Sei sulla pagina 1di 18

Reserve Reading: Do Not Redistribute

-'■'.

■^ * * * ^ • * <(mi £z*

PART
WEATHERING PROCESSES AND
PRODUCTS AND SEDIMENT
TRANSPORT

Chapter 2 Weathering and Soils


Chapter 3 Transport and Deposition
of Siliciclastic Sediment

Logan Butte, Weathered John Day Formation (Oligocene), eastern Oregon.


(Photograph by Gregory Iietallack.)
Reserve Reading: Do Not Redistribute

weathering of source rocks, erosion, sediment transport and deposition, and bur
Sedimentary
ial diagenesis.
rocks All
aresedimentary
formed through
rocksaare
complex
composed
set ofofprocesses
particles orthat
grains
include
of some
kind, which may range in size from microns to hundreds of centimeters. Many of these
grains are individual mineral crystals; others are composite grains made up of aggre
gates of crystals bonded or cemented together. Some grains are epiclastic particles,
derived by weathering of preexisting (source) rocks. Others are pyroclastic particles,
formed through explosive volcanism. A third group of particles are chemically/bio
chemically formed particles, produced by precipitation from water in lakes or the
ocean (e.g., the particles in limestones).
The process of forming sediments and sedimentary rocks begins with
weathering—the physical disintegration and chemical decomposition of rock that pro
duces solid (minerals, rock fragments) and dissolved chemical products. The solid
products of weathering may accumulate in situ to form soils, or they may be removed
eventually by erosion and transported to depositional basins. These transported weath
ering products, together with pyroclastic particles that originate through explosive vol
canism, are the source materials of sandstones, conglomerates, and shales. Because
they are derived from land, these clastic particles (broken fragments of preexisting
rock) are referred to as terrigenous grains. Most terrigenous grains are composed in
large part of silica; therefore, silicate terrigenous grains are commonly called siliciclas-
tic grains.
Siliciclastic grains are removed from weathering sites by erosion and are trans
ported as solids to depositional basins. Mass-transport processes such as slumps,
debris flows, and mud flows are important agents in initial stages of transport of sedi
ment from weathering sites to valley floors. Fluid-flow processes—which include mov
ing water, ice, and wind—move sediment from valley floors to depositional basins at
lower elevations. When transport processes are no longer capable of moving sediment,
deposition of sands, gravels, and muds takes place—subaerially (e.g., in desert dune
fields) or subaqueously in river systems, lakes, or the marginal ocean. Sediment
deposited at the ocean margin may be reentrained and retransported tens to hundreds
of kilometers into deeper water by turbidity currents or other transport processes. Sed
imentary rocks made up of siliciclastic grains are siliciclastic sedimentary rocks.
Weathering processes also release from source rocks soluble constituents such as
calcium, magnesium, and silica that make their way in surface water and groundwater
to lakes or the ocean. When concentrations of these chemical elements become suffi
ciently high, they are removed by chemical and biochemical processes to form lime
stones, cherts, and other chemical/biochemical sedimentary rocks.
Organic matter is an important part of many soils and sedimentary rocks. Plant
material on land is partially broken down by chemical and microbial processes to
yield organic residues. Some organic residues accumulate in soils; others are deposited
in swampy environments to form peats and coals; and still others may be transported
along with weathering detritus to depositional basins. Both plant and animal organic
residues are also generated in the ocean and may become deposited along with silici
clastic or biochemical sediment. Sedimentary rocks that contain substantial amounts
of organic material are called carbonaceous sedimentary rocks.
Chapter 2 focuses on the processes of weathering and soil formation and on the
characteristics of soils—particularly ancient soils or paleosols. Sediment transport and
deposition are treated in Chapter 3 and in some subsequent chapters in Part 5.
Reserve Reading: Do Not Redistribute

Weathering and Soils

2.1 INTRODUCTION

Weathering involves chemical, physical, and biological processes, although chemical


processes are by far the most important. Details of these processes may be found in
several books devoted to weathering and soil formation, such as those by Birkland
(1974), Drever (1985), Colman and Dethier (1986), Lerman and Meybeck (1988), and
Nahon (1991). Only a brief summary of weathering processes is presented here to illus
trate how weathering acts to decompose and disintegrate exposed rocks, producing
particulate residues and dissolved constituents. These weathering products are the
source materials of soils and sedimentary rocks.
It is important to understand how weathering attacks source rocks and what
remains after weathering to form soils and be transported as sediment and dissolved
constituents to depositional basins. The ultimate composition of soil and terrigenous
sedimentary rock bears a relationship to the composition of the source rock, but it is
clear from study of residual soil profiles that both the mineral composition and the
bulk chemical composition of soils may differ greatly from those of the bedrock on
which they form. Some minerals in the source rock are destroyed completely during
weathering, whereas more chemically resistant or stable minerals are loosened from
the fabric of the decomposing and disintegrating rock and accumulate as residues. Dur
ing this process, new minerals such as ferric oxides and clay minerals may form in situ
in the soils from chemical elements released during breakdown of the source rocks.
Thus, soils are composed of survival assemblages of minerals and rock fragments
derived from the parent rocks plus any new minerals formed at the weathering site.
Soil composition is governed not only by the parent-rock composition but also by the
nature, intensity, and duration of weathering and soil-forming processes.
In the following section we examine the principal processes of subaerial weath
ering and discuss the nature of the particulate residues and dissolved constituents that
17
Reserve Reading: Do Not Redistribute
18 WEATHERING PROCESSES AND PRODUCTS AND SEDIMENT TRANSPORT

result from weathering. We also look briefly at the processes of submarine weathering.
Submarine weathering includes both the interaction of seawater with hot oceanic
rocks along mid-ocean ridges—a process that leaches important amounts of chemical
constituents from hot crustal rocks—and the low-temperature alteration of volcanic
rocks and sediments on the ocean floor.

