Sunteți pe pagina 1din 17

Available online at www.sciencedirect.

com

Geoderma 144 (2008) 140 – 157


www.elsevier.com/locate/geoderma

Review
Quantitative models for pedogenesis — A review
Budiman Minasny a,⁎, Alex. B. McBratney a , Sébastien Salvador-Blanes b
a
Faculty of Agriculture, Food and Natural Resources, The University of Sydney, NSW 2006, Australia
b
Laboratoire de Géologie des Environnements Aquatiques Continentaux, Faculté des Sciences et Techniques, Université François Rabelais de Tours,
Parc de Grandmont, 37200 Tours, France

Available online 8 February 2008

Abstract

Pedogenesis models should give us answers to: how does soil form, how does it evolve, where does it come from and how long does it take to
reach this state? This paper reviews quantitative models that describe pedogenesis ab initio and the processes that directly lead to development or
evolution of soil. We review factorial, energy, and mass-balance models. An early conceptual model comes from James Hutton. The factorial
model of Jenny provides the first definition of soil system and quantitative approach in pedology. Much works in pedology were devoted to
proposing variations of the factorial model in a qualitative way, such as the pathways, and energy models. The energy model of Volubuyev
attempts to calculate the energy of soil formation at a macro-scale and the entropy of soil at the profile scale. The energy model is used mainly as
description of the state of a soil. From conceptual, empirical models, a move towards mechanistic models of soil formation followed at a slower
pace. The landscape model from geomorphology has made lots of progress in quantifying and modelling soil weathering and distribution in the
landscape. These models usually consider physical weathering and treat the soil as a single layer of regolith. Meanwhile mechanistic pedology
models consider weathering in a profile scale at a nearly level landscape. Approach to combine these two approaches has recently been proposed.
We demonstrate that a rudimentary mass-balance model can simulate soil thickness and organic carbon content variation in the landscape. A soil
profile can be created by applying fundamental physical and chemical processes. The mass-balance model provides a valuable platform to model
soil and link pedology to other modern earth science disciplines. We discuss some criteria for pedogenesis models and possible integration of the
factorial, energy, and mass-balance models.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Soil genesis; Pedological modelling; Soil formation; Landscape evolution; Non-linear dynamic systems

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
2. The early days . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
3. The pedogenesis model triad . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
4. The energy model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
4.1. Energy of soil formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
4.2. Thermodynamics and entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
5. Soil mass-balance models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
5.1. Landscape evolution models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
5.2. Weathering functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
5.3. Chemical weathering and mass-balance accounting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
5.4. Soil transport models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

⁎ Corresponding author. Tel.: +61 2 9351 3214; fax: +61 2 9351 3706.
E-mail address: b.minasny@usyd.edu.au (B. Minasny).

0016-7061/$ - see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.geoderma.2007.12.013
B. Minasny et al. / Geoderma 144 (2008) 140–157 141

5.5. Biological processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149


5.6. A soil landscape model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
5.7. Profile scale model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
5.8. A soil profile model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
6. Discussions and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

1. Introduction plants; and a soil is nothing but the materials collected from
the destruction of the solid land. Therefore, the surface of
Soil is a complex system composed of a set of interconnected this land, inhabited by man, and covered with plants and
physical, chemical and biological factors that function as a animals, is made by nature to decay, in dissolving from that
whole. To gain deeper understanding of this complexity, soil hard and, compact state in which it is found below the soil;
scientists have experimented with the application of modern and this soil is necessarily washed away, by the continual
mathematical and statistical models to its quantification. There circulation of the water, running from the summits of the
has been a growing movement in pedology to apply empirical mountains towards the general receptacle of that fluid. The
quantitative techniques to predict soil properties from landscape heights of our land are thus levelled with the shores; our
attributes at a specific site. However there is still little effort in fertile plains are formed from the ruins of the mountains;
mechanistic modelling directed towards improving the science and those travelling materials are still pursued by the
of soil formation. The idea of mechanistic modelling was moving water, and propelled along the inclined surface of
formalised by Jenny (1941). Kline (1973) proposed developing the earth.”
mathematical models of pedogenesis which attempt to synthe-
size soil data into a more comprehensive quantitative theory of If only the process of erosion happens continually in a
soil formation. The importance of quantitative modelling of soil landscape, then it will result in a barren “smooth” landscape and
formation is echoed again by Hoosbeek and Bryant (1992), soil that sustains life will eventually diminish.
however only a little research has been devoted to modelling
pedogenesis. Most mechanistic modelling comes from geomor- “If the vegetable soil is thus constantly removed from the
phology rather than pedology. surface of the land, and if its place is thus to be supplied
This paper will review and present various models for from the dissolution of the solid earth, as here represented,
pedogenesis. Reviews of pedogenesis models can be found in we may perceive an end to this beautiful machine; an end,
Amundson (2004), Hoosbeek et al. (1997), Schaetzl and arising from no error in its constitution as a world, but from
Anderson (2005), Olson (2006), and Yaalon (1975). Here we that destructibility of its land which is so necessary in the
focus on quantitative models, although initially conceptual system of the globe, in the economy of life and vegetation.”
models are essential to put soil into the context of a model.
Moreover we restrict our review to the ones that model Hutton therefore demanded that soil should never run out,
pedogenesis ab initio and the processes that directly lead to and new soil is formed through regenerative forces of uplifting
development or evolution of soil. and volcanism. This realisation led him to discover the
enormous magnitudes of time. Two hundred years later it was
2. The early days still being pursued by Johnson (1985) who wrote a model for
soil thickness evolution:
Perhaps the earliest scientist to put up a model for soil
formation is James Hutton (1726–1796). James Hutton is Soil Thickness ¼ f ðDeepening þ Upbuilding þ RemovalsÞ:
regarded as the father of modern geology, and known for his
theory on earth formation and the discovery of earth's deep time. 3. The pedogenesis model triad
In 1785 he proposed the theory of earth to account for the
magnitude of geological time of earth formation. Hutton grew Schaetzl and Anderson (2005) presented a summary of
his theory based on what Stephen Jay Gould (1987) called the pedogenetic models in terms of a triangle. The three vertices
“paradox of the soil”. Hutton, as a gentleman farmer, realised are: factors, processes, and pathways, where all models can fit
that soil is the substance of life, and must be rich and fertile to within the triangle. The state factor model is well-known in soil
fulfil life on earth. He realised that formation involves weath- science which defines a set of independent variables that define
ering of rocks, and erosion of soil materials. Hutton in Volume 1 or control the state or properties of the soil at any given time.
of his book “Theory of the Earth” (Hutton, 1788) described: The state factors are based on empirical field observations, and
originated from Vasilli Dokuchaev:
“A solid body of land could not have answered the purpose
of a habitable world; for, a soil is necessary to the growth of j ¼ f ð K; O; CÞ ð1Þ
142 B. Minasny et al. / Geoderma 144 (2008) 140–157

