrophobic soil, soil layering) also cause preferential flow.

💧 Hydrophobic soil

💧 What hydrophobic soil is

A soil is considered hydrophobic when its contact angle with water is greater than 90°.

  • Hydrophobic means "water-repelling."
  • Water does not readily enter a hydrophobic layer; instead, it tends to accumulate above the layer.

🌲 Where hydrophobicity occurs

  • Hydrophobic conditions can arise in surface or subsurface layers where:
    • Organic matter content is high, or
    • Soil minerals have extensive organic coatings.
  • Common locations: forest floors, golf course greens (places with substantial plant residue inputs).

🌀 How hydrophobic layers cause preferential flow

  • Water and solutes accumulate above the hydrophobic layer until reaching a critical point.
  • At that point, hydrophobicity is overcome in one or more distinct locations.
  • Water and solutes then flow preferentially through this small, newly-wetted portion of the layer, leaving much of the soil dry.
  • Example: Figure 9-4 shows a sand dune in the Netherlands with grass cover. After natural rainfall, red dye highlights the wetting pattern. The hydrophobic surface layer resulted in downward flow through small portions of the soil, increasing the total penetration depth while leaving much of the soil dry.

⏳ Transient hydrophobicity

  • Some soils are strongly hydrophobic when dry but transition to a hydrophilic (water-attracting) state when wetted.
  • This transient behavior further complicates the prediction of solute transport.
  • Don't confuse: hydrophobicity is not always permanent; it can change with soil moisture.

🪨 Soil layering

🪨 When layering causes preferential flow

  • Preferential flow can occur when a finer-textured layer has a sharp interface with a substantially coarser-textured layer.
  • The key is a sharp contrast in texture (pore size).

🔻 Why water doesn't readily pass into the coarser layer

  • The smaller pores in the overlying finer-textured layer hold water at a more negative pressure potential than the larger pores in the coarser layer.
  • In other words, water is held more tightly in the fine layer and does not easily move into the coarse layer until certain conditions are met.

🌊 Two types of layering-induced preferential flow

TypeInterface orientationFlow patternDescription
FunnelingSloping (not horizontal)Downslope channelingFlow is "funneled" in the downslope direction along the interface
Finger flowHorizontalVertical fingersWater and solutes penetrate the coarser layer only in fingers occupying a small portion of the total volume
  • In both cases, solutes can be transported to greater depths than they would otherwise.
  • Example: Figure 9-5 shows blue dye highlighting preferential flow caused by a sloping coarse-textured soil layer at a field site in New York. The dye is channeled along the sloping interface, demonstrating funneling.

🧪 Sampling challenges

  • Because flow paths are extremely heterogeneous, reliably sampling soil or groundwater to detect pollution becomes more difficult.
  • Example: if solutes move only through narrow fingers or along a sloping interface, a sample taken from the bulk soil may miss the contamination.
40

Groundwater Pollution

9.3. Groundwater Pollution

🧭 Overview

🧠 One-sentence thesis

Solute transport is a primary mechanism causing groundwater pollution worldwide, with agricultural nitrate and industrial contaminants posing significant threats that require understanding of transport mechanisms to manage effectively.

📌 Key points (3–5)

  • Primary mechanism: Solute transport drives groundwater pollution globally.
  • Agricultural source: Nitrate from fertilizers and manure leaches readily from agricultural soils; many U.S. locations exceed EPA's 10 mg/L drinking water limit.
  • Industrial sources: Mining, oil and gas production, industrial activities, and leaking underground storage tanks contribute serious contaminants like hydrocarbons and chlorinated solvents.
  • Fracking risk distinction: Surface spillage of fracking fluid or produced water poses greater groundwater risk than below-ground activities—reinforcing the importance of understanding soil solute transport.
  • Geographic patterns: High nitrate concentrations correlate with high nitrogen application rates, high water inputs, and well-drained or coarse-textured soils.

🌾 Agricultural groundwater pollution

💧 Nitrate contamination

  • What it is: Nitrate (NO₃⁻) is a common groundwater contaminant leached from agricultural soils.
  • Source: Nitrogen inputs through fertilizers or manure applied to farmland.
  • Why it matters: Nitrate can be "readily leached," meaning it moves easily through soil into groundwater.

📏 EPA drinking water standard

Maximum contaminant level for nitrate in drinking water: 10 mg per liter (mg L⁻¹).

  • The U.S. Environmental Protection Agency sets this limit to protect public health.
  • Statistical models and actual measurements show many U.S. locations exceed this threshold.

🗺️ Geographic hotspots

The excerpt identifies specific regions with elevated nitrate concentrations:

RegionNitrate status
Great PlainsExceeded EPA limit in portions
Upper MidwestExceeded EPA limit in portions
East CoastExceeded EPA limit in portions
California's Central ValleyExceeded EPA limit in portions

🧪 Factors driving high nitrate levels

Three conditions combine to create high groundwater nitrate concentrations:

  • High nitrogen application rates: More fertilizer or manure applied.
  • High water inputs: Rain plus irrigation increases leaching.
  • Well-drained or coarse-textured soils: Water (and dissolved nitrate) moves through these soils more easily.

Example: A farm in the Great Plains with heavy fertilizer use, irrigation, and sandy soil would be at high risk for nitrate leaching into groundwater.

🏭 Industrial and other human sources

⛽ Non-agricultural pollution sources

Agriculture is not the only contributor. The excerpt lists four major categories:

  • Mining: Serious groundwater pollution documented.
  • Oil and gas production: Contributes contaminants.
  • Industrial activities: Various industrial processes release pollutants.
  • Leaking underground storage tanks: Gas stations (which "many of us use on a regular basis") have tanks that can leak.

🧪 Types of industrial contaminants

The excerpt names specific chemical classes:

Contaminant typeExamplesSource
HydrocarbonsBenzene, tolueneOil/gas, industrial activities
Chlorinated solventsTrichloroethylene (TCE), tetrachloroethylene (PCE)Industrial activities
  • These are described as "serious groundwater contaminants."
  • Contamination from these sources "sometimes necessitate costly remediation efforts."

🔧 Hydraulic fracturing ("fracking")

🔍 Current evidence

  • Growing concern: Recent years have seen increased worry about fracking's potential threat to groundwater.
  • EPA findings: The U.S. EPA has "thus far not found widespread evidence of groundwater pollution due to 'fracking.'"
  • Isolated cases: Some individual cases have been reported, but not widespread contamination.

⚠️ Where the real risk lies

The available evidence indicates that accidental spillage of fracking fluid or produced water on the land surface constitutes a greater risk to groundwater than the associated below-ground activities.

  • Don't confuse: The underground fracturing process itself versus surface handling of fluids.
  • Greater risk: Surface spills of fracking fluid or produced water.
  • Lesser risk: The below-ground fracturing activities.

Example: If fracking fluid is spilled at the surface during transport or storage, it can infiltrate soil and reach groundwater through solute transport mechanisms—a bigger threat than the deep underground injection process.

🔬 Why understanding solute transport matters

🎯 Practical importance

The excerpt concludes by emphasizing that the fracking-risk finding "reinforces the importance of understanding the mechanisms of solute transport in the soil."

  • Why: Even when the primary industrial activity (deep fracturing) poses low risk, surface contamination can still threaten groundwater.
  • Implication: Predicting and preventing groundwater pollution requires understanding how chemicals move through soil.
  • Connection to earlier material: The excerpt references solute transport mechanisms (sorption, retardation, preferential flow) discussed in section 9.1.

🌍 Scale of the problem

  • Groundwater pollution is described as occurring "around the world," not just in the U.S.
  • Multiple human activities contribute, making this a complex, multi-source problem.
  • Statistical models and monitoring are used to estimate contamination extent across entire nations.
41

10.1. Necessary Conditions for Evaporation

10.1. Necessary Conditions for Evaporation

🧭 Overview

🧠 One-sentence thesis

Sustained evaporation requires three conditions—an energy supply, a water supply, and a vapor transport mechanism—and the rate is limited by either atmospheric demand or the soil's ability to deliver water to the evaporating surface.

📌 Key points (3–5)

  • Three necessary conditions: energy supply, water supply, and a transport mechanism must all be present for sustained evaporation.
  • Energy can come from multiple sources: primarily the sun, but also from the evaporating body itself (e.g., soil) or surrounding air, which is why evaporation can occur at night and why sweating cools us.
  • Water supply depends on soil hydraulic properties: liquid water must be transported from within the soil to the evaporating surface, so soil properties that affect water flow and retention influence evaporation.
  • Vapor transport requires a concentration gradient: water vapor concentration in the atmosphere must be lower than at the evaporating surface; wind speeds increase evaporation by moving vapor away through advection.
  • Common confusion—evaporative demand vs. soil limitation: evaporative demand (energy supply + atmospheric vapor transport capacity) is distinct from the soil's ability to supply water; the slower of the two limits the actual evaporation rate.

🔑 Definitions and distinctions

🔑 What evaporation means in this context

Evaporation: the process in which water changes phase from liquid to vapor and is transported to the atmosphere from the soil, the exterior surface of plants, plant residue, or surface water bodies.

  • This definition is narrower than everyday usage.
  • It excludes water vaporized from inside plants (that is transpiration).
  • It also excludes the combined process (evaporation + transpiration = evapotranspiration).

Don't confuse:

  • Evaporation (from soil, plant surfaces, residue, or water bodies)
  • Transpiration (vaporization from the interior of plants, exiting through stomates)
  • Evapotranspiration (the sum of evaporation and transpiration for a region)

⚡ The three necessary conditions

⚡ Energy supply

Latent heat of vaporization: the energy input required to overcome the molecular forces of attraction between water molecules in liquid form.

  • Water has an unusually high latent heat of vaporization (at 15°C, 2.5 × 10⁶ joules per kilogram).
  • "Latent heat" means energy absorbed during the phase change without changing temperature.
  • The energy can come from:
    • The sun (primary source)
    • The evaporating body itself (e.g., the soil)
    • Surrounding air
  • Example: Sweat evaporating from your body draws energy from your skin, cooling you down; similarly, evaporation can occur at night because energy can be drawn from the soil or air.

Unit note: One joule is approximately 4.18 calories; one food calorie (Cal) is approximately 4,180 joules.

💧 Water supply

  • Liquid water must be transported from within the soil to the location where vaporization occurs (typically at or near the soil surface).
  • Soil hydraulic properties—those that influence water flow and retention—directly affect the evaporation process.
  • If the soil cannot deliver water fast enough to the surface, the evaporation rate will be limited by the soil, not by atmospheric conditions.

🌬️ Vapor transport mechanism

  • A concentration gradient is required: water vapor concentration in the atmosphere must be less than the concentration at the evaporating surface.
  • This gradient drives diffusion of water vapor into the atmosphere.
  • Once in the atmosphere, vapor is moved by advection (bulk air flow).
  • Higher wind speeds increase evaporation by carrying vapor away more quickly.

🌡️ Evaporative demand and rate limitation

🌡️ What is evaporative demand

Evaporative demand: the combined effect of the above-ground energy supply and the atmosphere's capacity to transport vapor away from the land surface.

  • It represents the "pull" from the atmosphere.
  • It does not account for the soil's ability to supply water.

🚧 What limits the actual evaporation rate

The rate of evaporation from soil is limited by whichever is slower:

  1. Evaporative demand (energy + atmospheric vapor transport capacity), or
  2. The soil's ability to transport water to the location of vaporization.
Limiting factorWhat it means
Evaporative demandAtmosphere can remove vapor faster than soil can supply water
Soil water transportSoil cannot deliver water fast enough to meet atmospheric demand

Don't confuse: Evaporative demand is a property of the atmosphere and energy availability; soil limitation is a property of the soil's hydraulic characteristics. Both must be considered to predict actual evaporation.

42

10.2. Evaporation from a Water Table

10.2. Evaporation from a Water Table

🧭 Overview

🧠 One-sentence thesis

Evaporation from shallow water tables is a steady-state process that can sustain long-term water loss and drive soil salinization by leaving behind accumulated salts at the surface.

📌 Key points (3–5)

  • What shallow water table evaporation is: a process where evaporating soil water is continuously replenished by upward flow from groundwater, allowing evaporation to proceed for long periods.
  • Why it matters globally: this process is a major factor in soil salinization, which affects approximately 3.1% of Earth's land area and threatens agriculture and ecosystems.
  • How it differs from other evaporation: treated as steady-state (constant rate over time) because the water table continuously supplies water, unlike drying soil without a water source.
  • How to estimate the rate: apply the Buckingham-Darcy Law with Campbell's hydraulic conductivity function to calculate maximum steady evaporation rate for a given soil and water table depth.
  • Common confusion: evaporation rate is limited by either evaporative demand (atmospheric conditions) or the soil's ability to transport water upward—shallow water table scenarios focus on the latter.

🌊 Shallow water table evaporation as a steady-state process

🌊 What steady-state means here

Steady-state process: the rate of water movement is assumed constant over time, and the soil is neither drying nor wetting.

  • All evaporated water is replenished by upward flow from the underlying groundwater table.
  • This is the "simplest approximation" for understanding evaporation when a water table is relatively shallow.
  • The soil profile maintains a stable water distribution because supply matches loss.

🔄 Continuous replenishment mechanism

  • Water evaporates at the soil surface.
  • Groundwater moves upward through the soil to replace the lost water.
  • The cycle continues as long as the water table remains accessible.
  • Example: A field with a water table 2 meters below the surface can sustain evaporation indefinitely, unlike a deep water table where the soil would dry out.

🧂 Soil salinization and its global impact

🧂 How salinization occurs

  • Water transported to the soil surface carries dissolved salts.
  • When water evaporates, it leaves the salts behind.
  • Salts accumulate over time at or near the soil surface (as shown in the figure of salt-covered rangeland).

