Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Soil and Environmental Chemistry
Soil and Environmental Chemistry
Soil and Environmental Chemistry
Ebook955 pages7 hours

Soil and Environmental Chemistry

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Soil and Environmental Chemistry emphasizes the problem-solving skills students will need when they enter their chosen field. This revised reprint links valuable soil chemical concepts to the "big picture" by discussing how other soil and environmental factors affect soil chemistry. This broader environmental approach makes the text relevant to today’s soil science curriculums.

This book uses computer modeling for water and soil chemistry, providing students with the models used by practicing environmental chemists. It includes examples and complex problems with worked solutions, as well as examples based on real data that expose students to the real problems and data they will face in their careers. It also provides edits to formulas, numbers, and text.

This text will serve as a useful resource for upper-level undergraduate students studying soil chemistry without an extensive background in calculus and only limited background in physical chemistry, such as soil science majors and environmental science majors.

  • Use of computer modeling for water and soil chemistry provides students with the models used by practicing environmental chemists
  • Examples and complex problems with worked solutions included throughout the text
  • Examples based on real data provide exposure to the real problems and data students will face in their careers
LanguageEnglish
Release dateJul 28, 2011
ISBN9780124158627
Soil and Environmental Chemistry
Author

William F. Bleam

William Bleam is Professor of Soil Science at the University of Wisconsin, USA. His research interests include physical chemistry of soil colloids and sorption processes, chemistry of humic substances, factors controlling biological availability of contaminants to micro-organisms, magnetic resonance and synchrotron studies of adsorption and precipitation. He has taught an intermediate soil chemistry course (Soil Science 321, Soil & Environmental Chemistry) since 2006. Students taking this course include undergraduate and graduate students.

Related to Soil and Environmental Chemistry

Related ebooks

Agriculture For You

View More

Related articles

Reviews for Soil and Environmental Chemistry

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Soil and Environmental Chemistry - William F. Bleam

    Preface

    The Earth, atmosphere, water, and living things abide intimately and inseparably in a place we call the soil. Any attempt to describe the chemistry of this peculiar corner of the environment inevitably sets boundaries. Hydrology sits beyond the boundary for many soil chemists; but certain chemical processes take decades to develop, making residence time an important variable, to say nothing of belowground water transport. Again, soil microbiology may lay beyond the black stump separating chemistry from its cousin biology; yet respiratory processes drive redox chemistry wherever microbes are found in nature. Is there sufficient chemistry in risk assessment to warrant its inclusion despite differences in scientific dialect?

    Moving the boundaries outward to include Soil Moisture & Hydrology and Risk Assessment forces compensating choices. Herein Clay Mineralogy & Clay Chemistry embraces layer silicates and swelling behavior, but demurs the broader discussion of structural crystallography. Mineral weathering makes its appearance in Acid-Base Chemistry as the ultimate source of basicity, but detailed chemical mechanisms receive scant attention. Ion exchange is stripped to its bare essentials—the exchange isotherm and the factors determining its appearance. Natural Organic Matter & Humic Colloids adds content related to carbon turnover and colloidal behavior, but does not take on the humification process or the primary structure of humic molecules. Water Chemistry downplays algebraic methods in favor of model validation, reflecting my experience that students attain more success applying their chemical knowledge to validating simulation results than slogging through mathematically complex and symbolically unfamiliar coupled equilibrium expressions. The hallmark of Acid-Base Chemistry is integration—pulling together water chemistry, ion exchange, and fundamental acid-base principles to develop an understanding of two chemically complex topics: exchangeable acidity and sodicity.

    Each chapter includes insofar as possible: actual experimental data plotted in graphic form, one or more simple models designed to explain the chemical behavior manifest in experimental data, and quantitative examples designed to develop problem-solving skills. Examples and problems rely on actual experimental data when available, supplemented by data gleaned from the USDA Natural Resources Conservation Service Soil Data Mart, the United States Environmental Protection Agency PBT Profiler and Integrated Risk Information System, and the National Atmospheric Deposition Program, to name a few. These examples are designed to develop key problem-solving skills and to demonstrate methods for making sound estimates and predictions using basic chemistry principles, proven chemical models, and readily accessible soil, environmental, and chemical data.

