ArticlePDF Available

The Role of Plants in Controlling Rates and Products of Weathering: Importance of Biological Pumping

Authors:

Abstract and Figures

Key Words plant cycling, organometallic complexes, biological weathering, soil genesis, stability of soil minerals s Abstract The recycling of elements by plants and plant-induced biological ac-tivity cause the rates and products of weathering to be markedly different from what would result in abiotic processes. Plants directly control water dynamics, weathering, and the chemistry of weathering solutions, which is clearly exhibited in equatorial areas where old weathering mantles are greatly influenced by biological activity. Depending on the dynamics of plant-induced organometallic compounds, this weathering results in either clayey soils, which are in a dynamic equilibrium sustained by the forest's cycling of elements, or sandy soils. In most places (tropical as well temperate areas), the weathering mantle can be regarded as being in a dynamic equilibrium sustained by plants.
Content may be subject to copyright.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
Annu. Rev. Earth Planet. Sci. 2001. 29:135–63
Copyright c
2001 by Annual Reviews. All rights reserved
THE ROLE OF PLANTS IN CONTROLLING RATES
AND PRODUCTS OF WEATHERING:Importance
of Biological Pumping
Y Lucas
Laboratoire des Echanges Particulaires aux Interfaces, Universit´
e de Toulon et du Var,
BP 132, 83957 La Garde Cedex, France; e-mail: lucas@isitv.univ-tln.fr
Key Words plant cycling, organometallic complexes, biological weathering,
soil genesis, stability of soil minerals
Abstract The recycling of elements by plants and plant-induced biological ac-
tivity cause the rates and products of weathering to be markedly different from what
would result in abiotic processes. Plants directly control water dynamics, weathering,
andthe chemistryof weatheringsolutions, whichisclearly exhibitedin equatorialareas
where old weathering mantles are greatly influenced by biological activity. Depending
on the dynamics of plant-induced organometallic compounds, this weathering results
in either clayey soils, which are in a dynamic equilibrium sustained by the forest’s
cycling of elements, or sandy soils. In most places (tropical as well temperate areas),
the weathering mantle can be regarded as being in a dynamic equilibrium sustained by
plants.
INTRODUCTION
The characteristics of soils and weathering mantles reflect the integrated effects of
water, atmosphere, and organisms acting on rocks. Most rocks formed in environ-
mental conditions different from those that occur at the Earth’s surface. When ex-
posedatthesurfacetheyare generally subjected to a thermodynamic readjustment.
The result is the dissolution of primary minerals in the rocks and the formation
of newly generated secondary minerals that compose the overlying weathering
horizons. The weathering mantle deepens when the rate of downward advance of
the weathering front exceeds the rate of surface removal by agents of erosion. Be-
cause of many environmental problems, there is a great interest in understanding
the processes that control the interactions between minerals and aqueous solution
in the weathering mantle. Quantitative data are also necessary, for incorporation
into numerical models that couple transport of solutes with the chemical reactions,
thereby simulating the genesis and the dynamics of the weathering mantle.
0084-6597/01/0515-0135$14.00 135
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
136 LUCAS
In the upper decimeters of the weathering mantle, the role of living organisms is
obvious; for example, organic matter gives a darker color. Soil fauna, plant roots,
and organic matter directly affect soil structure and porosity. Sulfates, nitrates, and
organicacids released by bacterial activity control the acidity of soil solutions. The
role of plants in determining the rates of weathering at the mineral, the catchment,
or the continent scale was addressed in several studies (Likens et al 1977, Velbel
1985, Berner 1992, Bluth & Kump 1994, Cochran & Berner 1992, Drever 1994,
Moulton & Berner 1998). The role of plants in determining the mineral products of
weathering,that is, mostofthe soilminerals, is, however,notgenerally appreciated,
althoughLovering(1959) stressed this problem more than40yearsago. Byplants,
I consider both the direct effect of higher plants and the effect of the associated
living organisms (mainly bacteria, fungi, and microfauna). The purpose of this
paper is to examine how plants control the residual product of weathering, that is,
the soil profile.
The Products of Weathering from an Abiotic Point of View
On a global scale and particularly in the tropics, the spatial distribution of sec-
ondary minerals is frequently interpreted as a function of their stability in aqueous
solutions without consideration of biological processes. There is indeed a close
relationship between the availability of water, which depends on climate, and the
bulk result of weathering, that is, the average mineralogical composition of the
weathering mantle. This is obvious when considering aluminosilicate minerals
that are mainly precipitated in soils of the humid to the dry tropics. In warm and
humid climates the new products are mainly gibbsite [Al(OH)3] and kaolinite
[Si2O5Al2(OH)4], and almost all of the primary minerals are dissolved, including
quartz (SiO2). The soils that are formed under these conditions are called ferral-
sols, oxisols, or lateritic soils. With decreasing rainfall and increasing length of
the dry season, gibbsite ceases to be formed. With further drying, smectite [ap-
proximately Si4O10Al2(OH)2nH2O] rather than kaolinite becomes the dominant
clay, and quartz and many primary minerals are preserved. As a result, the Si/Al
ratio of the weathering cover progressively increases with increasing aridity of
the climate. This change with climate corresponds to the sequence of thermody-
namic stability of aluminosilicates, which is the order in which minerals reach
saturation in an Si-Al system (Figure 1), as pointed out by Pedro (1966) and
described by leaching models (Fritz & Tardy 1973, Fouillac & Michard 1977,
Tardy & Nahon 1985, Ambrosi 1990). Most of these models are based on the
assumption of local thermodynamic equilibrium between minerals and solution.
Some more recent models also consider the kinetics of the reactions (Mad´eetal
1994, Soler & Lasaga 1996). In simulations at 20Cor25
C, dilute rainwater
enters the topsoil and, as it percolates downward, reactions between solutions and
minerals lead to a progressive increase in the concentrations of Fe, Si, Al, and
other elements in the soil solution. Saturation is reached successively for goethite
(FeOOH), gibbsite, kaolinite, Fe-smectites, Fe-Al smectites, calcite (CaCO3),
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 137
Figure 1 Mineral stability in the Si-Al system at 25C. Considering a pH varying from
4.5 to 5.5 and all elements being free ions, the soil solutions from columns A through E of
Table 3 are plotted within the hatched rectangle.
and Mg-smectites. At lower temperatures the rates of reactions are lower, and the
equilibrium constants change, so that newly formed phases are different [e.g. illite,
approximately (Si(4x)Alx)O10Al2(OH)2Kx] and they are formed slowly or not
at all.
The results of such modeling are in agreement with the observed latitudinal
zonality of the weathering mantle, especially with respect to bulk mineralogical
composition in tropical soils. If water percolates into the weathering mantle at
a high rate, as in humid climates, solute concentrations in the soil water will be
low and the main newly generated minerals will be gibbsite and kaolinite. If water
percolates into the weathering mantle at a low rate, as in drier climates, the solute
concentrations increase, and the newly generated minerals are primarily smectite
and calcite.
The Limits of Abiotic Leaching Models
When leaching models are used without considering the role of plants and biolo-
gical activity, these models cannot describe some important features of the weath-
ering mantle. For example, the vertical sequence of horizons predicted by the
abiological models is as follows: (a) highly leached, aluminous, gibbsitic horizons
near the surface; (b) more siliceous kaolinitic horizons below; and (c) smec-
titic horizons at still greater depths. Actually, this structure is rarely obser-
ved, even in highly leached equatorial areas. There, kaolinite is the prevailing
secondary mineral in the topsoil horizons of most of the well-drained ferral-
sols. When more gibbsitic horizons are observed, they are situated beneath more
kaolinitic horizons, either as gibbsitic regoliths (Leneuf 1959, Delvigne 1965,
Sieffermann & Millot 1969, Novikoff 1974, Furian 1994), as bauxitic horizons
(Dennen & Norton 1977, Lucas et al 1989, Boulang´e & Carvalho 1997), or as
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
138 LUCAS
gibbsite accumulations in a kaolinitic matrix (Lucas et al 1986). Thus, observa-
tions of kaolinite in the uppermost soil horizons are inconsistent with the results
of thermodynamically based models in which the predominance of gibbsite is
predicted.
Another example of the inadequacy of the thermodynamically based mo-
dels is the frequent association of podzols with ferralsols in the humid tropics.
Podzols, also called spodosols, are characterized by thick, sandy horizons
composed of residual, resistant primary minerals, mainly quartz, and only scarce
secondary minerals. Podzols have been reported in South America (Klinge
1965, Sombroek 1966, Volkoff 1985), in Africa (Brammer 1973, Schwartz
1988), and in Asia (Brabant 1987, Siefferman 1990). In the Amazon Basin,
podzols cover >250,000 km2(Martinelli et al 1996). Despite this great ext-
ent in area, until recently, tropical podzols were commonly unmentioned on
soil maps of the world (Food and Agricultural Organization 1974). One reason
for this omission is that podzols, although mineralogically quite different
from them, are often closely associated with ferralsols, coexisting on the same
landscape and often developing from the same parent rock (Turenne 1977,
Lucas et al 1984, Chauvelet al 1987, Veillon 1990, Boulet et al 1997, Dubroeucq &
Volkoff 1998). This close association cannot be understood solely from
leaching models without considering the role of plants and associated biologic
activity.
In this paper, I consider more specifically the role of plants on (a) the mo-
bility at the profile or catchment scale of less mobile elements, mainly Si, Al,
and Fe, which play a major role on soil composition and properties because they
are the only constituents of many secondary minerals and the main constituent
of most of the others; and (b) the genesis of the profile, that is, the vertical
succession of horizons from the rock to the topsoil. After a brief review of the
main processes by which plants affect weathering and soil genesis, with a fo-
cus on the role of biological pumping (that is, the recycling of inorganic con-
stituents by plants) in the chemistry of soil solutions, a case study is discussed in
detail.