🌍 Scale and significance

Salinization: the accumulation of salts in the soil to a level that negatively impacts agricultural production, ecosystem health, and economic welfare.

  • Currently affects approximately 397 million hectares worldwide (3.1% of Earth's land area).
  • Contributed to the downfall of ancient societies in Mesopotamia.
  • One of the major factors in the global soil salinization crisis.
  • Understanding shallow water table evaporation physics helps predict and manage salinization.

⚠️ Don't confuse

  • Salinization is not caused by evaporation alone—it requires both evaporation and upward salt transport from a water source.
  • Without a shallow water table, salts would not continuously accumulate because the water supply would be exhausted.

🧮 Calculating maximum evaporation rate

🧮 The Buckingham-Darcy Law application

  • The excerpt applies the Buckingham-Darcy Law (Equation 4-5) to estimate evaporation rate.
  • This law describes water flow through soil based on hydraulic conductivity and potential gradients.

📐 Campbell's hydraulic conductivity function

  • The hydraulic conductivity K is defined using Campbell's function written in terms of pressure potential ψ.
  • The function includes the soil's air-entry potential ψ_e (a soil-specific parameter).

📊 Key variables and assumptions

Variable/AssumptionMeaning
zVertical position (positive downward)
LWater table depth
Pressure potential at surfaceAssumed to be negative infinity (very dry surface)
E_maxMaximum possible steady evaporation rate
NA parameter equal to 2 + 3/b (where b is a soil parameter)

🔢 The resulting equation

  • After integrating the Buckingham-Darcy Law with the stated assumptions, the maximum steady evaporation rate E_max is given by Equation 10-3.
  • The equation depends on:
    • Soil properties (air-entry potential, parameter b)
    • Water table depth L
  • The excerpt refers to a video for the full derivation and worked example.

🎯 What E_max represents

  • It is the maximum possible steady evaporation rate for a given soil type and water table depth.
  • Actual evaporation may be lower if evaporative demand (atmospheric conditions) is the limiting factor rather than soil transport capacity.
  • Example: If E_max is calculated as 5 mm/day but atmospheric demand is only 3 mm/day, actual evaporation will be 3 mm/day.

🔀 Evaporative demand vs. soil transport limitation

🔀 Two limiting factors

Evaporative demand: atmospheric factors (energy, humidity, wind) that transport vapor away from the land surface, often lumped together.

  • Evaporation rate is limited by whichever is smaller:
    1. Evaporative demand (how fast the atmosphere can remove water vapor)
    2. Soil's ability to transport water to the vaporization location

🌤️ When each factor dominates

  • Demand-limited: Soil can supply water faster than the atmosphere can remove it → evaporation rate = evaporative demand.
  • Soil-limited: Atmosphere can remove water faster than soil can supply it → evaporation rate = soil transport capacity (E_max in shallow water table case).

📍 Shallow water table context

  • The section focuses on calculating the soil transport capacity (E_max).
  • This is most relevant when the water table is shallow enough that soil transport becomes the bottleneck.
  • Don't confuse: E_max is not the actual evaporation rate unless soil transport is the limiting factor.
43

10.3. Evaporation in the Absence of a Water Table

10.3. Evaporation in the Absence of a Water Table

🧭 Overview

🧠 One-sentence thesis

Without a shallow water source, soil evaporation is transient and typically progresses through three stages—constant-rate, falling-rate, and low-rate—each governed by whether evaporative demand or soil water transmission is the limiting factor.

📌 Key points (3–5)

  • Why evaporation cannot be steady-state: without a near-surface water source like a shallow water table, the soil dries out as evaporation proceeds, causing the rate to eventually slow down.
  • Three stages of transient evaporation: first constant-rate (demand-limited), second falling-rate (soil-limited), and third low-rate (vapor transport only).
  • What limits each stage: stage 1 is limited by evaporative demand; stage 2 begins when soil cannot transmit water fast enough; stage 3 occurs when surface hydraulic conductivity is essentially zero.
  • Common confusion: the three-stage model assumes constant conditions, but in the field, daily cycles of evaporative demand cause repeated surface wetting and drying, making the three stages less distinct.
  • Duration depends on: soil hydraulic properties and evaporative demand—high demand shortens each stage (hours to days), low demand extends them (days to weeks).

🌊 Why evaporation is transient without a water table

🌊 No steady-state possible

  • When there is no near-surface water source (like a shallow water table), evaporation cannot maintain a steady rate.
  • As evaporation proceeds, the soil dries out progressively.
  • The drying process causes evaporation to eventually slow down—this is transient or time-varying evaporation.
  • This transient type is the most common form of evaporation in natural and agricultural settings.

📉 The three stages of transient evaporation

🔄 Stage 1: Constant-rate (demand-limited)

First constant-rate stage: evaporation is limited by the evaporative demand, and the soil can transmit water to the surface at a rate adequate to meet that demand.

  • During this stage, the soil's ability to move water is not the bottleneck.
  • If evaporative demand is constant (e.g., in a laboratory), the evaporation rate remains constant.
  • Example: Laboratory data in Fig. 10-3 show a flat evaporation rate during stage 1 under constant radiation and temperature.
  • Field reality: evaporative demand typically follows a daily cycle (peak near midday, minimum at night), so first-stage evaporation is not truly constant in the field.

📉 Stage 2: Falling-rate (soil-limited)

Second falling-rate stage: the soil is no longer able to transmit water to the surface at a rate adequate to keep up with the evaporative demand.

  • The evaporation process switches from demand-limited to soil-limited.
  • At this moment, the evaporation rate begins to fall (visible in Fig. 10-3).
  • The soil's hydraulic conductivity or water content has decreased enough that water cannot reach the surface fast enough.
  • Don't confuse: the evaporative demand has not changed; the soil's ability to supply water has become the constraint.

🐌 Stage 3: Low-rate (vapor transport only)

Third low-rate stage: the soil surface has become so dry that hydraulic conductivity is essentially zero, and evaporation can only proceed by sub-surface evaporation followed by water vapor transport through the soil surface layer to the atmosphere.

  • Liquid water can no longer be transmitted to the surface.
  • Evaporation now requires:
    1. Sub-surface evaporation (water turns to vapor below the surface).
    2. Vapor diffusion through the dry surface layer to the atmosphere.
  • Evaporation is reduced to a small fraction of the rate observed when the surface was wet.

⏱️ Duration and field complexity

⏱️ What controls stage duration

FactorEffect on duration
High evaporative demandFirst stage may last only a few hours; second stage only a few days
Low evaporative demandFirst stage may last for days; second stage for weeks
Soil hydraulic propertiesSoils with higher conductivity or water-holding capacity can sustain stage 1 longer

🌀 Daily cycles complicate the three-stage model

  • Strong daily cycles of evaporative demand result in repeated drying and rewetting of the soil surface.
  • This creates evaporation dynamics more complex than three distinct stages.
  • Classic field experiment (Phoenix, Arizona, 1971):
    • A loam soil was irrigated, then soil water content was measured every 30 minutes for approximately 15 days.
    • Fig. 10-4 shows pronounced daily wetting and drying of the soil surface (0–5 mm layer).
    • The author concluded that the three-stage concept appeared to have little meaning under these field conditions.
  • Don't confuse: the three-stage model is a useful simplification for constant conditions, but real field conditions often blur the boundaries between stages.

💧 Reducing evaporative losses

🌾 Crop residues reduce stage 1 evaporation

  • Keeping the soil covered with crop residues is an effective method to conserve water during the first stage.
  • How residues work:
    • Reduce the amount of solar radiation reaching the surface.
    • Reduce the rate of water vapor transport away from the surface.
    • Both effects lower the evaporative demand.
  • Data (Fig. 10-5): cotton, sorghum, and wheat residues have similar beneficial effects when present in equivalent thicknesses.
  • Trade-off: lowering stage 1 evaporation rates may prolong stage 1, potentially offsetting some water-conserving effects.

💦 Irrigation frequency and method

  • Decrease frequency of surface-applied irrigation:
    • Every time the surface is wetted, evaporation returns to first-stage levels.
    • Applying the same total irrigation amount in larger doses with greater time intervals reduces the time spent in first-stage evaporation.
    • This reduces cumulative evaporation.
  • Subsurface irrigation:
    • Even greater reductions in soil evaporation can be obtained.
    • Example: A Texas study found that using drip emitters at 30-cm depth (instead of surface drip emitters) might save up to 10% of total seasonal water inputs due to reduced evaporation.

🛡️ Other methods

  • A variety of other methods have been tried.
  • Some, like plastic mulches, have proven quite effective.
  • (The excerpt does not provide further details on other methods.)
44

10.4. Reducing Evaporative Losses

10.4. Reducing Evaporative Losses

🧭 Overview

🧠 One-sentence thesis

Farmers can reduce evaporative water losses through methods like crop residues and subsurface irrigation, but some strategies (like "dust mulch" tillage) have proven ineffective or even disastrous.

📌 Key points (3–5)

  • Why reduce evaporation: evaporation consumes a substantial portion of available water in cropping systems, so farmers seek methods to conserve water.
  • Crop residues work: covering soil with crop residues reduces solar radiation and vapor transport, lowering stage 1 evaporation rates.
  • Irrigation timing matters: less frequent, larger irrigation doses reduce cumulative evaporation by limiting time spent in first-stage evaporation; subsurface irrigation can save up to 10% of seasonal water.
  • Trade-off with residues: lowering stage 1 evaporation rates may prolong stage 1, offsetting some water-saving benefits.
  • Common confusion: "dust mulch" tillage was thought to reduce evaporation by creating a dry surface layer with low hydraulic conductivity, but it made soil vulnerable to massive wind erosion during drought.

🌾 Crop residue methods

🌾 How crop residues reduce evaporation

  • Keeping soil covered with crop residues is an effective method to conserve water during first-stage evaporation.
  • Two mechanisms:
    • Reduces solar radiation reaching the surface.
    • Reduces vapor transport away from the surface.
  • Both mechanisms lower the evaporative demand.

📊 Residue effectiveness

  • Cotton, sorghum, and wheat residues have similar beneficial effects when present in equivalent thicknesses.
  • The data in Fig. 10-5 show that residue thickness (calculated as residue mass per unit area divided by residue density) reduces first-stage evaporation rate.

⚖️ Trade-off: prolonging stage 1

  • The offset: lowering stage 1 evaporation rates may lead to prolonging stage 1.
  • This means the water-conserving effects may be partially offset because the first stage lasts longer.
  • Don't confuse: reducing the rate is not the same as reducing total evaporation if the duration increases.

💧 Irrigation strategies

💧 Frequency and timing

  • Surface irrigation: evaporative losses can be reduced by decreasing the frequency of surface-applied irrigation.
  • Every time the surface is wetted, evaporation returns to first-stage levels.
  • Strategy: apply the same total irrigation amount in larger doses with greater time intervals between applications.
  • Result: reduces the time spent in first-stage evaporation, thus reducing cumulative evaporation.

🚰 Subsurface irrigation

  • Even greater reductions in soil evaporation can be obtained by using subsurface irrigation instead of surface irrigation.
  • Example: a Texas study found that using drip emitters at 30-cm depth (instead of surface drip emitters) saved up to 10% of total seasonal water inputs due to reduced evaporation.
  • Subsurface placement keeps the surface drier, avoiding repeated returns to first-stage evaporation.

⚠️ Failed methods and historical lessons

⚠️ The "dust mulch" concept

  • The idea: some farmers believed that pulverizing and desiccating a shallow surface layer through tillage would create a layer with very low hydraulic conductivity.
  • The reasoning: this dry, pulverized "dust mulch" layer would protect underlying soil water from evaporation.
  • Partial success: it was effective in reducing evaporation in some cases.

🌪️ The Dust Bowl disaster

  • During the severe drought of the 1930s, aggressive tillage practices (including "dust mulch") made the land vulnerable to wind erosion on a massive scale.
  • Context: early 20th century settlers expanded cropland in the semi-arid western Great Plains; tillage was the primary weed control method (herbicides were not available).
  • Result: drought and poor land management combined to cause the "Dust Bowl," a time of severe wind erosion in the US Great Plains.
  • The iconic photograph from Oklahoma Panhandle (Fig. 10-6) shows dust drifts burying fences and farm buildings, with no crops or livestock remaining.
  • Don't confuse: a method that reduces evaporation in the short term may have catastrophic long-term consequences (soil erosion, loss of productivity).

🔍 Other methods

  • A variety of other methods for reducing evaporation have been tried.
  • Effective: plastic mulches have proven quite effective.
  • Ill-conceived: some methods (like "dust mulch") have proven ill-conceived, contributing to one of the greatest natural disasters in US history.
45

Wind Erosion

10.5. Wind Erosion

🧭 Overview

🧠 One-sentence thesis

Wind erosion, which involves detachment, transport, and deposition of soil particles by wind, became catastrophic during the 1930s Dust Bowl when aggressive tillage practices combined with drought left vast areas of the Great Plains vulnerable to massive soil loss.

📌 Key points (3–5)

  • Historical context: early 20th-century expansion of cropland in the semi-arid Great Plains, driven by favorable weather and prices, led farmers to plow huge tracts of prairie and rely heavily on tillage for weed control.
  • The "dust mulch" mistake: farmers believed pulverizing the surface soil would reduce evaporation by creating a low-conductivity dry layer, but this practice made land extremely vulnerable to wind erosion during drought.
  • Three stages of wind erosion: detachment (wind gusts separate particles), transport (surface creep, saltation, or suspension), and deposition (when wind velocity decreases).
  • Common confusion: tillage was thought to conserve soil water by reducing evaporation, and it did work in some cases, but it catastrophically backfired during the 1930s drought by leaving soil unprotected.
  • Why it matters today: severe dust storms still occur (as recently as 2012), so understanding and minimizing wind erosion remains important; management practices like cover crops and crop residue reduce wind speed near the soil surface.