where Π is soil, K is climate, O is organisms, and Γ is ground model soil mechanistically based on physical laws. Huggett
(parent materials). Shaw (1930) devised a formula for soil (1975) called this the systems approach.
formation: The pathways model (Johnson and Watson-Stegner, 1987;
Phillips, 1993) viewed soil as evolving along some genetic
s ¼ mðc þ vÞt þd ð2Þ
pathways that reflect variable internal and external factors.
which states that soil (s) is formed from parent material (m) These pathways ebb and flow not in a single direction towards
by the work of climatic factors (c) and vegetation (v) through an “organized” soil, but the processes can be progressive and
a period of time (t), and modification by erosion and regressive. Dokuchaev (as quoted in Nikiforoff, 1959) said
deposition (d). Shaw who also worked on the physics of soil “soil, like any other plant or animal organism, eternally lives
water defined the factors as a formula, however it is not for and changes, now progressively and then regressively”. The
quantitative prediction, but as a list of the potent factors in soil model is written as:
formation.
Jenny (1941) rigorously defined the soil as a system and s ¼ f ð P; RÞ ð6Þ
formalised the state factors of soil formation. Jenny described where P is progressive pedogenesis, including processes,
soil s as a function of state factors: factors and conditions that promote differentiated profiles
leading to physical–chemical stability. This includes; horizoni-
s ¼ f ðcl; o; r; p; t; N Þ ð3Þ
zation, leaching, developmental upbuilding, and soil deepening.
This is also called the “clorpt” model, where cl = climate, o = R is regressive pedogenesis, including processes, factors, and
organisms, r = relief, p = parent material, t = time, and … conditions that promote simplified profiles, leading to physi-
represents unspecified components. “The factors are not cal–chemical instability, such as haploidization/rejuvenation
formers, or creators, or forces; they are variables (state factors) processes, retardant upbuilding (impedance produced by sur-
that define the state of a soil system” (Jenny, 1961). The state face-accreted materials), and surface removals (erosion).
factors are not only qualitative but also quantitative, and are Huggett (1995) stated that in terms of quantitative modelling,
independent of the system, and may vary in space and time it is unduly complicated to single out processes that are
(Amundson and Jenny, 1997). Jenny (1941) suggested that there regressive from those that are progressive, and the designation
are two principal methods to solve this equation, i.e. in a of a process to be progressive or regressive is based on human
theoretical manner, and empirically from field observation. And value judgement and can vary with the scale of a system.
the solution suggested at that time is empirical and to use it as a Huggett (1991, 1995) formulated the brash model, which is
quantitative model a single factor can be defined by keeping the based on a general dynamic system equation:
other factors constant. This led to the development of empirical
dx
models to describe pedogenesis, such as climofunctions, ¼ f ð xÞ þ z ð7Þ
biofunctions, topofunctions, lithofunctions, and chronofunc- dt
tions (Yaalon, 1975), and the seeking for field situations with where x is a vector of variables describing the system, and f() is
unusual characteristics, i.e. where one of the state factors varies a transformation matrix defining the interactions between the
and the others remain constant. Huggett (1975) called this the state variables, and z is a vector of driving variables.
functional, factorial approach. Variable x is represented for terrestrial spheres: biosphere b,
Jenny (1961) re-defined his state factor model with a general toposphere r, atmosphere a, pedosphere s, and hydrosphere h.
form considering an open ecosystem: Huggett (1995) argued that this approach re-formulates the
factorial model into mathematically solvable equations, and
l; s; v; a ¼ f ðL0 ; Px ; t Þ ð4Þ
models soil properties as a function of processes. Applications
where l is ecosystem properties, s is soil properties, v is of these system equations can be found in Phillips (1993), where
vegetation, and a is animal properties which are related to three it was shown in a numerical example that changes in the initial
main state factors: L0 the initial state of a system, Px the external condition and parameter values can trigger the creation of
flux potentials, and t the age of the system. In pedogenesis, L0 chaotic behaviour of soil development.
can be the mineral and organic fraction of soil materials, p in Huggett (1998) viewed the nature of pedogenetic models
Jenny's clorpt model, and the configuration of the system r. Px reflected the development of thermodynamics as a subject:
are external environmental properties, which includes climate classical thermodynamics (1800s), open system dynamics (mid
cl, biotic factor o, and others. twentieth Century), and non-linear dynamic system (1970s).
The processes model (Simonson, 1959) concentrates on the Classical thermodynamics deals with closed system and
processes that have formed soil. Soil is viewed as an open equilibrium conditions. Jenny's (1941) theory is rooted in
system with additions and removals of materials to and from the classical thermodynamics. Open systems theory, or dynamical
profile, and translocation, transformation within the profile. The systems theory, was advocated into soil science by Nikiforoff
model is written as: (1959) and Simonson (1959). This theory saw the soil as an
open system and interpreted soil genesis as a dynamic process
s ¼ f ðaddition; removals; translocation; transformationÞ: ð5Þ
that does not always need to produce older or ‘mature’ soil
Although Simonson (1959) presented this as a qualitative (Nikiforoff, 1959). Jenny (1961) re-defined his state equation
conceptual model, this can be viewed as the first attempt to considering soil in an open ecosystem. Phillips (1993, 1998)
B. Minasny et al. / Geoderma 144 (2008) 140–157 143

embraced the non-linear dynamic systems approach to model


changes in soil system states. Phillips (1998) argued that
Jenny's reformulation of the state factor model (Jenny, 1961)
implicitly recognised the possibilities of self-organization and
deterministic chaos by assigning a strong role to initial
conditions and historical contingency. It is still arguable
whether the chaotic behaviour of mathematical model is due
to the non-linear dynamic of soil or numerical instability
(Pachepsky, 1998).
Although Jenny (1941) intended the model as the first
approach for quantitative pedology, much argument and
development in pedological modelling is focussed on qualita-
tive, narrative models. On the other hand, quantitative use of
factorial models evolves around deriving empirical relation-
ships between the soil-forming factors and soil attributes Fig. 1. The pedogenesis model triad.
(Birkeland, 1999).

4. The energy model


literature, and the papers have rarely been cited in any non-
The most important factor in pedogenesis is the transfer of Russian journals according to the ISI database (accessed
matter and energy. Yet the energy of soil formation is rarely November 2007).
quantified. We believe that the soil formation model triad is The energy model of Volubuyev works on a macro-scale, and
better described by: factors, processes, and energy (Fig. 1). can be derived based on the first law of thermodynamics. The
The well-known and most frequently cited work on an first law of thermodynamics concerns the conservation of
energy model in pedology is by Runge (1973) who presented energy, for a closed system it is written as:
the theory as a conceptual model and as an attempt to blend
DU ¼ Q  W ð9Þ
Jenny's factorial model with Simonson's process model. Runge
(1973) argued that soil is too complex to be described by The change in internal energy of a system (ΔU) is equal to
factors, as an alternative he considered energy as the principal the amount of heat added to the system (Q) minus the work done
drive for soil formation. The main source is gravity, encoura- by the system (W). This can describe laboratory studies of rock
ging water flow on and in the soil. Another source of energy is weathering. For an open system the energy (heat) required to
from plants that utilise water through transpiration. The model build a soil is (Volobuyev, 1964):
is written as:
E ¼ w1 þ w2 þ b1 þ b2 þ e1 þ e2 þ g þ v ð10Þ
S ¼ f ðo; w; t Þ ð8Þ
where
where o is organic matter production, w is the amount of water
available for leaching, and t is time. Although there are several E is the energy involved in soil formation
papers (Smeck et al., 1983; Schaetzl and Schwenner, 2006) that w1 is the energy of physical rock weathering
embraced this concept, they are used mainly in a conceptual and w2 is the energy for chemical weathering
qualitative way. No information is presented on the amount of b1 is the energy accumulating in soil organic matter
energy required for pedogenesis. b2 is the energy for soil organic matter transformation
e1 is the energy for evaporation from soil surface
4.1. Energy of soil formation e2 is the energy for transpiration
g is the energy losses in leaching of salts and fine
An important piece of work that seems to have been materials
overlooked by soil scientists is the work of Volobuyev in the v is the energy expended by the process of heat exchange
1960s, and his book, “Ecology of Soils” (Volobuyev, 1964). He between the soil and atmosphere (usually negligible).
presented a novel idea on the calculation of the energy
expended in the formation of soil. Tedrow (1965) wrote that The most significant energy is evapotranspiration (e1 + e2),
Volobuyev (sometimes quoted as Volobuyev) does not go into which is connected to the soil climate moisture regime. The
much detail on soil morphology, nor does he present many data. energy for evapotranspiration can be calculated as:
Volobuyev does not engage in small talk (soil morphology),
instead he discussed the complexities of the soil system. The e ¼ ET  Lv ð11Þ
works by Volobuyev are mainly published in Russian, some of where the amount of ET is evapotranspiration expressed in m of
the English works are published in Soviet Soil Science (1974a, water, and Lv is the latent heat of vaporisation of water
1968, 1977, 1980, 1983, 1984, 1985) and one in Geoderma (2260 MJ m− 3). Volobuyev (1964) defined the climate zones as
(1974b). This work is generally unknown in the soil science a function of mean annual temperature (MAT), and precipitation
144 B. Minasny et al. / Geoderma 144 (2008) 140–157