2.2 SUBAERIAL WEATHERING PROCESSES

Physical Weathering
Physical (mechanical) weathering is the process by which rocks are broken into
smaller fragments through a variety of causes, but without significant change in chemi
cal or mineralogical composition. Except in extremely cold or very dry climates, physi
cal weathering occurs together with chemical weathering, and it is difficult to separate
their effects.
Frost wedging, caused by freezing and thawing of water in rock fractures, is the
most important physical weathering process in climates where recurring freezing and
thawing take place. Water increases in volume by about 9 percent when it changes to
ice, creating enough pressure in tortuous rock fractures to crack most types of rock. To
be effective, water must be trapped within the rock body, and repeated freezing and
thawing are necessary to allow progressive disintegration of the rock. This process
commonly produces large, angular blocks of rock but may also cause granular disinte
gration of coarse-grained rocks such as granites. The presence of microfractures and
other microstructures exerts an important control on the sizes and shapes of shattered
blocks. Mechanically weak rocks such as shales and sandstones tend to disintegrate
more readily than do hard, more strongly cemented rocks such as quartzites and
igneous rocks.
Early workers suggested that alternate expansion and contraction of rock sur
faces as a result of diurnal changes in temperature caused weakening of bonds along
grain boundaries and subsequent flaking off of rock fragments or dislodging of mineral
grains. The quantitative importance of this postulated process is still being debated.
Griggs (1936) carried out experiments in the laboratory over a temperature range of
110°C to simulate the equivalent of 244 years of heating and cooling; he showed that
in the absence of water, little disintegration occurred. Oilier (1969) points out, how
ever, that Griggs's laboratory specimens were unconfined and could expand in all
directions, whereas a patch of rock on the surface of a boulder is confined by neighbor
ing patches and can expand only outward. Such a confined area of rock is more likely
to fracture under repeated stress. Oilier also suggested that small heating and cooling
stresses maintained for longer periods of time than the 15-minute heating and cooling
cycles in Griggs's experiments might lead to permanent strain. Thus, high tempera
tures and temperature changes from day to night may cause fracturing of rocks owing
to thermal changes alone (Kerr et al., 1984).
High temperatures in desert environments also tend to promote weathering
caused by the crystallization of salts in pore spaces and fractures (Sperling and Cooke,
1980). Evaporation of water acts to concentrate dissolved salts in saline solutions that
have access to rock fractures and pores. Growth of salt crystals generates internal pres
sures that can force cracks apart or cause disintegration of weakly cemented rocks.
This process is most common in semiarid regions but can occur also along seacoasts
where salt spray is blown onto sea cliffs.
Reserve Reading: Do Not Redistribute
W E AT H E R I N G AND SOILS 19

Release of overburden pressure owing to erosion of overlying strata causes the


development of rock fractures that are nearly parallel to the topographic surface. These
fractures divide the rock into a series of layers or sheets; hence, this process of crack
formation is called sheeting. These layers increase in thickness with depth and may
exist for several tens of meters below Earth's surface. Sheeting is most conspicuous in
homogeneous rocks such as granite but may occur also in layered rocks.
Other processes that may contribute to mechanical weathering under certain
conditions include volume increases owing to hydration of clay minerals or other min
erals (which may be more important than heating and cooling); volume changes owing
to alteration of minerals such as biotite and plagioclase to clay minerals; alternate wet
ting and drying of rocks (causing alternate expansion and contraction); growth of plant
roots in the cracks of rocks; plucking of mineral grains and rock fragments from rock
surfaces by lichens as they expand and contract in response to wetting and drying; and
burrowing and ingestion of soils and loosened rock materials by worms or other organ
isms.
The grain size of the particulate rock materials that result from physical weather
ing is a function of the thoroughness of the weathering process, but it is determined
ultimately by the grain size and degree of cementation of the parent rock and by the
abundance and spacing of large fractures and microfractures. Coarse-grained parent
rocks such as granites tend to yield grains of individual minerals upon disintegration,
whereas physical weathering of fine-grained sedimentary, volcanic, or metamorphic
rocks is more likely to produce rock fragments as disintegration products (Boggs, 1968;
Carrol, 1970).

Chemical Weathering
Chemical Weathering Processes. Chemical weathering involves changes that can alter
both the chemical and the mineralogical composition of rocks. Minerals in the rocks
are attacked by water and dissolved atmospheric gases (oxygen, carbon dioxide), caus
ing some components of the minerals to dissolve and be removed in solution. Other
mineral constituents recombine in situ and crystallize to form new mineral phases.
These chemical changes, along with changes caused by physical weathering processes,
disrupt the fabric of the weathered rock, producing a loose residue of resistant grains
and secondary minerals. Water and dissolved gases play a dominant role in every
aspect of chemical weathering. Because some water is present in almost every environ
ment, chemical weathering processes are commonly far more important than physical
weathering processes, even in very arid climates. Nevertheless, owing to the low tem
peratures of weathering environment (<30°C), chemical weathering occurs very slowly.
The principal processes of chemical weathering are listed and briefly described in
Table 2.1 along with selected examples of new minerals formed in situ during the
weathering processes.
Hydrolysis is an extremely important chemical reaction between silicate miner
als and acids (solutions containing H+ ions) that leads to breakdown of the silicate
minerals and release of metal cations and silica. If aluminum is present in the minerals
undergoing weathering, clay minerals such as kaolinite, illite, and smectite may form
as a byproduct of hydrolysis. Thus, orthoclase feldspar can break down to yield kaolin
ite or illite, albite (plagioclase feldspar) can decompose to kaolinite or smectite, and so
on, as illustrated by the reactions in Table 2.1. The H+ ions shown in Table 2.1 are
commonly supplied by the dissociation of C02 in water, increasing its acidity (COz +
HzO <-» H2C03 <-> H* + HC03~). Thus, the more C02 that is dissolved in water, the more
Reserve Reading: Do Not Redistribute
to
o

TABLE 2.1 Principal processes of chemical weathering

Principal types
ofrock
Name of materials
process Nature of process Examples affected