🌾 The Dust Bowl disaster

🌾 Expansion and tillage practices

  • In the early 20th century, settlers dramatically expanded cropland in the semi-arid western Great Plains.
  • Several years of favorable weather and commodity prices encouraged farmers to plow huge tracts of prairie to grow crops such as winter wheat.
  • Without herbicides, tillage was the primary means of weed control.

💨 The "dust mulch" theory

"Dust mulch": a shallow, dry, pulverized layer of soil on the surface, created by tillage, intended to reduce evaporative losses and conserve soil water.

  • The reasoning: if a shallow surface layer was pulverized and desiccated by tillage, its hydraulic conductivity would be so low that soil water in underlying layers would be protected from evaporation.
  • What actually happened: the dust mulch was effective in reducing evaporation in some cases, but during the severe drought of the 1930s, these aggressive tillage practices made the land vulnerable to wind erosion on a massive scale.
  • Don't confuse: the method worked for its intended purpose (reducing evaporation) under some conditions, but it created a far worse problem (catastrophic wind erosion) when drought struck.

📸 The devastation

  • Drought and poor land management practices combined to cause the "Dust Bowl" in the US Great Plains, a time of severe wind erosion.
  • Example: the iconic 1936 photograph from the Oklahoma Panhandle shows drifts of dust burying large parts of fences and farm buildings, with the sky filled with dust blowing in waves along the surface.
  • Farm families had no crops, no livestock, and perhaps no hope.
  • Between 1930-1940, millions of people packed up and moved out of the Great Plains states—one of the largest migrations in US history.

⚠️ Ongoing relevance

  • Wind erosion in the US has not reached Dust Bowl levels since that time.
  • However, severe dust storms have occurred in the Great Plains states as recently as 2012.
  • Understanding and minimizing wind erosion is important still today.

🔄 The three stages of wind erosion

🔄 Overview of the process

Wind erosion, like water erosion, includes three stages: detachment, transport, and deposition.

💥 Detachment

  • Detachment typically occurs due to wind gusts that reach adequate velocity to separate particles from the soil surface.
  • Why management matters: practices such as cover crops or standing crop residue can be effective preventive measures, since they reduce wind speed near the soil surface.

🚚 Transport

Once particles are detached, they may be transported by three processes:

Transport processDescription
Surface creepRolling along the surface
SaltationBouncing along the surface
SuspensionBeing carried up and away from the surface for a substantial distance

📍 Deposition

  • Deposition of windblown sediment is typically initiated by a decrease in wind velocity.
  • This is the same principle as for water-borne sediment.
  • Example: during the Dust Bowl, barbed wire fences provided enough of a wind break to allow large drifts to form, so that in some cases livestock simply walked right over the top of the fence and wandered away.

🔬 Prediction and modeling

🔬 Wind erosion models

  • Sophisticated models have been developed to predict wind erosion processes.
  • One of the most widely-used models is the USDA Wind Erosion Prediction System (WEPS).
46

11.1. Soil-Plant-Atmosphere Continuum

11.1. Soil-Plant-Atmosphere Continuum

🧭 Overview

🧠 One-sentence thesis

Transpiration is driven by the large water potential difference between soil and atmosphere along the soil-plant-atmosphere continuum, and plants must balance water loss through stomates with root water uptake to maintain healthy water status.

📌 Key points (3–5)

  • The continuum: water moves continuously from soil → through plants → to atmosphere, driven by water potential differences.
  • Transpiration is mostly passive: plants lose water vapor through stomates as a "concession" to atmospheric demand, not primarily as an active physiological function.
  • The stomate trade-off: plants must open stomates to take in CO₂ for photosynthesis, but unavoidably lose water vapor in the process.
  • Water balance determines plant health: when transpiration exceeds root uptake, plants wilt; when uptake exceeds transpiration (e.g., at night), plants recover.
  • Common confusion: transpiration vs evaporation—transpiration is vaporization from inside plants (exiting through stomates), while evaporation (previous chapter) is from soil/water surfaces.

🌊 The soil-plant-atmosphere continuum

🌊 What the continuum is

Soil-plant-atmosphere continuum: the continuous pathway by which water moves from the soil, through plants, to the atmosphere.

  • Water flows along this pathway in response to water potential differences.
  • The atmosphere typically has much lower (more negative) water potential than the soil, creating a strong driving force.

📉 Water potential gradient example

The excerpt gives a greenhouse tomato study in Shaanxi, China:

ConditionSoil water potentialAir water potentialPotential differenceEffect on transpiration
Normal greenhouse air-0.26 MPa-168 MPa~168 MPaBaseline
Humidified (fogging system)~-0.26 MPa (unchanged)-82 MPa~82 MPa (50% smaller)20% reduction in cumulative transpiration
  • The huge atmospheric demand (very negative potential) pulls water through the plant.
  • Humidifying the air reduced the driving force by ~50%, cutting transpiration by 20%.

🌱 Transpiration is mostly passive

The excerpt emphasizes:

Transpiration is, in one sense, not an active physiological function of plants but rather a primarily passive concession plants must make to the almost insatiable atmospheric demand for water.

  • Plants do not "choose" to transpire; they lose water because the atmosphere demands it.
  • The plant's role is to regulate (via stomates), not to actively pump water out.

🚪 Stomates and gas exchange

🚪 What stomates are

Stomates: pores or openings on the exterior surfaces of leaves and other plant organs.

  • A pair of specialized guard cells allows the plant to open and close each stomate.
  • Stomates regulate the flow of gases (water vapor, CO₂, O₂) into and out of the plant.

🔄 The stomate trade-off

  • To photosynthesize, the plant must open stomates to take in CO₂ from the atmosphere.
  • But opening stomates unavoidably allows water vapor to exit (transpiration).
  • Oxygen (O₂) produced by photosynthesis also exits through stomates.
  • Example: a plant in bright sun opens stomates wide to capture CO₂, but loses water rapidly; if soil water is limited, the plant may close stomates partially to conserve water, slowing photosynthesis.

🛡️ Don't confuse

  • Transpiration (this chapter): vaporization from inside the plant, exiting through stomates.
  • Evaporation (previous chapter): vaporization from soil or water surfaces, not involving plant tissues.

💧 Plant water status and balance

💧 Balancing transpiration and root uptake

  • Healthy balance: root water uptake matches transpiration → plant water stores remain stable.
  • Transpiration exceeds uptake: water stored in plant tissues is depleted → plant begins to shrink or wilt.
  • Uptake exceeds transpiration (e.g., at night or after rain): water stores are replenished → plant turgor is restored.

🥀 Wilting and water stress

When transpiration > uptake for a prolonged period:

  1. Water potential inside the plant decreases.
  2. The plant reduces stomatal openings to cut transpiration.
  3. Visible signs: rolled or wilted leaves.
  4. These are telltale signs of plant water stress.

🌙 Recovery conditions

  • Reduced evaporative demand: e.g., when the sun goes down, transpiration slows, allowing root uptake to catch up and restore turgor.
  • Increased soil water potential: rainfall or irrigation raises soil water potential, boosting root uptake even during the day.

⚠️ Permanent wilting point

Permanent wilting point: the soil volumetric water content at which the plant cannot recover even if evaporative demand is removed or the soil is rewetted.

  • At this critical level of water stress, the plant is permanently wilted.
  • The exact soil moisture depends on plant species and evaporative demand.
  • Common estimate: soil water content at a matric potential of -1500 kPa.
  • Don't confuse: temporary wilting (plant recovers at night) vs permanent wilting (plant cannot recover).

🌿 Root water uptake mechanisms

🌿 What limits root uptake rate

The excerpt states that root water uptake can be limited by:

  1. Hydraulic conductivity of the rhizosphere soil: the soil immediately adjacent to the roots; if conductivity is low, water cannot flow to roots fast enough.
  2. Water potential gradient between soil and roots: if the gradient is small, the driving force for water movement is weak.
  • Example: in very dry soil, hydraulic conductivity drops sharply, slowing root uptake even if the plant's water potential is low.
  • The excerpt mentions a "traditional conceptual model" but the text cuts off before explaining further details.
47

Water Status of Plants

11.2. Water Status of Plants

🧭 Overview

🧠 One-sentence thesis

Plants must balance water loss through transpiration with root water uptake to maintain healthy water status, and when transpiration exceeds uptake for too long, plants experience water stress that can lead to permanent wilting.

📌 Key points (3–5)

  • Transpiration is passive: transpiration is not an active physiological function but a passive concession to atmospheric demand for water, driven by water potential differences between soil, plant, and air.
  • Stomates create a trade-off: plants must open stomates to take in carbon dioxide for photosynthesis, but unavoidably lose water vapor in the process.
  • Water balance determines plant status: when transpiration exceeds root uptake, plants deplete internal water stores and wilt; when uptake exceeds transpiration (e.g., at night), plants recover.
  • Permanent wilting point: prolonged water stress can reach a critical severity where plants cannot recover even if conditions improve; this soil moisture state is called the permanent wilting point (often estimated at -1500 kPa matric potential).
  • Common confusion: temporary wilting vs permanent wilting—temporary wilting can be reversed by reducing evaporative demand or adding water, but permanent wilting cannot be reversed.

💧 The driving force behind transpiration

💧 Water potential gradients

  • Transpiration is driven by water potential differences between soil, plant, and atmosphere.
  • The atmosphere typically has very low (very negative) water potential, creating a strong gradient.
  • Example: In a Shaanxi greenhouse, air water potential was -168 MPa while soil was -0.26 MPa, creating a large gradient driving transpiration.

🌫️ How humidity affects transpiration

  • When air is humidified, its water potential rises (becomes less negative), reducing the gradient.
  • Example: Fogging the greenhouse raised air water potential from -168 MPa to -82 MPa, reducing the driving gradient by ~50% and cumulative transpiration by 20%.
  • This illustrates that transpiration responds passively to atmospheric demand rather than being actively controlled by the plant.

🚪 Stomates and the water-carbon dioxide trade-off

🚪 What stomates are

Stomates: pores or openings on the exterior surfaces of leaves and other plant organs.

  • A pair of specialized cells called guard cells allow the plant to open and close stomates.
  • Stomates regulate the flow of gases (water vapor, carbon dioxide, oxygen) into and out of the plant.

⚖️ The unavoidable trade-off

  • Plants must open stomates to take in carbon dioxide from the atmosphere—the first step in photosynthesis.
  • But opening stomates unavoidably causes water vapor loss to the atmosphere.
  • Oxygen gas generated during photosynthesis also exits through stomates.
  • Don't confuse: stomates are not optional—plants cannot photosynthesize without opening them, so water loss is an unavoidable consequence of carbon gain.

🌱 Plant water balance and stress

🌱 What determines water status

  • Water status depends on the balance between transpiration (water loss) and root water uptake (water gain).
  • When transpiration exceeds uptake, water stored in plant tissues is depleted.
  • When uptake exceeds transpiration, water stores are replenished.

😰 Signs of water stress

  • When water stores are depleted, plants begin to shrink or wilt.
  • If depletion is prolonged, water potential inside the plant decreases.
  • Plants respond by reducing stomatal opening size to reduce transpiration rate.
  • Visible signs: rolled or wilted leaves, especially under high evaporative demand.

🌙 Recovery from temporary wilting

Plants can recover from wilting in two ways:

ConditionMechanismResult
Reduced evaporative demandSun goes down, transpiration rate drops below uptake rateWater stores replenished, water potential rises, turgor restored
Increased soil waterRainfall or irrigation raises soil water potentialRoot uptake increases even with ongoing evaporative demand

⚠️ Permanent wilting point

Permanent wilting point: the soil volumetric water content at which plant water stress reaches a critical severity and the plant cannot recover even if evaporative demand is removed or soil is rewetted.

  • The exact soil moisture state depends on plant species and evaporative demand.
  • Commonly estimated as the soil water content at a matric potential of -1500 kPa.
  • Don't confuse: temporary wilting (reversible) vs permanent wilting (irreversible damage has occurred).

🌿 Root water uptake limitations

🌿 Two limiting factors

Root water uptake rate can be limited by:

  1. Hydraulic conductivity of the rhizosphere soil (soil immediately adjacent to roots)
  2. Water potential gradient between soil and roots

🔄 Traditional conceptual model

The traditional model describes a dynamic process:

  • Root water uptake lowers rhizosphere soil water content.
  • Lower water content increases the hydraulic gradient between rhizosphere and bulk soil → acts to increase water flow toward roots.
  • But lower water content also decreases soil hydraulic conductivity → acts to decrease water flow to roots.
  • As long as the increased gradient effect offsets the decreased conductivity effect, root uptake can proceed at a steady rate.

📉 Inevitable decline in uptake rate

  • Eventually, rhizosphere soil water potential reaches near-equilibrium with root water potential.
  • From that point on, the hydraulic gradient can only decrease while hydraulic conductivity continues to decrease.
  • Result: rate of root water uptake starts to decline sharply.
  • This is analogous to the transition from constant-rate evaporation to falling-rate evaporation.

📊 Transpiration and bulk soil water relationship

📊 Practical monitoring challenges

  • We cannot easily monitor rhizosphere soil water status directly.
  • For practical purposes, we seek to understand the relationship between transpiration and bulk soil water.

🌽 Classic corn study example

A field experiment with corn grown in 136 large containers illustrates key features:

  • Containers were individually watered to achieve a range of soil water contents each day.
  • Well-watered corn transpiration rates varied from 1.4 mm per day up to ~7 mm per day as weather and evaporative demand changed.
  • Relative transpiration rate was defined as the ratio of transpiration rate in a particular container to the transpiration rate of well-watered corn on the same day.
  • The study showed how transpiration responds to both soil water content and atmospheric demand.
48

Root Water Uptake

11.3. Root Water Uptake

🧭 Overview

🧠 One-sentence thesis

Root water uptake rate is controlled by both the hydraulic conductivity of the rhizosphere soil and the water potential gradient between soil and roots, and as uptake proceeds, declining conductivity eventually causes the uptake rate to fall sharply.