PROCESSES INVOLVING PLANTS IN WEATHERING
AND SOIL GENESIS
Plants absorb soil water through their roots and take in atmospheric CO2through
their leaves. The water is transferred by capillary and osmotic processes to the
leaves where it is evaporated and returns to the atmosphere. The uptake water
contains dissolved elements, which permit the plant to satisfy its nutrient needs.
In the plant cells, water, CO2, and nutrients are combined to produce the organic
matter of the plant tissues. As the plant or its parts die, the organic matter returns
to the topsoil, where it is consumed by other living organisms—fauna, fungi,
and bacteria. Most of the organic matter is thus oxidized. The CO2returns to the
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 139
atmosphere, and most other elements are released into the soil solution. A part
of the organic matter is incompletely oxidized and reacts to form the soil’s humic
substances.
The associated organisms are also involved in the processes of weathering and
soil genesis. Fungi and bacteria help the plants obtain nutrients from the soil
solution and the soil atmosphere, and they are indispensable to the decomposition
of dead organic matter. Plants also favor the development of a thicker soil by
diminishing the rate of surface removal by agents of erosion. There are close,
complex relationships and feedback processes between plants, microorganisms
and fauna, the soil solutions, and the structure and mineralogical composition
of the soil. Details on these processes can be found in generalized (Jenny 1980,
Duchaufour & Souchier 1991) or specialized literature (Callot et al 1982, Paul &
Clark 1989, Robert & Berthelin 1986, Robert & Chenu 1992). In this review,
three main consequences of these interactions are described: the enhancement of
weathering by plants, the control of water available for weathering processes, and
more specifically the geochemical consequences of the cycling of elements in the
upper part of the weathering mantle.
Enhancement of Weathering by Vegetation
Vegetation, with its associated fungi and bacteria, promotes mineral dissolution by
modifying the chemical characteristics of the soil solution. This occurs mainly in
the upper horizons, where the litterfall is decomposed, and in the volume immedi-
ately surrounding the fine roots (a few millimeters), called the rhizosphere, where
the activity of microorganisms is intensely stimulated by organic compounds ex-
creted by the roots. Plants alter the weathering conditions mainly by changing
the pH, the complexing capacity, concentrations of dissolved elements in the soil
solution, and, to a lesser extent, the Eh of the soil solutions.
The fine roots directly modify the pH in the rhizosphere by excreting pro-
tons (H+) or hydroxyl ions (OH). Jaillard et al (1996) mapped pH around
fine roots grown in pH-sensitive gel, and they showed that pH can be raised or,
more commonly, lowered by more than 2.5 pH units in the rhizosphere, depending
on the physiological state and the nutrient supply of the plant (B Jaillard, personal
communication). Acidification results from the combined effects of several pro-
cesses at the root-soil interface. The root actively excretes H+, which exchanges
for other nutrients (Mg2+,Ca
2+
,NH
4
+
,K
+
, and so on), and thus maintains a
charge balance (Briskin 1994). It takes up basic cations and excretes acid cations
[Al3+(Andersson 1988)] and organic acids (Jones & Brassington 1998). Other
processes related to the oxidation of the dead organic matter tend to lower the
pH. The release of CO2by respiration of the root and associated organisms
increases the carbonic acid content of the soil solution and, hence, its acidity.
The nitrogen and sulfur released from plant tissues are biologically oxidized to
form nitrate NO3and sulfate SO42, and organic acids are also excreted by
microorganisms.
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
140 LUCAS
The biota also enhances the formation of organometallic complexes, because
mostof the released organic acidshavethe ability to formcomplexes(Buffle 1988).
This enhances the mobility of metals whose solubility is low in abiologic systems,
such as Al3+,Fe
3+
, or other transition metals. It thereby enhances the dissolution
of metal-bearing minerals (Robert & Berthelin 1986, Tan 1991, Drever & Vance
1994). For example, Robert et al (1979) showed that the dissolution of a mica
(phlogopite) in a reactor is eightfold faster when a 104molar oxalic acid solution
is used than when pure water is used and threefold faster than when 104molar
sulfuric acid is used. The redox potential may be locally changed as the result
of depletion of dissolved O2through the oxidation of organic matter, which can
result in increased dissolution of redox-sensitive elements such as iron.
The effects of these processes on weathering have been quantified in the labo-
ratory as well as in some field experiments. Among numerous studies, Hinsinger
et al (1993) grew plants with a mineral (phlogopite) as the sole source for both
K and Mg, and they found that the plants were able to increase the release of K
and Mg by a factor of two to four after four days. In a similar experiment using
phosphate rock and alumina sand as the mineral substrate, dissolution was signi-
ficantly enhanced by the action of the roots (Hinsinger & Gilkes 1997). On a field
scale, Moulton & Berner (1998) measured the release of Ca and Mg into streams
and growing trees in two areas of basaltic rocks—one with vegetation and one
barren—in Iceland. The rate of release was two- to fivefold higher in the area with
vegetation than in the barren area.
The effect of plants on the weathering rates, defined as equivalent to the flux
of rock-derived base cations (Ca2+,Mg
2+
,Na
+
, and K+) leaving a catchment in
aqueous solutions, was addressed in detail by Drever (1994). The effect of plants
on weathering rates of silicate rocks is probably a factor between 1.5 and 10 and
likely no more than a factor of 2. Drever emphasized that a decrease of the pH
significantly enhances weathering of silicate only when it turns lower than 4.5,
and that the concentration of organic ligands is insufficient to cause a significant
increase in dissolution rate.
Anotherimportantfactor when considering the weathering rate at thecatchment
scaleis the thickness andnatureof the soil cover. Asemphasizedby Stallard (1985)
andfurther discussed byBluth& Kump (1994), weatheringratewill decrease asthe
soil gets deeper, isolating the bedrock from precipitation. Most of the aggressive
processes caused by plants are actually more strongly expressed in the upper
soil horizons and in the rhizosphere. Because the fine roots have a rapid turn-
over, all of the soil material is affected by rhizospheric processes at a rate that is
considered rapid with respect to the times of development of a soil profile. As
argued further, there is in the upper horizons great variation in biological activity,
and therefore great variations of pH and concentration of organic ligands. The
mobility of organic ligands is, however, very low in horizons containing clay,
oxide, or hydroxide minerals because organic compounds are easily adsorbed on
mineral surfaces. At greater depths, roots are scarce or missing, the biological
activity is reduced, the pH is buffered, and the concentration of the organic ligands
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 141
is quite low because those produced in the topsoil horizons were mostly retained
there.
Control of Water Dynamics
The water available for weathering processes depends on the balance between the
infiltration of rainwater into topsoil and the output of water by evapotranspiration.
The infiltration into the topsoil depends on the soil permeability, which by itself is
sustained by plant-controlled processes (Colin et al 1992). The evapotranspiration
is the sum of evaporation from topsoil and transpiration through the plant leaves.
In areas where vegetation cover is continuous, transpiration is the main process
of water loss. The corresponding uptake of water by roots has two main effects;
it reduces the deep drainage, that is, the flux of water that percolates downward,
and it creates a buffer that absorbs percolating free water, thus regulating deep
drainage, and contributes to reduction of the surface runoff.
Thereduction ofdeepdrainage byrootuptake isevident inmost dry ortemperate
areas, but is also quite significant in the humid tropics. In the Amazon Basin,
rainfall ranges from 1500 to 3500 mm year1, and the average evapotranspiration
for tropical forests is 1460 mm year1(Bruijnzeel 1990). Based on the values
listed by Fritsch (1992), it appears that most of the rainforest areas have deep
drainage of 500–800 mm. The reduction of the deep drainage is typically linked
to an increase of the residence time of the weathering solutions, so that these are
likely to reach higher concentrations for some dissolved elements such as Si, Al,
orFe. If concentrations getcloseto saturation withSi-,Al-,orFe-bearing minerals,
the dissolution of these minerals is reduced or stopped (Lasaga 1984).
Root uptake modifies the location of water in the soil voids, and thus the energy
state of water and itsmovement. In soil material, the pore diameters are not equally
distributed. The pore size distribution depends on the size and arrangement of soil
particles (Figure 2). Pores larger than 30 µm (equivalent void radius) allow a
fast downward percolation of the rainwater. In smaller pores, the water is retained
as capillary water. The plants are able to take up capillary water that is stored in
Figure 2 Void size dis-
tribution in a clayey fer-
ralsol. 1e, void index of
the pore size mode (ratio
of pore volume of the pore
sizemode tototalsolid vol-
ume); IW, immobile wa-
ter; WA, water available
for root uptake. Adapted
after Grimaldi et al (1993).
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
142 LUCAS
poresthatare>0.1to0.3µm, a limit corresponding to thewilting point. In pores
smaller than 0.1 µm, advection is negligible, and the water can be considered
immobile; elements dissolved in the water contained in these pores are exchanged
by diffusion with the advective water in larger pores. The capacity of soil material
to retain water (holding capacity) depends on the abundance of small pores, which
is greater in clay-rich than in sandy materials. The holding capacity, coupled with
the uptake by roots between rain events, provides a buffer to the moisture fronts
that percolate deeper. In clay ferralsols from Amazonia exposed to an average
2100 mm of rainfall, free water is never sampled in tension-free tray lysimeters
below a depth of 80 cm (Lucas et al 1996). In the upper decimeters of these soils,
the physical state of water is quite variable, alternating from free to capillary water
depending on the rain events. In deeper zones, water slowly percolates downward
as capillary water. Because there is a general correlation between water content
and the abundance and activity of microorganisms, to the extent that soil aeration
remains nonlimiting (Robert & Chenu1992), hydric pulses in the upper decimeters
are correlated with pulses of biological activity.
