You are on page 1of 14

Geoderma 124 (2005) 23 36

www.elsevier.com/locate/geoderma

Historical development of key concepts in pedology


J.G. Bockheim a,*, A.N. Gennadiyev b, R.D. Hammer c, J.P. Tandarich d
a

Department of Soil Science, University of Wisconsin, 1525 Observatory Drive, Madison, WI 53706-1299, USA
b
Faculty of Geography, M.V. Lomonosov Moscow State University, Moscow 119992, Russia
c
Department of Civil and Environmental Engineering, University of Missouri, Columbia, MO 65211-2200, USA
d
Hey and Associates, 1141 Commerce Drive, Geneva, IL 60134, USA
Received 1 May 2003; received in revised form 16 February 2004; accepted 18 March 2004
Available online 27 April 2004

Abstract
As a subdiscipline of soil science, pedology consists of an accepted body of laws and theories that cover a range of related ideas
and concepts. We have traced the history of these concepts as they pertain to the definition of the soil; soil horizons, profiles, and
pedons; soil-forming factors; pedogenic processes; soil classification; soil geography and mapping, and soil landscape
relationships. The presented concepts have proven to be useful in our careers and are offered here to generate discussion in the
pedology community. Because of space limitations, we have not attempted to critique these concepts. The concepts identified here
are useful not only for understanding the development of pedology, but also for identifying future areas of research and providing a
frame of reference from which pedologists can evaluate potential scientific contributions to a rapidly changing world.
D 2004 Elsevier B.V. All rights reserved.
Keywords: Pedology; Concepts; Soil history; Soil genesis; Soil landscape

1. Introduction
Pedology, a component of soil science, has evolved
through the creation and justification of ideas and
generation and rationalization of processes (Huggett,
1997). Defined by Joffe (1936) as the study of the
soil body in its natural position, pedology traditionally has been subdivided into soil morphology, soilforming factors, soil-forming processes, soil classification, and soil geography and mapping (Sokolov,
1996; Buol et al., 1997). Some have criticized that
pedology is too descriptive and overly dependent on

* Corresponding author. Fax: +1-608-265-2595.


E-mail address: bockheim@wisc.edu (J.G. Bockheim).
0016-7061/$ - see front matter D 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.geoderma.2004.03.004

the framework of soil classification systems (e.g.,


White, 1997); however, our review of the literature
shows a rich diversity of concepts that have led to the
development of pedology as an important component
of Earth sciences.
The literature1 contains several publications on
specific concepts in pedology (Arnold, 1965; Huggett,
1975; Yaalon, 1975; Sokolov, 1996; Targulian and
Sokolova, 1996; Nikitin, 2001). Many other substantive works on the history of soil science have

1
We have attempted to be unbiased in our selections of key
literature and have striven to provide an international balance to our
review. However, we recognize that the topic is conducive to
subjective interpretation.

24

J.G. Bockheim et al. / Geoderma 124 (2005) 2336

contributed to our understanding of some early concepts in pedology (Jenny, 1961b; Simonson, 1968;
Arnold, 1983; Tandarich et al., 1988, 2002; Hudson,
1992; Krupenikov, 1992; Gennadiyev et al., 1995,
1996; Buol et al., 1997; Yaalon, 1997; Brevik, 1999).
However, none of these works summarized key concepts in pedology.
Accordingly, the objective of this study was to
identify the key concepts in pedology. Some of the
identified concepts may be considered obsolete, dated,
or applicable to a particular soil system. However, as
Cline (1961) observed, each individuals concept of a
discipline represents the state of knowledge that the
individual has acquired, tempered by that individuals
unique experiences and perspectives. As knowledge is
acquired in a discipline, new models replace old
models. Cline (1961) further suggested that the best
scientific models are those that prompt sufficient,
focused research to set the template for their own
replacement by subsequent, more complete models.
An understanding of currently accepted concepts
often requires knowledge of the developmental history of ideas that led progressively from one concept to
another. We have made no attempt to critique the
concepts, as that is beyond the scope of this review.
1.1. Soil definition
Simonson (1968) provided a historical review of
the concept of soil. The earliest definition of the soil
was that it is a medium for plant growth. Known as
the edaphological concept of soil, this view prevailed in soil science early in the 20th century (e.g.,
Lyon and Buckman, 1922). This view is still emphasized where soil fertility is the primary concern and
the focus is on physical and chemical soil attributes
important for plant growth with little regard for
conditions external to the rooting medium.
Secondly, the soil can also be viewed as a mantle of
loose and weathered rock (Shaler, 1891; Hilgard,
1906; Coffey, 1912; Ramann, 1928; Targulian,
2001). A third view of the soil is that it is the excited
skin of the subaerial part of the Earths crust (Nikiforoff, 1959). Runge (1973) proposed an energy model
based on energy vectors operative in soils. Chesworth
(1973) viewed soils as systems spontaneously moving
toward a state of equilibrium. During soil development, the weathering process preferentially removes

mobile components from upper horizons and concentrates the relative immobile components in lower
horizons. Huggett (1975) viewed soil landscape systems as storing, transforming and transmitting power
plants whose inputs are material and energy and
whose outputs include clastic sediments, colloids,
and soluble material. An important component of
Huggetts model is the perspective of the soil as a
body larger than a pedon or polypedon, within which
the lateral down-slope migration of dissolved and
suspended constituents in temporally percolating water is a key component. Daniels and Hammer (1992)
expanded upon this concept in a perspective that
further welded geomorphic processes to soil-forming
conditions.
Dmitriev (1996) presented a detailed analysis of
existing definitions of soil within the scope of Dokuchaevs approach; he defined the soil as a natural
exon, which has properties developed as a result of
the influence of autotrophic and heterotrophic organisms on soil constituents resulting from exogenic
transformations. Targulian and Sokolova (1996) described the soil as a reactor, memory, and regulator of
biosphere interactions. Following the ideas of Vernadskii (1926) and Kovda (1990), Dobrovolskii et al.
(2001) considered the soil as a component of the
biosphere with ecological functions responsible for
biodiversity, productivity, etc. Nikitin (2001) considered the soil as an abiotic system with numerous
biospheric functions and emphasized that a soil acts
as a habitat, accumulator, and source of substances for
all terrestrial organisms, as a link between the biological and geological cycles of matter, and as a planetary
membrane (protective barrier and buffer system) that
maintains suitable conditions for normal development
of the biosphere.
These approaches have led to more recent applications of thermodynamics in modeling soil processes.
Today soils are recognized as complex systems that
change in entropy and are controlled by convergent
and divergent developmental pathways (Phillips,
1998, 2000).
The most common view of the soil from a pedological perspective is that it is an independent natural
evolutionary body that can be subdivided into subcompartments and that has formed under the influence
of the five soil-forming factorsclimate, organisms,
parent materials, relief, and time. This view was