📌 Key points (3–5)

  • Two limiting factors: root water uptake can be limited by either the hydraulic conductivity of the rhizosphere soil or the water potential gradient between soil and roots.
  • Traditional conceptual model: uptake lowers rhizosphere water content, which increases the gradient (boosting flow) but also decreases hydraulic conductivity (reducing flow).
  • Steady uptake condition: as long as the increased gradient offsets the decreased conductivity, uptake proceeds at a steady rate.
  • Common confusion: the rhizosphere soil (immediately adjacent to roots) vs. bulk soil—the rhizosphere is where uptake directly affects water content and conductivity, but we often monitor bulk soil for practical purposes.
  • Inevitable decline: when rhizosphere water potential nears equilibrium with root water potential, the gradient can only decrease while conductivity continues to drop, causing uptake rate to decline sharply.

🌱 What limits root water uptake

🚰 Two controlling factors

Root water uptake rate is limited by:

  • Hydraulic conductivity of the rhizosphere soil: how easily water moves through the soil immediately adjacent to the roots.
  • Water potential gradient: the difference in water potential between the soil and the roots.

Both factors interact to determine how much water the roots can take up at any moment.

🌾 Rhizosphere vs. bulk soil

Rhizosphere soil: the soil immediately adjacent to the roots.

  • The rhizosphere is the zone where root uptake directly affects soil water content and hydraulic properties.
  • For practical purposes, we often monitor bulk soil water (the soil farther from roots) because rhizosphere conditions are difficult to measure directly.
  • Don't confuse: changes in the rhizosphere drive uptake dynamics, but bulk soil measurements are what we typically use for management.

🔄 The traditional conceptual model

📉 How uptake changes the rhizosphere

When roots take up water:

  1. Water content of the rhizosphere soil decreases.
  2. This lowered water content has two opposing effects:
    • Increases the hydraulic gradient between the rhizosphere and the bulk soil → tends to increase water flow toward the roots.
    • Decreases the soil hydraulic conductivity (recall that conductivity falls as water content falls) → tends to decrease water flow to the roots.

⚖️ Steady uptake phase

  • As long as the increased gradient is adequate to offset the decreased conductivity, root water uptake can proceed at a steady rate.
  • In other words, the boost from a steeper gradient compensates for the loss of conductivity, so water continues to flow to the roots at a constant rate.

Example: Early in the uptake process, the rhizosphere dries slightly, creating a strong pull for water from the bulk soil, and conductivity is still high enough that water moves readily.

📉 Inevitable decline in uptake rate

🔻 Near-equilibrium and sharp decline

Eventually, a critical point is reached:

  • The water potential of the rhizosphere soil approaches near-equilibrium with the water potential of the roots.
  • From that time on:
    • The hydraulic gradient can only decrease (because the potentials are nearly equal).
    • The hydraulic conductivity continues to decrease (as the rhizosphere dries further).
  • Result: the rate of root water uptake starts to decline sharply.

🌊 Analogy to evaporation stages

The excerpt notes that this transition is analogous to the shift from the constant-rate stage of evaporation to the falling-rate stage:

  • Initially, uptake is steady (constant-rate).
  • Once the gradient can no longer compensate for falling conductivity, uptake drops off (falling-rate).

Don't confuse: this is not a sudden stop; it is a transition from steady uptake to declining uptake as the rhizosphere dries and equilibrates with the roots.

49

Transpiration and Soil Water

11.4. Transpiration and Soil Water

🧭 Overview

🧠 One-sentence thesis

Transpiration rate depends interactively on both soil water content and evaporative demand, so that the critical soil water content at which transpiration begins to decline is not fixed but varies with atmospheric conditions.

📌 Key points (3–5)

  • Interactive control: both soil water status and evaporative demand together determine transpiration rate, not soil water alone.
  • Relative transpiration rate: the ratio of actual transpiration to well-watered transpiration; it stays constant until soil water drops below a critical threshold, then declines linearly.
  • Critical threshold shifts: higher evaporative demand causes transpiration to decline at higher soil water content; lower demand allows plants to extract water down to lower soil water content before decline begins.
  • Common confusion: the "critical soil water content" is not a single fixed property of the soil—it depends on how much water the atmosphere is demanding.
  • Practical implication: monitoring bulk soil water content helps manage transpiration, but the relationship must account for day-to-day weather variation.

🌾 The classic corn experiment

🌾 Experimental design

  • Corn was grown in 136 large field containers, each watered individually to create a range of soil water contents every day.
  • Weather and evaporative demand varied day-to-day, so well-watered corn transpiration rates ranged from as low as 1.4 mm per day up to about 7 mm per day.
  • This setup allowed researchers to separate the effects of soil water content from the effects of atmospheric demand.

📏 Relative transpiration rate

Relative transpiration rate: the ratio of the transpiration rate for corn in a particular container to the transpiration rate of well-watered corn on the same day.

  • This metric normalizes for day-to-day weather differences.
  • It isolates the effect of soil water stress by comparing stressed plants to unstressed plants under identical atmospheric conditions.
  • Example: if well-watered corn transpires at 6 mm/day and a stressed plant transpires at 3 mm/day, the relative transpiration rate is 0.5 (or 50%).

🔄 How soil water and demand interact

🔄 Two-phase response pattern

  • Phase 1 (high soil water): relative transpiration rate stays constant at 1.0 (100%) as long as soil water content is above a critical threshold.
  • Phase 2 (low soil water): once soil water content falls below the critical level, relative transpiration rate declines approximately linearly with further decreases in soil water content.

🌡️ Critical threshold depends on evaporative demand

The excerpt provides two contrasting examples from the Colo silty clay loam soil:

Well-watered transpiration rateCritical soil water contentInterpretation
6.4 mm/day (high demand)0.34 cm³/cm³Transpiration begins to decline at relatively high soil water content
2.0 mm/day (low demand)0.25 cm³/cm³Transpiration does not decline until soil water content is much lower
  • Higher evaporative demand means the plant needs to extract water faster, so the hydraulic limitations of the soil kick in sooner (at higher water content).
  • Lower evaporative demand means slower extraction, so the plant can continue drawing water even when soil water content is lower.
  • Don't confuse: the critical threshold is not a fixed soil property like field capacity or wilting point—it shifts with atmospheric conditions.

🧩 Why the threshold shifts

  • The excerpt explains that root water uptake depends on both the hydraulic gradient (driving force) and the soil hydraulic conductivity (ease of flow).
  • As soil dries, conductivity drops sharply (recall the soil hydraulic conductivity curve).
  • When evaporative demand is high, the plant tries to extract water quickly, but the low conductivity of drier soil cannot supply water fast enough, so transpiration declines at higher water content.
  • When evaporative demand is low, the slower extraction rate can be sustained even at lower conductivity, so the plant can use water down to lower soil water content before hitting the limit.

🌍 Practical implications

🌍 Monitoring bulk soil water

  • The excerpt notes that "we cannot easily monitor the water status of the rhizosphere soil itself."
  • For practical purposes, we monitor bulk soil water content instead.
  • The relationship between bulk soil water and transpiration must account for:
    • Day-to-day variation in weather and evaporative demand.
    • Soil-specific hydraulic properties (water retention and conductivity curves).
    • Plant species differences (not detailed in this excerpt, but mentioned as a factor).

🔗 Connection to root water uptake dynamics

  • The excerpt draws an analogy to the transition from constant-rate to falling-rate evaporation.
  • As the rhizosphere soil dries, the hydraulic gradient initially increases (favoring flow), but conductivity decreases (opposing flow).
  • Eventually, the gradient can no longer increase enough to offset the conductivity decline, and uptake rate falls sharply.
  • This mechanism underlies the two-phase transpiration response seen in the corn data.

📊 Soil water availability indicators

📊 Plant available water (PAW)

Plant available water (PAW): the equivalent depth of water available for plant uptake within a layer of soil of specified thickness.

  • Formula (in words): PAW equals the current soil volumetric water content minus the volumetric water content at permanent wilting point, multiplied by the thickness of the soil layer.
  • Units: length (e.g., millimeters), the same as precipitation or irrigation depth.
  • Purpose: PAW normalizes soil water content to account for soil-specific differences in water retention, making it easier to compare water availability across different soils.

🔧 Why normalization matters

  • The excerpt emphasizes that "the specific soil water content at which the relative transpiration rate begins to decline depends... on the soil water retention characteristics."
  • Different soils hold water differently, so the same volumetric water content may represent very different availability to plants.
  • By subtracting the wilting point and scaling by layer thickness, PAW provides a more universal measure of how much water the plant can actually use.
  • Don't confuse: PAW is not the same as current soil water content—it is the usable portion of that water, adjusted for the soil's retention curve and the depth of the root zone.
50

Soil Water Availability Indicators

11.5. Soil Water Availability Indicators

🧭 Overview

🧠 One-sentence thesis

Soil water availability indicators like plant available water (PAW) and fraction of available water capacity (FAW) normalize soil water content across different soil types to better predict when plants will experience water stress.

📌 Key points (3–5)

  • Why normalization matters: the soil water content at which plants experience stress depends on both evaporative demand and soil-specific water retention characteristics, so raw volumetric water content alone is insufficient.
  • Plant available water (PAW): the depth of water available for plant uptake in a soil layer, calculated by subtracting the permanent wilting point from current water content and multiplying by layer thickness.
  • Fraction of available water capacity (FAW): the proportion of a soil's water storage capacity that is currently filled; FAW = 0.5 is often used as an approximate threshold for plant water stress.
  • Common confusion: available water capacity (AWC) vs. plant available water (PAW)—AWC is the capacity (maximum storage between field capacity and wilting point), while PAW is the current amount available.
  • Real-world application: FAW can predict plant water stress severity, such as the probability of large wildfires during the growing season.

📏 Core indicators and definitions

📏 Plant available water (PAW)

Plant available water (PAW): the equivalent depth of water available for plant uptake within a layer of soil of specified thickness.

Formula (in words):
PAW = (current volumetric water content − volumetric water content at permanent wilting point) × soil layer thickness

  • Units: length (e.g., mm), the same as precipitation, making it easy to understand.
  • Why it helps: by accounting for the soil-specific permanent wilting point, PAW provides a better indicator of water status across different soil types than raw volumetric water content.
  • Example: two soils may have the same volumetric water content, but if one has a higher wilting point, it has less water actually available to plants.

📏 Available water capacity (AWC)

Available water capacity (AWC): the capacity of a specific soil to store plant available water.

Formula (in words):
AWC = (volumetric water content at field capacity − volumetric water content at permanent wilting point) × soil layer thickness

  • What it represents: the maximum amount of water a soil can hold for plant use, not the current amount.
  • Practical value: despite shortcomings of the field capacity concept, AWC has proven to be a practical indicator of differences between soils in their capacity to store water for plant uptake.
  • Don't confuse: AWC is the storage capacity (a soil property), while PAW is the current available water (a state variable).

📏 Fraction of available water capacity (FAW)

Formula (in words):
FAW = (current volumetric water content − volumetric water content at permanent wilting point) ÷ (volumetric water content at field capacity − volumetric water content at permanent wilting point)

  • What it shows: the proportion of the soil's water storage capacity that is currently filled.
  • Range: 0 (at wilting point) to 1 (at field capacity).
  • Why it's useful: FAW normalizes across soil types, making it easier to compare water stress levels in different soils.

🌱 Using FAW to predict plant water stress

🌱 The FAW = 0.5 threshold

  • Approximate rule: FAW = 0.5 is often used as a threshold below which plant water stress occurs.
  • Variability: there is substantial variability in the exact FAW value at which stress begins, depending on plant species and conditions.
  • How to apply: if soil volumetric water content measurements and soil water retention properties are available, FAW can be used to estimate or monitor the degree of plant water stress.

🔥 Real-world example: wildfire prediction in Oklahoma

The excerpt describes how FAW was used to predict large growing-season wildfires in Oklahoma from 2000–2012:

  • Mechanism: large fires during the growing season only occur if vegetation is under severe water stress, reducing vegetation water content to levels low enough that it will readily burn.
  • Findings: the vast majority of large growing-season fires (159 of 174 fires) occurred when FAW dropped below 0.5, indicating water stress.
  • Severity breakdown:
    • 22 fires occurred under severe drought (FAW between some lower threshold and 0.5).
    • 134 fires occurred under extreme drought (FAW below the severe drought threshold).
  • Example: monitoring FAW allows fire managers to predict when conditions are ripe for large wildfires, even during the growing season when vegetation is normally green.

🔄 Why soil-specific adjustments are necessary

🔄 Interaction of evaporative demand and soil properties

The excerpt emphasizes that the specific soil water content at which relative transpiration rate begins to decline depends on:

  1. Evaporative demand: higher atmospheric demand means plants experience stress at higher soil water contents.
  2. Plant species: different species have different stress thresholds.
  3. Soil water retention characteristics: different soils hold water with different strengths at the same volumetric water content.
  • Why raw volumetric water content is insufficient: a volumetric water content of 0.30 cm³/cm³ might be adequate in one soil but cause stress in another, depending on how tightly that soil holds water.
  • Solution: normalizing by permanent wilting point and field capacity (via PAW, AWC, and FAW) accounts for soil-specific differences.