over potential evapotranspiration (P / ETp). The ratio K = P / ETp Based on this figure Volobuyev generated the weathering
is called the moisture index representing the availability of energy (Fig. 2c).
moisture. The energy of evapotranspiration as a function of The approximate total energy for soil formation is calculated
MAT and P / ETp is given in Fig. 2a. by summing energy for evapotranspiration, biomass production,
The second energy contributor is the biological component and mineral weathering, as shown in Fig. 2d. Most part of soil-
(b1 + b2), an estimate can be based on biomass production, forming energy is for evapotranspiration (95–99.5% of total
biomass can be defined as the mass of organisms per unit area energy). The energy for biological processes is only 0.5–5% of
and is usually expressed in units of energy (J m− 2). The the total energy. The energy for mineral weathering only 0.01%
assumption is 1 g photosynthesis or net primary productivity of the total energy required for soil formation. Volobuyev
(NPP) requires 15.7 kJ of solar energy, this energy can vary (1964) showed the energy of soil formation is variable for
between vegetation types. The amount of energy is approxi- different sites on the earth. In humid tropical wet areas, the
mated by the net primary production of biomass that is available annual energy is 2000 MJ/m2, tundras 40–100 MJ/m2, and for
for consumption by organisms (Fig. 2b). forests and steppes 500 to 1000 MJ/m2. Volobuyev then
The energy for rocks and minerals weathering is difficult to represented the energy as a simple equation:
estimate (w1 + w2). Volobuyev takes an approximate considering
a simple weathering reaction of K-Feldspar: Q ¼ Ra ð12Þ
where R is the energy of solar radiation, and a is the available
KAlSi3 O8 þ 8H2 O→AlðOHÞ3 þ KOH þ 3SiðOHÞ4 energy sources. For the climatic zones, he simplified it to:
where the Gibbs free energy of this reaction is − 64 kJ/mol at Q ¼ R exp ð1=mK Þ ð13Þ
25 °C. The energy involved in this reaction is 8 kJ/mol of water,
or 440 J/g of water. The amount of water participating in the where m is a dimensionless factor reflecting biological activity,
weathering process varies according to climatic conditions. and K is the moisture index (K = P / ETp).

Fig. 2. The effect of mean annual temperature (MAT in celsius) and precipitation over potential evapotranspiration (P / ETp) on: (a) energy of evapotranspiration,
(b) energy of biomass production as a function, (c) energy of mineral weathering, and (d) total energy of soil formation.
B. Minasny et al. / Geoderma 144 (2008) 140–157 145

This simple calculation provides an approximation for Table 1


energy at region, continent, and global scale. Although some Energy for various soil properties, assuming a clay soil with bulk density 1.4 g/
cm3 and depth 50 cm, 2% organic matter, and CEC = 300 mmol(+)/kg
might not agree to lump various energies together, this is an
elegant way to represent the climatic factors. This approach for State Energy Energy Energy Energy
variable conversion
quantifying the energy of soil formation is revisited by J/g J/m2
Rasmussen et al. (2005) applying it for predicting clay and Organic matter Heat Calorimetric energy of 400 2.8 × 104
organic matter content in the USA. content OM = 20 kJ/g
Another work on the energy of soil formation that is virtually (Reiners and Reiners,
1970)
unknown came from an ecological study by Regan (1977) who
Cation exchange Energy of Average bond 3 × 105 6.6 × 107
quantified energies in the ecosystem to model their role in the capacity exchange energy = 380 kJ/mole
economic development and urban growth in Florida. Regan bonds
calculated the energy of soil formation by considering Jenny's Field capacity Potential Water potential = 1 × 104 7.0 × 10− 1
factors of soil formation. This work lists the energy of the soil- energy of 10 J/kg
soil water
forming factors. Regan considered the solar energy as the main
Wilting point Potential Water potential = 1.5 × 106 1.1 × 105
energy related to climate, while Volobuyev considered the energy of 1500 J/kg
fraction of solar energy used to drive soil formation. Table 1 soil water
shows an example of calculation of the energy of soil variables. Energy of Energy Energy needed 500 3.5
Using an energy transfer model (Odum, 1971), Regan (1977) was aggregate needed to to rupture
formation disperse soil aggregates = 500 J/g
able to create an energy model for soil formation. He simulated
aggregates (Field and Minasny,
the energy of soil formation in Florida considering mineral and 1999)
organic soils. The model simulated the development of a mineral
soil in Florida that reached steady-state about 300 years with a rate
of 5.6 kJ m− 2 year− 1. This model clearly underestimates the rate
and time of soil formation; however this study is an excellent Smeck et al. (1983) described the concept of soil profile
example of linking soil, ecosystems and human system. With development as an energy consuming process with flow of
proper parametisation, this type of model can be executed with entropy to the external environment in an open system. They
better results and give a link between the amount of energy for soil hypothesized the relative change in entropy of soil order based
formation, vegetation, and population support. on positive (progressive pedogenesis of Johnson and Watson-
Stegner) and negative (regressive) processes. Positive values of
4.2. Thermodynamics and entropy ΔS are assigned to processes that result in the disorder of soil,
such as physical mixing, and mineral weathering. Negative
Thermodynamics is a potent principle which can be used to change in entropy is assigned to processes that sort soil
account for the flow and quality of energy in a system. Several constituents, such as weathering of secondary minerals,
reviews with relevance to soil formation have been given eluviation–illuviation, accumulation of organic materials,
(Smeck et al., 1983; Addiscott, 1995). They are mainly concepts leaching. Based on these simple assumptions, they calculated
trying to relate the thermodynamics to soil genesis. Thermo- the net entropy for the USDA soil taxonomy soil orders.
dynamics can potentially be used to quantify the energy of soil Standard Gibbs free energies G, enthalpy H, and entropy S
formation. Thermodynamics evaluates the change in Gibbs free for minerals have been compiled e.g. Lindsay (1979) and Yatsu
energy ΔG, which can be thought of as the maximum amount of (1988). They are based on calorimetric determination of mineral
work obtainable from a reaction. The energy state of a system formation. Most of them used thermodynamic stability
can be written as: diagrams to predict the stabilities, compositions, and the
direction of mineralogical reactions. Jenny (1980) postulated
DG ¼ DH  T DS: ð14Þ
that quantification of ΔG, ΔH and ΔS of the entire ecosystem is
The heat of reaction ΔH reflects the making and breaking of still far in the future. Volobuyev and Ponomarev (1977) took a
chemical bonds, and TΔS is the heat that the system takes or bold step by upscaling free energies and entropies of minerals to
gives to the environment. As well as quantifying the amount of soil. The free energy of a soil is calculated from:
energy, entropy measures the quality of the energy. This quality
is a measure of disorder in the system, and varies in different X
k
DG ¼ A i ni ð15Þ
kinds of energy and systems. Entropy determines the direction i¼1
of flow of energies. Generally energy always flows in such a
way that entropy increases, which implies that the entropy is where μi = chemical potentials of mineral component i, and ni =
very low initially (Price, 2006). number of moles of mineral component i. This method assumes
Entropy has always been the “holy grail” in science, to be the additivity of the thermodynamic potentials. The calculation of
able to give a state of energy in a soil system. Although Price the free energy and entropy for a soil is based on the mineralogical
(2006) argues the term entropy is inessential, we don't need to composition of the soil. Volobuyev and Ponomarev (1977) and
seek for the notion of entropy to see that there's an imbalance in Volobuyev et al. (1980) calculated Gibbs free energy and entropy
the system. for several soil types and presented it as in Fig. 3.
146 B. Minasny et al. / Geoderma 144 (2008) 140–157

simulate the formation of soil, this approach has not been


pursued further possibly due to lack of interest and data.