J.G. Bockheim et al. / Geoderma 124 (2005) 2336

initially proposed by Dokuchaev (1879a,b, 1883,


1893, 1899b) and supported by Sibirtsev (1900).
Glinka (1914, 1931), Coffey (1909, 1912) and Marbut
(1927, 1935) introduced this view of the soil to the
USA. According to Yaalon (1983), the realization that
soil profiles are a record of soil genesis and history
was possibly the most significant discovery in pedology. Brevik (1999) and Ableiter (1949) noted that
Coffey (1909) probably was the first American to
incorporate Dokuchaevs concept of soil as a natural
body into his perspective of soil, but that this idea was
ignored for nearly a decade. Ableiter (1949) suggested
that new ideas, no matter how sound, probably would
not be immediately accepted if they differ profoundly
from currently prevailing views.
According to Soil Taxonomy (Soil Survey Staff,
1960, 1975, 1999), the soil is a natural body composed of solids (minerals and organic matter), liquid,
and gases that occurs on the land surface, occupies
space, and is characterized by horizons or layers that
are distinguishable from the initial materials.
From this brief summary, key concepts regarding
the definition of the soil include the following:
1. The soil is a medium for plant growth.
2. The soil is a mantle of loose and weathered rock.
3. The soil is an independent, natural, evolutionary
body.
4. The soil is the excited skin of the subaerial part
of the Earths crust and a key component of the
biosphere.
5. The soil is a natural body composed of solids
(minerals and organic matter), liquids, and gases
that occurs on the land surface, occupies space, and
is characterized by horizons or layers that are
distinguishable from the initial material.
1.2. Soil horizons, profiles, pedons, and stratigraphic
units
Soil horizons originally were used solely for descriptive purposes (Dokuchaev, 1879a,b, 1883; Dokuchaev and Sibirtsev, 1893), but they were later
identified as genetic layers more or less parallel to
the earths surface that could be (1) distinguished on
the basis of morphological, chemical, and physical
properties and (2) used to interpret the developmental
history of the soil (Glinka, 1914; Marbut, 1921, 1927,

25

1928; Shaw, 1927). Marbut (1921) was the first


person to suggest that soil horizons should be used
to classify and differentiate soils. He proposed eight
descriptive criteria of the soil, seven of which were
soil-horizon attributes. This marked the beginning in
the USA of recognizing soils as natural bodies distinct
from geologic materials.
A soil profile was recognized as a vertical, twodimensional section of the soil from the surface to the
underlying unweathered material (Dokuchaev,
1879a,b, 1883, Glinka, 1914; Marbut, 1927, 1928;
Shaw, 1927). Tandarich et al. (1994, 2002) described
the history of the profile and the delineation of
particular horizons.
A question that arose early in the science of
pedology was what is the smallest unit that can be
recognized on the landscape as a soil? In the USA,
the pedon was identified as the smallest three-dimensional unit (volume) of soil that shows all of the soil
horizons present and their relationships (Simonson
and Gardner, 1960; Johnson, 1963). The cross-sectional area of the pedon later was defined as ranging
from 1 to 10 m2 depending on soil variability (Soil
Survey Staff, 1960, 1975, 1999).
Cline (1949) proposed the modal pedon as the
central individual profile within a mapping unit or soil
series. This concept was based on the statistical
mode of a population and was extended to be the
representative soil profile for a soil series. The
type location for a series or mapping area (individual soil survey) often is perceived to be the modal soil
for the population. Cline later regretted having proposed the modal pedon concept, because he thought
that it focused attention on profiles at the expense of
soil distributions.2 The polypedon was recognized as a
soil individual that is comprised of contiguous
pedons, all of which have characteristics lying within
the defined limits of a single soil series or soil
mapping unit (Johnson, 1963).
Broader views of the soil include recognition of the
geosol, a laterally traceable, mappable, geological
weathering profile that has consistent stratigraphic
position (Morrison, 1967), which is recognized by
the North American Commission on Stratigraphic
Nomenclature (1983) as the fundamental pedostrati-

Richard Cline, personal communication.

26

J.G. Bockheim et al. / Geoderma 124 (2005) 2336

graphic unit. The weathering profile, the geologic


homologue of the soil profile, is defined as a vertical
section that extends downward through the zones
affected by subaerial weathering to the unweathered
zones of unconsolidated or consolidated geologic
material, if present (see Tandarich et al., 1994,
2002). Tandarich et al. (1994) advanced the pedoweathering profile as a synthetic paradigm that
melded pedological and geological profile concepts.
Some key concepts relative to the subdivision of
the soil body include the following:
1. Soil horizons are genetic layers more or less
parallel to the earths surface that are distinguished
on the basis of morphological, chemical, and
physical properties.
2. A soil profile is a vertical, two-dimensional section
of the soil from the surface to the underlying
unweathered material.
3. The pedon is the smallest volume of soil that is
described and sampled; the horizontal area of the
pedon ranges from 1 to 10 m2 depending on soil
variability.
4. A polypedon is a collection of pedons within the
defined limits of a single soil series.
5. The modal profile is the central concept of soil
attributes in a soil series.
6. A geosol is a laterally traceable, mappable,
geological weathering profile that has consistent
stratigraphic position; it is the fundamental pedostratigraphic unit.
7. A weathering and pedo-weathering profile is a
vertical section that extends downward through the
zone affected by subaerial weathering that includes
the soil profile to the unweathered, unconsolidated
or consolidated rock.
1.3. Soil-forming factors
Dokuchaev (1879a,b, 1883, 1893, 1899b) recognized that the soil is a function of the interplay
of climate, organisms, relief, and parent material,
all operating over time. Dokuchaev, and later
Glinka (1914, 1927), Joffe (1936) and Marbut
(1927) considered soil-forming factors as the causes
of soil formation and soil properties as their effects.
Joffe (1936) stated that the soil-forming factors set
the conditions for internal soil-forming processes.

Shaw (1932) prepared the first soil-forming factor


equation:
S MC V T D

In this equation, M = parent material, C = climatic


factors, V = organic life, T = time, and D = modification of the soil by erosion and deposition. In the
discussion following Shaws (1932) presentation at
the International Congress of Soil Science, a scientist in the audience suggested that Shaw write the
equation as an integral function to more precisely
capture the interacting influences of the factors
upon an individual soil. Jenny (1941) stated that
the introduction of causality aspects to soil formation is not fruitful, because every property may be
considered a cause as well as an effect. According
to Jenny (1941), soil-forming factors are not forces
or causes, but are independent variables. Jenny
(1941) rewrote Shaws (1932) soil-forming factor
equation:
S f cl; o; r; p; t . . .

In this equation, cl = climate, o = organisms, r = relief, p = parent material, and t = time.


The soil-forming-factor equation has become a
popular concept in pedology. It serves to simplify
complex relations among the soil-forming factors in
mathematical terms and provides a basis for interpreting soil attributes in terms of genetic history. However, from a pragmatic standpoint, the equation has
never really been solved (Crowther, 1953; Kline,
1973; Phillips, 1998).
State-factor equations were intended to describe
the influence of a single factor on soil properties and
are useful in studying soil development (Jenny,
1961a). Yaalon (1975) reviewed quantitative solutions
of the univariant soil-forming functions and suggested
that computer simulation strategy can advance the
quantification of the process-oriented models of soil
dynamics. Richardson and Edmonds (1987) developed linear regression equations depicting Jennys
relative effectiveness of state factors equation.
Phillips (1998) showed the links between the application of soil state-factor theory and nonlinear dynamic systems for modeling complex systems.
Perhaps the most emphasized soil-forming factor is
soil climate; however, interactions among climate,