This correlation has a bearing on the significance of soil solutions that are sam-
pled in situ in the weathering profiles. Most commonly, soil solutions are sampled
using tension-free lysimeters or porous cups (Litaor 1988, Grossmann & Udluft
1991). The lysimeters sample freely percolating water situated in pores larger
than 30µm, and the porous cups sample water situated in pores >1.5 µm.
The water in smaller pores is inaccessible to these sampling devices, although it
may represent the greater part of the soil water in clayey materials. This water
surrounds mineral particles and interacts with them. Its chemistry may be sig-
nificantly different from that of the sampled water, because the thickness of the
diffuse layer at the mineral-water interface is significant when compared with the
distance between any point in the solution and the mineral surface (Fripiat et al
1971).
Plant-Induced Precipitation of Minerals
Plants can locally induce the precipitation of minerals, either in the root-
rhizosphere system or in their aerial parts, that is, stems, leaves, and flower
parts. The precipitation of clay (kaolinite) in root cells accompanying decomposi-
tion of cell tissues wasdemonstrated by Callot et al (1984, 1992). Theprecipitation
occurs when the secondary cell wallis still intact, which rules out the hypothesis of
a physical displacement of clay particles from the surrounding soil. The Si and Al
needed for precipitation may come from the soil solution and from the plant tissue
itself. Jaillard (1987) and Jaillard et al (1991) documented that the precipitation
of calcite in the root-rhizosphere system is an important process in areas subject
to alternately humid and dry conditions such as Mediterranean climates. The cal-
cite precipitation can occur inside or outside the plant tissue. When outside, it is
usually closely related to associated fungal activity. These processes can result in
the genesis of carbonate-indurated soil horizons called calcretes, which can serve
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 143
as paleoecological indicators (Wright et al 1995, Becze-De´ak et al 1997). The
precipitation of iron and manganese is frequently observed in the rhizosphere in
hydromorphic soils. Migration of Fe and Mn and localization of the phases are
related to the variations of EHin the region surrounding the roots due to the dif-
fusion of O2from the root cells (Chen et al 1980) and to the microbial activity
in the rhizosphere. FeIII is precipitated mainly as goethite and lepidocrocite, and
FeII mobility may be regulated by biologically precipitated green rust (G´enin et al
1998). Mn oxides occur as quite small crystals that are difficult to identify. The
oxidation of MnII and the precipitation of oxides containing MnIV, which are very
slow in laboratory conditions, are catalyzed in soils by bacteria or fungi (Herbillon
1994).
In plant tissues, biomineralization is related to the accumulation of elements
taken up by roots from the soil solution. Cystoliths are minute carbonate con-
cretions occurring in the aerial parts of plants. Phytoliths are small opal (SiO2)
concretions precipitated within the cells, cell walls, and intercellular spaces. Their
size and shape often are indicative of the plant family or genus that yielded them.
Because some phytoliths can be relatively stable in soils for thousands of years,
they may yield information on changes in vegetation and pedogenic processes
during late Quaternary time. Because they are more soluble than most siliceous
soil minerals, they constitute a source of labile Si in the upper part of the soil
cover (Alexandre et al 1994, Meunier et al 1999), which significantly affects the
chemical composition of the soil solution as well as the cycling of Si. Watteau
et al (2000) observed in beech leaves and roots that Si is stored within the cells
as coatings of polyphenolic substances and between and within the cell walls in a
close association with the pecto-cellulosic matrix.
Uptake of Elements by Plants
Plant tissues are mainly composed of H, C, and O, but they contain several other
elements in significant amounts. The average mineral composition of plants is
given in Table 1. Among the naturally occurring elements, only 17 are truly
essential for the growth of all plants. The incorporation of H, C, and O mainly
occurs during photosynthesis, and these are generally not considered to be mineral
nutrients. The other elements are mostly taken up from the soil solution by roots.
N, P, S, K, Mg, and Ca are found in relatively high concentrations in plant tissues.
They are considered majorelements or macronutrients. Fe, Mn, Zn, Cu, B, Mo, Cl,
and Ni are found in low concentrations in plant tissues and are considered minor
elements or micronutrients. Na, Si, Co, I, and V are needed by only some plant
species. Si has a special status because it is strictly essential only for some plant
species,butitisbeneficial formanyothers (Epstein1994). All of theother elements
are found in plant tissue in small quantities. Some of them, such as Al, although
potentially toxic to plants, may be found in significant concentrations because of
their abundance in acid soil solutions. The main source for most elements is the
soil solution. For some elements, such as N, S, and P or Cl, Na, and Ca in coastal
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
144 LUCAS
TABLE 1 Typical composition of plantsa
Dry matter Dry matter
Element (mg g1) Element (ppm)
C440 Fe 20–600
O440 B 2–800
H60 Mn 10–600
K 5–80 Zn 5–250
Ca <60 Cu 5–50
Mg <50bNi 0.05–5
P<50bCo 0.05–10
S<15 Mo 0.1–10
Na <10
Cl <8
Si <100
Al <10
aData are from Callot et al 1982 and Epstein 1994.
bValue usually <5mgg
1dry weight.
areas, a significant amount may come from rainfall and dry deposition of aerosols.
Direct leaf absorption may also occur and was postulated by Alfani et al (1996)
for Cu and Pb.
Becauseof theireconomicimportance, the uptakeandcycling ofmacronutrients
and many trace elements are well documented—in natural ecosystems as well as
in cultivated areas, in both tropical and temperate areas (Likens et al 1977, Mengel
& Kirkby 1987, Attiwill & Adams 1993, Marschner 1995). The uptake of toxic,
harmful, or environmentally hazardous elements such as heavy metals has been
the focus of numerous studies. Comparatively few studies have been done of the
cycling of those elements that constitute most of the secondary minerals and thus
are most involved in the results of weathering: Si, Al, and, to a lesser extent, Fe,
Mn, and Ti.
Si uptake is metabolically regulated and can be independent of the plant’s trans-
piration rate (Barber & Shone 1966, Leo & Barghoorn 1976). Si is precipitated
in phytoliths, although many plants do not have readily observable phytoliths yet
contain relatively large amounts of amorphous silica. Precipitation occurs near
transpiration termini, but it can also occur in the xylem vessels and in the endo-
dermis of roots (Raven 1983). Terrestrial plants contain Si in appreciable concen-
trations, ranging from a fraction of 1% of the dry matter to several percent and in
some plants >10%. Si concentrations in leaf litterfall in tropical forests range
from 0.05 to 25 mg g1dry weight. The lower value (0.05 mg g1) is from a forest
in Malaysia growing on limestone, a Si-depleted rock. Most values range between
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 145
4and8mgg
1(Klinge & Rodrigues 1968, Gautam-Basak & Proctor 1983, Lucas
et al 1993, Cornu et al 1995).
PlantscontainAl in lower concentrations thanSi. Alistoxic for plants when ab-
sorbed in excess, but Al tolerance varies greatly between species and even between
varieties of the same species. Al uptake is metabolically regulated by exclusion
mechanisms from root uptake and by active excretion (Andersson 1988). Al can
be precipitated with Si in phytoliths (Bartoli & Wilding 1980). Al concentrations
in leaf litterfall in tropical forests range from 0.12 to 6.2 mg g1dry weight, and
most values range between 0.2 and 0.8 mg g1(Webb et al 1969, F¨olster & de Las
Salas 1976, Gautam-Basak & Proctor 1983, Lucas et al 1993, Cornu 1995).
Element Cycling, Chemical Composition of the Soil Solution,
and Interaction with Soil Minerals
The elements that have accumulated in plant tissues constitute a reservoir of those
elements in and above the soil profile. Assuming that the amount and composition
of the vegetation cover are in steady state, all of the elements taken up annually
from the soil return annually to the top of the soil profile. This return occurs mainly
by means of litterfall or decaying plants; however, foliar leaching by rainfall
(especially during leaf senescence in autumn) and stem flow (the leaching of
elements along the branches and trunks by the downward flowing water) can also
besignificant. If themassvegetationincreaseswith time orif vegetationis cropped,
there is a net loss of elements from the soil volume influenced by the roots.
Litterfall is decomposed within the organic horizons in the upper part of the
soil, so that most of the mineral elements contained in plant tissues are released
into the soil solution that percolates downward. At a few centimeters in depth,
the soil solution may be potentially enriched with Si, Al, Fe, and other elements
that control the stability of secondary minerals. Direct interaction of an element
with a soil mineral is significant if the considered element is a constituent of the
mineral and if its concentration in the soil solution is sufficient with regard to
thermodynamic equilibrium. Grimaldi (1987, 1988) described, for example, the
control of Al in catchment waters in French Guyana by precipitation with
a low rate of a secondary mineral, likely gibbsite. Three main types of plant
cycling are possible (Figure 3). If the element is not a constituent of any mineral
of the considered soil, the biogeochemical cycling of this element involves only
atmosphere, soil solution, and plants (Figure 3A). This happens to N and S in most
soils and to Ca, Mg, or K in highly leached soils where all primary minerals have
been weathered and no secondary minerals contain these elements. If the element
is a constituent of minerals of the considered soil, but its concentration in the
soil solution remains below saturation with any secondary minerals, plant cycling
only increases the dissolution of the element-bearing minerals (Figure 3B). This
happens, for example, to Ca, Mg, or K inmany soils. If the elementis a constituent
of minerals of the considered soil and its concentration in the soil solution reaches
saturation with secondary minerals, plant cycling directly interacts with the result
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
146 LUCAS
Figure 3 Plant cycling interaction with soil minerals. For explanations of the processes
illustrated in parts A,B, and C, see text.
of pedogenesis (Figure 3C). This may happen for Si, Al, or Fe, because of the low
solubility of oxides or clay minerals.