5. Soil mass-balance models

The mass-balance models are based on a mechanistic


approach and are in line with the “processes” approach. It
generally attempts to formulate a continuity equation to account
for the change in soil properties over time. Modelling
pedogenesis as processes can be interpreted in various meanings
by different people. Two broad categories can be identified, the
first is modelling soil formation ab initio, starting from a
bedrock, dealing with processes that can produce a soil with its
properties. The second is modelling changes in soil properties
(or processes) in the soil. For the latter model, soil is already in
Fig. 3. Gibbs free energy (ΔG) and entropy (S) for different soils, rocks, and place and the model deals with physical, chemical and
minerals. Soils and rocks data are from Volobuyev and Ponomarev (1977) and biological processes that influence the development of soil.
Volobuyev et al. (1980). Minerals data are from Yatsu (1988). Kline (1973) recognised the complexity of the soil system
but proposed to translate the conceptual model into a flow
diagram that can be modelled mathematically. He proposed a
Volobuyev and Ponomarev (1977) identified two groups: model for simulating the origin of chemical stratification in soil
(1) decrease in Gibbs' free energy ΔG and increase in entropy horizons. Kirkby (1977) realised the complexity in modelling
S, (2) a more “reactive” soil, increase in ΔG and a decrease in pedogenesis, and strived a balance between simplicity and the
entropy. Soils with large quantities of residual minerals and need to representing the important processes. While many
oxides SiO2, Al2O3, and Fe2O3 have low Gibbs free energy and physical, chemical, and biological models are available to
high entropy. The energy has been “lost” in minerals weath- simulate individual soil processes at short (daily to annual) time
ering. This is followed (in the order of higher energy and lower scales, models which simulate soil weathering processes at
entropy) by phyllosilicate minerals, carbonates, and soluble geological time scales are still lacking. Hoosbeek and Bryant
salts. The increase in entropy and decrease in Gibbs free energy (1992) attempt to distinguish the multiplicity of soil genesis and
reflects minerals that have higher intensity of leaching and more process models based on their spatial scale.
resistance to weathering. There are two main schools in mechanistic modelling of
Volobuyev et al. (1980) also showed that ΔG is related to pedogenesis: the landscape evolution model and the soil profile
water infiltration capacity, the lower the values of ΔG the higher model. The landscape evolution model mainly comes from
the infiltration capacity. Samedov and Nadirov (1989) showed geomorphology, where soil is modelled as a single layer of
that the activity of earthworms in the soil decreases ΔG because regolith, where weathering is mostly physical with materials
of the change in chemical composition and enrichment of the transport in the landscape is the main process. The soil profile
organic component. model from pedology and geochemistry mostly view it
Volobuyev's soil entropy is different from the entropy of weathering and as vertical transport of materials within the
Smeck et al. (1983). Weathering of primary minerals according profile, assuming a level landscape with minimum runoff and
to Smeck et al. (1983) increases the entropy, but accompanying lateral transport. No successful model has been able to link the
it the internal profile processes of illuviation, accumulation of two of them.
organic matter, formation of secondary minerals, and leaching
decreases the entropy. Thus highly weathered soils such as 5.1. Landscape evolution models
Oxisols and Ultisols have negative entropy change (ΔS
negative) while Vertisols have a positive entropy change (ΔS In the fields of geology and geomorphology, mechanistic
positive). The difference is Smeck's model is based on a profile models describing slope development have been developed
and qualitative description of main processes. Meanwhile successfully (Ahnert, 1977; Willgoose et al., 1991). Models in
Volobuyev's calculation is based on the mineral composition geomorphology usually simulate the evolution of a landscape
of a soil layer. It will be useful to calculate Volobuyev's entropy by the processes of erosion and deposition. The simplest model
for each horizon of a soil profile. of modelling soil in a landscape is considering the change in
Much work on quantitative energy of soil formation has been elevation over time is a function of material transport, the
made by Volobuyev, however this work is virtually unknown in continuity equation is:
soil science literature. At present, the energy model is mainly
∂z
used to describe the state of a soil, and is used at most for ¼ jqs ð16Þ
empirical correlations between the amount of energy and soil ∂t
properties (Rasmussen et al., 2005). However an example by where z is elevation [L], t is time [T], andqs is material flux
Regan (1977) showed the possibility on the use of energy to [L3 T− 1 ], and ∇ is partial derivative vector ∂x∂ þ ∂y∂ . This model
B. Minasny et al. / Geoderma 144 (2008) 140–157 147

forms the bases of landscape evolution models (Ahnert, 1967). reduction of weathering rate with thickening of soil is related to
Carson and Kirkby (1972) and later Dietrich et al. (1995) the exponential decrease of temperature amplitude with
introduced soil in the continuity equation: increasing depth below the soil surface (Ahnert, 1967, 1977),
and also the exponential decrease in average water penetration
∂h qr ∂e
¼  jqs ð17Þ (for freely-drained soils). Heimsath et al. (1997) called this the
∂t qs ∂t soil production function, which refers to the conversion of
where h is soil thickness [L], e is the boundary between soil bedrock to soil, or the lowering of bedrock–soil boundary. Note
and bedrock (e = z − s) [L] (see Fig. 4). ρs is the density of soil that soil production is not strictly equal to weathering rate, soil
[M L− 3] and ρr is the density of rock [M L− 3] (Fig. 4). This production refers to the conversion of bedrock into soil as
equation considers weathering of bedrock into soil materials, distinguished by lacking relict bedrock structure. Most bedrock
and transport of the materials. The first application in soil is weathered in place before it gets disrupted and incorporated
science is probably by Huggett (1975) who considers soil into the soil materials. Dietrich et al. (1995) argued that the
landscape as a homomorphic model. Dietrich et al. (1995) and primary mechanism in many places for this production is
Heimsath et al. (1997) applied this model in a watershed in the disturbance by biogenic activity. Heimsath et al. (1997, 1999,
Tennessee valley, California. Minasny and McBratney (1999) 2000, 2001) measured the production rate using cosmogenic
employed this in a rudimentary one-dimensional soil-land- nuclides and parameterized the function.
scape model, and expanded it to two dimensions in Minasny An alternative soil production function is the so called
and McBratney (2001) with chemical weathering. Rosenbloom humped model (Humphreys and Wilkinson, 2007), which
et al. (2001) further expanded this model (Eq. (17)) separating implies the presence of chemical weathering as well as physical
the soil variables (h) into four components: sand, silt, clay and weathering. Initially hypothesized by Grove Karl Gilbert in
organic matter. 1877 (Humphreys and Wilkinson, 2007), the soil production
reaches a maximum under an optimal soil depth. Weathering of
5.2. Weathering functions bedrock depends on water, thus decomposition of bedrock is
fastest under an intermediate thickness of soil and slower under
Studies on weathering functions in pedology begin with exposed bedrock or under thick mantled soil. This is because
investigation of chronofunctions, the rate of soil physical and chemical weathering requires the presence of water (Ahnert,
chemical weathering over time. Colman (1981) reviewed 1967). Under thin soil or exposed bedrock water tends to run
various studies of rock weathering, and suggested that the rate off, reducing the chance of the decomposition of bedrock.
of weathering decreases with time and weathering rate can be Various forms of the function are given by Cox (1980). Furbish
approximated as a logarithmic function of time. and Fagherazzi (2001) presented a function:
Ahnert (1967) and Heimsath et al. (1997) suggested that the  
rate of soil production (∂e / ∂t) can be represented as an ∂e 1 þ jh
¼ P0 expðkhÞ: ð19Þ
exponential decline with soil thickness: ∂t 1 þ j=k
∂e κ [L− 1] is a coefficient that controls the intercept and rate of
¼ P0 exp ðkhÞ ð18Þ
∂t increasing weathering rate with increasing soil thickness, with
where P0 [L T− 1] is the potential (or maximum) weathering rate k b κ and when κ = 0, it reduces to the physical weathering
of bedrock and k [T− 1] is an empirical constant. This equation model. The critical soil thickness where maximum weathering
mainly describes the physical weathering of bedrock, where the rate occurs at:

1 1
hc ¼  : ð20Þ
k j

As an alternative, Minasny and McBratney (2006) presented


a double exponential model:

∂e
¼ ðP0 ½ exp ðk1 hÞ  exp ðk2 hÞ þ Pa Þ: ð21Þ
∂t

where k1 represents the rate of mechanical breakdown of the


rock materials, and k2 is the rate of chemical weathering, and Pa
is the weathering rate at steady-state condition [m year− 1] with
condition k1 b k2. The critical thickness where weathering is
maximum is given by:

Fig. 4. A mass-balance model for soil formation in a landscape. Symbol e refers


ln ðk2 =k1 Þ
to the interface between bedrock and soil, h is the thickness of soil layer, z is hc ¼ : ð22Þ
elevation, and qs is soil transport. k2  k1
148 B. Minasny et al. / Geoderma 144 (2008) 140–157