J.G. Bockheim et al. / Geoderma 124 (2005) 2336

vegetation and soils are complex and are not easily


distinguished on some landscapes. Hilgard (1882)
stressed the role of climate in soil formation, emphasizing soil temperature and moisture conditions. Marbut (1935) subdivided soils of the USA into two
climatic groups, Pedalfers and Pedocals, which were
based on leaching and non-leaching soil water
regimes, respectively. Bonneau (1982) identified
pedoclimate as the driving factor of soil formation;
indeed, it is used to delineate soils in global soil
taxonomic systems at the suborder (second) level.
Butler (1959) and Hack and Goodlett (1960)
provided evidence that soil development and erosion
have been periodic and are driven by episodic geomorphic processes (e.g., Butlers K cycles), rather
than being continuous in concert with climate
change. Daniels and Hammer (1992) emphasized
the integrating roles of water and gravity in synergistic soil-geomorphic processes. The recognition that
geomorphic processes are periodic and related to soil
development (Lobeck, 1939) was an important conceptual difference from the geologic concept of
uniformitarianism upon which Marbut (1935) based
his ideas of normal soils. Marbuts normal soil
concept was based on Davis (1909) slope-forming
geographic principles.
The role of vegetation was recognized early in the
development of pedology as a reflection of climate
and as being important in soil formation (Dokuchaev,
1883). Rode (1947) proposed that soil development is
theoretically possible only within the soil plant system, thereby emphasizing the edaphologic perspective
of the soil. The role of organisms in soil development
was emphasized by Hole (1961) and Johnson et al.
(1987) in terms of pedoturbation, or soil mixing,
which represents a local cyclic movement of soil
materials.
Parent material was recognized early in the history
of pedology as a key soil-forming factor particularly
at the regional scale (Dokuchaev, 1883; Hilgard,
1906; Coffey, 1912). Early approaches to distinguishing soil individuals were based on the geologic
substratum or major geologic events (i.e., glaciation).
Topography was initially recognized as being an
important factor in the vertical zonation of soils. The
topographic factor also has received special attention
in connection to the soil catena concept. A catena
was recognized as a drainage sequence of soils

27

(Milne, 1935; Bushnell, 1942; Fridland, 1972; Sommer and Schlichting, 1997). Dan and Yaalon (1968)
identified a pedomorphic surface as a landscape in
which the soils are genetically and evolutionarily
interdependent and a pedomorphic surface is synonymous with a geomorphic surface. The soil-geochemical catena concept was developed in Russia (Polynov,
1953; Glazovskaya, 1964; Kasimov and Perelman,
1992).
Early ideas in soil science recognized that soils
mature with time (Dokuchaev, 1883; Kossovich,
1911; Shaw, 1932; Marbut, 1935). Stevens and
Walker (1970) illustrated the importance of chronosequences in studying soil development. Soil
chronosequences have different combinations of
isochronism (isochrony) and/or time transgression
of incipience and cessation of development of their
encompassed end members (Vreeken, 1975). Bockheim (1980) prepared chronofunctions from published data and showed that linear and curvilinear
models effectively depicted changes in soil properties over time. Gennadiyev (1990) and Birkeland
(1999) analyzed various approaches to the role of
time in soil development and different schemes of
soil formation that included the time factor.
In summary, some key concepts with regards to
soil-forming factors include the following:
1. The soil is a function of the interplay of climate,

2.
3.

4.
5.
6.

7.
8.

organisms, relief, and parent material, all operating over time.


Soil-forming factors set the conditions for soilforming processes.
Soil-forming equations simplify complex relationships among the soil-forming factors in mathematical terms and provide a basis for interpreting
soil attributes in terms of genetic history.
Climate is a driving factor of soil formation.
Parent material is a key soil-forming factor,
especially at the regional scale.
Topography is a key soil-forming factor at the
landscape level; a catena is the interlocking of
soils on a landscape.
Organisms play a key role in soil development
from a microscopic to the continental scale.
Soil chronosequences are valuable for understanding rates and directions of soil and landscape
evolution.

28

J.G. Bockheim et al. / Geoderma 124 (2005) 2336

9. State-factor equations are intended to describe the


influence of a single factor on soil properties (i.e.,
the other factors are assumed to be more or less
univariant over time) and are useful in studying
soil development.
10. Soil development and erosion have been periodic
rather than continuous.
1.4. Soil-forming processes
Perhaps the earliest recognition of soil-forming
processes was posited in the concepts of eluviation
and illuviation (Glinka, 1914; Marbut, 1927). Eluviation is the removal of material in suspension or
solution primarily from A and E horizons; and illuviation is the deposition of these weathering products in
B-horizons. Huggett (1997) adopted a similar view
that soils represent a balance between soil formation
and dynamic denudation, which he identified as a key
paradigm in pedology.
Beginning with the pioneering efforts of Dokuchaev (1879a,b, 1883, 1899a) and continuing with his
Russian successors (Polynov, 1923; Neustruev, 1930),
soil scientists paid considerable attention to soil-forming processes. Soil processes were recognized in the
USA until the late 1950s (Marbut, 1927, 1928; Joffe,
1936; Baldwin et al., 1938). For example, Joffe
(1936) and Byers et al. (1938) identified seven predominant soil-forming processes, including podzolization, calcification, laterization, gleization,
salinization, solonization, and solodization.
Three general kinds of soil-forming processes have
been recognized in Russia, including microprocesses,
elementary processes, and general soil-forming processes (Rode, 1971). Microprocesses include various
biological, chemical, and physical processes that collectively define the elementary soil-forming processes; elementary soil-forming processes are responsible
for soil-horizon formation; and general soil-forming
processes are responsible for formation of the entire
soil profile. Gerasimov and Glazovskaya (1960) identified 10 elementary soil-forming processes that contribute to the development of genetic horizons and the
entire profile. Gerasimov (1975) systematized elementary pedogenic processes as a basis for the genetic
classification of soils, using a hierarchy: soil-forming
factors>soil-forming processes>diagnostic soil properties. Rode (1971) showed how successive cycles of

soil microprocesses were transformed into directed


irreversible, elementary soil-forming processes that
result in a natural body. Targulian and Sokolov
(1976) developed the idea of a soil as a moment,
which was represented by labile short-term soil-forming processes; soil as a memory was represented by
consequences of long-term processes. Rozanov
(1983) grouped about 70 elementary soil-forming
processes into nine categories.
The movement away from an emphasis on soilforming processes in the USA was predicated on the
assumption that soil properties result from soil processes and are more readily quantifiable than the
processes (Bockheim and Gennadiyev, 2000). Soil
processes were viewed as being poorly understood;
and specific pedogenic processes were recognized as
occurring simultaneously in a given soil, reinforcing
or contradicting one another.
Simonson (1959) proposed that soil horizons form
from two overlapping steps, the accumulation of
parent materials and the differentiation of horizons
in the profile. Horizon differentiation was ascribed to
the internal (soil profile) balance among additions,
losses, transfers, and transformation of energy and
matter. This concept has been the underpinning of soil
genesis for the past 45 years. Crompton (1960)
suggested that soil formation represented the balance
between release of elements from weathering (richness of weathering) and loss of elements due to
leaching (intensity of leaching), as moderated by
new materials that form from the products of weathering. The mass-balance approach to soil development, originally proposed by Haseman and Marshall
(1942) and expanded upon by Chadwick et al. (1990),
enables the calculation of gains and losses in weathering products based on the distribution of index
minerals.
As soil properties approach a dynamic steady state,
they adjust at a decreasing rate (Yaalon, 1971). Muhs
(1984) recognized that intrinsic thresholds that explain instability in the absence of environmental
change exist in soil systems. Yaalon (1983) identified
and discussed feedback systems in soils that slowly
change internal processes. For example, as argillic
horizons form, they gradually reduce the internal
vertical percolation of soil water.
Bockheim and Gennadiyev (2000) synthesized
approaches of different investigators and identified