Hydrolysis Reaction between H+ and OH" 2KAlSi308 + 2H+ + 9H20 -> H4Al2Si209 + 4H4Si04 + 2K+ Silicate minerals
ions of water and the ions of (orthoclase) aq (kaolinite) (silicic acid) aq
silicate minerals, yielding
soluble cations, silicic acid and 2NaAlSi308 + 2H+ + 9H20 -» H4Al2Si209 + 4H4Si04 + 2Na+
clay minerals (if Al present) (albite) aq (kaolinite) (silicic acid)
Hydration and Gain or loss of water molecules CaS04 • 2H20 <-» CaS04 + 2H20 (Dehydration) Evaporites
dehydration from a mineral, resulting in (gypsum) (anhydrite)
formation of a new mineral Fe203 + H20 <-» 2FeOOH (Hydration) Ferric oxides
(hematite) (goethite)
Oxidation Loss of an electron from an 4FeSi03 + 02 -» 2Fe203 + 4Si02 Iron and manganese-
element (commonly Fe or Mn) (pyroxene) (hematite) (quartz) bearing silicate
in a mineral, resulting in the minerals, sulfur
formation of oxides or, if water MnSi03 + 1/2 Oz + 2H20 -> Mn02 + H4Si04
is present, hydroxides (rhodonite)
2FeS2 + 15/2 Oz + 4H20 -» Fe203 + 4S042" + 8H+
(pyrite) (hematite)
Solution Dissolution of soluble minerals, H20 + C02 + CaC03 <-» Ca2+ + 2HC03" (carbonation) Carbonate rocks
commonly in the presence of (calcite) (bicarbonate)
C02, to yield cations and CaS04 • 2H20 -» Ca2++ S042" + 2HzO (direct solution) Evaporites
anions in solution (gypsum)
Ion exchange Exchange of ions, principally Na-clay + H* -> H-clay + Na+ Clay minerals
cations, between solutions and
minerals
Chelation Bonding of metal ions to organic Metal ions (cations) + chelating agent [excreted by lichens] -» H+ ions Silicate minerals
molecules having ring + chelate [in solution]
structures
Reserve Reading: Do Not Redistribute
W E AT H E R I N G AND SOILS 21

aggressive the hydrolysis reaction. Most of the silica set free during hydrolysis goes
into solution as silicic acid (HiSiOJ; however, some of the silica may separate as col
loidal or amorphous Si02 and be left behind during weathering to combine with alu
minum to form clay minerals or crystallize into minute grains of quartz (Krauskopf,
1979). Hydrolysis is the primary process by which silicate minerals decompose during
weathering. A more rigorous and detailed discussion of this process is given by Nahon
(1991, p. 7).
Hydration is the process by which water molecules are added to a mineral to
form a new mineral. Common examples of hydration are the addition of water to
hematite to form goethite, or to anhydrite to form gypsum. Hydration is accompanied
by volume changes that may lead to physical disruption of rocks. Under some condi
tions, hydrated milierals may lose their water, a process called dehydration, and be
converted to the anhydrous forms, with accompanying decrease in mineral volume.
Dehydration is relatively uncommon in the weathering environment because some
water is generally present in this environment.
Oxidation of iron and manganese in silicate minerals such as biotite and pyrox
enes, owing to oxygen dissolved in water, is also an important weathering process
because of the abundance of iron in the common rock-forming silicate minerals. An
electron is lost from iron during oxidation (Fe2+ -> Fe3+ + e~, where e" = electron trans
fer), which causes loss of other cations such as Si4+ from crystal lattices to maintain
electrical neutrality. Cation loss leaves vacancies in the crystal lattice that either bring
about the collapse of the lattice or make the mineral more susceptible to attack by
other weathering processes (Birkland, 1974). Oxidation of manganese minerals to form
oxides and silicic acid or other soluble products is a less important but common
weathering process. Another element that oxidizes during weathering is sulfur. For
example, pyrite (FeS2) is oxidized to form hematite (Fe203), with release of soluble sul
fate ions. Under some conditions where material undergoing weathering is water satu
rated, oxygen supply may be low and oxygen demand by organisms high. These condi
tions can bring about reduction of iron (gain of an electron) from Fe3+ to Fe2+. Ferrous
iron (Fe2+) is more soluble, and thus more mobile, than ferric iron (Fe3+) and may be
lost from the weathering system in solution.
Simple solution of highly soluble minerals such as calcite, dolomite, and gyp
sum owing to exposure to meteoric water (rainwater) during weathering can result in
decomposition of these minerals. If carbon dioxide is dissolved in the rainwater
through interaction with atmospheric or soil COz, the usual case in the weathering
environment, the solubilizing ability of water is enhanced (because of increased acid
ity), particularly for carbonate minerals. Simple solution of this type is an important
weathering process only in moderately wet climates where carbonate rocks or evapor
ites are present near the surface or at the water table.
Ion exchange is a weathering process that is particularly important in alteration
of one type of clay mineral to another. It is the reaction between ions in solution and
those held in a mineral, for example, the exchange of sodium for calcium. Most ion
exchange takes place between cations, but anion exchange also occurs.
Chelation involves the bonding of metal ions to organic substances to form
organic molecules having a ring structure (see Boggs, Livermore, and Seitz, 1985, for
additional discussion of chelates). During weathering, chelation, or organic complex-
ing, performs the dual role of removing cations from mineral lattices and also keeping
the cations in solution until they are removed from the weathering site. Chelated metal
ions will remain in solution under pH conditions and at concentration levels at which
nonchelated ions would normally be precipitated. The bonding of aluminum or iron
with a complexing agent and the subsequent removal of these elements from a rock are
Reserve Reading: Do Not Redistribute
22 WEATHERING PROCESSES AND PRODUCTS AND SEDIMENT TRANSPORT

of particular importance. A good example of natural chelation is provided by lichens


that cause an increase in the rate of chemical weathering on the rock surfaces on
which they grow by secreting organic chelating agents. In addition to the role of plant
organic matter as chelating agents, plants also enhance chemical weathering processes
by retaining soil moisture and by acidifying waters by release of C02 and various types
of organic acids during decay.