Table 2 gives values, in kilograms per hectares per year, of measured annual
turnover resulting from litterfall in some tropical and temperate forest ecosys-
tems. The values for Si are high, comparable with the values for Ca, which is
the main macronutrient. The values for Al and Fe are lower but significant with
respect to the solubility of these elements in water. Assuming that all of the
elements contained in the annual litterfall are dissolved in the water that annually
percolates through the litter, we can obtain average concentrations in the soil so-
lution beneath the litter. The water that annually percolates through the litter was
considered equal to the trough flow, which is annual rainfall minus rain intercep-
tion by the canopy. The results of these calculations are given in Table 3. Assu-
ming that the elements contained in the litter are released in the soil solution
as free ions, these waters are plotted in the solubility diagrams of Figure 1.
With such anassumption, the waters are supersaturated with Al andSi with respect
to gibbsite and kaolinite for a pH >4.5 at 25C (Figure 1). These waters are also
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 147
TABLE 2 Annual turnover caused by the litterfall in some forest ecosystems
Turnover (Kg/ha year1) of element:
Type of forest Referencea
ecosystem (location) Si Al Fe Mn Ti Ca
A
Rainforest, ferrasols (Carajas, Brazil) 9 2 2.6 1 0.05
B
Rainforest, ferrasols (Manaus, Brazil) 33.3 3.9 1.2 0.79 0.22 36.7
C
Rainforest, ferrasols (Manaus, Brazil) 48.6 6.7 0.4
D
Rainforest, podzols (Manaus, Brazil) 21.1 4.2 0.28
E
Deciduous forest, cambisols (Vosges, France) 22 ±7 1.5
F
Coniferous forest, podzols (Vosges, France) 5 ±2
G
Deciduous forest, cambisols (Vosges, France) 35.5
H
Deciduous forest, alfisols (Vosges, France) 45
I
Mixed forest, cambisols (Vosges, France) 21
J
Coniferous forest, alfisols (Vosges, France) 22.5
K
Bamboo forest, volcanic ash soil (Island of Reunion) 970–1380
aA, Rose et al 1993; B, Lucas et al 1993; C and D, Cornu 1995; E and F, Bartoli 1986; G to J, Bartoli & Souchier 1978; K, Meunier et al 1999.
´
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
148 LUCAS
TABLE 3 Calculated average composition of the soil solution beneath litter
Composition (mM) of element in forest ecosystema[annual trough flow (mm)]
Element A(1530) B(1790b)C (1790b)D (1790b)E (910) F(910)
Si 2.1 ×1026.7 ×1029.7 ×1024.2 ×1028.6 ×1022.0 ×102
Al 4.8 ×1038.1 ×1031.4 ×1028.7 ×1036.1 ×103
Fe 3.0 ×1031.2 ×1034.0 ×1042.8 ×104
aLetters above columns correspond to ecosystems in Table 2.
bCalculated by the method of Roche (1990).
saturated with respect to amorphous ferric hydroxide [Fe(OH)3] in the same range
of pH.
These considerations (a) assume that all of the elements contained in the lit-
ter are released into the soil solution as free ions and (b) ignore the role of the
complexing compounds released by biological activity. In fact, biominerals such
as phytoliths are partially resistant to dissolution within the litter. In a study of
Hawaiian ecosystems, Kelly et al (1998) considered that a pasture ecosystem
maintains lower amounts of more soluble silica than the forest because phytoliths
produced by grasses are more resistant to weathering. Watteau et al (2000), study-
ing the release of Si from decomposing beech leaves, observed that Si associated
with polyphenolic substances is more labile than Si associated with the pecto-
cellulosic membranes. In a rainforest ferralsol from the Congo (Alexandre et al
1994, 1997), phytoliths are selectively dissolved in the litter. Some resistant phy-
toliths are translocated downward into the underlying horizons where they are
slowlydissolved. The measured amount of Si stored in that profile is high (11 ×
103kg ha1in the first 30 cm) and progressively decreases with depth. Higher
values for Si storage as soil phytoliths (45 ×103to 104 ×103kg ha1) were
found in soil from the Island of R´eunion (Meunier et al 1999). Assuming that only
5%–10% of the phytoliths from the litterfall are dissolved annually in the litter,
the concentrations of Si in Table 3 can be divided by 10–20, giving concentrations
that are still sufficient to maintain supersaturation for kaolinite in the solutions
percolating beneath the litter.
When complexing agents are released in the soil water by the biota, they com-
bine with dissolved Al, Fe, or other metallic elements to form organometallic com-
plexes that either remain in solution or are adsorbed onto the surfaces of minerals
or organic matter. These complexed metals are no longer available to precipitate
as secondary minerals. The stability constants of organic complexes with Fe and
Al are relatively high (Buffle 1988). Cornu et al (1998) calculated that >90%
of the dissolved Al and Fe in tropical ferralsol and podzols was present as organic
complexes in the soil solutions. The existence of dissolved organic complexes is
thus able to keep the soil solutions undersaturated with secondary minerals, even
when total Al, Si, and Fe concentrations are relatively high (Stevenson & Vance
1989, Duchaufour 1990).
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 149
The stability of secondary minerals in the weathering mantle depends on inter-
actions between the biological cycling of elements, the rates of release of elements
within the litter and translocation of biogenic minerals downward in the soil pro-
file, the complexing capacity of soil solutions, and the dynamics of water within
the weathering mantle. These interactions are described below in the case study
of an exhaustively investigated weathering cover in central Amazonia.
PLANT-CONTROL OF WEATHERING COVER:
EXAMPLE OF AN AMAZONIAN SOIL SYSTEM
The Ferralsol-Podzol Systems
Ferralsols are highly weathered, yellow to red, clayey to sandy-clay soils. The
organic residues decompose quickly on the topsoil, and the litter forms a thin hori-
zon. Most elements are leached, including Si, and there is a relative accumulation
of Al and Fe. The remaining primary minerals are those that are the most resistant
to weathering, mainly quartz. The secondary minerals that accumulate are mainly
1:1 clays (kaolinite), Al-hydroxides (gibbsite), Fe-oxi-hydroxides (goethite) and
Fe-sesquioxydes (hematite). Podzols are highly weathered, bleached, sandy soils.
The organic residues decompose slowly on the topsoil, and the litter forms a thick,
peat-like horizon. All elements are leached out of the upper horizons, including Al
and Fe, which migrate as organometallic complexes. The upper horizons consist,
of residual white sand, mainly quartz and other resistant minerals. Organic matter,
Al, and Fe can accumulate at depth.
The weathering mantle of the Central Amazon Basin, 80 km north of
Manaus, Brazil, has been described in detail by Chauvel et al (1987) and Lucas
et al (1996). The geomorphic landscape consists of low plateaus dissected by flat-
bottomed valleys; the topographic relief between plateau and valley bottom is
typically 30 to 40 m. The underlying rocks are sandy-clayey, Cretaceous-Tertiary
sediments that consist mainly of quartz and kaolinite. The annual mean rain-
fall is 2100 mm, and the annual mean temperature is 26C. The soils on the
plateaus and on the upper parts of the slopes are clayey, deeply weathered ferral-
sols that support a typical rainforest flora. On the lower parts of gradual slopes,
the soils are thick, sandy podzols. Along the slopes, the ferralsols and the pod-
zols grade into one another. As the soils become progressively sandier down-
slope, the trees become smaller, and the species diversity becomes lower. The
long-term evolution of the soil mantle is a progressive replacement of the clayey-
ferralsol weathering mantle by a sandy podzolic mantle. As a slope develops
backward at the expense of a plateau, the podzols progress upslope at the ex-
pense of ferralsols (Lucas et al 1987). This forms a transformation system (Boulet
et al 1997). Under natural forest, the surface runoff is very weak (Franken &
Leopoldo 1984). The mechanical surficial erosion is small with regard to geo-
chemical leaching, and the evolution of the whole system is mainly geochemi-
cal. Stallard & Edmond (1983) stressed that the weathering rates in the lower parts
of the Amazon Basin are extremely slow, which applies to the studied area. The
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
150 LUCAS
old, thick weathering cover (>30 m) contains a very small amount of base cations
(Irion 1984), and more than 98% of the constitutive minerals are residual quartz
and secondary minerals (kaolinite, gibbsite, hematite, goethite). This weathering
cover thus configures an outstanding geochemical model for studying the con-
ditions of the stability of the secondary minerals and particularly the effect of
vegetation.
The plateau ferralsols consist of a 3- to 8-m-thick kaolinite-rich horizon overly-
ing a more gibbsitic 3-m-thick horizon, followed by an 8-m-thick progressive
transition to saprolite. The depth to the unweathered sediment is unknown. Water
percolates vertically downward to the deep-water table, which is at 20 to 30 m
below the surface. From the saprolite to the top of the gibbsitic horizon, quartz
grains show features of increasingly strong corrosion. The quartz abundance de-
creases,from55% in clayey layersand80% in sandy layers intheparent saprolite
to 5% at 2 m below the surface. The ferralsols develop by progressive dissolu-
tion of quartz and neo-formation of kaolinite. Petrographic evidence attests that
gibbsite is currently formed in the more gibbsitic horizon, underneath the thick
kaolinitic upper horizon (Irion 1984, Lucas et al 1986).