🔄 Comparison of indicators

IndicatorWhat it measuresUnitsUse case
Volumetric water content (θ)Current water contentcm³/cm³ (dimensionless)Basic measurement, but not comparable across soils
Plant available water (PAW)Current depth of water available to plantsmmEasy to compare with precipitation; accounts for wilting point
Available water capacity (AWC)Maximum storage capacity for plant-available watermmCharacterizes soil's storage potential
Fraction of available water capacity (FAW)Proportion of storage capacity currently filleddimensionless (0–1)Best for predicting stress across soil types
  • Don't confuse: θ is a raw measurement; PAW adjusts for wilting point; AWC is the maximum PAW; FAW is the current PAW as a fraction of AWC.
51

Water Use Efficiency

11.6. Water Use Efficiency

🧭 Overview

🧠 One-sentence thesis

Water use efficiency (WUE)—the ratio of carbon or biomass accumulation to water use—is easily misunderstood and may be overemphasized, because increasing intrinsic WUE does not necessarily lead to higher crop yields or reduced agricultural water use.

📌 Key points (3–5)

  • What WUE measures: the ratio of carbon or biomass accumulation to water use or supply during a given time span, but it can be defined in many different ways (carbon vs biomass, transpiration vs total water, instantaneous vs seasonal).
  • Fundamental constraint: stomates are not selective—when they open to let CO₂ in, they necessarily let water vapor out, so WUE depends primarily on the gradients of CO₂ and water vapor.
  • Three main factors: WUE is controlled by (1) photosynthetic pathway (C₄ > C₃), (2) atmospheric CO₂ concentration, and (3) vapor pressure deficit.
  • Common confusion: higher intrinsic WUE does not automatically mean higher yields or lower water use; increasing transpiration (not WUE) may be more effective for improving crop productivity in water-limited environments.
  • Rising atmospheric CO₂: increasing CO₂ concentration is raising plant WUE, especially in C₃ plants and trees, though other climate changes may counteract this effect.

🔗 The transpiration–photosynthesis linkage

🔗 Why WUE matters across Earth systems

  • Transpiration links three critical Earth cycles:
    • Soil water balance
    • Surface energy balance
    • Atmospheric carbon balance
  • This linkage has attracted significant research attention and public interest.
  • The excerpt notes that WUE is "an important concept that can be easily misunderstood and overemphasized."

🌿 The stomate constraint

Stomates are not selective. When they open to let carbon dioxide in, they necessarily let water vapor out.

  • CO₂ flows into the leaf driven by the difference between atmospheric CO₂ concentration and internal leaf CO₂ concentration.
  • Water vapor flows out of the leaf driven by the difference between internal leaf water vapor concentration and atmospheric water vapor concentration.
  • Don't confuse: WUE is not about "how much water the plant needs" in isolation; it is about the unavoidable trade-off between carbon gain and water loss through the same openings.

📐 Defining and measuring WUE

📐 Multiple definitions

Water use efficiency can be expressed in many combinations:

ComponentOptions
Accumulation• CO₂ assimilation<br>• Above-ground biomass<br>• Harvested yield
Water consumed• Transpiration<br>• Evapotranspiration<br>• Total water supply
Time period• Instantaneous<br>• Seasonal<br>• Annual
  • This variety of definitions is one reason WUE is easily misunderstood.
  • Example: comparing "instantaneous leaf-level CO₂ assimilation per transpiration" vs "seasonal harvested yield per total water supply" can lead to very different conclusions.

🧮 Leaf-level instantaneous WUE formula

The excerpt provides an equation (Eq. 11-4) for instantaneous leaf-level WUE:

  • WUE (ratio of carbon assimilated A to transpiration T) is approximately:
    • c × (atmospheric CO₂ partial pressure P_a) / (saturation vapor pressure at leaf temperature e*_l − atmospheric water vapor pressure e)
  • c is a constant depending on the ratio of internal to external CO₂ concentrations.
  • This formula shows that WUE depends on CO₂ concentration and vapor pressure deficit.

🌾 Factors controlling WUE

🌾 Photosynthetic pathway (C₃ vs C₄ vs CAM)

  • C₃ plants (about 85% of plant species, including wheat, rice, soybean):
    • The carbon-fixing enzyme rubisco is exposed to relatively low internal CO₂ concentrations.
    • Carbon fixation is relatively inefficient.
    • Constant c ≈ 0.3–0.4.
    • Lower WUE.
  • C₄ plants (about 3% of plant species, including corn, sorghum, sugarcane):
    • Special cells ensure rubisco is exposed to higher internal CO₂ concentrations.
    • Carbon fixation efficiency is enhanced.
    • Constant c ≈ 0.7.
    • Substantially higher WUE than C₃ plants.
  • CAM plants (e.g., pineapple):
    • Open stomates at night.
    • Achieve high WUE by avoiding daytime vapor pressure deficits.

Implication: Crop WUE can be increased by using more C₄ crops and fewer C₃ crops.

🌍 Atmospheric CO₂ concentration

  • As atmospheric CO₂ concentration (P_a) increases, WUE increases.
  • Higher external CO₂ concentration drives a greater rate of CO₂ diffusion into leaves when stomates are open.
  • Due mainly to fossil fuel combustion, atmospheric CO₂ is steadily rising (Fig. 11-6 shows the Mauna Loa record).
  • Evidence: Northern Hemisphere forests have shown a substantial increase in WUE over the past two decades.
  • Who benefits more:
    • Trees may benefit more than other plant types.
    • C₃ crops may benefit more than C₄ crops (Table 11-1 shows relative WUE increase with CO₂ enrichment).
  • Caveat: Other climate changes (e.g., increasing air temperatures) could reduce or prevent the expected WUE increase.

💧 Vapor pressure deficit

  • WUE is higher when the atmospheric vapor pressure deficit (e*_l − e) is relatively low.
  • Lower deficit → smaller gradient driving water vapor diffusion out of leaves → higher WUE.
  • Agricultural implications:
    • Focus crop production in more humid climates (lower vapor pressure deficit).
    • Shift growing seasons earlier so more growth occurs during cooler parts of the year (lower vapor pressure deficit).
    • Trade-off: earlier growing seasons may increase freeze damage risk or slow growth due to low temperatures.

⚠️ Limits and misconceptions about WUE

⚠️ Little scope for fundamental alteration

  • The factors controlling WUE in plants are relatively well-understood.
  • There appears to be little scope for altering WUE at a fundamental level.
  • The stomate constraint is unavoidable: CO₂ in and water vapor out are coupled.

⚠️ Weak justification for breeding for higher intrinsic WUE

  • Major research investments have been made to study intrinsic WUE and increase it through plant breeding or genetic engineering.
  • However: the logical justification for this work may not be altogether solid.
  • No clear evidence that increasing intrinsic WUE will result in:
    • Increased crop production, or
    • More drought-resistant crops, or
    • Reduced water use in agriculture.

🚜 A better focus: increasing transpiration, not WUE

  • A stronger case can be made for focusing research on improving crop yields by increasing crop water use (transpiration), not WUE per se.
  • Why: There is a strong, linear relationship between plant biomass accumulation and transpiration.
  • Crop varieties that sustain higher rates of transpiration through effective use of soil water—especially during critical reproductive periods—will be more productive in water-limited environments.
  • Don't confuse: "water use efficiency" sounds like it should always be maximized, but in practice, using more water effectively (higher transpiration) can lead to better yields than simply raising the ratio.
52

12.1. Modes of Heat Transfer

12.1. Modes of Heat Transfer

🧭 Overview

🧠 One-sentence thesis

The surface energy balance depends on three fundamental modes of heat transfer—radiation, conduction, and convection—with radiation often being the largest single energy flux term at the land surface.

📌 Key points (3–5)

  • Three modes of heat transfer: radiation (electromagnetic waves/particles), conduction (random thermal motion within a system), and convection (bulk fluid motion).
  • Radiation is dominant: in the surface energy balance, the energy flux via radiation is often the largest single term.
  • Heat transfer direction: heat moves from higher temperature to lower temperature, analogous to how water flows from higher to lower water potential.
  • Temperature as a measure: temperature measures the average kinetic energy of random microscopic movements of molecules or particles.
  • Context: understanding these modes is essential for understanding the surface energy balance, which accounts for heat transfer between the land surface and its surroundings.

🌡️ Heat transfer fundamentals

🌡️ What temperature means

Temperature: a measure of the average kinetic energy of the random microscopic movements of the molecules or particles of a substance.

  • Temperature is not just "how hot" something feels; it is a physical quantity tied to molecular motion.
  • Temperature determines the direction of heat transfer between two substances in thermal contact.
  • Analogy: temperature is to heat transfer what water potential is to water flow—both determine direction of movement.

⚡ What heat transfer is

Heat transfer: energy transfer that depends on the temperatures of the objects or systems involved.

  • Heat always moves from regions of higher temperature to regions of lower temperature.
  • The surface energy balance is essentially a way of accounting for heat transfer between the land surface and its surroundings.

🔄 The three modes of heat transfer

📡 Radiation

Radiation: the emission or transfer of energy in the form of electromagnetic waves or particles.

  • This mode involves energy moving through space without requiring a physical medium.
  • In the surface energy balance, radiation is often the largest single energy flux term.
  • The excerpt emphasizes that this mode will be considered first because of its importance.

🧱 Conduction

Conduction: heat transfer due to the random thermal motion of the molecules and particles within a system.

  • Energy moves through direct contact within a material or between materials in contact.
  • This mode relies on molecular-scale random motion, not bulk movement.

🌊 Convection

Convection: heat transfer due to bulk fluid motion.

  • Unlike conduction, convection involves the physical movement of fluid (liquid or gas).
  • Energy is carried by the moving fluid itself.
  • Don't confuse: conduction is random molecular motion within a system; convection is organized bulk motion of the fluid.

🌍 Role in the surface energy balance

🌍 Energy balance context

The excerpt places these three modes within the broader framework of the surface energy balance, which includes:

  • Radiation received at the land surface
  • Exchange of sensible heat between surface, atmosphere, and soil
  • Consumption of energy (latent heat) during evapotranspiration

🔑 Why all three matter

ModeRole in surface energy balance
RadiationOften the largest single energy flux term
ConductionTransfers heat within soil and between surface and soil
ConvectionTransfers heat between surface and atmosphere via bulk air motion

Each mode plays an important role, but the excerpt highlights radiation as the starting point for detailed examination because of its dominant contribution.

53

Radiation Basics

12.2. Radiation Basics

🧭 Overview

🧠 One-sentence thesis

The sun and Earth constantly emit electromagnetic radiation at different wavelengths determined by their temperatures, with the sun emitting shortwave radiation and the Earth emitting longwave radiation, both of which are critical components of the surface energy balance.

📌 Key points (3–5)

  • Temperature determines wavelength: hotter objects emit radiation at shorter wavelengths (Wien's displacement law); the sun peaks in visible light, Earth in infrared.
  • Shortwave vs longwave distinction: solar radiation is called "shortwave," while radiation from Earth's surface and atmosphere is called "longwave."
  • Intensity depends strongly on temperature: the Stefan-Boltzmann Law shows that radiant flux increases with the fourth power of absolute temperature.
  • Emissivity varies by surface: a measure (0 to 1) of how effectively a surface emits radiation; snow/ice/sun are nearly perfect (~0.99), while soil varies with moisture (0.86–0.96).
  • Common confusion: only about half of solar radiation reaching the top of the atmosphere actually arrives at Earth's surface—the rest is reflected or absorbed by clouds, air molecules, dust, and aerosols.

🌡️ How temperature controls radiation

🌡️ Wien's displacement law

Wien's displacement law: describes the relationship between the wavelength of maximum radiation intensity (λ max) and an object's absolute surface temperature (T) for a perfect emitter (blackbody).

  • The relationship is inverse: hotter objects emit at shorter wavelengths.
  • Formula in words: λ max equals b divided by T, where b = 2,900 μm K.
  • This law explains why different objects emit radiation in different parts of the electromagnetic spectrum.

☀️ Sun vs Earth emission

ObjectSurface TemperaturePeak WavelengthSpectrum Region
Sun~5,780 K~0.5 μmGreen (visible)
Earth's surface~287 K~10 μmInfrared
  • The sun's much higher temperature means it emits at much shorter wavelengths.
  • Example: The sun peaks in the visible green range, which is why we can see sunlight; Earth peaks in infrared, which is invisible to our eyes.
  • Don't confuse: "shortwave" and "longwave" are context-specific terms in surface energy balance—shortwave refers to solar radiation, longwave to terrestrial radiation.

🔥 Radiation intensity and emissivity

🔥 Stefan-Boltzmann Law

Stefan-Boltzmann Law: describes the relationship between an object's temperature and the intensity of its radiation.

  • Formula in words: total radiant flux (J_t, in W per square meter) equals emissivity (ε) times the Stefan-Boltzmann constant (σ = 5.67 × 10^-8 W per square meter per K to the fourth power) times absolute temperature (T in K) to the fourth power.
  • The intensity depends on temperature raised to the fourth power, meaning even small temperature changes produce large changes in emitted radiation.
  • This relationship is much stronger than the wavelength-temperature relationship.

📏 What emissivity means

Emissivity: a measure, ranging from 0 to 1, of the effectiveness of a surface in emitting radiation.

  • A perfect emitter (blackbody) has emissivity = 1.
  • Higher emissivity = more effective at emitting radiation for a given temperature.
  • Emissivity is not a fixed property—it can vary with surface conditions (e.g., soil moisture).

🌍 Emissivity of different surfaces

Surface TypeEmissivity RangeNotes
Sun, snow, ice~0.99Nearly perfect emitters
Water0.98–0.99Very high
Vegetated surfaces0.95–0.98For infrared (8–13 mm wavelengths)
Soil0.86–0.96Positively related to water content
  • The composite emissivity of Earth's land surface is highest at the north and south poles (snow/ice) and lowest in mid-latitude deserts (dry soil).
  • Example: Wet soil emits more effectively than dry soil at the same temperature because its emissivity is higher.
  • Don't confuse: emissivity with temperature—a cooler surface with high emissivity can emit more radiation than a slightly warmer surface with low emissivity.