Rates of Chemical Weathering. Chemical weathering proceeds at different rates


depending upon the climate and the mineral composition and grain size of the rocks.
Weathering processes are more rapid in humid, hot climates than in cold or very dry
climates. Average rainfall is known to be a controlling factor in the weatering rate
(Nahon, 1991, p. 4); however, the influence of temperature on weathering rate is diffi
cult to quantify. A general qualitative rule is that chemical reaction rates are approxi
mately doubled with a 10°C increase in temperature (Hunt, 1979, p. 127). The rate of
weathering of silicate rocks of a given grain size is related to the relative stabilities of
the common rock-forming silicate minerals. Chemical stability refers to the resistance
of minerals to alteration or destruction by chemical processes. Table 2.2 shows the
order of relative stability to weathering of the most important mafic and felsic miner
als, as determined by Goldich (1938) through empirical study of sand- and silt-size
particles in soil profiles. Readers will recognize this order as the same as that in which
minerals crystallize in Bowen's reaction series. Minerals that crystallize at high tem
peratures (e.g., olivine) have the greatest degree of disequilibrium with surface weath
ering temperatures and thus tend to be less stable than minerals that crystallize at
lower temperatures (e.g., quartz). Furthermore, the high-temperature minerals are
bonded with weaker ionic or ionic-covalent bonds, whereas quartz is bonded with
strong covalent bonds. Jackson (1968) suggests that the stability of very fine-size (clay-
size) particles may differ somewhat from that of larger particles (Table 2.2).
Owing to the preponderance of low-stability minerals in basic igneous rocks,
these rocks tend to weather faster than acid igneous rocks of the same grain size. Thus,
gabbro weathers faster than granite, and basalt weathers faster than rhyolite. Fine-

TABLE 2.2 Relative stability of common sand-size minerals and various clay-size minerals
under conditions of weathering
Sand- and silt-size minerals* Clay-size minerals**
Mafic minerals Felsic minerals 1. Gypsum, halite
2. Calcite, dolomite, apatite
Olivine 3. Olivine, amphiboles, pyroxenes
Ca plagioclase 4. Biotite
Pyroxene 5. Na plagioclase, Ca plagioclase,
Ca-Na plagioclase K-feldspar, volcanic glass
Amphibole Na-Ca plagioclase 6. Quartz
Na plagioclase 7. Muscovite
Biotite 8. Vermiculite (clay mineral)
K-feldspar, muscovite, 9. Smectite (clay mineral)
quartz 10. Pedogenic (soil) chlorite
11. Allophane (clay mineral)
12. Kaolinite, halloysite (clay minerals)
1 13. Gibbsite, boehmite (clay minerals)
14. Hematite, goethite, magnetite
(Increasing stability) 15. Anatase, titanite, rutile, ilmenite (all,
titanium-bearing minerals), zircon
Source: *Goldich (1988); * "Jackson (1968).
Reserve Reading: Do Not Redistribute
WEATHERING AND SOILS 23

grained basic igneous rocks such as basalt may, however, weather more slowly than
coarse-grained granitic rocks (Birkland, 1974). There is no rule of weathering suscepti
bility that can be applied generally to sedimentary rocks. Rates of weathering of these
rocks are a function of the mineralogy, the amount and type of cement in the rocks,
and the climate. Limestones, for example, weather rapidly by solution in wet climates
and much more slowly in very arid or very cold climates. Quartz-rich sandstones
cemented with silica cement weather very slowly under most climatic conditions. The
general order of weathering of some common rocks, in order of increasing stability to
weathering in temperate to tropical climates, is probably limestone, dolomite, silt-
stone, sandstone, basalt, granite, chert, and quartzite (Birkland, 1974).
Finally, it is likely that rates of weathering have varied throughout geologic time
depending upon climatic conditions and vegetative cover. Prior to the development of
land plants in early Paleozoic time, absence of plant cover to hold soil moisture and
contribute organic acids probably slowed rates of chemical weathering while con
tributing to increased rates of physical erosion. The rates of chemical weathering are
now being investigated extensively through theoretical, empirical, and experimental
studies; however, many unresolved problems remain, such as mechanisms and rates of
individual processes, the relative roles of equilibrium and nonequilibrium kinetics,
and the influence of macro- and microenvironmental variables (Colman and Dethier,
1986).

Products of Subaerial Weathering


Subaerial weathering generates three types of weathering products (Table 2.3):
1. source-rock residues consisting of chemically resistant minerals and rock fragments
2. secondary minerals formed in situ by chemical recombination and crystallization,
largely as a result of hydrolysis and oxidation
3. soluble constituents released from parent rocks by hydrolysis and solution
Until they are removed by erosion, residues and secondary minerals accumulate at the
weathering site to form a soil mantle composed of particles of various compositions
and of grain sizes ranging from clay to gravel. Grain size and composition depend
upon the grain size and composition of the parent rock and upon the nature and inten
sity of the weathering process. These characteristics of the weathering environment are
in turn functions of climate, topography, and duration of the weathering process.

TABLE 2.3 Principal kinds of products formed by subaerial weathering processes and the types of sedimentary
rocks ultimately formed from these products

Type of weathering Ultimate


Weathering process product Example depositional product

Physical weathering Particulate residues Silicate minerals such as quartz Sandstones, conglom
and feldspar; all types of rock erates, mudrocks
Chemical weathering fragments
Hydrolysis Secondary minerals Clay minerals; fine quartz Mudrocks; mud matrix
Soluble constituents Silicic acid; K+, Na+, Mg2+, Ca2+, Chert, limestones,
etc. evaporites, etc.
Oxidation Secondary minerals Fine-grained Si02 minerals; Mudrocks; mud matrix
ferric oxides
Solution Soluble constituents Silicic acid, S042", etc. Chert, evaporites, etc.
Soluble constituents Bicarbonate, S042~, Ca2+, Mg2+, Limestones, evaporites,
etc. etc.
Reserve Reading: Do Not Redistribute
24 WEATHERING PROCESSES AND PRODUCTS AND SEDIMENT TRANSPORT