J.G. Bockheim et al. / Geoderma 124 (2005) 2336

17 elementary soil-forming processes that could be


linked to soil taxa and diagnostic horizon, properties,
and materials in Soil Taxonomy (Soil Survey Staff,
1999) and the World Reference Base (WRB) for Soil
Resources (FAO, 1998).
A summary of main ideas about soil-forming
processes includes the following:
1. Eluviation represents the removal of soil material
in suspension or solution; and illuviation represents
the deposition of weathering products in the
subsoil.
2. Soil horizons form as a result of the accumulation
of parent materials and the differentiation of
horizons due to balances among additions, losses,
transfers, and transformations of energy and matter.
3. Soils represent a balance between the richness of
weathering and the intensity of leaching.
4. Soil properties may rapidly or slowly approach a
dynamic steady state and upon burial may be
persistent or reversible.
5. Intrinsic thresholds exist in soil systems that
explain instability in the absence of environmental
change.
6. Soils may develop feedback systems that change
internal processes over time.
7. Three general kinds of soil-forming processes have
been recognized. Microprocesses include various
biological, chemical, and physical processes that
collectively define the elementary soil-forming
processes; elementary soil-forming processes are
responsible for the formation of specific soil
horizons; and general soil-forming processes are
responsible for formation of the entire soil profile.
8. Soil-forming processes can be linked to diagnostic
surface and subsurface horizons and higher categories of soil classification systems.
1.5. Soil classification
Joel (1926) explained that one of the fundamental
needs of any natural science is to classify the
proposed bodies of study. This was a particularly
daunting challenge for soil science because the soil,
unlike mammals, reptiles, insects, birds and plants,
does not exist as discrete entities. Rather, the soil is a
continuous multivariable natural body that changes at
different rates and along different pathways into

29

diverse combinations of the defining attributes (Joffe,


1936). Crowther (1953) emphasized the temporally
dynamic natures of many key soil attributes as yet
another dimension of complexity in an already
conceptually bewildering entity. Joel (1926) and
Marbut (1935) believed that the study of soils could
not advance as a science until a generally accepted
classification system, based on the soil itself, was
developed.
The earliest soil classification systems were based
mainly on geomorphological and geological concepts,
such as mineralogical and chemical properties of
parent materials. Soil classification schemes, beginning with Dokuchaev (1883, 1949) and continuing
with those used in the USA until 1960 (Marbut, 1928;
Baldwin et al., 1938), were based on the zonality
concept. For example, in the USA soils were divided
into three broad orders. Soils in the zonal (normal
soils) order contained well-developed characteristics
reflecting the predominant influence of climate and
vegetation on well-drained stable sites; soils in the
intrazonal order were well developed but had characteristics reflecting local factors such as relief, parent
material, or age; soils in the azonal order were poorly
developed.
The evolutionary approach to soil classification
was developed by Polynov (1923) and Kovda et al.
(1967). Higher categories in the latter scheme were
based on different geochemical mass and energy
transfers, as evidenced by major stages and trends in
weathering and humus and clay mineral formation.
Elements of the evolutionary approach are contained
in the French soil classification scheme (Duchaufour,
1968).
Some early and the current soil classification
schemes used in Russia are based on the concept of
genetic profiles (Gerasimov, 1975; Glazovskaya,
1983; Dobrovolsky and Trofimov, 1996; Goryachkin
et al., 2003). The highest categories in the current
Russian system were distinguished on the basis of a
conjugated system of genetic soil horizons that make
up the soil profile. The definitions of soil horizons
were based on the integrity of substantive soil properties dictated by soil-forming processes.
Cline (1949) summarized the basic principles of
soil classification that were the foundation of global
soil classifications such as Soil Taxonomy (1999) and
the WRB (FAO, 1998). In these systems, soils are

30

J.G. Bockheim et al. / Geoderma 124 (2005) 2336

classified on the basis of diagnostic surface horizons


(epipedons), subsurface horizons, and other characteristics. In Soil Taxonomy, soils are classified at six
levels in a hierarchical system, from highest to lowest:
orders, suborders, great groups, subgroups, families,
and series (Soil Survey Staff, 1999). Soil orders are
differentiated on the basis of soil processes (i.e., the
threads of genesis principle of Cline and Johnson,
1963) as indicated by the presence or absence of
diagnostic horizon and materials; in 8 of the 12 orders,
suborders are distinguished primarily from soil climate; and great groups are subdivided on the basis of
several criteria. Subgroups are subdivided according
to the central concept of the great group vs. intergradations to other taxa or extra-gradations to not soil.
Families are separated into particle-size, mineralogy,
and soil temperature classes. Fundamental concepts of
Soil Taxonomy were presented in a special symposium
of the Soil Science Society of America Proceedings in
1963, and Volume 96 of the journal Soil Science was
dedicated to the philosophical underpinnings and
applications of Soil Taxonomy.
The lowest category in Soil Taxonomy (Soil Survey
Staff, 1999) is the series, which is based on the kind
and arrangement of horizons. Soil series are divided
into phases on the basis of surface stoniness, slope
steepness, amount of previous erosion or other attributes that are not diagnostic in Soil Taxonomy, but
which are important to land use.
The soil series concept has changed with time and
summarizes the evolution of pedology from an observational, geology based discipline to a more quantitative science. The initial definition of soil series
(Whitney, 1906) was:
. . . a given set of soil classes. . . so evidently related
through source of material, method of formation,
topographic position, and coloration that the
different types constitute merely a gradation in the
texture of an otherwise uniform material. Soils of
different classes that are thus related constitute a
series. A complete soil series consists of material
similar in many other characteristics, but grading
in texture from stones and gravel. . . through the
sands and loams to a heavy clay. . .
Thus, a soil type was a textural class within the
series. This concept persisted (Marbut, 1910; Kruse-

kopf, 1921) until Marbut (1921) suggested that the


soil profile should be classified on the basis of
morphologic attributes of the horizons. Simonson
(1968) cited the Sassafras soil as an example of the
effect of the taxonomic change on distributions of soil
series. The original Sassafras soil (fine-loamy, siliceous, semiactive, mesic Typic Hapludults) became
more than 50 separate soil series by 1968.
It should be noted that the concepts of Soil Taxonomy were not universally accepted when they were
introduced and was debated in the literature. Webster
(1968) reviewed and discussed the most consistently
expressed objections and proposed a statistically
based system of soil classification.
Some key concepts pertaining to soil classification
include the following:
1. A zonal (normal) soil has well-developed characteristics reflecting the influence of climate and
vegetation; an intrazonal soil is more or less well
developed but has characteristics reflecting local
factors such as relief, parent material, or age; an
azonal soil is one that lacks well-developed
characteristics.
2. Evolutionary or process-based systems distinguish
soils at the higher levels on the basis of the
collection of genetic soil horizons comprising the
soil profile.
3. In Soil Taxonomy (Soil Survey Staff, 1999) and the
WRB for Soil Resources (FAO, 1998), soils are
classified the basis of diagnostic surface horizons
(epipedons), subsurface horizons, and other diagnostic characteristics.
4. In Soil Taxonomy (Soil Survey Staff, 1999), an
example of a global soil classification system, soils
are classified at six levels, from broadest to
narrowest: orders, suborders, great groups, subgroups, families, and series.
5. The fundamental soil entity mapped in most soil
surveys is the soil series.
1.6. Soil geography and mapping
The factors accounting for the distribution of soils
on the landscape have been of interest in pedology for
more than a century. Dokuchaev (1899a) distinguished
two levels of soil-geographical regularities. The first
level deals with the variation of soils across continental