☁️ What happens to solar radiation in the atmosphere

☁️ Only half reaches the surface

  • On average, just over half (~53%) of solar radiation arriving at the top of Earth's atmosphere actually reaches the land surface.
  • The rest is reflected or absorbed by atmospheric components before it can reach the ground.

🌥️ Breakdown of solar radiation fate

ProcessPercentageWhat happens
Reflected by clouds~25%Bounced back to space
Reflected by air/dust/aerosols~5%Bounced back to space
Absorbed by air/dust/aerosols~14%Heats the atmosphere
Absorbed by clouds~3%Heats clouds
Direct to surface~31%Reaches ground directly
Diffuse to surface~22%Reaches ground after scattering/reflection
  • Total reaching surface: 31% (direct) + 22% (diffuse) = 53%.
  • Example: On a cloudy day, most of the radiation reaching the surface is diffuse rather than direct.
  • Don't confuse: "reflected" (sent back to space) with "scattered" (redirected but may still reach the surface as diffuse radiation).
54

Net Radiation

12.3. Net Radiation

🧭 Overview

🧠 One-sentence thesis

Net radiation—the sum of all incoming and outgoing radiation fluxes at the land surface—is typically the largest term in the surface energy balance and determines how much energy is available to drive evapotranspiration, heat the air, and warm the soil.

📌 Key points (3–5)

  • What net radiation is: the sum of all incoming and outgoing radiation fluxes (shortwave and longwave) at the land surface.
  • Albedo controls shortwave reflection: the fraction of incoming solar radiation reflected by the surface; dark surfaces (low albedo) absorb more, light surfaces (high albedo) reflect more.
  • Longwave radiation matters too: the atmosphere emits longwave radiation downward to the surface, with magnitude depending on atmospheric temperature, emissivity, clouds, and water vapor.
  • Common confusion: net radiation is not just incoming solar radiation—it includes reflected shortwave, incoming longwave from the atmosphere, and outgoing longwave from the surface.
  • Why it matters: net radiation is the energy source for the surface energy balance, partitioned into latent heat (evapotranspiration), sensible heat (air warming), and soil heat fluxes.

☀️ Shortwave radiation and albedo

☀️ What happens to incoming solar radiation

  • When direct and diffuse shortwave (solar) radiation reach the land surface, some is reflected by the surface.
  • The reflected portion depends on the surface's albedo.

🪞 Albedo: the reflection fraction

Albedo: the fraction of incoming shortwave radiation that is reflected by the surface.

  • Albedo is a dimensionless number (0 to 1).
  • Dark surfaces have low albedo (absorb more radiation).
  • Light-colored surfaces have high albedo (reflect more radiation).

🌨️ Albedo values for different surfaces

SurfaceAlbedo rangeNotes
Fresh snowUp to 0.9Reflects large portions of incoming shortwave radiation
WaterTypically < 0.1Can be much higher when sun angle is low
Soils and vegetation0.1 to 0.4Intermediate values
  • Moist soil surfaces are typically darker and have lower albedo than dry soil surfaces.
  • Soil organic matter darkens soil, resulting in lower albedo.
  • Example: a wet, organic-rich soil absorbs more solar radiation than a dry, light-colored soil.

🌡️ Longwave radiation exchange

🌡️ The atmosphere emits longwave radiation downward

  • The land surface emits longwave radiation upward to the atmosphere (as seen previously).
  • But the atmosphere also emits longwave radiation both upward into space and downward toward the Earth's surface.
  • This downward longwave radiation is often overlooked but is essential for the net radiation calculation.

🌫️ What controls atmospheric longwave emission

  • The magnitude of downward longwave radiation depends on:
    • Atmospheric temperature
    • Atmospheric emissivity (according to Eq. 12-2 in the excerpt)
  • Atmospheric emissivity can range from 0.5 to nearly 1.

☁️ Clouds and water vapor increase emissivity

  • Clouds significantly increase the emissivity of the atmosphere.
    • This is why cloudy winter nights are often warmer than clear winter nights—more longwave radiation is emitted downward, warming the surface.
  • Water vapor in the atmosphere also strongly affects emissivity.
    • Higher humidityhigher emissivity → more downward longwave radiation.
  • Don't confuse: atmospheric emissivity is not constant; it varies with cloud cover and humidity.

🧮 Calculating net radiation

🧮 The net radiation equation

Net radiation (R_n): the sum of all incoming and outgoing radiation fluxes at the land surface.

  • Mathematically (Eq. 12-3 in the excerpt):
    • Net radiation = (1 minus albedo) times incoming solar radiation, plus incoming longwave, minus outgoing longwave.
    • In words: R_n = (1 - α) × R_s + R_li - R_lo
    • Where:
      • α = albedo
      • R_s = incoming solar radiation (direct and diffuse)
      • R_li = incoming longwave radiation
      • R_lo = outgoing longwave radiation

🔍 Breaking down the terms

  • (1 - α) × R_s: the portion of incoming solar radiation that is absorbed (not reflected) by the surface.
  • + R_li: add the longwave radiation emitted downward by the atmosphere.
  • - R_lo: subtract the longwave radiation emitted upward by the land surface.
  • Example: if albedo is 0.2, then 80% of incoming solar radiation is absorbed; the net radiation also includes the longwave exchange.

⚠️ Don't confuse net radiation with incoming solar radiation

  • Net radiation is not just the solar radiation hitting the surface.
  • It accounts for:
    • Reflection (via albedo)
    • Incoming longwave from the atmosphere
    • Outgoing longwave from the surface
  • This is why net radiation can be negative at night (when there is no incoming solar radiation but the surface still emits longwave radiation).

🔋 Net radiation in the surface energy balance

🔋 Net radiation is the largest energy term

  • The excerpt states: "Typically the largest term in the surface energy balance is the net radiation."
  • Net radiation provides the energy that is partitioned into other fluxes.

⚖️ The surface energy balance equation

  • The surface energy balance (Eq. 12-4 in the excerpt) is:
    • R_n = LE + H + G
    • Where:
      • R_n = net radiation
      • LE = latent heat flux (energy used for evapotranspiration)
      • H = sensible heat flux (heat transfer to the atmosphere)
      • G = soil heat flux (heat transfer into the soil)
  • Fluxes are commonly expressed in watts per square meter (W m⁻²).

🔄 Sign convention

  • Net radiation is defined as positive towards the surface.
  • All other terms (LE, H, G) are defined as positive away from the surface.
  • Daytime: net radiation is typically towards the surface; LE, H, and G are away from the surface.
  • Nighttime: fluxes are in the opposite directions.
  • Latent heat flux is typically near 0 at night, although negative values (indicating condensation) are possible.

🌞 Example: daytime and nighttime patterns

  • The excerpt describes hypothetical daytime and nighttime directions for a moist land surface in summer (Fig. 12-5):
    • Daytime: net radiation is positive (towards surface); LE, H, and G are positive (away from surface).
    • Nighttime: net radiation is negative (away from surface); H and G reverse direction (towards surface); LE is near 0.
  • Example measurements (Fig. 12-6): net radiation ranged from -50 W m⁻² at night to +300 W m⁻² around noon; sensible heat flux peaked around noon at +200 W m⁻².
55

Surface Energy Balance

12.4. Surface Energy Balance

🧭 Overview

🧠 One-sentence thesis

The surface energy balance partitions net radiation into latent heat, sensible heat, and soil heat fluxes, with the direction and magnitude of each flux depending on whether it is day or night and on surface conditions like moisture and vegetation cover.

📌 Key points (3–5)

  • Net radiation is the largest term: it is the sum of all incoming and outgoing radiation fluxes at the land surface.
  • Four main components: the energy balance equation includes net radiation (Rn), latent heat flux (LE), sensible heat flux (H), and soil heat flux (G).
  • Sign convention matters: net radiation is positive toward the surface; the other three fluxes are positive away from the surface.
  • Day vs night reversal: during daytime, net radiation is typically toward the surface and other fluxes away; at night, directions reverse.
  • Common confusion: the latent heat flux is not simply "evaporation"—it is the energy absorbed during evaporation/transpiration, equal to the latent heat of vaporization multiplied by the evapotranspiration rate.

🌐 Net radiation

🌐 What net radiation is

Net radiation: the sum of all incoming and outgoing radiation fluxes at the land surface.

  • It is typically the largest term in the surface energy balance.
  • Mathematically: net radiation (Rn) = (1 − α) × Rs + Rli − Rlo
    • α = albedo
    • Rs = incoming solar radiation (direct and diffuse)
    • Rli = incoming longwave radiation
    • Rlo = outgoing longwave radiation

☁️ Atmospheric longwave radiation

  • The atmosphere emits longwave radiation both upward into space and downward toward the Earth's surface.
  • Downward longwave radiation depends on atmospheric temperature and emissivity.
  • Atmospheric emissivity ranges from 0.5 to nearly 1.
  • Clouds increase emissivity: cloudy winter nights are often warmer than clear winter nights because clouds increase downward longwave radiation.
  • Water vapor increases emissivity: more humid conditions lead to higher emissivity values.

⚖️ The surface energy balance equation

⚖️ The equation and its terms

The surface energy balance is:

Rn = LE + H + G

TermNameWhat it representsUnits
RnNet radiationSum of all incoming and outgoing radiationW m⁻²
LELatent heat fluxEnergy absorbed during evaporation/transpirationW m⁻²
HSensible heat fluxHeat transfer between surface and atmosphereW m⁻²
GSoil heat fluxHeat transfer between surface and underlying soilW m⁻²

🔄 Sign convention

  • Net radiation (Rn): positive toward the surface.
  • All other terms (LE, H, G): positive away from the surface.
  • This convention is important for interpreting measurements and understanding energy flow direction.

🌞 Daytime vs nighttime patterns

🌞 Daytime behavior

  • Net radiation is typically toward the surface (positive).
  • The other fluxes (LE, H, G) are away from the surface (positive).
  • Example: on a summer day over moist land, energy arrives at the surface via net radiation and is partitioned into evapotranspiration (LE), heating the air (H), and warming the soil (G).

🌙 Nighttime behavior

  • Fluxes are in the opposite directions compared to daytime.
  • Net radiation is typically away from the surface (negative) due to radiative cooling.
  • Latent heat flux is typically near 0 during the night, although negative values (indicating condensation) are possible.
  • Sensible heat flux is often toward the surface because the surface cools faster than the atmosphere.

🔥 The three energy flux components

💧 Latent heat flux (LE)

Latent heat flux: the energy that is absorbed by water at the Earth's surface during evaporation or transpiration apart from any change in temperature.

  • It is a flux (transfer of energy) because the resulting water vapor is transported away from the surface by diffusion and advection.
  • How to calculate: LE = latent heat of vaporization for water × evapotranspiration (ET) rate.
  • Don't confuse: LE is not the amount of water evaporated; it is the energy consumed by that evaporation.

🌡️ Sensible heat flux (H)

Sensible heat flux: the heat transfer between the surface and the atmosphere by conduction and convection.

  • Daytime: the land surface is often warmer than the atmosphere, so the surface heats the air.
    • Energy is transferred from surface to air by conduction (direct contact).
    • Wind moves that air away, transferring energy to the atmosphere by convection.
  • Nighttime: the surface is often cooler than the atmosphere (due to radiative cooling), so the processes reverse and sensible heat flux is toward the surface.

🪨 Soil heat flux (G)

Soil heat flux: the heat transfer between the surface and the underlying soil, predominantly by conduction.

  • Radiation is negligible for subsurface heat transfer.
  • Convective heat transfer by flowing water can be important during high water flow rates, but these are not the norm.
  • The excerpt notes that soil heat flux will be examined in more detail in the next chapter on soil temperature.

📊 Real-world example: Iowa corn field

📊 Bowen ratio measurements

  • Researchers used energy balance measurements (based on a Bowen ratio technique) to study how crop residues affected heat transfers, soil moisture, and temperature in an Iowa corn field during fall and spring.
  • This example illustrates how the energy balance equation is applied in practice.

📊 Observed flux magnitudes

Flux componentRange observedNotes
Net radiation (Rn)−50 to +300 W m⁻²Negative at night, peaked around noon
Sensible heat flux (H)Peaked at +200 W m⁻² around noonSecond-largest component
Latent heat flux (LE)Did not exceed +100 W m⁻²Suppressed by crop residue
Soil heat flux (G)Did not exceed +100 W m⁻²Buffered by crop residue
  • Effect of crop residue: suppressed evaporation (lower LE) and buffered soil surface temperature (lower G).
  • Example: the presence of residue changed the partitioning of energy, reducing the latent and soil heat fluxes compared to bare soil.
56

Reference Evapotranspiration

12.5. Reference Evapotranspiration

🧭 Overview

🧠 One-sentence thesis

Reference evapotranspiration (ET₀) provides a standardized estimate of water loss from a well-watered surface under prevailing atmospheric conditions, enabling site-specific actual evapotranspiration calculations through adjustments based on local management and crop characteristics.

📌 Key points (3–5)

  • What ET₀ represents: the evapotranspiration rate expected for a standard well-watered surface, independent of site-specific management practices.
  • Why ET₀ is useful: it can be determined for large regions and then adjusted to estimate actual ET for specific locations when adequate information is available.
  • Two main estimation methods: the Hargreaves method (simple, low data requirements) and the Penman-Monteith method (more precise, more data-intensive).
  • Common confusion: actual ET vs reference ET—actual ET can be substantially less than ET₀ during dry periods but can exceed ET₀ during wet periods.
  • How to distinguish methods: Hargreaves requires only temperature data and is suited for 5-day or longer periods; Penman-Monteith requires net radiation, vapor pressure, wind speed, and other variables but provides more reliable estimates.