Source-Rock Residues. The residual particles in young or immature soils developed


on igneous or metamorphic rocks may include, in addition to rock fragments, assem
blages of minerals with low chemical stability, for example, biotite, pyroxenes, horn
blende, and calcic plagioclase. Mature soils, developed after more prolonged or inten
sive weathering of these rocks, tend to contain only the most stable minerals: quartz,
muscovite, and perhaps potassium feldspars. Because the silicate minerals that make
up siliciclastic sedimentary rocks such as sandstones have already passed through a
weathering cycle before the siliciclastic rocks were formed, the weathering products of
these rocks tend to be depleted in easily weathered minerals. Thus, even young soils
developed on siliciclastic sedimentary rocks may have assemblages of mature miner
als. Weathering of limestones by solution produces thin soils composed of the fine-size
insoluble silicate and iron oxide residues of these rocks.

Secondary Minerals. Secondary minerals developed at the weathering site are domi-
nantly clay minerals, iron oxides or hydroxides, and aluminum hydroxides. The com
mon secondary iron minerals include goethite, limonite, and hematite. The weathering
products that form are a function of both the nature and the intensity of the weathering
process and the composition of the parent rock. Clay minerals formed in immature
soils under only moderately intense chemical weathepjng conditions may be illites or
smectites. More prolonged and intense leaching conditions lead to formation of kaolin
ite. Under extremely intense chemical weathering conditions, aluminum hydroxides
such as gibbsite and diaspore are formed. These latter clay minerals are aluminum
ores. As an example of the influence of parent rock composition, potassium feldspars
tend to weather to kaolinite or, with more intense weathering, gibbsite. Biotite weath
ers to chlorite, smectite, vermiculite, and kaolinite, and plagioclase feldspars weather
to sericite, vermiculite, smectite, kaolinite, and gibbsite (Tardy et al., 1973).
Comparison of the chemical composition of unweathered silicate rocks with that
of the weathering products of these rocks shows a net loss owing to weathering of all
major cations except aluminum and iron (Krauskopf, 1979). In the oxidized state, alu
minum and ferric iron (Fe3+) are both relatively insoluble. Although considerable silica
is lost during weathering as soluble silicic acid, loss of Mg, Ca, Na, and K is compara
tively much greater. Therefore, the relative abundance of silica, aluminum, and ferric
iron in the particulate weathering residues of silicate rocks is greater than that in the
parent source rocks.

Soluble Materials. Soluble materials extracted from parent rocks by chemical weather
ing processes are removed from the weathering site in surface water or soil groundwa
ter more or less continuously throughout the weathering process. Ultimately these sol
uble products make their way into rivers and are carried to the ocean. The most
abundant inorganic constituents of rivers, representing the principal soluble products
of weathering, are, in order of decreasing abundance, HC03" (bicarbonate), Ca2+,
H4Si04 (silicic acid), S042" (sulfate), CI", Na+, Mg2+, and K+ (Garrels and McKenzie,
1971). These constituents are the raw materials from which chemically and biochemi
cally deposited rocks such as limestones and cherts are formed in the oceans.

2.3 SUBMARINE WEATHERING PROCESSES AND PRODUCTS

Geologists have long recognized that sediments and rocks on the seafloor are altered by
reaction with seawater, a process called halmyrolysis (submarine weathering). Hahny-
Reserve Reading: Do Not Redistribute
W E AT H E R I N G AND SOILS 25

rolysis includes alteration of clay minerals of one type to another, formation of glau-
conite from feldspars and micas, and formation of phillipsite (a zeolite mineral) and
palagonite (altered volcanic glass) from volcanic ash. Dissolution of the siliceous and
calcareous tests of organisms may also be considered a type of submarine weathering.
Prior to the 1970s, submarine weathering processes had not received a great deal of
research, and that they might have a significant effect on the overall chemical composi
tion of the oceans was not recognized. Our concept of the importance of submarine
weathering has changed dramatically since the middle 1970s because studies of vol
canic rocks and weathering processes on the seafloor show that submarine weathering
of basalts, particularly on mid-ocean ridges, is an extremely important chemical phe
nomenon. This process results in both widespread hydration and leaching of basalts
and changes in composition of reacting seawater owing to ion exchange during the
reaction of seawater with basalt.
Alteration of oceanic rocks occurs both at low temperatures (less than 20°C) and
at higher temperatures ranging to 350°C. Low-temperature alteration takes place as sea
water percolates through fractures and voids in the upper part of the ocean crust, per
haps extending to depths of 2 to 5 km. Olivine and interstitial glass in the basalts are
replaced by smectite clay minerals, and further alteration may lead to formation of zeo
lite minerals and chlorite. As a result of these changes, an exchange of elements
between rock and water takes place, and large volumes of seawater become fixed in the
oceanic crust in hydrous clay minerals and zeolites.
As a result of the discovery in 1977 of submarine thermal springs along the Gala
pagos Rift (Corliss et al., 1979), we are now aware that large-scale hydrothermal activ
ity is taking place along the crests of mid-ocean ridges. Since that initial discovery, sci
entists using submersible vehicles and water-sampling techniques have located many
additional hot springs along the East Pacific Rise and the Juan de Fuca Ridge in the
Pacific (e.g., McConachy et al., 1986; Baker and Massoth, 1987). Hydrothermal systems
have also been reported along the Mid-Atlantic Ridge (e.g., Rona et al., 1986) and even
on midplate volcanoes in the Hawaiian chain (Karl et al., 1988). These springs origi
nate from the activity of seawater in areas of active or recent volcanism along the ridge
crests. Seawater enters the ocean crust along fractures or other voids and comes in con
tact with hot volcanic rock. The heated water then flows out into the ocean through
vents on the ocean floor and mixes with the overlying water. The heated water rises as
hydrothermal plumes 100 to 300 m above the vent field. Exceptional plumes rising to
heights of 1000 m have also been reported (Cann and Strens, 1989).
On the top of the East Pacific Rise, for example, investigators have found spectac
ular vents composed of sulfide, sulfate, and oxide deposits up to 10 m tall and dis
charging plumes of hot solutions. These vents or chimneys are called black smokers if
they discharge water containing suspended, fine-grained, dark-colored minerals or
white smokers if the water contains no suspended dark minerals (McDonald, Spiess,
and Ballard, 1980). The temperature of the water when it emerges from the vents may
exceed 350°C. When these hot solutions mix with seawater of ambient temperature,
they precipitate various minerals, particularly pyrite (FeS) and chalcopyrite (CuFeS2),
to build sulfide deposits around the vents (Fig. 2.1). The deposits of fossil hydrothermal
systems have now been observed in ancient oceanic ophiolite complexes exposed on
land (Cann and Strens, 1989). Study of these ancient complexes suggests that the struc
ture of seafloor hydrothermal systems may be something like that shown in Figure 2.2.
Reactions between hot basalt and seawater may play a major role in regulating
the chemical composition of seawater. Magnesium and sulfate ions are removed from
seawater during this exchange, whereas other elements—such as calcium, manganese,
silicon, potassium, lithium, rubidium, and barium—are enriched in the seawater
Reserve Reading: Do Not Redistribute
26 WEATHERING PROCESSES AND PRODUCTS AND SEDIMENT TRANSPORT