    I wish to acknowledge several individuals who contributed to the planning, writing, and completion of this book. Philip Helmke and Phillip Barak, my chemistry colleagues at Madison, influenced uncounted choices of content and emphasis. Birl Lowery, John Norman, and Bill Bland, my soil physics colleagues, guided my choice to include hydrology. Robin Harris and Bill Hickey fed my interest and passion in environmental microbiology. Dr. Beat Müller of EAWAG (Das Schweizer Eidgenössischen Technischen Hochschulen und Forschung) was most helpful on all questions regarding ChemEQL. Robert W. Taylor, a friend dating from my Philadelphia days, exemplified encouragement. My graduate advisors—Roscoe Ellis, Murray B. McBride, and Roald Hoffman—guided my development as a young scientist, setting me on a course that ultimately led to writing this book. Edna J. Cash, a dear friend, diligently edited final drafts and proofs. Sharon—my wife—was always patient, always supportive.

    William F. Will Bleam

    October 2010

    Chapter 1

    Elements

    Their Origin and Abundance

    Chapter Outline

    1.1 Introduction

    1.2 A Brief History of the Solar System and Planet Earth

    1.3 The Composition of Earth’s Crust and Soils

    1.4 The Abundance of Elements in the Solar System, Earth’s Crust, and Soils

    1.5 Elements and Isotopes

    1.6 Nuclear Binding Energy

    1.7 Enrichment and Depletion during Planetary Formation

    1.8 Planetary Accretion

    1.9 The Rock Cycle

    1.10 Soil Formation

    1.11 Concentration Frequency Distributions of the Elements

    1.12 Estimating the Most Probable Concentration and Concentration Range Using the Logarithmic Transformation

    1.13 Summary

    Appendix 1A Factors Governing Nuclear Stability and Isotope Abundance

    Appendix 1B Nucleosynthesis

    Appendix 1C Thermonuclear Fusion Cycles

    Appendix 1D Neutron-Emitting Reactions that Sustain the S-Process

    Appendix 1E Random Sequential Dilutions and the Law of Proportionate Effect

    Appendix 1F The Estimate of Central Tendency and Variation of a Log-Normal Distribution

    1.1 Introduction

    Popular Science published an electronic version of the Periodic Table of the Elements in November 2006 (http://www.popsci.com/files/periodic_popup.html) that contained pictures of virtually all of the elements in pure form. Notably absent are the radioactive elements: promethium Pm, astatine At, radon Rn, francium Fr, actinium Ac, protactinium Pa, and elements beyond uranium U. There is a reason images of these elements are absent: every one of them is unstable and, therefore, extremely rare. The Periodic Table of the Elements asserts that all elements exist in principle, but this particular table correctly implies that all of the elements from hydrogen to uranium exist on planet Earth. In fact, every sample of water, rock, sediment, and soil contains every stable element—and probably most of the unstable elements—from hydrogen to uranium.

    The Environmental Working Group published an article in October 2008 entitled Bottled Water Quality Investigation: 10 Major Brands, 38 Pollutants (Naidenko, Leiba et al., 2008). Among the contaminants found in bottled water sold in the United States were the radioactive strontium isotope Sr-90 (0.02 ), radioactive radium (isotopes Ra-286 plus Ra-288; 0.02 ), boron (60–90 ), and arsenic (1 ). These concentrations, combined with a commentary listing the potential and actual toxicity of these substances, can be alarming. The important question is not whether drinking water, food, air, soil, or dust contains toxic elements; it most certainly does! The important question is: is the level of any toxic element in drinking water, food, or soils harmful?

    The United States Environmental Protection Agency (USEPA) has established a maximum contaminant level (MCL) for beta emitters—such as strontium-90 or radium—in public drinking water: 0.296 (8 ). The current USEPA drinking water MCL for arsenic is 10 ; the mean arsenic concentration in U.S. groundwater (based on over 20,000 samples) is 2 . The USEPA does not have a drinking water standard for boron, but the World Health Organization (WHO) recommends boron levels in drinking water less than 500 , and in 1998 the European Union adopted a drinking water standard of 100 . Typical boron concentrations in U.S. groundwater fall below 100 and 90% below 40 . Evaluated in context, the contaminants found in bottled water are less threatening and raise the question of whether it is reasonable to entirely eliminate trace elements from water and food.