The kaolinite in the profileis heterogeneous (Lucas at al 1986, Giral-Kacmarc´ık
et al 1998). In the saprolite, kaolinites are hexagonally shaped and arranged in
well-formed booklets. They have low Fe substitution and high crystallinity; the
Hinckley Index values, which are different in each particle size fraction, range
from 1.1 to 1.4, and infrared (IR) spectra have well-defined bands at 3672 and
938 cm1, which indicate well-ordered lattice. All of these features progressively
change in an upward direction. In the uppermost part of the profile, kaolinites
are highly rounded. They have higher Fe substitution and low crystallinity; the
Hinckley Index values range from 0.4 to 0.9, and the 3672- and 938-cm1bands
are lacking in IR spectra, which indicate disordered lattice. The well-defined
upward decrease in kaolinite crystallinity indicates that kaolinite is progressively
modified in response to changing microenvironments as weathering fronts move
progressively downward.
The δ18O values of kaolinites along the profile are in the range of the δ18O
values of kaolinites in equilibrium with present-day percolating waters. Variations
inδ18O valuesof kaolinites fromdifferent-sizedfractions atasingle depthrepresent
climatic variations on a timescale yet to be determined, but perhaps one that is
seasonal and on the order of centuries or millennia (Giral-Kacmarcik et al 1998).
The kinetics of precipitation and dissolution of secondary minerals were also
estimated from petrographic features (Callot et al 1992) and from experiments
in which mineral separates (kaolinite, gibbsite, and amorphous silica) enclosed
within a permeable membrane were buried in the upper horizons (Cornu et al
1995). Kaolinite was precipitated in the cells of decomposing roots. Dissolution
ofkaoliniteand/orprecipitation of new minerals in the buriedmembranewas noted
within 6 months of burial. The reactivity of minerals is thus very rapid with regard
to the duration of pedogenesis, and kaolinite is indeed in dynamic equilibrium in
the uppermost part of the profile. The studied weathering mantle represents a case
study regarding the stability of the kaolinite in the upper horizons.
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 151
Various explanations for the persistence of kaolinite in the upper horizons have
already been proposed: uplift of deep material by the fauna (Nye 1955), rapid
percolation of the rainwater through the kaolinitic horizons so that the water-
mineral interaction does not reach equilibrium (Kronberg et al 1982), or decreased
activity of water in the topsoil during the dry seasons displacing the kaolinite-
gibbsite equilibrium (Tardy & Novikoff 1988). These hypotheses do not apply to
the studied soil however. Faunal activity is restricted to the uppermost part of the
thick kaolinitic horizons. Percolating free water is absorbed in the first 80 cm,
and soil solution slowly percolates deeper as capillary water at a rate estimated at
3–4 m year1by isotopic measurements (Rozanski et al 1991, Giral-Kacmarc´ık
et al 1998). If, during exceptionally dry periods, soil suction reaches 15 ×105Pa,
which is the limit for plant wilting, maximum water activity coefficient changes
would range from 1 to 0.989, which correspond to a very small (+0.55%) shift in
the dissolved [Si(OH)4] concentration at the gibbsite-kaolinite equilibrium.
The podzols that develop downslope consist of a quartz sand horizon with a
thickness ranging from 1 to >4 m, overlying more clayey layers inherited from
the parent sediments. Organic matter accumulates at the interface between sandy
and more clayey material. The sandy horizons have a high hydraulic conductivity.
Biogeochemistry of the System
To investigate the processes that determine whether the products of weathering
become either stable kaolinitic ferralsol or podzols, the biogeochemical cycling
of elements was exhaustively determined in both environments. The litterfall was
sampled in a 2-year experiment (Table 2, line B) and a 1-year experiment (Table 2,
lines C and D). The soil water was sampled by tension-free cylindrical and tray
lysimeters installed beneath the litter and at 10, 20, 40, and 80 cm. The groundwa-
ter was sampled in boreholes at the water table. The annual balance of dissolved
Si, Fe, and Al that is transferred by rain, throughfall, stem flow, soil water, and
groundwater was determined throughout a hydrological year (Chauvel et al 1989,
Cornu et al 1997a). The springwaters were sampled in small streams issuing from
two well-defined areas, a ferralsol area and a podzol area. The dissolved fraction
in springwater was separated by tangential ultrafiltration with a threshold at 5 kDa
(Eyrolle 1994). The speciation of Si, Fe, and Al and the state of equilibrium of the
solutions with respect to the dominant minerals (kaolinite, gibbsite, and quartz)
were determined by using the MINTEQ (Center for Exposure Assessment Model-
ing1993)program.Thereleaseofelementsbydecomposinglitterwasinvestigated
insituthroughlitter-bag experiments (Cornu et al 1997b).Thechemistryofthesoil
solutions was studied within the soil by using vermiculite and chelating resins as
test minerals (Righi et al 1990); vermiculite undergoes de-alumination in the pres-
ence of complexing agents, and forms hydroxy-Al vermiculites in simply acidic
systems, the chelating resins exchange cations from the soil solution. The main
results of these studies are summarized below.
The main source of Si, Al, Fe, and macronutrients in the topsoil is litterfall.
Inputs from rain, dust, and biological release in the canopy are small but not
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
152 LUCAS
negligible. As seen above (Table 2), the input of Si and Al by litterfall is high
in both ferralsol and podzol forests. The concentrations of these elements in soil
solution exceeded what was necessary to reach saturation with kaolinite, although
moreSi was recycledinthe ferralsol forest than inthepodzol forest. In the ferralsol,
the total Si input in the topsoil (40–50 kg/ha year1) was about three- to fourfold
greaterthan the Sileachedout of thesystem by deepdrainage(11–17 kg/ha year1).
When the water percolates through the litter in both environments, it becomes
enriched with dissolved Si, Al, and Fe. The annual balance, however, indicates a
deficit with regard to the annual input from throughfall and litterfall. That means
that a part of the Si and Al released by the litter is not transferred into the soil as
dissolved compounds in free water but as translocated particles (phytoliths) and
dissolved compounds in capillary water, which was not sampled by the lysimeter.
In the uppermost 40 cm of the soils, the concentrations in nutrients in the soil
waters in both the ferralsol and the podzol are of the same order of magnitude.
Uptake of nutrients by roots is high in the uppermost 20 cm of soil. The Si and
Al concentrations in that part of the profile are high but vary greatly with time.
The Al concentrations are correlated with concentrations of NO3or SO42and
are higher in the hours after a rainfall, which indicates that the release of Al is
related to pulses of microbial activity. In the ferralsol soil solutions, the Si and Al
concentrations range from 26 to 54 µM and 4.1 to 20 µM, respectively, and the pH
ranges from 3.9 to 4.8. In the ferralsol water table, the Si and Al concentrations
range from 39 to 64 µM and 1.9 to 4.8 µM, respectively, and the pH ranges from
4.6 to 5. In the podzol soil solutions, the Si concentrations are slightly higher,
and the Fe and Al concentrations are considerably higher than in the ferralsol soil
solution, although these latter elements are less abundant in the podzol material
than in the ferralsol. The pH ranges from 3.2 to 4.6. In both environments, most of
the dissolved Al and Fe are present as organometallic complexes, and the waters
are undersaturated with kaolinite, gibbsite, and quartz. Nevertheless, the in situ
experiment with test minerals enclosed in permeable bags revealed differences
between the two environments. The ferralsols are characterized by a simple acidic
system, whereas podzols are more affected by organic complexation, especially in
the rainy periods. The waters sampled in the ferralsols are thus not representative
oftheprevailingprocesses in the horizons. This may be understood by considering
the water dynamics in the horizons.
Plant-Sustained Stability of Kaolinite
In the topsoil, the peak of biological activity that follows rainfall events enhances
the mineralization of the litter, increases acidity, and produces complexing organic
compounds that attack minerals aggressively. This results in high concentrations
of Si, Al, and Fe in the soil solutions. Most Al and Fe occur as organic complexes,
and the soil solution is undersaturated with kaolinite.
In the ferralsols, the dissolved organometallic Fe and Al complexes cannot
migrate deeply into the profile, because they are adsorbed on surfaces of minerals
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 153
Figure 4 Processes con-
trolling Al transfer in the
topsoil. Dashed arrow, oc-
curs in both environments;
black arrow: occurs in fer-
ralsols; white arrow: oc-
curs mainly in podzols.
with high exchange capacities, until the complexes are biologically decomposed
(Figure 4). Metals are then released to the soil solution, which reaches saturation
withkaolinite. The bimodal distributionof pores (Figure2)favorsabufferingeffect
between the percolating water and the water actually in contact with minerals.
Kaolinite thus remains in the topsoil because (a) the recycling of Si by biological
pumping provides high concentrations of Si in the soil solutions and (b) organic
compounds, inclusive of those that form Al complexes, are not able to migrate
through the clayey materials, so that Al is not leached out. The reason gibbsite
precipitates at greater depths is still unclear. It may be due to one or more of the
following processes: acidification by CO2produced in place by root respiration
(Grimaldi & Pedro 1996) or CO2accumulated after transferring from the upper
horizons; increased Al concentrations in the soil solution because of exclusion
mechanisms of Al from root uptake; transfer of Al by inorganic complexes (Taylor
1988); or shifting ofthermodynamic equilibrium due tochanges in nanopore fabric
and abundance. Water infiltrating the surface takes some years to reach the deep
water table and longer to emerge in springs. The long residence time of water in
the weathering mantle provides enough time for it to approach Si saturation for
quartz. The spring waters that flow out of the system are thus clear (colorless),
poor in organic matter, Fe, and Al, and relatively enriched in Si (Table 4).