J.G. Bockheim et al. / Geoderma 124 (2005) 2336

expanses, and the second level explains spatial changes


of soils at the landscape scale. Dokuchaev (1899a)
identified the factors controlling global soil distribution
and developed the doctrine of soil zones, including the
laws of horizontal and vertical bio-climatic zonation of
soils. These concepts were explained in the first schematic soil map of the Northern Hemisphere presented
at the World Exhibition in Paris in 1900 (see Gerasimov and Glazovskaya, 1960).
Prasolov (1922) and Gerasimov (1933) divided
Dokuchaevs zones into provinces and facies, which
are controlled mainly by climate and geomorphology.
Bioclimatic zonation and provinciality of soil cover
were expressed on a continental scale and were used
by Glinka (1927), Gerasimov (1933), and Prasolov
(1937) to prepare the first world soil maps.
It was apparent, however, that while the regional
controls of climate and vegetation were important at
continental scales, local soil variability within
smaller areas could often be attributed to changes in
relief and topography and to local parent materials
(Glinka, 1927; Joffe, 1936; Marbut, 1935; Ramann,
1928). An important component of this concept was
the definition of basic soil-geographic units, or soil
individual. The soil series is now recognized as a
group of polypedons; it is deemed the most homogeneous category in Soil Taxonomy (Soil Survey Division Staff, 1993). Some other definitions included the
three-dimensional soil body (Hole, 1953), the
soil-area unit (Muir, 1962), the artificial soil
body (van Wambeke, 1966), and pedomorphic
forms (Dan and Yaalon, 1968).
Fridland (1972) introduced the concept of the elementary soil areals (ESA), defining them as soil
bodies without internal pedogeographic boundaries.
The soil cover pattern as defined by Fridland (1972)
pertains to the detailed arrangement of ESAs, CSAs
(combinational soil areas), and associated bodies of
non-soil. In 1972, he wrote: A certain pattern of soil
cover is characterized by a set of soil areals which
repeats across space, resulting in a consistent character
and structure of soil cover and geochemical and geophysical linkages between soils within the sets (p. 12).
Each soil cover pattern has the same history of development. His approach was based on the earlier concept
of soil combinations (Sibirtsev, 1900; Neustruev,
1930). Fridland (1972) suggested indices to describe
and classify ESAs and soil cover patterns. They are

31

taxonomic soil units of definite size and composition,


spatial characteristics, and environmental condition.
According to Hole and Campbell (1985), there are
two contrasting views of soil geography: (1) the
traditional view employed for soil surveying and
mapping and (2) Fridlands (1972) soil cover patterns
approach (i.e., the other soil geography). The
geography of this soil geography denotes those
properties that pertain to soil bodies, specifically as
areal entities possessing characteristic sizes, shapes,
volumes, slopes, and internal variability (p. 97). He
denoted as a point of departure: Just as an elementary soil body is a natural, relatively distinct cluster of
pedons, so a combination soil body is a distinct unit in
the soil cover (p. 4).
The geomorphic approach to the distribution of soil
bodies on landscapes recognizes that geomorphic
processes are important agents of distributing parent
materials (Ruhe, 1956) and that the sorting and
arrangement of particles during transportation and
deposition creates landforms and deposits with unique
and distinguishing features (Gerrard, 1981; Selby,
1982; Daniels and Hammer, 1992; Paton et al.,
1995). Fundamental to this linkage of soils and landforms is that they are co-evolving (Hall, 1983). The
synergism between soils and geomorphology is important for predicting the areal distribution of soils on
the landscape and the flow of water through the
watershed (Daniels and Hammer, 1992). This perspective links sediment processes with soil formation
in ways that complement and complete Simonsons
(1959) process model and the soil cartographers
soil landform paradigm (Hudson, 1992). This approach also provides a conceptual linkage with the
soil to its broader landscape in ways not formally
discussed or recognized in Soil Taxonomy (Soil Survey Staff, 1960, 1975, 1999).
In the USA, broader soil map units include soil
associations and complexes, which were used as
groups of soils occurring together in a characteristic
and unique patterns on the landscape (Kellogg, 1933;
Simonson et al., 1952). Soil associations were mostly
formed by the different soil series of significant
genetic and morphologic diversity. Whereas the components of an association can be mapped separately at
a scale of about 1:24,000 and are distributed in
predictable patterns in the landscape, the components
of a complex cannot be differentiated at this scale and

32

J.G. Bockheim et al. / Geoderma 124 (2005) 2336

may be distributed randomly. Hole (1978) identified


soilscapes as distinct subdivisions of the landscape
with unique sizes, shapes, and arrangements of geographic units.
Rozanov (1977) differentiated hierarchical levels
of the worlds soil cover according to: mega-, macro-,
meso-, micro-, and nanostructures. Megastructure
refers to the genetic-morphostructural regions, e.g.,
soils of mountains, plains, lowlands, etc., that are
produced by the geological peculiarities of the continents. Macrostructures are bioclimatic zones, facies,
and districts. Mesostructures are areas consisting of
soil combinations related to mesotopographical forms.
Microstructures are related to microrelief; and nanostructures are the elementary soil areals.
Fridland (1984) did not support the idea of zonalprovincial structure of soil cover as the upper levels of
an overall hierarchical system of soil-geographical
organization. He proposed six levels of soil cover
patterns: elementary soil area, elementary soil cover
pattern (microstructure), mesostructure, soil region,
soil district, and soil country. He also recognized
several levels in the zonal-provincial (bio-climatic)
structure of soil cover: province, facies, subzone,
zone, and region (Fridland, 1984).
Key assumptions relating to soil geography and
mapping include the following:
1. The continental-scale distribution of soils is
controlled by the effect of climate and its effect
on organisms and soils; at the landscape level, the
distribution of soils is influenced by topography,
i.e., soil catenas.
2. Soils are linked to their resident landforms and
their landscapes, by geomorphic processes with
which they co-evolve.
3. Basic soil geographic units include three-dimension soil bodies (Hole, 1953); soil-area units
(Muir, 1962); artificial soil bodies (van Wambeke, 1966), pedomorphic forms (Dan and
Yaalon, 1968; and elementary soil areals (Fridland, 1972).

2. Conclusions
As a subdiscipline of soil science, pedology consists of an accepted body of laws and theories that

cover a range of related ideas and concepts pertaining


to the definition of the soil, soil horizons and profiles,
soil-forming factors and processes, soil classification,
and soil geography and mapping.
In this review, we are able to identify several key
milestones in pedology, including the following:

Pre-1880

Concept of soil as a medium for plant growth and


as a weathered rock layer.
1880 1900
Appearance of fundamental pedology concepts:
soil as a natural body; soil horizons/profiles;
soil-forming factors; early ideas of soil geography.
1900 1940
Global acceptance of concepts of soil as a natural
body and soil-forming factors; development of
first regional soil classification systems; soil
surveys initiated; identification of key
soil-forming processes.
1940 1960
Factors of soil formation and genesis of soils
clarified; development of global soil taxonomic
systems; intensified soil mapping.
1960 1990
Refinement of global soil taxonomic systems;
identification of pedon concept; development of
early soil models and soil cover pattern concept;
recognition of co-evolution of soils and
landforms.
1990 present Increased understanding of soil processes;
refinement of global soil models; further
refinement of global soil taxonomic systems;
development of statistical and computer-based
soil information systems.

These presented concepts and milestones have


proven to be useful in our careers and are offered
here to generate discussion in the pedological
community.

Acknowledgements
We appreciate the constructive reviews of R.J.
Huggett and L.R. Follmer.