    1.2 A Brief History of the Solar System and Planet Earth

    The gravitational collapse of a primordial cloud of gas and dust gave birth to the present-day Solar System. The conservation of angular momentum in the primordial dust cloud explains the rotation of the Sun (25-day rotation period at the equator) and a primordial accretion disk that spawned the planets and other bodies that orbit the Sun. The primordial gas and dust cloud was well mixed and uniform in composition. Gravitational accretion and radioactive decay released sufficient heat to melt the early Earth, leading segregation into a solid metallic core, a molten mantle, and a crystalline crust. Planetary formation and segregation altered the composition of Earth’s crust relative to the primordial cloud and imposed variability in the composition of each element as a direct consequence of each separation process.

    Throughout its entire history, planet Earth has experienced continual transformation as plate tectonics generate new inner (oceanic) crust and shift plates of outer (continental) crust around like pieces of a jigsaw puzzle. Plate tectonics and the hydrologic cycle drive a rock cycle that reworks portions of the outer crust through weathering, erosion, and sedimentation. The rock cycle imposes a new round of geochemical separation processes that alters the composition of the terrestrial land surface relative to the outer crust from which it derives. The rock cycle and soil development, much like planetary formation and segregation early in Earth’s history scale, impose additional variability in the composition of each element. Transformations in the overall composition of planet Earth and the imposition of composition variability relative to the primordial gas cloud are the central themes of this chapter.

    1.3 The Composition of Earth’s Crust and Soils

    Most books on geology, soil science, or environmental science include a table listing the composition of Earth’s crust or soil. There are several reasons for listing the elemental composition of Earth materials. The composition of soil constrains the biological availability of each element essential for living organisms. Soils develop from the weathering of rocks and sediments and inherit much of their composition from the local geology. The most common elements—those accounting for 90–95% of the total composition—determine the dominant mineralogy of rocks and materials that form when rocks weather (residuum, sediments, and soils).

    The soil at a particular location develops in starting or parent material—rock or sediments—under a variety of local influences—landform relief, climate, and biological community over a period of decades or centuries. Not surprisingly, the composition of soil is largely inherited from the parent material. The rocks or sediments in which a soil develops are usually not the basement rocks that compose the bulk of Earth’s crust. Regardless of geologic history, virtually every rock exposed at the Earth’s terrestrial surface owes its composition to the crystalline igneous rock of the continental crust. The Earth’s outer crust inherits its composition from the overall composition of the planet and, by extension, the gas and dust cloud that gave rise to the Solar System as a whole.

    How much do the composition of soil, Earth’s crust, and the Solar System have in common? What can we learn about the processes of planetary formation, the rock cycle, and soil development from any differences in composition? What determines the relative abundance of elements in soil?

    1.4 The Abundance of Elements in the Solar System, Earth’s Crust, and Soils

    We are looking for patterns, and, unfortunately, data in tabular form usually do not reveal patterns in their most compelling form. Patterns in the elemental abundance of the Solar System, Earth’s crust, and soils are best understood when abundance is plotted as a function of atomic number Z. Astronomers believe the composition of the Sun’s photosphere is a good representation of the primordial gas and dust cloud that gave rise to the Solar System. Usually the composition of the Solar System is recorded as the atom mole fraction of each element, normalized by the atom mole fraction of silicon.

    The composition of Earth’s crust and soil is usually recorded as the mass fraction of each element. If we are to compare the compositions of the Solar System, Earth’s crust, and soils, the abundance data must have the same units. Since atoms combine to form compounds based on atomic mole ratios, the atom mole fraction is the most informative. The atom mole fraction is found by dividing the mass fraction of each element—in crust or soil—by its atomic mass and then normalized by the atom mole fraction of silicon.