Genesis of Podzols versus Ferrasols
In the sandy uppermost horizons of the podzols, the low moisture-storage capacity
results in transient waterlogging followed by dry conditions. This heterogeneous
moisture regime combined with soil acidity inhibits a complete mineralization of
organic matter and favors the production of organic compounds. The resulting
dissolved organo-Fe and organo-Al complexes are rapidly leached out of the sys-
tem, because of the high hydraulic conductivity, the low hydraulic buffering ca-
pacity, and the low exchange capacities of the surfaces of the sandy materials.
Because of the low residence times of the percolating waters and because the rate
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
154 LUCAS
TABLE 4 Concentrations of elements in spring watersa,b
Concentration (mg liter1) in soil type area
Ferralsol (state of element) Podzol (state of element)
Element Dissolved Bulk Dissolved Bulk
Si 1.7 1.9 0.8 0.91
Al 0.12 0.23 0.31 0.63
Fe 0.6 0.06 0.08 0.28
MO 0.42 1.2 4.4 22
aAdapted from Lucas et al 1996.
bThe dissolved fraction was separated by tangential ultrafiltration with a threshold at 5 kDa.
of Si recycling by plantsis lower than in ferralsol, the Siconcentrations in the solu-
tions remain lower than in the ferralsols solutions. The spring water thus has high
concentrations of dissolved organic matter, Fe, and Al and lower concentrations
of Si than ferralsol spring waters (Table 4).
The soil profile, that is, the product of the pedogenic process, is thus entirely
controlled by biological pumping and system hydrodynamics (Figure 5). The bio-
logical pumping allows the maintenance of a Si stock in the upper horizons, but the
stability of the kaolinite is ensured only where the organo-Al complexes cannot be
leached out of the topsoil. In such places, the final result of weathering is a clayey,
kaolinitic material. The soil solutions slowly drain downward, quartz is dissolved,
Si is leached, and there is a relative accumulation of Al and Fe. Where topsoil
solutions can drain rapidly out of the system, all of the elements are leached at
a high rate, and a relative accumulation of residual quartz grains develops. This
process creates positive feedback by increasing the hydraulic conductivity of the
Figure 5 Dynamic of the podzol-ferrasol soil system. OM, organic matter.
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 155
Figure6 Comparison of thedynamics of Amazonianferralsol (A), Amazonianpodzol (B),
eastern-France cambisol or alfisol (C), and eastern-France podzols (D). Phytolith storage
is given for the 0- to 50-cm horizons. Illustrations are based on data from Cornu (1995),
Lucas et al (1996), Bartoli & Souchier (1978), Bartoli (1986), and Berthelin et al (1990).
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
156 LUCAS
soil material and favoring a heterogeneous moisture regime in the topsoil. The on-
set of podzolization is likely to be favored by a sandy parent material that ensures
easy leaching downward, ultimately toward the forest stream. Downslope alluvial
deposits are particularly conducive to this scenario. After this beginning, positive
feedback leads to the development of podzols at the expense of the ferralsol cover.
A sandy parent material, however, is not necessary to initiate such a process. In
French Guyana, podzols develop at the expense of sandy-clay ferralsols, the whole
system being developed on the same saprolite issued from basement rock (Boulet
et al 1997).
Theresults described abovemustbecomparedwith those obtained in temperate
areas by Bartoli & Souchier (1978) and Bartoli (1986) regarding plant cycling of
Si (Table 2) and by Berthelin et al (1990) regarding soil processes. Most of these
resultsconcernsoilsformedongraniteorsandstoneintheVosges (eastern France).
Results are summarized in Figure 6. Cambisols or alfisols have high Si plant cy-
cling, moderate weathering of minerals, low residence time of phytoliths, and low
Si leaching out of the system. Organic matter is fully mineralized. Dissolved or-
ganic compounds cannot migrate through horizons, and acidity of the soil solution
is mainly due to nitrates. Free cations in the soil solution are mainly Ca2+,Mg
+
,
and K+if soil is formed on base-rich parent material, and Al3+if soil is formed on
base-poorparent material. Podzols havelowSibiological cycling, highweathering
of minerals, and high Si leaching from the system. Organic matter is incompletely
mineralized. Dissolved organic compounds migrate freely through eluvial hori-
zons, and acidity of the soil solution is mainly due to complexing organic acids.
Cations in the soil solution form organometallic complexes that migrate toward
lower depths. The role of plant cycling and the biogeochemistry of the systems
are, thus, close to those related to Amazonian soils. The Si storage as phytoliths is,
however, much higher in tropical ferralsols than in temperate soils. Another differ-
enceisthat Si leaching from the system ishigherinthe temperate podzol than in the
moreclayey temperatesoil, which is theopposite of whatoccurs in Amazonia. This
may be due to the low residence time of percolating water in temperate alfisols or
cambisolsand to permanencyof easily weatherable minerals in temperate podzols.
The change in soil dynamics after a change of vegetation may give rapidly
noticeable results when soil is already sandy. The evolution of nonpodzolic sandy
soilsto podzols (by removalof clay and Feoxidesandbleaching) in French Guyana
occurred during a few years after forest clearing and cropping. The same was
observed in France after substitution of deciduous forest by coniferous (Y Lucas,
unpublished data).
CONCLUSIONS
Biologicalactivity,mainly related to plant growth,maymodifyrates and especially
products of weathering relative to those that would result from abiotic processes.
Plants have significant actions at the scale of microsites, mainly due to peculiar
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 157
chemistry in the rhizosphere, and at the scale of bulk soil, mainly due to plant
cycling of elements, control of water dynamics, and plant-induced release of acid
and complexing compounds into the soil solution. The effect of plants is well
expressed when considering most of the secondary minerals (e.g. kaolinite, gibb-
site, goethite, hematite) for which the soil solution is likely to reach saturation. It
is likely less significant with respect to primary minerals bearing base cations for
which the soil and weathering solutions usually remain far for equilibrium (Drever
1994).
The results of weathering on Si-Al secondary minerals, the main constituents
of the weathering mantles, depend on the balance between two main processes:
(a) the stability of the clay minerals, sustained by the plant Si cycling; and (b) the
leachingof plant-induced organo-Al compounds.Theseresults are evidentinequa-
torialareaswhere old weathering mantles are highly affectedbybiologicalactivity,
and they can also be observed in temperate areas. A change of vegetation may sig-
nificantly change the stability of secondary minerals owing to changes in water
dynamics as well as in the nature of the litterfall and the resultant changes in
Si cycling. Other minerals, such as iron-bearing minerals, are also affected by
these processes. It is thus important to continue to obtain more data on the cy-
cling of mineral-forming elements and to better understand the biogeochemical
cycles.
From a global and geologic point of view, the cycling of elements, mainly
Si, by plants is a key factor in the stability of secondary minerals; in most
places, the weathering mantle can be regarded as a dynamic-equilibrium sys-
tem sustained by plants. The plants maintain not only the nutrients they need
for growing but also a mineral substrate favorable to life. These considerations
raise questions for long-term results of changes in the vegetation, for example
replacement of high-recycling deciduous forest by low-recycling coniferous for-
est or crops. The expression of these changes in the soil profile, however, de-
pends on the soil type. The results of the replacement of forest by crops on
the mineral composition of the soil may be apparently insignificant on clayey
soils, but rapidly noticeable on sandy soils. The processes by which plants control
the weathering mantle have thus to be considered when studying global changes
through geological times as well as during present time. They must be included
in predictive models of soil water chemistry, stream water chemistry, and soil
genesis.
ACKNOWLEDGMENTS
The author would like to gratefully acknowledge financial support through the
French National Research Programs, Programme d’Etude de la G´eosph`ere In-
tertropicale (PEGI) and Programme de Recherche Sols Erosion (PROSE). The au-
thorswouldalsoliketoacknowledgehelpful reviews from SM Savinandunknown
reviewers.
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
158 LUCAS
Visit the Annual Reviews home page at www.AnnualReviews.org
LITERATURE CITED
Alexandre A, Colin F, Meunier JD. 1994.
Les phytolithes, indicateurs du cycle bio-
eochimique dusilicium enforˆet ´equatoriale.
C. R. Acad. Sci. Paris (II)319:453–58
Alexandre A, Meunier JD, Colin F, Koud JM.
1997. Plant impact on the biogeochemical
cycle of silicon and related weathering pro-
cesses. Geochim. Cosmochim. Acta 61:677–
82
Alfani A, Maisto G, Iovenio P, Rutigliano FA,
Bartoli G. 1996. Leaf contamination by at-
mospheric pollutants as assessed by elemen-
tal analysis of leaf tissue, leaf surface deposit
and soil. J. Plant Physiol. 148:243–48
Ambrosi JP. 1990. Mod´
elisation thermody-
namique de l’alt´
eration lat´
eritique dans le
syst`
emeFe2O3-Al2O3-SiO2-H2O.PhDthesis.
Univ. Poitiers, France. 205 pp.
Andersson M. 1988. Toxicity and tolerance of
aluminum in vascular plants: a literature re-
view. Water Air Soil Pollut. 39:439–62
Attiwill P, Adams MA. 1993. Nutrient cycling
in forests. New Phytol. 124:561–82
Banfield JF, Barker WW, Welch SA, Taunton
A. 1999. Biological impact on mineral dis-
solution: application of the lichen model to
understanding mineral weathering in the
rhizosphere. Proc. Natl. Acad. Sci. USA
96:3404–11
BarberD, ShoneM. 1966.The absorption ofsil-
ica from aqueous solutions by plants. J. Exp.