🌱 What reference evapotranspiration means

🌱 Definition and purpose

Reference evapotranspiration (ET₀): the evapotranspiration rate expected for a standard well-watered surface under the prevailing atmospheric conditions.

  • Actual evapotranspiration for a given time and location can be substantially influenced by site-specific management practices.
  • ET₀ removes these local variations by defining a standard surface.
  • Once ET₀ is determined, it can be adjusted to estimate actual ET for a specific location if adequate information is available.

🔄 Relationship to actual ET

  • The excerpt shows that actual ET can differ significantly from ET₀ depending on water availability.
  • Example: During dry periods (summers of 2011 and 2012 in Oklahoma), actual ET was substantially less than ET₀; during wet periods (summer of 2013), actual ET exceeded ET₀.
  • Methods for estimating ET₀ or ET are commonly based on the surface energy balance.

🧮 The Hargreaves method

🧮 Design and purpose

  • One of the simplest approaches for estimating ET₀.
  • Developed to facilitate irrigation management and water planning.
  • Initial work began in California; the importance of having a simple approach with low data requirements was solidified when the work expanded to Haiti.

📊 How it works

The Hargreaves equation is:

ET₀ = 0.0023 × Rₐ × (T̄ + 17.8) × √(Tₘₐₓ - Tₘᵢₙ)

Where:

  • Rₐ is the cumulative extraterrestrial radiation for the calculation period converted to units of mm of water
  • T̄ is the mean temperature (°C) for the calculation period
  • Tₘₐₓ is mean daily maximum temperature for the calculation period
  • Tₘᵢₙ is mean daily minimum temperature for the calculation period

🔑 Key advantages

  • Low data requirements: Since Rₐ can be accurately estimated based on site latitude and day of year, and air temperature measurements are widely available, the Hargreaves method can be used virtually anywhere.
  • When to use it: This is a good choice for ET₀ estimation when the only meteorological data available are air temperatures.
  • Time scale: Designed for use at 5-day or larger calculation periods; frequently used with monthly calculation periods.

📈 Development rationale

  • Hargreaves found that the product of incoming solar radiation (Rₛ) and air temperature (T) explained much of the variance in measured ET₀.
  • The problem was that measurements of Rₛ are often unavailable.
  • Solution: Hargreaves and colleagues developed a way to estimate Rₛ from knowledge of extraterrestrial radiation (Rₐ, solar radiation at the top of Earth's atmosphere for a specific place and time) and air temperature.

🌍 Application example

  • The Hargreaves method was used to estimate a 45-year time series of ET₀ for a large watershed stretching across southwest Oklahoma and part of west Texas, USA.
  • Compiling long time series like this would be more difficult, or in some cases impossible, for more data-intensive ET₀ estimation methods.

🔬 The Penman-Monteith method

🔬 Development history

  • A more precise approach for estimating ET.
  • In 1948, Penman combined the equations for the surface energy balance and for turbulent transport of water vapor away from a saturated evaporating surface.
  • In 1965, Monteith expanded the Penman approach to surfaces where the vapor pressure was less than the saturated vapor pressure (i.e., less than 100% relative humidity).

🧪 The general equation

The Penman-Monteith equation is:

λET = [Δ(Rₙ - G) + ρₐcₚ(eₛ - eₐ)/rₐ] / [Δ + γ(1 + rₛ/rₐ)]

Where:

  • λ is the latent heat of vaporization for water (J per kg)
  • Rₙ is the net radiation (W per m²)
  • G is the soil heat flux (W per m²)
  • eₛ - eₐ is the vapor pressure deficit of the air (kPa)
  • ρₐ is the density of the air (kg per m³)
  • cₚ is the specific heat capacity of the air at constant pressure (J per kg per °C)
  • Δ is the slope of the relationship between saturated vapor pressure and temperature (kPa per °C)
  • γ is called the psychrometric constant (kPa per °C)
  • rₛ is the surface resistance to vapor transport (s per m)
  • rₐ is the aerodynamic resistance to vapor transport (s per m)

🌬️ Resistance terms

  • The two resistance terms (rₛ and rₐ) provide a simplified but effective representation of water vapor transport processes near the land surface.
  • Surface resistance (rₛ): controls vapor transport at the surface itself.
  • Aerodynamic resistance (rₐ): controls vapor transport through the air above the surface.

🌾 FAO standardized version

A panel of experts from the United Nations Food and Agricultural Organization (FAO) has defined the reference surface as:

"A hypothetical reference crop with an assumed height of 0.12 m, a fixed surface resistance of 70 s per m, and an albedo of 0.23."

Based on this definition and employing some approximations for the resistance terms, the FAO Penman-Monteith equation for ET₀ is:

ET₀ = [0.408Δ(Rₙ - G) + γ(900/(T + 273))u₂(eₛ - eₐ)] / [Δ + γ(1 + 0.34u₂)]

Where:

  • ET₀ is the reference evapotranspiration (mm per day)
  • Rₙ and G are in MJ per m² per day
  • T is the mean daily air temperature at 2-m height (°C)
  • u₂ is the mean daily wind speed at 2 m (m per s)

🔧 Converting to actual ET

  • Once ET₀ is determined, site- and day-specific "crop coefficients" (Kc) can be estimated to calculate actual ET from ET₀.
  • Formula: ET = ET₀ × Kc

✅ Reliability and applications

  • With judicious use, the FAO Penman-Monteith equation has proven to provide reliable ET₀ estimates for use in crops, grasslands, and forests around the world.
  • Example: The FAO Penman-Monteith equation was used to estimate daily ET₀ during a 3-year bioenergy crop experiment in Oklahoma, showing that actual ET can be substantially less than ET₀ during dry periods but can exceed ET₀ during wet periods.

⚖️ Comparing the two methods

AspectHargreaves MethodPenman-Monteith Method
Data requirementsOnly air temperatures (max, min, mean) and latitude/day of year for RₐNet radiation, soil heat flux, vapor pressure deficit, air density, specific heat, wind speed, temperature
PrecisionSimpler, less preciseMore precise
When to useWhen only temperature data are availableWhen comprehensive meteorological data are available
Time scaleDesigned for 5-day or larger periods; frequently monthlyCan be used for daily estimates
ComplexityOne of the simplest approachesMore complex, requires more variables
Long-term studiesEasier to compile long time series (e.g., 45-year datasets)More difficult or impossible for long time series in data-sparse regions
ReliabilityGood for low-data situationsProven reliable for crops, grasslands, and forests worldwide

🎯 Don't confuse

  • Both methods estimate ET₀ (reference evapotranspiration), not actual ET.
  • The choice between methods depends on data availability, not on which is "better" in absolute terms.
  • Hargreaves is not "worse"—it is optimized for situations where comprehensive meteorological data are unavailable.
57

Heat Transfer in Soil

13.1. Heat Transfer in Soil

🧭 Overview

🧠 One-sentence thesis

Heat transfer in soil occurs primarily by conduction (governed by Fourier's Law), and the rate and capacity of heat movement depend on soil water content, bulk density, and the interplay of thermal conductivity, volumetric heat capacity, and thermal diffusivity.

📌 Key points (3–5)

  • Primary mechanism: Heat transfer in soil occurs mainly by conduction, though convection can be important in some cases.
  • Three key thermal properties: thermal conductivity (how well heat flows), volumetric heat capacity (how much energy is needed to warm soil), and thermal diffusivity (how fast temperature changes spread).
  • Water content matters differently: volumetric heat capacity increases strictly linearly with water content, while thermal diffusivity is less sensitive to water content than the other two properties.
  • Common confusion: don't confuse thermal conductivity (ability to conduct heat) with volumetric heat capacity (energy needed to change temperature)—they combine to determine thermal diffusivity.
  • Practical implication: wetter, denser soils require more energy to warm up, contributing to lower soil temperatures and delayed crop development in no-tillage management.

🔥 Heat transfer mechanisms

🔥 Conduction as the primary mode

  • Heat transfer in soil occurs primarily by conduction.
  • Convective heat transfer can be important in some cases, but conduction dominates.
  • Conduction is governed by Fourier's Law, first documented in 1807 and published in 1822 in France.
  • The excerpt notes that Fourier's Law may have influenced the later development of Darcy's Law (1856).
  • Just as Darcy's Law says water flux is proportional to the hydraulic gradient, Fourier's Law relates heat flux to temperature gradient.

🌡️ Volumetric heat capacity

🌡️ What it measures

Volumetric heat capacity: the amount of energy required to increase a unit volume of soil by one degree (J m⁻³ °C⁻¹).

  • It is not about how well heat flows; it is about how much energy is needed to change the temperature.
  • The more energy required, the harder it is to warm or cool the soil.

📈 How water and density affect it

  • Volumetric heat capacity increases strictly linearly as soil water content increases (Fig. 13-4).
  • It is also a linear function of bulk density.
  • The excerpt provides a calculation formula (Eq. 13-2):
    • Depends on soil bulk density (ρ_b, g cm⁻³), specific heat of soil solids (c_s, J g⁻¹ °C⁻¹), density of water (ρ_w, g cm⁻³), specific heat of water (c_w), and volumetric water content (θ, cm³ cm⁻³).

🌾 Practical consequence

  • Wetter, denser soil requires more energy to increase temperature than drier, less dense soil (which has lower volumetric heat capacity).
  • This is one factor that can contribute to lower soil temperatures and delayed crop development in soils managed with no tillage.
  • Example: A wet, compacted field in spring will warm more slowly than a dry, loose field, delaying planting and early growth.

⚡ Thermal diffusivity

⚡ What it measures

Thermal diffusivity: the ratio of thermal conductivity to volumetric heat capacity (m² s⁻¹).

  • It is an indicator of the rate at which a temperature change will be transmitted through the soil by conduction.
  • When thermal diffusivity is high, temperature changes are transmitted rapidly through the soil.
  • When thermal diffusivity is low, temperature changes spread more slowly.

🔗 How it relates to other properties

  • Thermal diffusivity is influenced by all the factors that influence thermal conductivity and heat capacity.
  • It combines two opposing effects:
    • Higher thermal conductivity → heat flows faster → higher diffusivity.
    • Higher volumetric heat capacity → more energy needed to change temperature → lower diffusivity.
  • The ratio balances these two properties.

💧 Sensitivity to water content

  • Thermal diffusivity is somewhat less sensitive to soil water content than are thermal conductivity and volumetric heat capacity (Fig. 13-4).
  • Don't confuse: even though water content strongly affects both conductivity and capacity, the ratio (diffusivity) changes less because both numerator and denominator increase.

🛠️ Why it is useful

  • Thermal diffusivity is a particularly useful parameter to aid in understanding and modeling soil temperatures.
  • The excerpt indicates that soil temperatures are the next topic to be considered.
  • Example: Predicting how quickly a cold front will penetrate into the soil, or how fast the soil will warm after sunrise, depends on thermal diffusivity.

📊 Summary comparison

PropertyWhat it measuresRelationship to water contentUnits
Thermal conductivityHow well heat flows through soilIncreases with water (nonlinear)(not given in excerpt)
Volumetric heat capacityEnergy needed to warm a unit volume by one degreeIncreases strictly linearlyJ m⁻³ °C⁻¹
Thermal diffusivityRate at which temperature change spreadsSomewhat less sensitive (ratio of the above two)m² s⁻¹
58

Soil Thermal Properties

13.2. Soil Thermal Properties

🧭 Overview

🧠 One-sentence thesis

Soil thermal properties—volumetric heat capacity, thermal conductivity, and thermal diffusivity—together determine how quickly temperature changes propagate through soil, with water content and bulk density playing key roles.

📌 Key points (3–5)

  • Volumetric heat capacity: measures the energy needed to raise a unit volume of soil by one degree; increases linearly with water content and bulk density.
  • Thermal diffusivity: the ratio of thermal conductivity to volumetric heat capacity; indicates how fast temperature changes travel through soil by conduction.
  • Water content effects: wetter, denser soils require more energy to warm up, which can delay crop development (e.g., in no-tillage systems).
  • Common confusion: thermal diffusivity is less sensitive to water content than thermal conductivity or heat capacity alone, even though it depends on both.
  • Why it matters: thermal diffusivity is particularly useful for understanding and modeling soil temperature behavior.

🔥 Volumetric heat capacity

🔥 What it measures

Volumetric heat capacity: the energy required to raise the temperature of a unit volume of soil by one degree (J m⁻³ °C⁻¹).

  • Unlike thermal conductivity, volumetric heat capacity increases strictly linearly as soil water content increases.
  • It is also a linear function of bulk density.

🧮 How it is calculated

The excerpt provides Equation 13-2:

  • Volumetric heat capacity depends on:
    • ρ_b: soil bulk density (g cm⁻³)
    • c_s: specific heat of soil solids (J g⁻¹ °C⁻¹)
    • ρ_w: density of water (g cm⁻³)
    • c_w: specific heat of water
    • θ: volumetric water content (cm³ cm⁻³)
  • The formula combines contributions from both soil solids and water.

💧 Practical implications

  • Wetter, denser soil requires more energy to increase temperature than drier, less dense soil.
  • This lower heat capacity contributes to lower soil temperatures and delayed crop development in no-tillage management systems.
  • Example: A field with high water content and compaction will warm up more slowly in spring, potentially delaying planting or germination.

🌡️ Thermal diffusivity

🌡️ What it measures

Thermal diffusivity: the ratio of thermal conductivity to volumetric heat capacity (m² s⁻¹).

  • It indicates the rate at which a temperature change will be transmitted through the soil by conduction.
  • When thermal diffusivity is high, temperature changes are transmitted rapidly through the soil.