    An abundance plot of the Solar System, Earth’s crust, and soil using a linear atomic mole fraction scale reveals little because 99.86% of the Solar System consists of hydrogen and helium. A plot using a logarithmic atom mole fraction scale for the Solar System, Earth’s crust, and soil reveals three important features. First, the abundance of the elements decreases exponentially with increasing atomic number Z (the decrease appears roughly linear when plotted using a logarithmic scale). Second, a zigzag pattern is superimposed on this general tread, known as the even-odd effect: elements with an even atomic number are consistently more abundant than elements with an odd atomic number. Third, while the data set for soil composition is missing many of the elements recorded for the Solar System and Earth’s crust, the exponential decrease in abundance with atomic number (Figure 1.1) and the even-odd effect (Figure 1.2) appear in all three data sets.

    Figure 1.1 Elemental abundance of the Solar System, Earth’s crust, and soil decreases exponentially with atomic number Z ( Shacklette, 1984 ; Lide, 2005 ).

    Figure 1.2 Elemental abundance of the Solar System, Earth’s crust, and soil from calcium (atomic number Z = 20) to zirconium (atomic number Z = 40), showing the even-odd effect ( Shacklette, 1984 ; Lide, 2005 ).

    Cosmological processes, beginning with the so-called Big Bang, determined Solar System composition. The cosmological imprint—the exponential decrease in abundance with atomic number and the even-odd effect—is clearly discernible in the composition of the Earth’s crust and terrestrial soils despite the profound changes resulting from planetary formation, the rock cycle, and soil development.

    1.5 Elements and Isotopes

    The Periodic Table of the Elements organizes all known elements into groups and periods. Elements in the same chemical group have the same number of electrons in the outermost electronic shell but differ in the number of occupied electronic shells. Elements in the same chemical period have the same number of occupied electronic shells but differ in the number of electrons in the partially filled outermost electronic shell. Period 1 contains elements H and He, the filling of atomic shell 1s. Atomic shell 2s is filled in groups 1 and 2 of period 2—elements Li and Be—while 2p is filled in groups 13–18 of the same period. Elements of the same period exhibit similar chemical properties because their outermost or valence electronic shell has the same number of electrons. It is this characteristic—the valence electron configuration—that has the greatest influence on the chemistry of each element.

    The Periodic Table of the Elements always lists the symbol, and usually the atomic number Z and atomic weight, of each element. The atomic number Z is an integer equal to the number of positively charged protons in the nucleus of that element. The atomic weight, unlike Z, is not an integer but a decimal number larger than Z. The atomic weight is the mass of one mole of the pure element and accounts for the relative abundance of the stable and long-lived radioactive isotopes for that element (see Appendix 1A).

    Figure 1.3 is an unusual Periodic Table of the Elements because it plots the relative abundance of each element in the Earth’s outermost layers—the mantle and the crust—on a logarithmic vertical scale. We will have more to say about the segregation of the Earth later, but for now we are most concerned with the processes that determine why some elements are more abundant than others. You will notice that the abundance of elements in a given period—say, period 4—fluctuate; Ca is more abundant than K or Sc, and Ti is more abundant than Sc or V. This is the even-odd effect mentioned earlier and results from the nuclear stability of each element, a topic discussed further in Appendix 1A.

    Figure 1.3 Periodic Table of the Elements: the abundance of each element in the bulk silicate Earth (i.e., mantle plus crust) plotted on the vertical scale in logarithmic units.

    Source: Reproduced with permission from Anders, E., and N. Grevesse, 1989. Abundance of the elements: meteoric and solar. Geochim. Cosmochim. Acta. 53, 197–214.

    1.6 Nuclear Binding Energy

    The even-odd effect fails to explain the abundance of iron (Z = 26) or, to be more precise, . Figure 1.4 is a plot of the nuclear binding energy per nucleon for all known isotopes. It clearly shows is the isotope with the greatest binding energy per nucleon. The fusion of lighter nuclei to form heavier nuclei is exothermic when the product has a mass number A ≤ 56 but is endothermic for all heavier isotopes. You will also notice that the energy release from fusion reactions diminishes rapidly as the mass number increases to 20 and then gradually in the range from 20 to 56.

    Figure 1.4 Binding energy per nucleon of each isotope as a function of the mass number A .

    The mass of an electron (0.000549 unified atomic mass units u) is negligible compared to that of protons (1.007276 u) and neutrons (1.008665 u). The mass number A of each isotope is an integral sum of protons Z and neutrons N in the nucleus, while the isotope mass is the mass of one mole of pure isotope. The mass of an isotope with mass number A is always lower than the mass of Z-independent protons and N-independent neutrons. The nuclear binding energy of hydrogen isotope (Example 1.1) illustrates the so-called mass defect.