References
Ableiter, J.K., 1949. Soil classification in the United States. Soil
Sci. 67, 183 191.
Arnold, R.W., 1965. Multiple working hypotheses in soil science.
Soil Sci. Soc. Am. Proc. 29, 717 724.
Arnold, R.W., 1983. Concepts of soils and pedology. In: Wilding,
L.P., Smeck, N.E., Hall, G.F. (Eds.), Pedogenesis and Soil

J.G. Bockheim et al. / Geoderma 124 (2005) 2336


Taxonomy: I. Concepts and Interactions. Elsevier, Amsterdam,
pp. 1 21.
Baldwin, M., Kellogg, C.E., Thorp, J., 1938. Soil classification.
Soils and Men. U.S. Dep. Agric. Yearbook. U.S. Govt. Print.,
Washington, DC, pp. 979 1001.
Birkeland, P.W., 1999. Soils and Geomorphology, 3rd edn. Oxford
Univ. Press, NY.
Bockheim, J.G., 1980. Solution and use of chronofunctions in
studying soil development. Geoderma 23, 71 85.
Bockheim, J.G., Gennadiyev, A.N., 2000. The role of soil-forming
processes in the definition of taxa in Soil Taxonomy and the
World Soil Reference Base. Geoderma 95, 53 72.
Bonneau, M., 1982. The concept of pedoclimate. In: Bonneau, M.,
Souchier, B. (Eds.), Constituents and Properties of Soil.
Academic Press, London, pp. 372 376.
Brevik, E.C., 1999. George Nelson Coffey, early American pedologist. Soil Sci. Soc. Am. J. 63, 1485 1493.
Buol, S.W., Hole, F.D., McCracken, R.J., Southard, R.J., 1997. Soil
Genesis and Classification, 4th edn. Iowa State Univ. Press,
Ames.
Bushnell, T.M., 1942. Some aspects of the soil catena concept. Soil
Sci. Soc. Am. Proc. 7, 466 476.
Butler, B.E., 1959. Periodic phenomena in landscapes as a basis for
soil studies. Aust. CSIRO Soil Publ. 14, 1 10.
Byers, H.G., Kellogg, C.E., Anderson, M.S., Thorp, J., 1938. Formation of soil. Soils and Men. U.S. Dep. Agric. Yearbook. U.S.
Govt. Print., Washington, DC, pp. 948 978.
Chadwick, O.W., Brimhall, G.H., Hendricks, D.M., 1990. From a
black to a gray boxa mass balance interpretation of pedogenesis. Geomorphology 3, 369 390.
Chesworth, W., 1973. The residual system of chemical
weathering: a model for the chemical breakdown of silicate rocks at the surface of the earth. J. Soil Sci. 24,
69 81.
Cline, M.G., 1949. Basic principles of soil classification. Soil Sci.
67, 87 91.
Cline, M.G., 1961. The changing model of soil. Soil Sci. Soc. Am.
Proc. 25, 2 446.
Cline, A.J., Johnson, D.D., 1963. Threads of genesis in the Seventh
Approximation. Soil Sci. Soc. Am. Proc. 27, 220 222.
Coffey, G.N., 1909. Physical principles of soil classification. Proc.
Am. Soc. Agron. 1, 175 185.
Coffey, G.N., 1912. A Study of the Soils of the United States.
USDA Bur. of Soils Bull., vol. 85. U.S. Govt. Print. Office,
Washington, DC.
Crompton, E., 1960. The significance of the weathering/leaching
ratio in the differentiation of major soil groups, with particular reference to some very strongly leached brown earths
of the hills of Britain. Trans. 7th Internat. Congress of Soil
Sci. vol. 4. Internat. Congress of Soil Sci., Madison, WI,
pp. 406 412.
Crowther, E.M., 1953. The skeptical soil chemist. J. Soil Sci. 40,
107 122.
Dan, J., Yaalon, D.H., 1968. Pedomorphic forms and pedomorphic
surfaces. Trans. 9th Internat. Congress of Soil Sci. vol. IV.
Internat. Congress of Soil Sci., Adelaide, South Australia,
pp. 577 584.

33

Daniels, R.B., Hammer, R.D., 1992. Soil Geomorphology.


Wiley, NY.
Davis, W.M., 1909. Geographical Essays. Harvard Univ. Press,
Harvard, MA.
Dmitriev, E.A., 1996. Soils and soil-like bodies. Eurasian Soil Sci.
29, 275 282.
Dobrovolskii, G.V., Nikitin, E.D., Karpachevskii, L.O., 2001. New
approaches to the concept of soil place in the biosphere. Eurasian Soil Sci. 34 (Suppl. 1), S1 S5.
Dobrovolsky, G.V., Trofimov, S.Y., 1996. Soil Systematics and
Classification (History and Up-to-Date Status). Moscow Univ.
Press, Russia.
Dokuchaev, V.V., 1879a. Mapping the Russian Soils (in Russian).
Imperial Univ. of St. Petersburg, St. Petersburg, Russia.
Dokuchaev, V.V., 1879b. Chernozeme (terre noire) de la Russie
DEurope. Societe Imperiale Libre Economique. Imprimeric
Trenke and Fusnot, St. Petersburg.
Dokuchaev, V.V., 1883. The Russian Chernozem Report to the Free
Economic Society (in Russian). Imperial Univ. of St. Petersburg,
St. Petersburg, Russia.
Dokuchaev, V.V., 1893. The Russian steppes/study of the soil in
Russia, its past and present. Dept. of Agric. Ministry of Crown
Domains for the Worlds Colombian Exposition at Chicago, St.
Petersburg.
Dokuchaev, V.V., 1899a. A Contribution to the Theory of Natural
Zones: Horizontal and Vertical Soil Zones (in Russian). Mayors
Office Press, St. Petersburg, Russia. 62 pp.
Dokuchaev, V.V., 1899b. The Place and Role of Contemporary
Pedology in Science and Life (in Russian). Mayors Office
Press, St. Petersburg, Russia. 28 pp.
Dokuchaev, V.V., 1949. Natural Historical Classification of Russian Soils (in Russian). From Selected Works. vol. 3. Publishing
House for Agric. Literature, Moscow, Russia. 19 pp.
Dokuchaev, V.V., Sibirtsev, N.M., 1893. Short Scientific Review of
Professor Dokuchaevs and His Pupils Collection of Soils Exposed in Chicago in the Year 1893. E. Evdokimova, St. Petersburg, Russia.
Duchaufour, P., 1968. LEvolution des Sols. Masson et Cie, Paris,
France. 93 pp.
FAO, 1998. World Reference Base for Soil Resources. World Soil
Resources Rep. vol. 84. Food and Agric. Organization, Rome,
Italy. 161 pp.
Fridland, V.M., 1972. Pattern of the Soil Cover. Transl. from the
Russian. Israel Program for Scientific Translations, Jerusalem.
Fridland, V.M., 1984. Patterns of the Soil Cover of the World (in
Russian). Mysl, Moscow.
Gennadiyev, A.N., 1990. Soils and Time: Models of Development.
Moscow Univ. Press, Moscow.
Gennadiyev, A., Gerasimova, M., Arnold, R., 1995. Evolving
approaches to soil classification in Russia and the United States
of America: their divergence and convergence. Soil Surv. Horiz.
36, 104 111.
Gennadiyev, A.N., Olson, K.R., Chernyanskii, S.S., 1996. Soil
science in the United States and the doctrine of V.V. Dokuchaev.
Eurasian Soil Sci. 29, 133 138.
Gerasimov, I.P., 1933. On soil-climatic facies of the USSR. Proc.
Dokuchaev Soil Inst. 8 (5), 29 37.