FIGURE 2.1 A multiple-orifice


black smoker issuing from its con
structional chimney of chalcopyrite-
sphalerite-anhydrite on the East
Pacific Rise. The temperature of the
water issuing from the chimney is
350°C. The "smoke" is caused by the
presence of fine-grained sulfide pre
cipitates that form by reaction of the
hot waters with cold, ambient sea
water. (Photo by Fred Spiess. Cour
tesy of Scripps Institution of
Oceanography, University of Califor
nia, San Diego.)

(Edmond et al., 1982; Von Damm et al, 1985). The magnitude of hydrothermal alter
ation of basalts along mid-ocean ridges is still being investigated; however, Edmond
(1980) suggests that a volume of seawater equivalent to the entire ocean could be circu
lated through crustal rocks along ridges every 8 million years or so. This flow rate is
only about 0.5 percent of that of all the world's rivers, but the concentration of many
elements in these hot waters is 100 to 1000 times greater than in the average river.
Therefore, the fluxes of dissolved material out of and into ridge axes could be compa
rable to those derived from continental weathering. Clearly, both seafloor hydrother
mal reactions and continental weathering are important processes that supply salts to
the ocean.

2.4 SOILS

Soil-forming Processes
The subaerial physical, chemical, and biologic weathering processes discussed in the
preceding section generate a mantle of soil above bedrock. The characteristics and
thickness of this soil mantle are a function of the bedrock lithology, the climate (rain
fall, temperature), and the slope of the bedrock surface. These factors govern the inten
sity of weathering and determine which minerals survive to become part of the soil
profile, what new minerals are created in the soil, and the length of time soil materials
remain before being eroded and transported to depositional basins. On very steep
slopes, for example, the weathered mantle may be removed so rapidly by erosion that
little soil accumulates.
In addition to the weathering processes discussed, several other processes oper
ate within soils to produce their overall characteristics. Among the more important of
these processes are humification, gleization, podzolization, lessivage, ferrallitization,
Reserve Reading: Do Not Redistribute
WEATHERING AND SOILS 27

FIGURE 2.2 Hypothetical struc


ture of a hydrothermal black
smoker system underlying a
spreading ridge. (From J. R. Cann sulfide deposit
and M. R. Strens, 1989, Modeling
periodic megaplume emissions by ocean floor
black smoker systems: Jour. Geo-
phys. Research, v. 94, Fig. 1, p. 12,
228. Reproduced by permission.)

magma chamber 1km

calcification, and salizination and desalizination. Each of these processes is explained


and briefly described in Table 2.4 and illustrated graphically in Figure 2.3. Additional
details of these soil-forming processes may be found in Buol, Hole, and McCracken
(1989), Fanning and Fanning (1989), Fitzpatrick (1980), Nahon (1991), Retallack
(1990), and Ross (1989).

Soil Profiles and Soil Classification


Soils are classified on the basis of the characteristic horizontal layers or horizons that
are visible in roadcuts, pits, and so on. The thickness and nature of these soil horizons
are determined by the various soil-forming processes discussed and may vary widely.
Soil profiles can be divided crudely into three major horizons: A, B, and C. The A hori
zon is the dark-colored upper zone of organic accumulation composed of leaf litter that
is decaying and mixing with mineral soil. The B horizon is composed dominantly of
minerals with minimal organic content; most of the original rock structures have been
obliterated by soil-forming processes. The C horizon, which lies above bedrock, can be
Reserve Reading: Do Not Redistribute
28 WEATHERING PROCESSES AND PRODUCTS AND SEDIMENT TRANSPORT