    Example 1.1 Calculate the Nuclear Binding Energy for Deuterium , an Isotope of Hydrogen.

    The nucleus of a deuterium atom consists of one proton and one neutron. The masses of the constituents, in unified atomic mass units u (1 u = 1.66053886 10-27 kg):

    The mass of one mole of pure , however, is 2.014102 u—0.001839 u less than the mass found by adding the rest mass of a proton and a neutron. The mass difference, multiplied by 931.494 MeV u-1 and divided by 2 to account for the two nucleons of , is the binding energy per nucleon for : 0.8565 MeV.

    Exothermic fusion reactions are the source of the energy output from stars, beginning with nuclei and other nuclei produced during the early stages of the universe following the Big Bang. Figure 1.4 shows that exothermic fusion becomes increasingly inefficient as an energy source. Simply put: a star’s nuclear fuel is depleted during its lifetime and is eventually exhausted. Appendix 1B describes the nuclear processes that formed every isotope found in our Solar System and the Earth today. Some of those processes occurred during the early moments of the universe under conditions that no longer exist. Other processes continue to occur throughout the universe because the necessary conditions exist in active stars or in the interstellar medium. Nucleosynthesis is an accumulation process that gradually produces heavier nuclides from lighter nuclides, principally by exothermic fusion or neutron capture. Lighter nuclides are generally more abundant than heavier nuclides. The extraordinary energies required for exothermic nuclear fusion and the complex interplay of neutron-capture and radioactive decay, the principal nucleosynthesis processes, translate into an abundance profile (see Figure 1.1) that embodies the relative stability of each nuclide.

    1.7 Enrichment and Depletion during Planetary Formation

    Gravitational collapse of a primordial cloud of gas and dust and subsequent planetary formation in the Solar System was a process of differentiation—the transformation from a homogeneous state to a heterogeneous state—on scales ranging from the scale of the Solar System itself to the smallest microscopic scale we can characterize experimentally.

    Homogeneity in the primordial gas cloud that ultimately became the Solar System resulted from the random motion of gas molecules and dust particles. If we collected samples from the primordial cloud for elemental analysis, the sample population would yield a frequency distribution consistent with the conditions existing in the gas cloud. A homogenous, well-mixed cloud of gas and dust would yield a normal distribution of concentrations for each element (Figure 1.5). The average (or central tendency) concentration of each element is estimated by calculating the arithmetic mean concentration for the sample population.

    Figure 1.5 The frequency of concentrations from a sample population (bar graph) closely approximates a normal probability density function (line). The mean (or central tendency) for a normal probability distribution equals the arithmetic mean for the sample population.

    Source: Image generated using Sample versus Theoretical Distribution (Wolfram Demonstration Project) using 40 bins and a sample size n = 4118.

    A consequence of the various separation processes during planetary accretion and segregation, the rock cycle on planet Earth, and, ultimately, soil formation is more than the enrichment and depletion of individual elements, altering the abundance profile (see Figures 1.2 and 1.3). Elemental analysis shows that the weathered rock materials and soils that blanket the Earth’s continental crust bear the imprint of numerous separation processes and, as we will presently discover, yield abundance distributions that differ markedly from the normal distribution shown in Figure 1.5.

    1.8 Planetary Accretion

    The Solar System formed during the gravitational collapse of a gas cloud that ultimately became the Sun. Preservation of angular momentum within the collapsing gas cloud concentrated dust particles into a dense rotating central body—the Sun—and a diffuse disc (Figure 1.6, left) perpendicular to the axis of the Sun’s rotation. Gravity led to the accretion of planetesimals within the disc that grew in size as they collided with one another.

    Figure 1.6 Planetesimal formation in the accretion disc of a protosun (left) and the present Solar System (right) resulted in the differentiation and segregation of the primordial gas cloud.