34

J.G. Bockheim et al. / Geoderma 124 (2005) 2336

Gerasimov, I.P., 1975. Elementary pedogenic processes as the basis


for genetic diagnostics of soils. Sov. Soil Sci. 5, 3 9.
Gerasimov, I.P., Glazovskaya, M.A., 1960. Fundaments of Soil
Science and Geography. Geografgiz, Moscow. In Russian.
Gerrard, A.J., 1981. Soils and Landforms: An Integration of
Geomorphology and Pedology. George Allen and Unwin,
Boston, MA.
Glazovskaya, M.A., 1964. Geochemical Foundations of the Typology and Methodology of Investigating Natural Landscapes (in
Russian). Moscow Univ. Press, Moscow, Russia. 237 pp.
Glazovskaya, M.A., 1983. Soils of the World, vol. 1. Amerind
Publishing, New Delhi.
Glinka, K.D., 1914. Die Typen der Bodenbildung, irhe Klassification und Geographische Verbreitung. Gebruder Borntraeger,
Berlin.
Glinka, K.D., 1927. Dokuchaevs ideas in the development of pedology and cognate sciences. Russian Pedological Investigations. USSR Acad. Sci., vol. 1. Publishing Office of the
Academy of Science U.S.S.R., Leningrad.
Glinka, K.D., 1931. Treatise on soil science. In: Gourevitch, A.
(Ed.), Transl. from Russian, 4th edn. Israel Program for Scientific Translation, Jerusalem. 674 pp.
Goryachkin, S.V., Tonkonogov, V.D., Gerasimova, M.I., Lebedeva, I.I., Shishov, L.L., Targulian, V.O., 2003. Changing
concepts of soil and soil classification in Russian. In:
Eswaran, H., Rice, T., Ahrens, R., Stewart, B.A. (Eds.), Soil
Classification: A Global Desk Reference. CRC Press, Boca
Raton, FL, pp. 187 199.
Hack, J.T., Goodlett, J.C., 1960. Geomorphology and forest ecology of a mountain region in the central appalachians. Geol.
Surv. Prof. Pap., vol. 347. U.S. Gov. Print. Office, Washington,
DC. 65 pp.
Hall, G.F., 1983. Pedology and geomorphology. In: Wilding, L.P.,
Smeck, N.E., Hall, G.F. (Eds.), Pedogenesis and Soil Taxonomy: I. Concepts and Interactions. Elsevier, NY, pp. 117 140.
Haseman, J.F., Marshall, C.E., 1942. The quantitative evaluation of
soil formation and development by heavy mineral studies: a
Grundy silt loam profile. Soil Sci. Soc. Am. Proc. 7, 448 453.
Hilgard, E.W., 1882. Report on the relations of soil to climate. U.S.
Dep. Agric. Weather Bull. 3, 1 59.
Hilgard, E.W., 1906. Soils: Their Formation, Properties, Composition, and Relations to Climate and Plant Growth in the Humid
and Arid Regions. Macmillan, NY.
Hole, F.D., 1953. Suggested terminology for describing soils
as three-dimensional bodies. Soil Sci. Soc. Am. Proc. 17,
131 135.
Hole, F.D., 1961. A classification of pedoturbations and some other
processes and factors in soil formation in relation to isotropism
and anisotropism. Soil Sci. 91, 375 377.
Hole, F.D., 1978. An approach to landscape analysis with special
emphasis on soils. Geoderma 21, 1 23.
Hole, F.D, Campbell, J.B., 1985. Soil Landscape Analysis. Rowman and Allanheld, Totowa, NY. 196 pp.
Hudson, B.D., 1992. The soil survey as paradigm-based science.
Soil Sci. Soc. Am. J. 56, 836 841.
Huggett, R.J., 1975. Soil landscape systems: a model of soil genesis. Geoderma 13, 1 22.

Huggett, R.J., 1997. Environmental Change: The Evolving


Ecosphere. Routledge, London.
Jenny, H., 1941. Factors of Soil Formation. McGraw-Hill, NY.
Jenny, H., 1961a. Derivation of state factor equations of soils and
ecosystems. Soil Sci. Soc. Am. Proc. 25, 385 388.
Jenny, H., 1961b. E.W. Hilgard and the Birth of Modern Soil
Science. Collana Della Revista Agrochimica, vol. 33. CDRA,
Pisa, Italy.
Joel, A.H., 1926. Changing viewpoints and methods in soil classification. Sci. Agric. 6, 225 232.
Joffe, J.S., 1936. Pedology. Rutgers Univ. Press, New Brunswick,
NJ.
Johnson, W.M., 1963. The pedon and polypedon. Soil Sci. Soc.
Am. Proc. 27, 212 215.
Johnson, D.L., Watson-Stegner, D., Johnson, D.N., Schaetzl, R.J.,
1987. Proisotropic and proanisotropic processes of pedoturbation. Soil Sci. 143, 278 292.
Kasimov, N.S., Perelman, A.I., 1992. The geochemistry of soils.
Eurasian Soil Sci. 24, 59 76.
Kellogg, C.E., 1933. A method for the classification of rural lands
for assessment in North Dakota. J. Land Public Util. Econ. 9,
10 14.
Kline, J.R., 1973. Mathematical simulation of soil plant relationships and soil genesis. Soil Sci. 115, 240 249.
Kossovich, P.S., 1911. Basics of Soil Science (in Russian). St.
Petersburg, Russia. 168 pp.
Kovda, V.A., 1990. The role and functions of the soil cover in the
earths biosphere. Problems of Soil Science. Nauka, Moscow,
pp. 5 10.
Kovda, V.A., Lobova, E.V., Rozanov, B.G., 1967. Problems of
world soil classification. Sov. Soil Sci. 7, 3 13.
Krupenikov, I.A., 1992. History of Soil Science from Its Inception
to the Present. Acad. Sci. of the USSR. Transl. from the Russian. Amerind Publishing, New Delhi.
Krusekopf, H.H., 1921. The Missouri soil survey. Univ. Mo. Coll.
Agric. Exp. Stat. Circ. 104. Columbia, MO. 20 pp.
Lobeck, A.K., 1939. Geomorphology: An Introduction to the Study
of Landscapes. McGraw-Hill, NY.
Lyon, T.L., Buckman, H.O., 1922. The Nature and Properties of
Soils: A Textbook of Edaphology. Macmillan, NY.
Marbut, C.F., 1910. Soils of the Ozark region: a preliminary report
on the general character of the soils and the agriculture of the
Missouri Ozarks. Univ. Mo. Agric. Exp. Stat. Res. Bull. 3.
Columbia, MO. 273 pp.
Marbut, C.F., 1921. The contribution of soil surveys to soil science.
Soc. Prom. Agric. Sci. 41.
Marbut, C.F., 1927. In: Glinka, K.D. (Ed.), The Great Soil Groups
of the World and their Development. Edwards Bros., Ann Arbor, MI. Transl.
Marbut, C.F., 1928. A scheme for soil classificationProc. 1st Int.
Cong. Soil Sci. Comm., vol. 5. Internat. Congress of Soil Sci.,
Washington, DC, pp. 1 31.
Marbut, C.F., 1935. Soils of the United States. U.S. Dep. Agric.,
Atlas Am. Agric. (Pt. 3).
Milne, G., 1935. Some suggested units of classification and
mapping particularly for East African soils. Soil Res. 4,
183 198.