Bot. 17:569–78
Bartoli F. 1983. The biogeochemical cycle of
silicon in two temperate forest ecosystems.
Environ. Biogeochem. Ecol. Bull. (Stock-
holm) 35:469–76
Bartoli F. 1986. Les cycles biog´eochimiques
dans les ´ecosyst`emes forestiers temp´er´es.
Sci. G´
eol. Bull. 39(2):195–209
BartoliF,SouchierB. 1978.Cycle etrˆoledu sili-
cium d’origineeg´etale dans les ´ecosyst`emes
forestiers temp´er´es. Ann. Sci. For. 78:187–
202
Bartoli F, Wilding LP. 1980. Dissolution of bio-
genic opal as a function of its physical and
chemical properties. Soil Sci. Soc. Am. J. 44:
873–78
Becze-De´ak J, LangohrR, VerrecchiaEP. 1997.
Small scale secondary CaCO3accumula-
tions in selected sections of the European
loess belt: morphological forms and poten-
tial for paleoenvironmental reconstruction.
Geoderma 76:221–52
Berner RA. 1992. Weathering, plants and the
long term carbon cycle. Geochim. Cosmo-
chim. Acta 56:3225–31
Berthelin J, Munier-Lamy C, Wedraogo FX,
Belgy G, Bonne M. 1990. M´ecanismes mi-
crobiens d’acidification et d’alt´eration in-
tervenant dans les sols bruns acides et les
podzols forestiers. Synth`ese et discussion
de travaux ant´erieurs. Sci. Sol 28, 4:301–
18
Bluth GJS, Kump LR. 1994. Lithologic and
climatologic controls of river chemistry.
Geochim. Cosmochim.Acta 58:2341–59
Boulang´e B, Carvalho A. 1997. The bauxites
of Porto Trombetas. In Amazonian Bauxites,
ed. A Carvalho, B Boulang´e, AJ Melfi, Y Lu-
cas, pp. 55–73. S˜ao Paulo, Brazil: Univ. S˜ao
Paulo/Fund. Amparo a Pesquisa do Estado
de S˜ao Paulo/Off. Rech. Sci. Tech. Outre-
Mer
Boulet R, Lucas Y, Fritsch E, Paquet H. 1997.
Geochemical processes in tropical land-
scapes: role of the soil covers. In Soils and
Sediments—Mineralogy and Geochemistry,
ed. H Paquet, N Clauer, pp. 67–96. Heidel-
berg, Germany: Springer-Verlag
Brabant P. 1987. La r´epartition des podzols
`a Kalimantan (Ile de Born´eo). In Podzols
et Podzolisation, ed. D Righi, A Chauvel,
pp. 13–24. Plaisir/Paris: Assoc. Fr. d’Etude
Sols-Inst. Nat. Rech. Agron.
Brammer H. 1973. Podzols in Zambia. Geo-
derma 10:249–60
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 159
Briskin DP. 1994. Membranes and transport
systems in plants: an overview. Weed Sci.
42:255–62
Bruijnzeel LA. 1990. Hydrology of Moist
Tropical Forests and Effects of Conver-
sion: A State of Knowledge Review. Paris-
Amsterdam: Int. Hydrol. Progr. UNESCO
& Vrije Univ. 224 pp.
Buffle J. 1988. Complexation Reactions in
Aquatic Systems: An Analytical Approach,
ed. RA Chalmers, JN Millers. New York:
Wiley & Sons. 692 pp.
Callot G, Chamayou H, Bullock P, Murphy
CP. 1984. Observations de concentrations
argileuses dans les structures racinaires au
cours des transformations p´edologiques. C.
R. Acad. Sci. Paris (II)299:825–30
Callot G, Chamayou H, Maertens C, Salsac
L. 1982. Les Interactions Sol-Racine: Inci-
dences sur la Nutrition Min´
erale. Paris: Inst.
Nat. Rech. Agron. 325 pp.
Callot G, Chauvel A, Arvieu JC, Chamayou
H. 1992. Mise en ´evidence de kaolinite et
de silice dans les structures cellulaires de
l’´epiderme etdu cortex deracines de palmier,
en forˆet d’Amazonie. Bull. Soc. Bot. Fr.
139:7–14
Center for Exposure Assessment Modeling.
1993. Matel Speciation Equilibrium Model
for Surface and Ground Water. MINTEQA2
v3.11. Athens, Georgia: US EPA
Chauvel A, Andreux F, Cerri CC, Lucas Y.
1989. Superficial evolution of Amazonian
bauxite deposits. Trav. Int. Comm. Study
Bauxites, Alumina Alum. 19:45–53
Chauvel A, Lucas Y, Boulet R. 1987. On the
genesis of the soil mantle in the region
of Manaus, Central Amazonia. Experientia
43(3):285–99
Chen C, Dixon JB, Turner FC. 1980. Iron coat-
ingson riceroots: morphology andmodels of
development. Soil Sci. Soc. Am. J. 44:1113–
19
Colin F, Brimhall GH, Nahon DB, Lewis CJ,
Baronnet A, Danti K. 1992. Equatorial rain-
forest lateritic soils: a geomembrane filter.
Geology 20:523–26
CochranMF,BernerRA. 1992.Thequantitative
role of plants in weathering. In Water-Rock
Interaction,WR1-7, ed. YK Kharaka, AS
Maest, pp. 473–76. Rotterdam, The Nether-
lands: Balkema
CornuC. 1995. Cycles biog´
eochimiquesdu sili-
cium, du fer et de l’aluminium en forˆ
et ama-
zonienne.PhD thesis. Univ.Aix-Marseille III
Cent. Eur. Rech. G´eosci. Environ. 411 pp.
Cornu S, Lucas Y, Ambrosi JP, Desjardins T.
1997a. A comparative study of the soil solu-
tion chemistry of two Amazonian forest soils
(Central Amazonia, Brazil). Hydrol. Earth
Syst. Sci. 1:313–24
Cornu S, Lucas Y, Ambrosi JP, Desjardins T.
1998. Transfer of dissolved Al, Fe and Si in
twoAmazonian forest environmentin Brazil.
Eur. J. Soil Sci. 3:77–84 (In French)
Cornu S, Lucas Y, Desjardins T, Nitsche S.
1995. Rapidit´e de la vitesse d’alt´eration des
min´eraux du sol en conditions ferrallitiques.
C. R. Acad. Sci. Paris (II)321:311–16 (In
French)
Cornu S, Luiz˜ao FJ, Rouiller J, Lucas Y. 1997b.
Comparative study of the litter decomposi-
tionand mineral elementrelease in twoAma-
zonian forest ecosystems: litter bag experi-
ments. Pedobiologia 41:456–71
Delvigne J. 1965. P´
edogen`
ese en Milieu Tro-
pical. Paris: ORSTOM. 177 pp. (In French)
Dennen WH, Norton HA. 1977. Geology and
geochemistry of bauxite deposits in the
LowerAmazon Basin. Econ. Geol. 72:82–89
Drever JI. 1994. Effect of plants on chemical
weathering rates. Geochim. Cosmochim.
Acta 58:2325–2332
Drever JI, Vance GF. 1994. Role of soil or-
ganic acids in mineral weathering processes.
In The Role of Organic Acids in Geological
Processes, ed. MD Lewan, ED Pittman. New
York: Springer-Verlag. 138 pp.
DreverJI,Zobrist J. 1992.Chemical weathering
of silicate rocks as a function of elevation
in the southern Swiss Alps. Geochim. Cos-
mochim. Acta 56:3209–16
Dubroeucq D, Volkoff B. 1998. From oxisols to
spodosols and histosols: evolution of the soil
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
160 LUCAS
mantles in the Rio Negro Basin (Amazonia).
Catena 32:245–80
DuchaufourP.1990.Laformationet l’´evolution
descomplexesorgano-min´eraux dans lessols
et leur rˆole dans la p´edogen`ese. Sci. Sol
28:273–84
Duchaufour P, Souchier B, eds. 1991. P´
edo-
logie: Sol, V´
eg´
etation, Environnement.Paris:
Masson. 289 pp.
EpsteinE. 1994. Theanomaly of siliconin plant
biology.Proc.Natl.Acad. Sci. USA91:11–17
Eyrolle F. 1994. La fraction collo¨
ıdale or-
ganique dans les processus de transport des
m´
etauxdans les eauxde surface: application
aux syst`
emes d’alt´
eration en milieu tropical
(Br´
esil). PhD thesis. Univ. Aix-Marseille III,
France. 334 pp.
olster H, de Las Salas G. 1976. Litter fall and
mineralizationin threetropical evergreenfor-
est stands. Acta Cient. Venez. 27:196–202
Food and Agricultural Organization. 1974. Soil
Map of the World. Rome/Paris: Food Agric.
Org. UNESCO
Fouillac C, Michard G. 1977. Une m´ethode
de simulation de l’´evolution des profils
d’alt´eration. Geochim. Cosmochim. Acta 41:
207–13
Franken W, Leopoldo PR. 1984. Hydrology of
catchment areas of central-Amazonian forest
streams. See Sioli 1984, pp. 501–19.
Fripiat J, Chaussidon J, Jelli A. 1971. Chimie
Physique des Ph´
enom`
enes de Surface. Paris:
Masson. 387 pp.
Fritsch JM. 1992. Les Effets du D´
efrichement
de la Forˆ
et Amazonienne et de la Mise en
Culture sur l’Hydrologie de Petits Bassins
Versants. Op´
eration Ecerex en Guyane
Franc¸aise. Paris: ORSTOM. 392 pp.