    The Solar System formed about 4 billion years ago and contains 8 major planets and 153 confirmed moons, asteroids, meteors, comets, and uncounted planetesimals at its outer fringe. The innermost planets—Mercury, Venus, Earth, and Mars—are rocky (density ≈5 Mg m-3), while the outermost planets—Jupiter, Saturn, Uranus, and Neptune—are icy (density <2 Mg m-3). Earth is the largest of the rocky planets, with a radius of nearly 6370 km. Mars, mean radius 3390 km, has the most moons of the rocky planets: two. The icy planets are much larger—ranging from roughly 4 to 11 times the radius of Earth; are more massive—from 17 to 318 times the mass of Earth; and are surrounded by far more moons—ranging from 13 to 63.

    Goldschmidt (1937) identified several element groups based on geochemical behavior: lithophiles, chalcophiles, atmophiles, and siderophiles. The atmophiles are hydrogen, carbon, nitrogen, and the noble gases helium, neon, argon, krypton, and xenon. The noble gases do not combine with other elements and easily escape the relatively weak gravitational field of the rocky planets. Hydrogen, carbon, and nitrogen react with oxygen, but the products tend to be volatile gases. The gaseous planets, however, are cold enough and massive enough to capture atmophilic gases. Earth and the other rocky planets are depleted of the noble gas elements, hydrogen and other volatile elements relative to the Solar System, as shown in Figure 1.7.

    Figure 1.7 Segregation of the Earth’s crust relative to the Solar System: atmophilic elements (crosses), ferromagnetic metals (diamonds), noble metals (triangles), chalcophilic elements (filled circles), and lithophilic elements (open circles). Ferromagnetic transition metals and noble metals are grouped as siderophilic elements ( Goldschmidt, 1937 ; Lide, 2005 ).

    Planet Earth began segregating into layers as the heat released by accretion and radioactive decay triggered melting. Modern Earth consists of a metallic core composed of ferromagnetic elements and an outer silicate-rich layer—sometimes called the basic silicate Earth—that further separated into a molten mantle and crust of crystalline silicate rock (Figure 1.8). The silicate-rich crust is further segregated into an inner or oceanic crust and an outer or continental crust. The inner crust has a composition and density very similar to the mantle, while the outer crust is less dense, thicker, and far more rigid than the inner crust. The composition of the outer crust reflects the effect of segregation on the planetary scale but bears little imprint from the rock cycle because the mean age of the outer crust is roughly the same as the age of planet Earth.

    Figure 1.8 Segregated Earth: core, mantle, and crust.

    The ferromagnetic elements comprise four transition metals from the fourth period: manganese Mn, iron Fe, cobalt Co, and nickel Ni. These four metals are depleted in Earth’s crust (see Figure 1.7), having become major components of the Earth’s metallic core. The noble metals—ruthenium Ru, rhodium Rh, palladium Pd, silver Ag, rhenium Re, osmium Os, iridium Ir, platinum Pt, and gold Au—have little tendency to react with either oxygen or sulfur. These elements are depleted from the basic silicate Earth, enriching the metallic core. Goldschmidt (1937) grouped the ferromagnetic transition metals and noble metals together as siderophilic metals.

    The lithophilic elements (Goldschmidt, 1937) consist of all elements in the periodic table characterized by their strong tendency to react with oxygen. The chalcophilic elements—which include sulfur and the elements from periods 4 (copper through selenium), 5 (silver through tellurium), and 6 (mercury through polonium)—combine strongly with sulfur. The chalcophilic and lithophilic elements would be enriched in basic silicate Earth after the segregation of the metallic core.

    Given the atomic mass of oxygen and sulfur and the relative atomic mass of lithophilic chalcophilic elements in the same period; it should not be surprising that sulfide minerals are significantly denser than oxide minerals. Buoyancy in the Earth’s gravitational field would cause lithophilic magma to migrate toward the Earth’s surface, while chalcophilic magma would tend to sink toward the Earth’s core, thereby depleting the crust of chalcophilic elements (see Figure 1.7).

    1.9 The Rock Cycle

    The Earth’s crust is in a state of continual flux because convection currents in the mantle form new inner crust along midoceanic ridges. The spreading of newly formed oceanic crust drives the edges of the inner crust underneath the more buoyant outer crust along their boundaries, propelling slabs of outer crust against each other. This movement—plate tectonics—drives the rock cycle (Figure 1.9).