J.G. Bockheim et al. / Geoderma 124 (2005) 2336


Morrison, R.B., 1967. Principles of Quaternary stratigraphy. In:
Morrison, R.B., Wright, H.E. (Eds.), Quaternary Soils. Proc.
Internat. Assoc. Quat. Res. 7th Cong., vol. 9. Desert Research
Institute, Univ. of Nevada, Reno, pp. 1 69.
Muhs, D.R., 1984. Instrinsic thresholds in soil systems. Phys.
Geogr. 5, 99 110.
Muir, J.W., 1962. The general principles of classification with
reference to soils. J. Soil Sci. 13, 22 30.
Neustruev, S.S., 1930. An essay of soil process classification. Elements of Soil Geography (in Russian). Selkhozgiz, MoscowLeningrad.
Nikiforoff, C.C., 1959. Reappraisal of the soil. Science 129,
186 196.
Nikitin, E.D., 2001. Soil as a bio-abiotic polyfunctional system.
Eurasian Soil Sci. 34, S6 S12.
North American Commission on Stratigraphic Nomenclature, E.D.,
1983. North American Stratigraphic Code. Am. Assoc.
Petroleum Geol. Bull. 67, 841 875.
Phillips, J.D., 1998. On the relations between complex systems and
the factorial model of soil formation (with discussion). Geoderma 86, 1 21.
Phillips, J.D., 2000. Signatures of divergence and self-organization
in soils and weathering profiles. J. Geol. 108, 91 102.
Polynov, B.B., 1923. Soils and Their Formation (in Russian). Mysl
Publishing House, Petrograd, Russia. 172 pp.
Polynov, B.B., 1953. Doctrine of Landscapes (in Russian). Geogr.
Proc., vol. 33. USSR Acad. Sci. Press, Moscow.
Prasolov, L.I., 1922. Soil Regions of European Russia (in Russian).
Gosizdat, Petrograd, Russia. 156 pp.
Prasolov, L.I., 1937. On nomenclature and fundamentals of genetic
soil classification. Soil Sci. 8 (in Russian).
Ramann, E., 1928. The Evolution and Classification of Soils
(Translated by C.L. Whittles). W. Heffer & Sons, London.
Richardson, J.L., Edmonds, W.J., 1987. Linear regression estimations of Jennys relative effectiveness of state factors equation.
Soil Sci. 144, 203 208.
Rode, A.A., 1947. The Soil-forming Process and Soil Evolution.
(Transl. by J.S. Joffe). Israel Program for Scientific Transl.,
Jerusalem. Office of Tech. Serv., U.S. Dep. Commerce, Washington, DC.
Rode, A.A., 1971. System of Research Methods in Soil Science (in
Russian). Nauka, Novosibirsk.
Rozanov, B.G., 1977. Soil Cover of the Earth (in Russian). Moscow
Univ. Press, Moscow, Russia. 248 pp.
Rozanov, B.G., 1983. Soil Morphology (in Russian). Moscow
State Univ., Moscow, Russia. 320 pp.
Ruhe, R.V., 1956. Geomorphic surfaces and the nature of soils. Soil
Sci. 82, 441 455.
Runge, E.C.A., 1973. Soil development sequences and energy
models. Soil Sci. 115, 183 193.
Selby, M.J., 1982. Hillslope Materials and Processes. Oxford Univ.
Press, Oxford.
Shaler, N.S., 1891. The origin and nature of soils. In: Powell, J.W.
(Ed.), 12th Annual Report of the U.S. Geol. Surv. 1890 1891.
U.S. Govt. Print. Office, Washington, DC, pp. 219 345.
Shaw, C.F., 1927. Report of committee on soil terminology. Am.
Soil Surv. Assoc. Bull. 8, 66 98.

35

Shaw, C.F., 1932. A soil formation formula. In: Prassolov, L.F.,


Vilensky, D.G. (Eds.), Proceedings and Papers of the Second
International Congress of Soil Science. Commission V, vol. 5,
pp. 7 14.
Sibirtsev, N.M., 1900. . In: Skorokhodva, Y.N. (Ed.), Pochvovedenie. Y.N. Skorokhodov, St. Petersburg, Russia. 212 pp.
Simonson, R.W., 1959. Outline of a generalized theory of soil
genesis. Soil Sci. Soc. Am. Proc. 23, 152 156.
Simonson, R.W., 1968. Concept of soil. Adv. Agron. 20,
1 47.
Simonson, R.W., Gardner, D.R., 1960. Concepts and function of
the pedon. 7th Int. Cong. Soil Sci., Trans., vol. 4. Internat.
Congress of Soil Sci., Madison, WI, pp. 127 131.
Simonson, R.W., Riecken, F.F., Smith, G.D., 1952. Understanding
Iowas Soils. Iowa State Univ. Press, Ames.
Soil Survey Division Staff, 1993. Soil Survey Manual. U.S.
Dept. Agric. Handbook, vol. 18. U.S. Govt. Print. Office,
Washington, DC.
Soil Survey Staff, 1960. Soil Classification, A Comprehensive
System, 7th Approximation. Soil Conserv. Serv., U.S. Dept.
Agric. U.S. Govt. Printing Office, Washington, DC. 265 pp.
Soil Survey Staff, 1975. Soil Taxonomy, A Basic System of Soil
Classification for Making and Interpreting Soil Surveys. Soil
Conserv. Serv., U.S. Dept. Agric. Handbook, vol. 436. U.S.
Govt. Printing Office, Washington, DC. 754 pp.
Soil Survey Staff, 1999. Soil Taxonomy: A Basic System of
Soil Classification for Making and Interpreting Soil Surveys,
2nd edn. Agric. Handbook, vol. 436. U.S. Govt. Print. Office,
Washington, DC. 869 pp.
Sokolov, I.A., 1996. Paradigm of pedology from Dokuchaev to the
present day. Eurasian Soil Sci. 29, 222 231.
Sommer, M., Schlichting, E., 1997. Archetypes of catenas in
respect to matter; a concept for structuring and grouping
catenas. Geoderma 76, 1 33.
Stevens, P.R., Walker, T.W., 1970. The chronosequence concept
and soil formation. Quart. Rev. Biol. 45, 333 350.
Tandarich, J.P., Darmody, R.G., Follmer, L.R., 1988. The development of pedological thought: some people involved. Phys.
Geog. 9, 162 164.
Tandarich, J.P., Darmody, R.G., Follmer, L.R., 1994. The pedoweathering profile: a paradigm for whole regolith pedology
from the glaciated midcontinental United States of America.
In: Cremeens, D.R. (Ed.), Whole Regolith Pedology. Soil Sci.
Soc. Am., Spec. Pub., vol. 34, pp. 97 117.
Tandarich, J.P., Darmody, R.G., Follmer, L.R., Johnson, D.L., 2002.
Historical development of soil and weathering profile concepts
from Europe to the United States of America. Soil Sci. Soc. Am.
J. 66, 335 346.
Targulian, V.O., 2001. Pedogenesis and the lithosphere. Eurasian
Soil Sci. 34, S21 S27.
Targulian, V.O., Sokolov, I.A., 1976. Structural and functional
approaches to soil: soil memory and soil moment. Mathematical Modeling in Ecology (in Russian). Nauka, Moscow,
pp. 17 34.
Targulian, V.O., Sokolova, T.A., 1996. Soil as a biotic/abiotic
natural system: a reactor, memory and regulator of biospheric
interactions. Eurasian Soil Sci. 29, 30 41.

36

J.G. Bockheim et al. / Geoderma 124 (2005) 2336

Vernadskii, V.I., 1926. Biosphere. Chem-Tech. Publ. House,


Leningrad.
Vreeken, W.J., 1975. Principal kinds of chronosequences and
their significance in soil history. J. Soil Sci. 26, 378 394.
Webster, R., 1968. Fundamental objections to the 7th Approximation. J. Soil Sci. 19, 354 366.
White, R.E., 1997. Soil science: raising the profile. Aust. J. Soil
Res. 35, 961 977.
Whitney, M., 1906. Soil Survey Field Book. U.S. Dept. Agriculture,
Bureau of Soils. 319 pp.
Yaalon, D.H., 1971. Soil-forming processes in time and space. In:

Yaalon, D.H. (Ed.), PaleopedologyOrigin, Nature and


Dating of Paleosols. Int. Soc. Soil Sci. and Israel Univ. Press,
Jerusalem, pp. 29 39.
Yaalon, D.H., 1975. Conceptual models in pedogenesis: can
soil-forming functions be solved? Geoderma 14, 189 205.
Yaalon, D.H., 1983. Climate, time and soil development. In:
Wilding, L.P., Smeck, N.E., Hall, G.F. (Eds.), Pedogenesis
and Soil Taxonomy: I. Concepts and Interactions. Elsevier,
NY, pp. 233 251.
Yaalon, D.H., 1997. History of soil science in context: international
perspective. Adv. Geo. Ecol. 29, 1 13.

You might also like