TABLE 2.4 Some important soil-forming processes

Process Description
Humification The transformation of raw organic matter into soil humus
(dark, partially decomposed, more or less stable organic
material) and soluble organic acids (e.g., humic and
fulvic). Transformation takes place through biogenic
comminution of larger organic structures to fine,
amorphous organic matter and chemical oxidation of
organic matter to simpler compounds such as carbon
dioxide.
Gleization The reduction of iron under anoxic or anaerobic (low-
oxygen) soil conditions to produce bluish to greenish
gray waterlogged soil (gley). Anaerobic microorganisms
play an important role in reducing oxidized minerals.
The process of gleying tends to retard mineral
weathering.
Podzolization The downward (commonly) chemical migration of
aluminum, iron, and/or organic matter within a soil
profile. This process results in the relative
concentration of silica in the upper layer from which
these elements are removed and concentration of
aluminum, iron, and organic matter in a deeper layer.
Podzolization causes destruction of clay minerals and
leaching of exchangeable cations such as Ca2+, Mg2+,
K+, and Na+.
Lessivage The mechanical migration of clay-size mineral particles
from the A (surface) to B horizons of a soil, producing
in the B horizon relative enrichment in clay (argillic
horizons). This process produces a near-surface horizon
that is light in color and a subsurface clayey zone that
is darker in color.
Ferrallitization Intense, deep weathering resulting in thick, uniform soil
profiles depleted of exchangeable cations. The soil as a
whole is enriched in clay and sesquioxides (Fe203,
Al203) in the form of fine-grained, weather-resistant
minerals such as kaolinite, gibbsite, and hematite.
Calcification The accumulation of calcium in subsurface horizon near
the depth of average rainfall wetting in well-drained
soils of semiarid to subhumid regions. Sufficient water
may be present in these subsurface horizons to remove
some exhchangeable cations (Na+, K+, Mg2+) but not
Ca2+; thus, calcium becomes relatively enriched.
Salizination and Desalinization The accumulation (salizination) of soluble salts such as
sulfates and chlorides of calcium, magnesium sodium,
and potassium in salty horizons, or the leaching and
removal (desalizination) of salts from these horizons.
Source: Buol, Hole, and McCracken (1989); Retallack (1990).

deeply weathered but is relatively unaffected by soil-forming processes. Studies of soil


profiles show, however, that soil layers are commonly much more complex than indi
cated by this simple scheme. Up to 16 different kinds of soil horizons have been
described (Fanning and Fanning, 1989). Several systems for more detailed classifica
tion of soils are in existence: the British classification, the Australian handbook classi
fication, the U.S. Soil Taxonomy, and the FAO (UNESCO) world map classification
Reserve Reading: Do Not Redistribute
WEATHERING AND SOILS 29

GLEIZATION PODZOLIZATION LESSIVAGE FERRALLITIZATION CALCIFICATION SALINIZATION

FIGURE 2.3 Schematic illustration of the major regimes in which the common soil-forming
processes take place. (From G. J. Retallack, 1990, Soils of the past: Unwin Hyman (Chapman &
Hall), Fig. 4.12, p. 87, reproduced by permission.)

(Courtney and Trudgill, 1984, Chapter 7; Retallack, 1990, Chapter 5). One of the more
widely used soil classifications in the United States is the U.S. Soil Taxonomy (Soil
Survey Staff, 1975).

Paleosols
Paleosols, sometimes referred to as fossil soils, are buried soils or horizons of the geo
logic past. Most soil horizons that developed in the past on elevated landscapes were
eventually destroyed as erosion lowered the landscape. Nonetheless, some soils, pre
sumably those formed mainly in low-lying areas, escaped erosion to become part of the
stratigraphic record. Quaternary soils that formed particularly on glacial or fluvial
deposits are most common (e.g., Catt, 1986). Such soils that have not been buried are
called relict soils. Many buried soils of Quaternary and much older age are also
known. Old paleosols occur in the stratigraphic record at major unconformities,
including unconformities in Precambrian rocks, where their presence may reflect the
combined processes of soil formation, erosional landscape lowering, reorganization of
preexisting soil horizons, and changing flow of groundwater (Retallack, 1990, p. 14).
Paleosols are also present as interbeds in sedimentary sequences that are at least as old
as the Ordovician (e.g., Reinhardt and Sigleo, 1988). Geologists are becoming increas
ingly interested in paleosols as indicators of paleoenvironments and ancient climatic
conditions.
Reserve Reading: Do Not Redistribute
30 WEATHERING PROCESSES AND PRODUCTS AND SEDIMENT TRANSPORT

Recognition of Paleosols
Because interbedded paleosols in sedimentary sequences superficially resemble sedi
ments or sedimentary rocks, many paleosols have unquestionably gone unrecognized
in the past. Many of us have simply identified them as gray, rod, or green mudstones or
shales. As awareness of paleosols has increased, however, more and more paleosols are
being recognized. How can the ordinary geologist, not specifically trained in soil sci
ence, recognize paleosols in the field? Fenwick (1985) suggests a number of possible
criteria for recognition, including the following:
1. the presence of surface horizons enriched in organic matter
2. red-colored horizons that become more intense in color toward the top
3. marked decline in weatherable minerals toward the top of the soil profile
4. disruption of original structures by organisms such as earthworms or by physical
processes such as frost action and solifluction
Retallack (1988) discusses three main features of paleosols that are particularly
useful in field recognition: root traces, soil horizons, and soil structures. Root traces
provide diagnostic evidence that rock was exposed to the atmosphere and colonized
by plants, thus forming a soil. The top of a paleosol is the surface from which root
traces emanate. Root traces mostly taper and branch downward (Fig. 2.4), which helps
to distinguish them from burrows. On the other hand, some root traces spread laterally
over hardpans in soils, and some kinds branch upward and out of the soil. Root traces

FIGURE 2.4 An example of root


traces in a paleosol. The original
organic matter has been partially
replaced by iron oxides. Early
Miocene, Molalla Formation, west
ern Oregon. (Photograph courtesy
ofG. J. Retallack.)
Reserve Reading: Do Not Redistribute
WEATHERING AND SOILS 31

are most easily recognized when their original organic matter is preserved, which
occurs mostly in paleosols formed in waterlogged, anoxic lowland environments. Root
traces in red, oxidized paleosols consist mainly of tubular features filled with material
different from the surrounding paleosol matrix.
The presence of soil horizons is a second general feature of paleosols. The top of
the uppermost horizon of a paleosol is commonly sharply truncated by an erosional
surface, but boundaries between underlying horizons are typically gradational. Differ
ences in grain size, color, reaction with weak hydrochloric acid (to test for the pres
ence of carbonates), and the nature of the boundaries must all be examined to detect
soil horizons (Retallack, 1988).
Owing to bioturbation (disruption) by plants and animals, wetting and drying,
and other soil-forming processes, paleosols develop characteristic soil structures at the
expense of the original bedding and structures in the parent rock. One of the character
istic kinds of soil structure is a network of irregular planes (called cutans) surrounded
by more stable aggregates of soil material called peds. This structure gives a hackly
appearance to the soil. Peds occur in a variety of sizes and shapes (Fig. 2.5). Their
recognition in the field depends upon recognition of the cutans that bound them,
which commonly form clay skins around the peds. Other kinds of soil structure
include concentrations of specific minerals to form hard, distinct, calcareous, ferrugi
nous, or sideritic lumps called glaebules (a general term including nodules and con
cretions). More diffuse, irregular, or weakly mineralized concentrations are called mot
tles.
This short, generalized description of paleosols is intended only to pique reader
interest in fossil soils. Several books, many of which are listed under Further Readings
below, are available to anyone who wishes to pursue further research on paleosols.