Furbish and Fagherazzi (2001) performed stability analysis and Drosdoff (1943) presented a model to account for gains and
to confirm the qualitative analysis of Carson and Kirkby (1972) losses of substances in a volume of soil. Brimhall and Dietrich
that perturbations of soil thickness in the initial condition will be (1987) introduced a mass-balance model that quantitatively
dampened when soil thickness is greater than hc, and instability links chemical composition to bulk density, mineral density,
will be amplified when soil thickness is less than hc. Decreasing volumetric properties, porosity, and amount of volume loss or
weathering rate with increasing soil thickness reinforces the strain. It uses elements in rocks and soils as geochemical tracers
stabilising effect of the diffusion process. Meanwhile increasing indicative of specific physico-chemical transport processes
weathering rate with soil thickness has a destabilising effect, during weathering. The model is:
most probably due to the thin soil and the need to maintain its
qr Ci;r
thickness to produce more soil. Dietrich et al. (1995) modelled e¼ 1 ð25Þ
both the humped and exponential soil production functions in a qs Ci;s
soil development model for a watershed in the Tennessee where ε is strain, or change in thickness over original
Valley, California. Their results showed that using the humped thickness [L L − 1 ], Ci is mass concentration of the immobile
model resulted in the emergence of bedrock in the convex crests element [M M − 1 ], ρ is bulk density [M L − 3 ], subscripts r and
and pointed out that there can be no soil thickness between bare s represent rock and soil material. This provides a means of
bedrock and the peak in the hump (the unstable region). using chemical data to determine the amount of deformation
Meanwhile the exponential function resulted in a soil thickness in residual weathering profiles. Strain is positive for dilation
pattern that is more consistent with field observations. and negative for collapse of the profile.
Rather than using a single weathering function for a The work by Brimhall and Dietrich (1987) is the cornerstone of
landscape, Saco et al. (2006) varied the exponential weathering the last 20 years of soil formation work by many people, and
function spatially according to soil moisture conditions: quantitatively opened up soil formation from a black to grey box
∂e   (Chadwick et al., 1990). However most studies on soil chemical
ð x; y; hÞ ¼ P0 1 þ a ½F ð x; yÞd expðbhÞ: ð23Þ weathering rates at profile scale are assuming residual weathering
∂t on a nearly level landscape, where soil erosion is minimal and
The weathering potential parameter k (Eq. (19)) is modified lateral water flux is at a minimum. On hillslopes, geomorphol-
to include the term F(x,y), a spatial variable of soil moisture ogists implicitly assumed that soil production and removal was
which can be a function of topographic wetness index, an entirely physical process. Chemical weathering at landscape
subsurface water depth, with a and d as empirical parameters. scale on sloping terrain was just recently measured and modelled
They found that varying the weathering function spatially can using soil production rate and geochemical composition of
result in well-defined soil spatial patterns. They also showed the soil (Riebe et al., 2003, Yoo et al., 2007). Riebe et al. (2003)
that varying the weathering potential as a function of subsurface considered the soil production rate at a site with steady-state
water can create a weathering function with a maximum value formation, erosion, and weathering. The soil production rate then
that is non-zero soil depth. equal to the sum of the rates of chemical weathering and erosion.
Based on the chemical mass-balance (Eq. (25)), the chemical
5.3. Chemical weathering and mass-balance accounting weathering can be calculated by considering immobile elements
such as Zr. The ratio between Zr concentration in the rock and soil
Chemical weathering rates at watershed to global scale are indicates the soil chemical weathering rate W [L T− 1]:
usually studied by analysing the concentrations of solute in the  
W ¼ D 1  ½Zrr =½Zrs ð26Þ
streams. There are many factors controlling weathering rates,
however many studies at a particular scale have narrowed down where D is the soil production rate [L T− 1] obtained from in-situ
to a simple relationship between weathering rates and a major produced cosmogenic nuclide concentrations measured in soil
controlling factor. At a global scale, a simple linear relationship samples. Assuming steady-state conditions in soil production and
between runoff, precipitation and chemical weathering rates (as transport and chemical weathering, Yoo et al. (2007) combined
measured by chemical element fluxes) is observed using the the chemical mass-balance (Eqs. (25) and (26)) and the soil mass-
watershed dataset from the world (Wakatsuki and Rasyidin, balance model (Eq. (17)) to come up with a model for soil physical
1992; White and Blum, 1995). Meanwhile weathering fluxes w and chemical weathering rate as a function of hillslope position.
(mol m− 2 s− 1) and temperature follow the Arrhenius law: Research in the measurement of soil age and weathering rate
is lacking in pedology. Most studies and data come from
w ¼ A exp ðEa =RT Þ ð24Þ geomorphology and geochemistry. Various studies have explored
where A is potential weathering rate, Ea is the activation energy, the influence of climate and erosion on chemical weathering.
R is a gas constant, and T is temperature. Lasaga et al. (1994) West et al. (2005) suggested a model for silicate weathering rates
attempted to present a model to account for various controlling considering their limiting factors: erosion and reaction kinetics.
factors: temperature, pH, and solution composition. They At low erosion rates mineral supply limits weathering, and at high
proposed a model that incorporates ΔG, temperature and erosion rates there is abundant material but the kinetic and
adsorption of H+, OH− , and other ions. climatic factors limit weathering. From cosmogenic nuclides
Another chemical weathering model generally used in soil and geochemical measurement, Riebe et al. (2004) found that
and geochemistry is the geochemical mass-balance. Nikiforoff chemical weathering rates decrease rapidly with increasing
B. Minasny et al. / Geoderma 144 (2008) 140–157 149

altitude, they argued that the faster the erosion rate, the faster the It is originally developed for fluvial transport. This model
chemical weathering. Burke et al. (2007) measured soil simulates surface wash processes induced by overland flow
production and the chemical weathering rate in a granitic hillslope (Kirkby, 1985; Willgoose et al., 1991; Dietrich et al., 2003).
at Point Reyes, California. They showed that chemical weathering • Soil thickness-dependent flow (Braun et al., 2001; Heimsath
rates decrease with increasing soil thickness and account for 13 to et al., 2005)
51% of the soil production rate. They also found that spatial
variation in chemical weathering appears to have a good qH ¼ K H hp Sq ð29Þ
correlation with topography: high chemical weathering rates qH erosion flux [L2 T− 1],
correspond to gentle slopes, and the rate and intensity decrease KH transport coefficient [L2 T− 1],
with steeper slopes. Gabet et al. (2006) considered the effect of h soil thickness [L],
hydrology and developed a mathematical model describing that p and q empirical parameters.
when weathering is not limited by the supply of parent materials,
the weathering rates are enhanced with decreasing hydraulic Braun et al. (2001) suggested p = 1.67 and q = 0.5 analogous
conductivity and longer soil water residence time. to Manning's equation for open channel flow. They showed that
this model better captures spatial variations of soil thickness on
5.4. Soil transport models hillslopes. An increasing flux with increasing soil thickness is
hypothesized due to increased materials to be transported,
The transport of soil materials in the landscape can be porosity and bioturbation. Heimsath et al. (2005) showed that
modelled as the following processes: the thickness-dependent transport law is more broadly applic-
able than the linear diffusion model.
• The linear diffusion process, formulated as:
5.5. Biological processes
qD ¼ DS ð27Þ
Biological processes are an important part of the soil
where: formation processes. Activities of animals and plants can affect
the soil formation factors. These activities are termed
qD volume of material that flows across an area per unit bioturbation, which is generally defined as “the churning and
time by diffusive transport [L2 T− 1], stirring of sediment by organisms” (Gabet et al., 2003). It is
D diffusivity [L2 T− 1], applied in pedology and geomorphology for an array of
S slope gradient [L L− 1]. biomechanical processes of soil profile reorganization (Johnson
et al., 2005b). Bioturbation in soil is mainly the result of
Diffusive transport simulates slow mass movement or creep of earthworm/ant/termite activity, and results in the detachment,
soil materials moving downhill by the force of gravity. This transport, sorting and deposition of material, both within the
model which is also called linear transport is the oldest and has soil mantle and on its surface resulting in the homogenisation of
been the most widely used in landscape evolution models. the topsoil, and material transport between the subsoil and the
However it has limited applications, field evidence of Dietrich topsoil.
et al. (2003) indicates that outside of freeze thaw activity and Carson and Kirkby (1972, pp. 289–291) regarded bioturba-
clay rich soils, most of the soil ”creep“ is due to biogenic tion as secondary in importance to systematic moisture and frost
activity. From field measurements, Heimsath et al. (2005) cycling in producing soil creep. Meanwhile Johnson et al.
calculated sediment fluxes against the product of soil thickness (2005b) advocated bioturbation as the primary process of soil
and hillslope gradient and showed that soil transport is a non- formation. Johnson et al. (2005a) gave an animation of the
linear, thickness-dependent process. Their data suggest that the process to explain the evolution of soil thickness. However it is
widely used linear diffusion equation is only appropriate for only a conceptual model without any physical formulation.
shallow gradient, convex regions. Gabet et al. (2003) provided the first comprehensive review of
• The overland water erosion process, where the transport is quantitative models of bioturbation on soil formation and
proportional to local water discharge and slope. It is defined as: sediment transport. In this review the authors provided physical
models to quantify sediment flux for tree throw and root growth
qA ¼ KA Am Sn ð28Þ and decay. Heimsath et al. (2002) used natural quartz grains in a
mature soil to determine grain movements from the time elapsed
where: since each grain last visited the ground surface, measured by
single-grain optical dating. Using a Monte Carlo simulation of
qA erosion flux [L2 T− 1], particle transport, they suggested that soil creep involves
KA water discharge [L2 T− 1], independent movements of mineral grains throughout the soil
A upstream catchment area, body and that grains are reburied or eroded by overland flow
m and n empirical parameters, with 1.0 ≤ m ≤ 1.8 and 0.9 ≤ upon reaching the surface.
n ≤ 1.8,with the best single combination of values Yoo et al. (2005) modelled soil mass transport by soil-
being m = n = 1.4 (Prosser and Rustomji, 2000). burrowing animals in terms of the population density and energy
150 B. Minasny et al. / Geoderma 144 (2008) 140–157