    Figure 1.9 The rock cycle.

    If we take a global view, the outer crust is largely composed of granite, while the inner crust is primarily basalt. Granite is a coarse-grained rock that crystallizes slowly and at much higher temperatures than basalt. Basalt rock has a density comparable to the mantle; it solidifies along midoceanic ridges, where convection currents in the mantle carry magma to the surface. Geologists classify granite, basalt, and other rocks that solidify from the molten state as igneous rocks.

    A more detailed view of the crust on the scale of kilometers to tens of kilometers reveals other rocks that owe their existence to volcanic activity, weathering, erosion, and sedimentation. Magma welling up from the mantle melts through the crust. Rock formations along the margin of the melted zone are annealed; minerals in the rock recrystallize in a process geologists call metamorphism. Both the mineralogy and texture of metamorphic rocks reveal radical transformation under conditions just short of melting. Metamorphic rocks provide geologists with valuable clues about Earth history, but they are far less abundant than igneous rocks.

    Inspection on the kilometer scale reveals a zone of rock weathering and sediment deposits covering both the continental and oceanic crusts. Stress fractures, caused by thermal expansion and contraction, create a pathway for liquid water to penetrate igneous and metamorphic rock formations. Freezing and thawing accelerate fracturing, exfoliating layers of rock and exposing it to erosion by flowing water, wind, gravity, and ice at the Earth surface. The water itself, along with oxygen, carbon dioxide, and other compounds dissolved in water, promotes reactions that chemically degrade minerals formed at high temperatures in the absence of liquid water, precipitating new minerals that are stable in the presence of liquid water.

    Burial, combined with chemical cementation by the action of pore water, eventually transforms sedimentary deposits into sedimentary rocks. Primary minerals—minerals that crystallize at high temperatures where liquid water does not exist—comprise igneous and metamorphic rocks. Secondary minerals—those that form through the action of liquid water—form in surface deposits and sedimentary rocks.

    The rock cycle (see Figure 1.9) is completed as plate tectonics carry oceanic crust and their sedimentary overburden downward into the mantle at the continental margins, or magma melting upward from the mantle returns rock into its original molten state.

    1.10 Soil Formation

    The rock cycle leaves much of the outer crust untouched. Soil development occurs as rock weathering alters rocks exposed on the terrestrial surface of the Earth’s crust. Weathered rock materials, sediments and saprolite, undergo further differentiation during the process of soil formation. Picture freshly deposited or exposed geologic formations (sediments deposited following a flood or landslide, terrain exposed by a retreating glacier, fresh volcanic ash and lava deposits following a volcanic eruption). This fresh parent material transforms through the process of chemical weathering, whose intensity depends on climate (rainfall and temperature), vegetation characteristic of the climate zone, and topography (drainage and erosion). Over time these processes and unique characteristics of the setting lead to the development of soil.

    The excavation of a pit into the soil reveals a sequence of layers or soil horizons with depth (Figure 1.10). This sequence of soil horizons is known as the soil profile for that site. Soil horizons vary in depth, thickness, composition, physical properties, particle size distribution, color, and other properties. Soil formation takes hundreds to thousands of years, and, provided the site has remained undisturbed for that length of time, the soil profile reflects the natural history of the setting.

    Figure 1.10 A northern temperate forest soil developed in glacial till (Northern Ireland).

    Plotting the abundance of each element in the Earth’s crust divided by its abundance in the Solar System reveals those elements enriched or depleted during formation of planet Earth and the formation of the outer crust during the segregation of the Earth (see Figure 1.7). Plotting the abundance of each element soil divided by its abundance in the Earth’s crust reveals those elements enriched or depleted during the rock cycle and soil development. When both are displayed together on the same graph (Figure 1.11), plotting the logarithm of the abundance ratio of crust to the Solar System and the ratio of soil to crust, an important and not surprising result is immediately apparent: enrichment and depletion during planetary formation and segregation are orders of magnitude greater than those that take place during the rock cycle and soil formation. This explains the findings (Helmke, 2000) that a characteristic soil composition cannot be related to geographic region or soil taxonomic group.

    Enjoying the preview?
    Page 1 of 1