ANGULAR SUBANGULAR GRANULAR CRUMB


TYPE PLATY PRISMATIC COLUMNAR
BLOCKY BLOCKY

SKETCH

tabular and elongate with flat elongate with equant with sharp equant with dull spheroidal with rounded and
DESCRIPTION horizontal to land top and vertical domed top and interlocking edges interlocking edges slightly interlocking spheroidal but not
surface to land surface vertical to surface edges interlocking
USUAL HORIZON E.Bs.K.C Bt Bn Bt Bt
initial disruption of as for prismatic, cracking around as for angular active bioturbation as for granular-,
swelling and
relict bedding; shrinking on but with greater roots and burrowsj blocky, but with and coating of soil including fecal
MAIN more erosion and with films of cloy,
accretion of wetting and erosion by swelling and pellets and relict
LIKELY soil clasts
cementing material drying percolating water, shrinking on deposition of sesquioxides and
CAUSES and greater wetting and material in cracks organic matter
swelling of clay drying
very thin < I mm very fine<l cm very fine< I cm very fine < 05 cm very fine<05cm very fine<lmm very fine<l mm

thin I to 2 mm fine I to 2 cm fine I to 2 cm fine 0.5 to I cm fine 05 to I cm fine I to 2 mm fine I to 2 mm


SIZE
medium 2 to 5 mm medium 2 to 5 cm medium 2 to 5 cm medium I to 2 cm medium I to 2 cm medium 2 to 5 mm medium 2 to 5 mm
CLASS
thick 5 to 10 mm coorse 5 to 10 cm coarse 5 to 10 cm coarse 2 to 5 cm coarse 2 to 5 cm coorse 5 to 10 mm not found

not found
very thick > 10 mm very coarse > 10 cm very coarse> 10cm very coarse> 5cm very coarse*5cm very coarse»IOmm

FIGURE 2.5 Characteristics of various kinds of soil peds. (From G. J. Retallack, 1988, in J. Rein-
hardt and W. R. Sigleo, eds., Field recognition of paleosols: Geol. Soc. America Spec. Paper 216.
Fig. 9, p. 216.)
Reserve Reading: Do Not Redistribute
32 WEATHERING PROCESSES AND PRODUCTS AND SEDIMENT TRANSPORT

FURTHER READINGS
Weathering Bronger, A., and J. A. Catt (eds.), 1989, Paleopedology:
Balasubramaniam, D. S., et al. (eds.), 1989, Weathering; Nature and application of paleosols: Catena Verlag,
its products and deposits, v. 1, Processes, 462 p., v. Destedt, Germany, 232 p.
II, Deposits, 671 p.: Theophrastus Publicatione, Buol, W. W„ F. D. Hole, and R. J. McCracken, 1989, Soil
Athens, Greece.
genesis and classification, 3rd ed.: Iowa State Uni
Birkland, P. W., 1974, Pedology, weathering and geo- versity Press, Ames, 446 p.
morphological research. Oxford University Press, Catt, J. A., 1986, Soils and Quaternary geology: A hand
New York, 285 p. book for field scientists: Clarendon Press, Oxford,
Carroll, D., 1970, Rock weathering: Plenum Press, New 267 p.
York, 203 p. Courtney, F. M., and S. T. Trudgill, 1984, The soil, 2nd
Colman, S. M., and D. P. Dethier, 1986, Rates of chemi ed.: Edward Arnold, London, 123 p.
cal weathering of rocks and minerals: Academic Duchaufour, P., 1982, Pedology: George Allen & Unwin,
Press, Orlando, 603 p. London, 448 p.
Drever, J. I. (ed.), 1985, The chemistry of rock weather Fanning, D. S., and M. C. B. Fanning, 1989, Soil: Mor
ing: D. Reidel, Hingham, Mass., 336 p. phology, genesis, classification: John Wiley & Sons,
Lerman A., and M. Meybeck (eds.), 1988, Physical and New York, 395 p.
chemical weathering in geochemical cycles: Kluwer Reinhardt, J., and W. R. Sigleo (eds.), 1988, Paleosols
Academic, Dordrecht, 375 p. and weathering through geologic time: Principles
Loughnan, F. C, 1969, Chemical weathering of silicate and applications: Gecil. Soc. America Spec. Paper
minerals: Elsevier, New York, 154 p. 216,181 p.
Nahon, D. B„ 1991, Introduction to the petrology of Retallack, G. J., 1990, Soils of the past: Unwin Hyman,
soils and chemical weathering: Jphn Wiley & Sons, Boston, 520 p.
New York, 313 p.
Wright, V. P., 1986, Paleosols: Their recognition and
Rona, P. A. and R. P. Lowell (eds.), 1980, Seafloor interpretation: Princeton University Press, Prince
spreading centers: Hydrothermal systems: Bench ton, N.J., 315 p.
mark Papers in Geology, v. 56, Dowden, Hutchinson Yaalon, D. H., 1971, Paleopedology: Origin, nature, and
and Ross, Stroudsburg, Pa., 424 p.
dating of palesols: International Society of Soil Sci
ence and Israel University Press, Jerusalem, 350 p.
Soils and Paleosols
Boardman, J. (ed.), 1985, Soils and Quaternary land
scape evolution: John Wiley & Sons, Chichester,
391 p.

Potrebbero piacerti anche