expenditure of the animals. They also calculated the average


energy input by the animals for soil transport is 120 J m− 2 year− 1,
only a small fraction of the energy of soil formation.
The effect of biological processes may create a unique
topographic signature of life at the profile or horizon scale.
Dietrich and Perron (2006) emphasized the effect of biota on
landscape evolution, arguing that it even influences the height,
width and symmetry of mountains, but they concluded that
there appear to be no unique signature of life in the topography.
Istanbulluoglu and Bras (2005) suggested that vegetation can
affect landscape form by changing the dominant erosion process
from overland flow under abiotic conditions to land sliding in
well-vegetated states.
Biological processes occur mainly in soil, and linking soil as
a profile in the landscape is the key to be able to give us a
mechanistic understanding of the role of biotic processes in
landscape evolution.
Fig. 5. Biomass productivity as a function of soil thickness and organic carbon
content.
5.6. A soil landscape model

Minasny and McBratney (2006) presented a mass-balance


model for soil formation in the landscape. The model uses the kc = kc0 exp(−bkh). Transport of carbon is modelled as simple
mass-balance Eq. (18), considering a soil production function diffusion process
(Eq. (22)) and transport models (Eqs. (27)–(29)). In this review, qC ¼ r D k S ð34Þ
we have expanded the model incorporating soil organic carbon
evolution (Fig. 4). The following example attempts to integrate r is a reduction factor in soil transport due to improvement in
the landscape mass-balance model to simulate soil evolution in soil structure due to organic matter addition, Dk is a transport
the landscape. coefficient [L T− 1] and S is slope.
In addition to the soil mass-balance, we model the carbon We applied the model to a digital elevation model of the
component separately but allow interaction with the mineral Hunter Valley in New South Wales, Australia (Minasny and
component similar to the model by Yoo et al. (2006): McBratney, 2006). The area was assumed to have a uniform
parent material and initial thickness of 10 cm. Parameters of the
∂Cz
¼ Iz þ qm  kCz  qC : ð30Þ model were obtained from the literature. Fig. 6 shows the results
∂t of the model from a simulation of 40 000 years. The model can
The change in carbon concentration C [L L− 1] over depth z, simulate the distribution of soil thickness and organic carbon in
Cz, is a function of production Iz [L L− 1], within profile (vertical) a landscape. Within this landscape we also recognise three main
mixing qm, decomposition k and transport qC. Soil carbon soil groups according to the topographical position (Paton,
production is modelled as a function of depth (Yoo et al., 2006): 1978): residual, transportational, and depositional. Residual
soils (Fig. 7) mostly occur on gentle topography, the long
Iz ¼ Ci expðbzÞdz ð31Þ
residence time of soil material meaning weathering processes
−1
where Ci is the carbon production at soil surface [L T ], and b are dominant. Transportational soils (Fig. 8) take place in
[L− 1] is an empirical constant. The production is influenced by convex topography where lateral movement is dominant near
soil thickness, and productivity (Fig. 5). Productivity is defined as the surface and weathering processes are dominant near the
a sigmoid function of organic carbon, suggesting a feedback of bedrock surface. Depositional soils (Fig. 9) occur in concave
productivity with increasing soil carbon level. Vertical mixing is topography where lateral movement is dominant, and weath-
modelled as a diffusion process (Kirkby, 1977): ering processes are less important. Residual soils (Fig. 7) show
slow build-up of soil thickness. Carbon is slowly accumulating
∂2 C in the soil in the initial period (8000 years) and starts to increase
qm ¼ DC ð32Þ
∂z2 with increasing soil thickness. Transportational soils have very
where DC is a mixing constant [L2 T− 1]. This represents mixing little accumulation of soil materials, and carbon (Fig. 8). After
of organic materials by soil fauna. Decomposition is based on a 25 000 years with sufficient soil build-up, the carbon begins to
simple two compartment model (Hénin and Dupuis, 1945): accumulate. Depositional soils are characterised with rapid
build-up of soil thickness and carbon storage (Fig. 9). This
∂C
¼ hI  kc C ð33Þ example shows that the simulated patterns of soil formation
∂t follow the general soil–landscape relationship. The challenge is
where h is a humification factor, and kc is the rate constant of to build a realistic and computationally feasible soil profile
decomposition. The rate constant is dependent on soil depth model that can be incorporated in the landscape model.
B. Minasny et al. / Geoderma 144 (2008) 140–157 151

Fig. 6. Simulation of the distribution of (a) soil thickness and (b) organic carbon in a landscape after 40 000 years.

5.7. Profile scale model the amount of proportion of bedrock converted to soil. The
mass-balance model is given as:
Kirkby (1977) realised the importance of soil in landscape  
evolution, and presented the first mechanistic soil profile model ∂w ð1  P s Þ
¼ j J  qs ð36Þ
as a component of hillslope models. His model is based on the ∂t Ps
quantity called soil deficit:
where J is chemical sediment transport, and Ps is properties
Z h remaining in the soil at the surface. The model is developed in
w¼ ð1  PÞdz ð35Þ three parts: weathering profile, organic profile, and inorganic
z¼0 profile. The inorganic minerals are treated as mixtures of
elementary oxides. Soil organic materials include production,
where P is the proportion of bedrock or substance remaining decomposition, and vertical mixing. The soil CO2 produced by
in a soil profile with values between 0 to 1.0 (unweathered organic matter decomposition is simulated via diffusion process,
materials), w is the accumulated “soil deficit” which represents and soil pH is calculated with reference to the CO2 distribution.
152 B. Minasny et al. / Geoderma 144 (2008) 140–157

Fig. 7. The evolution of soil thickness, carbon storage in the profile, and surface carbon content for residual soils.

Kirkby (1985) improved his model considering nutrient regolith (Eq. (18)). The regolith layer is then subjected to
uptake, leaf-fall and decomposition of nutrients in organic physical and chemical weathering creating a particle-size
matter, mechanical mixing by soil fauna, leaching of solutes. distribution and chemical composition with time. The weath-
These processes are shown to be sufficient to produce shallow ering of the coarse fraction in the profile is considered as a
and deep weathering profiles, and the formation of an organic- physical weathering process only. The primary minerals
rich surface horizon and a subsurface horizon with base released are subsequently subjected to both physical and
depletion. No further work has been developed after this. chemical weathering. The physical weathering of the minerals
consists in breaking a given mineral particle into smaller
5.8. A soil profile model particles. The physical fractionation model is based on the
model of Legros and Pedro (1985). The chemical weathering is
Salvador-Blanes et al. (2007) devised a soil profile model modelled according to White et al. (1996):
which is envisaged to be incorporated into a landscape model.
The main processes are physical and chemical weathering, mw ¼ kr S ð37Þ
−1
and reorganization process due to bioturbation. The profile where mw is the mineral weathering rate [mol s ], kr the mineral
starts with the physical disintegration of bedrock into a layer of weathering rate constant [mol m− 2 s− 1], and S the mineral surface

Fig. 8. The evolution of soil thickness, carbon storage in the profile, and surface carbon content for transportational soils.
B. Minasny et al. / Geoderma 144 (2008) 140–157 153

Fig. 9. The evolution of soil thickness, carbon storage in the profile, and surface carbon content for depositional soils.