🔗 What influences it

  • Thermal diffusivity is influenced by all the factors that affect thermal conductivity and heat capacity.
  • However, it is somewhat less sensitive to soil water content than either thermal conductivity or volumetric heat capacity alone.
  • Don't confuse: even though diffusivity depends on both conductivity and capacity, its sensitivity to water is dampened because both numerator and denominator change with water content.

🛠️ Why it is useful

  • Thermal diffusivity is a particularly useful parameter for understanding and modeling soil temperatures.
  • It captures the combined effect of how well soil conducts heat (conductivity) and how much energy is needed to change its temperature (capacity).
  • Example: Two soils may have different conductivities and capacities, but similar diffusivities—meaning temperature waves propagate at similar speeds through both.

🔄 Heat transfer context

🔄 Conduction in soil

  • In soil, heat transfer occurs primarily by conduction, though convective heat transfer can be important in some cases.
  • Heat conduction is governed by Fourier's Law, first documented in 1807 and published in 1822 in France.
  • The excerpt notes that Fourier's Law may have influenced the later development of Darcy's Law (1856).
  • Just as Darcy's Law states that water flux is proportional to the hydraulic gradient, Fourier's Law relates heat flux to the temperature gradient.

📊 Summary comparison

PropertyWhat it measuresSensitivity to water contentKey use
Volumetric heat capacityEnergy to raise temperature by 1° per unit volumeStrictly linear increaseExplains energy requirements for warming soil
Thermal conductivity(Not detailed in this excerpt)(Varies, not strictly linear)Governs heat flow rate
Thermal diffusivityRatio of conductivity to capacitySomewhat less sensitiveModels temperature change propagation speed
59

Soil Surface Temperature

13.3. Soil Surface Temperature

🧭 Overview

🧠 One-sentence thesis

Soil thermal properties—thermal conductivity, volumetric heat capacity, and thermal diffusivity—together determine how quickly temperature changes propagate through soil, with wetter and denser soils requiring more energy to warm and responding differently to temperature changes.

📌 Key points (3–5)

  • Volumetric heat capacity increases linearly with both soil water content and bulk density, meaning wetter and denser soils need more energy to warm up.
  • Thermal diffusivity (the ratio of thermal conductivity to volumetric heat capacity) indicates how fast temperature changes move through soil by conduction.
  • Common confusion: thermal diffusivity is less sensitive to soil water content than thermal conductivity or volumetric heat capacity are individually.
  • Practical implication: higher volumetric heat capacity in wetter, denser soils contributes to lower soil temperatures and delayed crop development in no-tillage management.
  • Heat transfer mechanism: in soil, heat moves primarily by conduction (governed by Fourier's Law), though convection can matter in some cases.

🌡️ Volumetric heat capacity

🔢 What it measures

Volumetric heat capacity: the energy required to raise the temperature of a unit volume of soil by one degree (J m⁻³ °C⁻¹).

  • It quantifies how much energy input is needed for a given temperature rise.
  • Unlike thermal conductivity, volumetric heat capacity increases strictly linearly as soil water content increases.
  • It is also a linear function of bulk density.

🧮 How to calculate it

The excerpt provides Equation 13-2:

  • Volumetric heat capacity depends on:
    • ρ_b: soil bulk density (g cm⁻³)
    • c_s: specific heat of soil solids (J g⁻¹ °C⁻¹)
    • ρ_w: density of water (g cm⁻³)
    • c_w: specific heat of water
    • θ: volumetric water content (cm³ cm⁻³)
  • The formula combines the contributions of soil solids and water.

🌾 Why wetter, denser soils warm more slowly

  • To increase the temperature of wetter, denser soil requires more energy than for drier, less dense soil.
  • Wetter and denser soils have higher volumetric heat capacity.
  • Example: This is one factor contributing to lower soil temperatures and delayed crop development in no-tillage management systems.

🚀 Thermal diffusivity

📐 Definition and meaning

Thermal diffusivity: the ratio of thermal conductivity to volumetric heat capacity (m² s⁻¹).

  • It indicates the rate at which a temperature change will be transmitted through the soil by conduction.
  • When thermal diffusivity is high, temperature changes propagate rapidly through the soil.

🔍 Sensitivity to soil water content

  • Thermal diffusivity is influenced by all factors that affect thermal conductivity and heat capacity.
  • However, thermal diffusivity is somewhat less sensitive to soil water content than thermal conductivity and volumetric heat capacity are individually.
  • Don't confuse: even though both conductivity and capacity change with water content, their ratio (diffusivity) changes less dramatically.

🛠️ Usefulness for modeling

  • Thermal diffusivity is a particularly useful parameter for understanding and modeling soil temperatures.
  • The excerpt indicates this is the next topic to be considered.

🔥 Heat transfer mechanisms

🌊 Conduction vs convection

  • In soil, heat transfer occurs primarily by conduction.
  • Convective heat transfer can be important in some cases (the excerpt does not detail when).

📜 Fourier's Law

  • Heat conduction is governed by Fourier's Law.
  • First documented in 1807 and published in 1822 in France.
  • Historical note: Fourier's Law may have influenced the later development of Darcy's Law (1856).
  • Just as Darcy's Law states that water flux is proportional to the hydraulic gradient, Fourier's Law relates heat flux to the temperature gradient.

📊 Comparison of thermal properties

PropertyRelationship to water contentRelationship to bulk densityWhat it tells us
Volumetric heat capacityStrictly linear increaseLinear increaseEnergy needed to warm soil
Thermal conductivityIncreases (not strictly linear)AffectedHow well heat moves through soil
Thermal diffusivitySomewhat less sensitiveAffected via both conductivity and capacityHow fast temperature changes propagate
60

Sub-Surface Soil Temperatures

13.4. Sub-Surface Soil Temperatures

🧭 Overview

🧠 One-sentence thesis

Soil thermal properties—conductivity, heat capacity, and diffusivity—together determine how quickly temperature changes move through soil, with wetter and denser soils requiring more energy to warm and transmitting temperature changes at different rates.

📌 Key points (3–5)

  • Volumetric heat capacity: increases linearly with both water content and bulk density; wetter, denser soil needs more energy to warm up.
  • Thermal diffusivity: the ratio of conductivity to heat capacity; indicates how fast a temperature change spreads through soil by conduction.
  • Common confusion: thermal diffusivity is less sensitive to water content than conductivity or heat capacity alone, even though it depends on both.
  • Practical impact: lower heat capacity in no-till soils (wetter, denser) can delay warming and slow crop development.
  • Heat transfer mechanism: conduction (governed by Fourier's Law) is the primary mode; convection can matter in some cases.

🔥 Volumetric heat capacity

🔥 What it measures

Volumetric heat capacity: the energy required to raise the temperature of a unit volume of soil by one degree (J m⁻³ °C⁻¹).

  • It is not about how well heat moves, but how much energy the soil can store per unit volume.
  • The excerpt emphasizes that it increases strictly linearly with soil water content (Fig. 13‑4).

📐 How it is calculated

The excerpt provides Equation 13-2:

  • Volumetric heat capacity = (bulk density) × (specific heat of soil solids) + (density of water) × (specific heat of water) × (volumetric water content).
  • In symbols: ρ_b · c_s + ρ_w · c_w · θ.
  • Both bulk density (ρ_b) and water content (θ) increase heat capacity linearly.

🌾 Why wetter, denser soil warms more slowly

  • To raise the temperature of wetter, denser soil requires more energy than for drier, less dense soil.
  • Lower volumetric heat capacity in drier soil → easier to warm.
  • Example: No-till soils are often wetter and denser, so they have higher heat capacity, leading to lower soil temperatures and delayed crop development.

🚀 Thermal diffusivity

🚀 What it measures

Thermal diffusivity: the ratio of thermal conductivity to volumetric heat capacity (m² s⁻¹).

  • It indicates the rate at which a temperature change is transmitted through the soil by conduction.
  • High thermal diffusivity → temperature changes spread rapidly.
  • Low thermal diffusivity → temperature changes spread slowly.

🔗 How it relates to conductivity and capacity

  • Thermal diffusivity = (thermal conductivity) / (volumetric heat capacity).
  • It depends on all the factors that influence both conductivity and capacity.
  • Don't confuse: even though diffusivity is calculated from conductivity and capacity, it is somewhat less sensitive to soil water content than either property alone (Fig. 13‑4).

🛠️ Why it is useful

  • The excerpt calls thermal diffusivity "a particularly useful parameter" for understanding and modeling soil temperatures.
  • It combines the effects of how well heat moves (conductivity) and how much energy is needed (capacity) into a single rate measure.

🌡️ Heat transfer mechanisms

🌡️ Conduction as the primary mode

  • In soil, heat transfer occurs primarily by conduction.
  • Conduction is governed by Fourier's Law (first documented 1807, published 1822 in France).
  • The excerpt notes that Fourier's Law may have influenced the later development of Darcy's Law (1856).
  • Just as Darcy's Law says water flux is proportional to the hydraulic gradient, Fourier's Law relates heat flux to the temperature gradient.

💨 Convective heat transfer

  • Convective heat transfer can be important in some cases, but the excerpt does not elaborate on when or how.
  • The primary focus remains on conduction.

📊 Summary comparison

PropertyWhat it measuresSensitivity to water contentUnits
Volumetric heat capacityEnergy to warm soil per unit volumeStrictly linear increaseJ m⁻³ °C⁻¹
Thermal conductivityHow well heat moves through soil(Not detailed in this excerpt)(Not given)
Thermal diffusivityRate of temperature change transmissionSomewhat less sensitive than conductivity or capacitym² s⁻¹
61

Measured Soil Temperatures

13.5. Measured Soil Temperatures

🧭 Overview

🧠 One-sentence thesis

Soil thermal properties—thermal conductivity, volumetric heat capacity, and thermal diffusivity—together determine how quickly temperature changes move through soil, with wetter and denser soils requiring more energy to warm up and transmitting temperature changes at different rates.

📌 Key points (3–5)

  • Volumetric heat capacity increases linearly with both soil water content and bulk density, meaning wetter and denser soils need more energy to change temperature.
  • Thermal diffusivity (the ratio of thermal conductivity to volumetric heat capacity) indicates how fast a temperature change travels through soil by conduction.
  • Common confusion: thermal diffusivity is less sensitive to soil water content than either thermal conductivity or volumetric heat capacity alone.
  • Practical implication: higher volumetric heat capacity in wetter, denser soils can lead to lower soil temperatures and delayed crop development (e.g., in no-tillage systems).
  • Heat transfer mechanism: in soil, heat moves primarily by conduction (governed by Fourier's Law), though convection can matter in some cases.

🌡️ Volumetric heat capacity

🌡️ What it measures

Volumetric heat capacity: the energy required to raise the temperature of a unit volume of soil by one degree (J m⁻³ °C⁻¹).

  • It is not about how hot the soil is, but how much energy is needed to change its temperature.
  • The excerpt emphasizes that this property increases strictly linearly with soil water content and bulk density.

📐 How to calculate it

The excerpt provides Equation 13-2:

  • Volumetric heat capacity = (bulk density × specific heat of soil solids) + (density of water × specific heat of water × volumetric water content)
  • Variables:
    • ρ_b = soil bulk density (g cm⁻³)
    • c_s = specific heat of soil solids (J g⁻¹ °C⁻¹)
    • ρ_w = density of water (g cm⁻³)
    • c_w = specific heat of water (J g⁻¹ °C⁻¹)
    • θ = volumetric water content (cm³ cm⁻³)

💧 Why wetter and denser soils need more energy

  • Wetter soil has more water; denser soil has more mass per unit volume.
  • Both factors increase volumetric heat capacity linearly.
  • Example: to warm up a wet, dense soil by 1°C requires more energy than warming a dry, loose soil by the same amount.

🌾 Practical consequence for crop development

  • Higher volumetric heat capacity → more energy needed to warm the soil → lower soil temperatures in spring.
  • The excerpt notes this is one factor contributing to delayed crop development in no-tillage systems, where soils tend to be wetter and denser.
  • Don't confuse: this is not the only factor—thermal conductivity and diffusivity also play roles.

🚀 Thermal diffusivity

🚀 What it measures

Thermal diffusivity: the ratio of thermal conductivity to volumetric heat capacity (m² s⁻¹).

  • It indicates how fast a temperature change will be transmitted through the soil by conduction.
  • High thermal diffusivity → temperature changes spread rapidly through the soil.
  • Low thermal diffusivity → temperature changes spread slowly.

🔗 How it relates to other properties

  • Thermal diffusivity depends on both thermal conductivity (numerator) and volumetric heat capacity (denominator).
  • The excerpt states it is influenced by "all the factors which influence thermal conductivity and heat capacity."
  • However, thermal diffusivity is somewhat less sensitive to soil water content than either thermal conductivity or volumetric heat capacity alone.

🧮 Why it is useful

  • The excerpt calls thermal diffusivity "a particularly useful parameter to aid in understanding and modeling soil temperatures."
  • It combines two properties into one indicator of temperature-change speed, simplifying analysis.

🔥 Heat transfer mechanisms in soil

🔥 Conduction vs convection

MechanismRole in soilGoverning law
ConductionPrimary heat transfer modeFourier's Law (1807/1822)
ConvectionCan be important in some cases(Not detailed in excerpt)
  • The excerpt emphasizes that conduction is the main way heat moves through soil.
  • Convection (heat carried by moving fluids) can matter but is secondary.

📜 Fourier's Law

  • Fourier's Law governs heat conduction, first documented in 1807 and published in 1822 in France.
  • The excerpt notes it may have influenced the later development of Darcy's Law (1856), which describes water flow.
  • Just as Darcy's Law says water flux is proportional to the hydraulic gradient, Fourier's Law relates heat flux to the temperature gradient.
  • Don't confuse: Fourier's Law is for heat; Darcy's Law is for water—but both describe flux proportional to a gradient.