area [m2]. Primary minerals are subjected to weathering into bioturbation, is able to simulate the formation of a soil with a
secondary minerals according to known chemical weathering stone layer, and thus a first attempt to simulate simple soil
pathways. horizon formation.
The reorganization of soil materials within the profile was
modelled as bioturbation, where a given proportion of the fine 6. Discussions and conclusions
fraction (b2 mm) of soil from each layer in the subsoil is
transported to the topsoil layer due to biogenic activities. Thus, From field observations, soil scientists have come up with
as the soil forms, each layer is subjected to increase weathering conceptual models trying to capture the main factors and
that resulted in a loss of chemical elements, while bioturbation processes that are responsible for the genesis of soil and its
leads to a removal of soil material from subsoil to topsoil layers. horizons. From conceptual models, empirical relationships are
The model was applied to a study of an in-situ weathering constructed representing climofunctions, biofunctions, topo-
profile from meta-gabbro, which is mainly composed of functions, lithofunctions, and chronofunctions (Birkeland,
plagioclase, hornblende and quartz (Schroeder et al., 2000). 1999). These functions are useful to explain soil distribution
Results for the evolution of particle-size distribution for various at catena, landscape or higher spatial scales. However most
ages ranging from 10 000 to 80 000 years are shown in Fig. 10. research in pedological modelling focussed on developing the
This example of a simple in-situ, soil profile model with 3 major factorial narrative models, which are descriptive, based on field
pedogenic processes: physical and chemical weathering, and observations, and not for quantitative prediction (Fig. 11). A

Fig. 10. The evolution of particle-size distribution of a simulated soil profile forming in-situ with physical, chemical and biological processes.
154 B. Minasny et al. / Geoderma 144 (2008) 140–157

Fig. 11. A summary of the development of factorial models.

Fig. 12. A summary of the energy models.

Fig. 13. A summary of the mass-balance pedogenesis models.


B. Minasny et al. / Geoderma 144 (2008) 140–157 155

move towards mechanistic models of soil formation followed at formation of soil horizons or soil profiles. Not many models can
a slower pace. create these qualitative descriptions, and it is the challenge to be
Jenny (1980) distinguished two approaches in modelling able to model these processes. This is extremely challenging for
pedogenesis, those based on soil processes and state-factor most, so many will dismiss it as irrelevant, however this is the
sequences. The process approach is reductionist utilising outlook of pedology that allows us to test and better understand
physics, chemistry and biology to study individual soil various soil formation processes.
processes. The state factor approach is phenomenological, We have demonstrated that simple mechanistic models based
seeking the explanation of how soil varies in space and time. on mass-balance are able to simulate realistically the distribution
Various quantitative pedogenesis models have been con- of a layer of soil and organic matter in the landscape. A profile
structed by various authors. In geochemistry, these include model can simulate the development of simple soil horizons. The
modelling continental basalt weathering at geological time scale big challenge is to be able to link these two together. The factorial
(Dessert et al., 2003; Goddéris et al., 2007), mineral weathering model allows us to isolate and to identify potent factors in soil
at catchment scale (Goddéris et al., 2006), profile scale over formation, and to confirm them with reality based on empirical
years (Suarez and Goldberg, 1994). These models focussed observations (Fig. 11). The energy model accounts for the flow of
mainly on the weathering and formation of minerals and soil energy and indicates which direction the soil is going. As shown
solution chemistry, while soil itself is a medium of reaction. Soil in Regan (1977) it is possible to model the energy of soil
physical and chemical models include transformation with formation in an ecosystem (Odum, 1971) (Fig. 12). The mass-
physical and chemical processes over decades in a profile balance approach offers the best way to mechanistically model
(Hoosbeek and Bryant, 1994). These models attempt to soil processes (Fig. 13). All of these models are complementary,
simulate current or previous soil conditions to predict what is and the ultimate goal is the ability to link all these three
going to happen in the future. Soil is assumed to have been approaches, and to be able to model how soil horizons and profiles
formed. We might call this predictive pedology. This is a useful evolve and change in the landscape. The mechanistic platform
pursuit especially considering climate change and land-use allows a solid link between pedology and modern earth sciences.
pressures. Here time scales of decades and centuries are relevant
for humanity and its survival. The real challenge is to be able to Acknowledgements
model soil development ab initio. This is useful to give us a
better understanding of pedogenesis and help us answers This work is funded by the Australian Research Council
questions such as: How does soil form, how does it evolve, Discovery project “How do soils grow?” The authors thank
where does it come from and how long does it take to form and Dr. Bill Dietrich and an anonymous reviewer for their helpful
reach this state? There appear only a few models that are reviews.
designed for the purpose. Most of them come from geomor-
phology such as those by Kirkby, Dietrich, Heimsath and References
others, which have made substantive progress in modelling soil
evolution in the landscape based on mass-balance equations. Addiscott, T.M., 1995. Entropy and sustainability. European Journal of Soil
Science 46, 161–168.
We tried to define some criteria for pedogenetic models. The Ahnert, F., 1967. The role of the equilibrium concept in the interpretation of
first is physical basis; the model being based on physical laws, landforms of fluvial erosion and deposition. In: Macra, P. (Ed.), L'evolution
consistent with the conservation of mass and energy. Although des versants. L' Université de Liège, Liège, pp. 23–41.
some relationships between soil processes are still not Ahnert, F., 1977. Some comments on the quantitative formulation of
geomorphological process in a theoretical model. Earth Surface Processes
quantifiable, empirical relationships can be placed within the
2, 191–201.
framework. The second is it can simulate the present condition, Amundson, R., 2004. Soil formation. Chap. 5.01. In: Holland, H.D.,
given an initial condition (also called nowcasting); it can Turekian, K.K. (Eds.), Treatise on Geochemistry. Elsevier Press,
simulate realistically the soil thickness and soil properties Amsterdam, pp. 1–35.
variation in the landscape. Ultimately, it should simulate soil Amundson, R., Jenny, H., 1997. On a state factor model of ecosystems.
profile formation in a landscape, where each voxel represents a BioScience 47, 536–543.
Birkeland, P.W., 1999. Soils and Geomorphology, 3rd edn. Oxford University
horizon characteristic. And third, the model can be extended to Press, New York.
simulate what is the likely past condition given the present Bockheim, J.G., Gennadiyey, A.N., 2000. The role of soil-forming processes in
condition (also called backcasting), and for scenario modelling the definition of taxa in Soil Taxonomy and the World Soil Reference Base.
what will be the likely soil condition in the future (forecasting). Geoderma 95, 50–72.
Pedogenesis modelling may not be able to replicate present Bockheim, J.G., Gennadiyev, A.N., Hammer, R.D., Tandarich, J.P., 2005.
Historical development of key concepts in pedology. Geoderma 124, 23–36.
condition accurately, but it should be able to model the trend Braun, J., Heimsath, A.M., Chappell, J., 2001. Sediment transport mechanisms
and this is the essential part of the modelling exercise. on soil mantled landscapes. Geology 29, 683–686.
One of the criteria is to be able to create soil horizons. The soil Brimhall, G.H., Dietrich, W.E., 1987. Constitutive mass balance relations
horizon is the unique property that defines a soil from pedological between chemical composition, volume, density, porosity, and strain in
point of view. Soil horizons here do not refer to descriptive soil metasomatic hydrochemical systems: results on weathering and pedogen-
esis. Geochimica et Cosmochima Acta 51, 567–587.
morphology, but are genetic layers having characteristic chemical, Burke, B.C., Heimsath, A.M., White, A.F., 2007. Coupling chemical weathering
and physical properties (Bockheim et al., 2005). Bockheim and with soil production across soil-mantled landscapes. Earth Surface Processes
Gennadiyev (2000) identified 17 main processes that lead to the and Landforms 32, 853–873.
156 B. Minasny et al. / Geoderma 144 (2008) 140–157

Carson, M.A., Kirkby, M.J., 1972. Hillslope Form and Process. Cambridge Huggett, R.J., 1975. Soil landscape systems: a model of soil genesis. Geoderma
University Press, p. London. 13, 1–22.
Chadwick, O.A., Brimhall, G.H., Hendricks, D.M., 1990. From a black to a gray Huggett, R.J., 1991. Climate, Earth Processes and Earth History. Springer,
box — a mass balance interpretation of pedogenesis. Geomorphology 3, Heidelberg.
369–390. Huggett, R.J., 1995. Geoecology. An Evolutionary Approach. Routledge, London.
Colman, S.M., 1981. Rock-weathering rates as a function of time. Quaternary Huggett, R.J., 1998. In Discussion of: J.D. Phillips, On the relation between
Research 15, 250–264. complex systems and the factorial model of soil formation. Geoderma 86,
Cox, N.J., 1980. On the relationship between bedrock lowering and regolith 23–25.
thickness. Earth Surface Processes 5, 271–274. Humphreys, G.S., Wilkinson, M.T., 2007. The soil production function: a brief
Dessert, C., Dupre, B., Gaillardet, J., Francois, L.M., Allegre, C.J., 2003. Basalt history and its rediscovery. Geoderma 139, 73–78.
weathering laws and the impact of basalt weathering on the global carbon Hutton, J., 1788. Theory of the Earth or an investigation of the laws observable
cycle. Chemical Geology 202, 257–273. in the composition, dissolution, and restoration of land upon the globe.
Dietrich, W.E., Perron, T., 2006. The search for a topographic signature of life. Transactions of the Royal Society of Edinburgh I, 209–304 part II.
Nature 439, 411–418. Istanbulluoglu, E., Bras, R.L., 2005. Vegetation-modulated landscape evolution:
Dietrich, W.E., Reiss, R., Hsu, M., Montgomery, D.R., 1995. A process-based effects of vegetation on landscape processes, drainage density, and
model for colluvial soil depth and shallow landsliding using digital elevation topography. Journal of Geophysical Research 110, F02012.
data. Hydrological Processes 9, 383–400. Jenny, H., 1941. Factors of soil formation. A System of Quantitative Pedology.
Dietrich, W.E., Bellugi, D., Heimsath, A.M., Roering, J.J., Sklar, L., Stock, J.D., McGraw-Hill, New York.
2003. Geomorphic transport laws for predicting the form and evolution of Jenny, H., 1961. Derivation of state factor equations of soils and ecosystems.
landscapes. In: Wilcock, P., Iverson, R. (Eds.), Prediction in Geomorphol- Soil Science Society of America Proceedings 25, 385–388.
ogy. AGU Geophysical Monograph Series, vol. 135, pp. 103–132. Jenny, H., 1980. The soil resource. Origin and Behavior. Springer-Verlag, New
Field, D.J., Minasny, B., 1999. A description of aggregate liberation and York.
dispersion in A horizons of Australian Vertisols by ultrasonic agitation. Johnson, D.L., 1985. Soil thickness processes. In: Jongerius, P. (Ed.), Soils and
Geoderma 91, 11–26. Geomorphology. Catena Supplement, vol. 6. Catena Verlag, Braunschweig,
Furbish, D.J., Fagherazzi, S., 2001. Stability of creeping soil and implications pp. 29–40.
for hillslope evolution. Water Resources Research 37, 2607–2618. Johnson, D.L., Watson-Stegner, D., 1987. Evolution model of pedogenesis. Soil
Gabet, E.J., Reichman, O.J., Seabloom, E.W., 2003. The effects of bioturbation Science 143, 349–366.
on soil processes and sediment transport. Annual Review of Earth and Johnson, D.L., Domier, J.E.J., Johnson, D.N., 2005a. Animating the
Planetary Sciences 31, 249–273. biodynamics of soil thickness using process vector analysis: a dynamic
Gabet, E.J., Edelman, R., Langner, H., 2006. Hydrological controls on chemical denudation approach to soil formation. Geomorphology 67, 23–46.
weathering rates at the soil–bedrock interface. Geology 34, 1065–1068. Johnson, D.L., Domier, J.E.J., Johnson, D.N., 2005b. Reflections on the nature
Goddéris, Y., Francois, L.M., Probst, A., Schott, J., Moncoulon, D., Labat, D., of soil and its biomantle. Annals of the Association of American
2006. Modelling weathering processes at the catchment scale: the WITCH Geographers 95, 11–31.
numerical model. Geochimica et Cosmochimica Acta 70, 1128–1147. Kirkby, M.J., 1977. Soil development models as a component of slope models.
Goddéris, Y., Donnadieu, Y., Dessert, C., Dupre, B., Fluteau, F., Francois, L. Earth Surface Processes 2, 203–230.
M., Meert, J., Nedelec, A., Ramstein, G., 2007. Coupled modeling of Kirkby, M.J., 1985. A basis for soil profile modelling in a geomorphic context.
global carbon cycle and climate in the Neoproterozoic: links between Journal of Soil Science 36, 97–121.
Rodinia breakup and major glaciations. Comptes Rendus Geoscience 339, Kline, J.R., 1973. Mathematical simulation of soil–plant relationships and soil
212–222. genesis. Soil Science 115, 240–249.
Gould, S.J., 1987. Time's Arrow, Time's Cycle: Myth and Metaphor in the Lasaga, A.C., Soler, J.M., Ganor, J., Burch, T.E., Nagy, K.L., 1994. Chemical
Discovery of Geological Time. Harvard University Press, Cambridge. weathering rate laws and global geochemical cycles. Geochimica et
Heimsath, A.M., Dietrich, W.E., Nishiizumi, K., Finkel, R.C., 1997. The soil Cosmochimica Acta 58, 2361–2386.
production function and landscape equilibrium. Nature 388, 358–388. Legros, J.P., Pedro, G., 1985. The causes of particle size distribution in soil
Heimsath, A.M., Dietrich, W.E., Nishiizumi, K., Finkel, R.C., 1999. profiles derived from crystalline rocks, France. Geoderma 36, 15–25.
Cosmogenic nuclides, topography, and the spatial variation of soil depth. Lindsay, W.L., 1979. Chemical Equilibrium in Soils. Wiley, New York.
Geomorphology 27, 151–172. Minasny, B., McBratney, A.B., 1999. A rudimentary mechanistic model for soil
Heimsath, A.M., Chappell, J., Dietrich, W.E., Nishiizumi, K., Finkel, R.C., production and landscape development. Geoderma 90, 3–21.
2000. Soil production on a retreating escarpment in southeastern Australia. Minasny, B., McBratney, A.B., 2001. A rudimentary mechanistic model for soil
Geology 28, 787–790. formation and landscape development II; a two-dimensional model
Heimsath, A.M., Chappell, J., Dietrich, W.E., Nishiizumi, K., Finkel, R.C., incorporating chemical weathering. Geoderma 103, 161–179.
2001. Late Quaternary erosion in southeastern Australia: a field example Minasny, B., McBratney, A.B., 2006. Mechanistic soil-landscape modelling as
using cosmogenic nuclides. Quaternary International 83–85, 1169–1185. an approach to developing pedogenetic classifications. Geoderma 133,
Heimsath, A.M., Chappell, J.C., Spooner, N.A., Questiaux, D.G., 2002. 138–149.
Creeping soil. Geology 30, 111–114. Nikiforoff, C.C., 1959. Reappraisal of the soil. Science 129, 186–196.
Heimsath, A.M., Furbish, D.J., Dietrich, W.E., 2005. The illusion of diffusion: Nikiforoff, C.C., Drosdoff, M., 1943. Genesis of claypan soil. Soil Science 53,
Field evidence for depth dependent sediment transport. Geology 33, 949–952. 459–482.
Hénin, S., Dupuis, M., 1945. Essai de bilan de la matière organique des sols. Odum, H.T., 1971. Environment, Society and Power. John Wiley and Sons, New
Annales Agronomiques 15, 161–172. York.
Hoosbeek, M.R., Bryant, R.B., 1992. Towards the quantitative modeling of Olson, C.G., 2006. Geomorphological soil-landscape models. In: Grunwald, S.
pedogenesis — a review. Geoderma 55, 183–210. (Ed.), Environmental Soil-Landscape Modeling. Geographic Information
Hoosbeek, M.R., Bryant, R.B., 1994. Developing and adapting soil process Technologies and Pedometrics. Taylor and Francis, Boca Raton, pp. 105–124.
submodels for use in the pedodynamic Orthod model. In: Bryant, R.B., Pachepsky, Y.A., 1998. In Discussion of: J.D. Phillips, On the relation between
Arnold, R.W. (Eds.), Quantitative Modeling of Soil Forming Processes. . complex systems and the factorial model of soil formation. Geoderma 86,
SSSA Special Publication, vol. 39. Soil Science Society of America, Inc., 31–32.
Madison, WI, pp. 111–128. Paton, T.R., 1978. The Formation of Soil Material. George Allen & Unwin,
Hoosbeek, M.R., Amundson, R.G., Bryant, R.B., 1997. Pedological modeling. London.
In: Sumner, M.E. (Ed.), Handbook of Soil Science. CRC Press, Boca Raton, Phillips, J.D., 1993. Progressive and regressive pedogenesis and complex soil
FL, pp. E-77–E-116. evolution. Quaternary Research 40, 169–176.

S-ar putea să vă placă și