Fritz B, Tardy Y. 1973. Etude thermody-
namique du syst`eme gibbsite, quartz, kaoli-
nite, gaz carbonique: application `a la gen`ese
des podzols et des bauxites. Sci. G´
eol. Bull.
26(4):339–67
Furian S. 1994. Morphogen`
ese/P´
edogen`
ese en
milieu tropical humide de la Serra do Mar,
Br´
esil: contribution de l’alt´
eration `
a une dy-
namique active de glissement. PhD thesis.
Univ. Caen, France-Univ. S˜ao-Paulo, Brazil.
178 pp.
Gautam-Basak M, Proctor J. 1983. Micronutri-
ents, aluminum, silicon and ash in a leaf lit-
terfallfrom forests in Gunung Mulu National
Park, Sarawak. Malays. For. 46(2):224–32
enin JMR, Bourri´e G, Trolard F, Abdelmoula
M, Jaffrezic A, et al. 1998. Thermodynamic
equilibria in aqueous suspensions of syn-
thetic and natural Fe(II)-Fe(III) green rusts:
occurrences of the mineral in hydromorphic
soils. Environ. Sci. Technol. 32:1058–68
Giral-Kacmarcik S, Savin SM, Nahon DB,
Girard JP, Lucas Y, et al. 1998. Oxygen
isotope geochemistry of kaolinite in laterite-
forming processes, Manaus, Amazonas,
Brazil. Geochim. Cosmochim. Acta 62(11):
1865–79
Grimaldi C. 1987. Chemical composition and
transfer of water in two small watersheds
(tropical zone): study applied to the varia-
tions of (Al), (Si), pH. In Proc. Int. Workshop
Geochem. Monit. Represent. Basins (GE-
OMON), Prague, ed. B Moldan, T Paces ,
pp. 83–84, Prague, Czechoslov: Geol. Surv.
Grimaldi C. 1988. Origine de la composi-
tion chimique des eaux superficielles en mi-
lieu tropical humide. Exemple de deux pe-
tits bassins versants sous forˆet en Guyane
Fran¸caise. Sci. G´
eol. Bull. 41:247–62 (In
French)
Grimaldi C, Pedro G. 1996. Importance de
l’hydrolyse acide dans les syst`emes
edologiques des r´egions tropicales hu-
mides: rˆole de la forˆet et cons´equences sur
la gen`ese des sables blancs tropicaux. C. R.
Acad. Sci. Ser. B 323:483–92 (In French)
Grimaldi M, Sarrazin M, Chauvel A, Luiz˜ao
FJ, Nunes N, et al. 1993. Les effets de la
eforestation et des cultures sur la structure
des sols argileux d’Amazonie br´esilienne.
Agric. Cah. Et. Rech. Francoph. 2:36–47 (In
French)
Grossmann J, Udluft P. 1991. The extraction
of soil water by the suction-cup method: a
review. J. Soil Sci. 42:83–93
Herbillon AJ. 1994. Oxydes de fer, gibbsite,
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 161
oxydes de mangan`ese et min´eraux allo-
phaniques. In P´
edologie. 2. Constituants et
Propri´
et´
es du Sol, ed. M Bonneau, B Souch-
ier, pp. 17–45. Paris: Masson
Hinsinger P, Elsass F, Jaillard B, Robert M.
1993. Root-induced irreversible transforma-
tionof a trioctahedral micain the rhizosphere
of rape. J. Soil Sci. 44:535–45
Hinsinger P, Gilkes RJ. 1997. Dissolution of
phosphate rock in the rhizosphere of five
plant species grown in an acid, P-fixing min-
eral substrate. Geoderma 75:231–49
Irion G. 1984. Clay minerals of Amazonian
soils. See Sioli 1984, pp. 537–79
Jaillard B. 1987. Les Structures Rhizomorphes
Calcaires: Mod`
eles de R´
eorganisation des
Min´
eraux du Sol par les Racines. Montpel-
lier,France: Inst. Nat.Rech. Agron.Lab. Sci.
Sol. 221 pp.
Jaillard B, Guyon A, Maurin AS. 1991. Struc-
ture and composition of calcified roots, and
their identification in calcareous soils. Geo-
derma 50:197–210
Jaillard B, Ruiz L, Arvieu JC. 1996. pH map-
ping in transparent gel using color indicator
videodensitometry. Plant Soil 183:85–95
Jenny H. 1980. The Soil Resource: Origin and
Behavior. New York: Springer-Verlag. 377
pp.
Jones DL, Brassington DS. 1998. Organic acid
sorption in acid soils and its implications
in the rhizosphere. Eur. J. Soil Sci. 49:447–
55
Kelly EF, Chadwick OA, Hill CA. 1998. The
effect of plants on mineral weathering. Bio-
geochemistry 42:21–43
Klinge H. 1965. Podzol soils in the Amazon
Basin. J. Soil Sci. 16:96–103
Klinge H, Rodrigues WA. 1968. Litter produc-
tion in an area of Amazonian terra firme for-
est. Part II. Mineral nutrient content of the
litter. Amazoniana 1:303–10
KronbergBI, Fyfe WS, McKinnon BJ, Couston
JF,StilianidiFB,et al. 1982. Model for baux-
ite formation: Paragominas (Brazil). Chem.
Geol. 35:311–20
Lasaga AC. 1984. Chemical kinetics of water-
rock interactions. J. Geophys. Res. 89:4009–
25
Leneuf N. 1959. L’Alt´
eration des Granites
Calco-alcalins et des Granodiorites en Cˆ
ote
d’Ivoire Foresti`
ere et les Sols qui en sont
D´
eriv´
es. PhD thesis, Univ. Paris
Leo RF, Barghoorn ES. 1976. Silica in the bio-
sphere. Acta Cient. Venez. 27:231–34
Likens GE, Bormann FH, Pierce RS, Eaton
JS, Johnson NM. 1977. Biogeochemistry of
a Forested Ecosystem. New York: Springer-
Verlag. 146 pp.
Litaor MI. 1988. Review of soil solution sam-
plers. Water Resour. Res. 24:727–33
Lovering TS. 1959. Significance of accumula-
torplants in rockweathering. Bull. Geol.Soc.
Am. 70:781–800
Lucas Y, Boulet R, Veillon L. 1987. Syst`emes
sols ferrallitiques—podzols en r´egion ama-
zonienne. In Podzols et Podzolisation, ed. D
Righi, A Chauvel, pp. 53–65. Plaisir/Paris:
Assoc. Fr. d’Etude Sols—Inst. Natl. Rech.
Agron.
Lucas Y, Chauvel A, Ambrosi JP. 1986. Proce-
sses of aluminium and iron accumulation in
latosols developed on quartz rich sediments
from Central Amazonia (Manaus, Brazil). In
Geochemistry and Mineral Formation in the
Earth Surface, ed. R Rodriguez-Clemente, Y
Tardy, pp. 289–99. Madrid/Paris: Cons. Sup.
Cien.-Cent. Natl. Rech. Sci.
Lucas Y, Chauvel A, Boulet R, Ranzani G, Sca-
tolini F. 1984. Transi¸ao latossolos-podzois
sobre a forma¸ao Barreias na regi˜ao de
Manaus, Amazˆonia. Rev. Bras. Cienc. Solo
8:325–35 (In Portuguese)
Lucas Y, Kobilsek B, Chauvel A. 1989. Struc-
ture, genesis and present evolution of amazo-
nian bauxites developed on sediments. Trav.
Int. Comm. Study Bauxites, Alumina and
Alum. 19:81–94
Lucas Y, Luiz˜ao FJ, Chauvel A, Rouiller J,
Nahon D. 1993. The relation between bio-
logical activity of the rainforest and mineral
compositionof thesoils. Science 260:521–23
Lucas Y, Nahon D, Cornu S, Eyrolle F. 1996.
Gen`ese et fonctionnement des sols en milieu
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
162 LUCAS
´equatorial. C. R. Acad. Sci. Paris (IIa)322:1–
16
Mad´eB,Cl´ement A, Fritz B. 1994. Modeling
mineral/solution interactions: the thermody-
namic and kinetic code KINDISP. Comput.
Geosci. 20:1347–63
Marschner H. 1995. Mineral Nutrition in
Higher Plants. London: Academic. 674 pp.
2nd ed.
Martinelli LA, Lucas Y, Cerri CC, Victoria RL.
1996. Conf. Int. Um. Estrategia Lat. Am.
Amaz. S˜
ao Paulo, ed. C Pavan, pp. 209–26.
ao Paulo, Brazil: Univ. Estad. Paul.
Mengel K, Kirkby EA. 1987. Principles of Pl-
ant Nutrition. Bern: Int. Potash Inst. 4th ed.
MeunierJD, ColinF,AlarconC. 1999.Biogenic
storage of silica in soils. Geology 27:835–38
Moulton KL, Berner RA. 1998. Quantification
of the effect of plants on weathering: studies
in Iceland. Geology 26:895–98
Novikoff A. 1974. L’Alt´
eration de Roches dans
le Massif du Chaillu (R´
epublique Populaire
du Congo). Strasbourg: Univ. Louis Pasteur.
298 pp.
Nye P. 1955. Some soil-forming processes in
the humid tropics. IV. The action of soil
fauna. J. Soil Sci. 6:73–83
PaulEA,Clark FE.1989. SoilMicrobiologyand
Biochemistry. San Diego, CA: Academic.
275 pp.
Pedro G. 1966. Essai sur la caract´erisation
eochimiquedes diff´erents processuszonaux
esultants de l’alt´eration des roches superfi-
cielles. C. R. Acad. Sci. Ser. D 262:1828–31
Raven JM. 1983. The transport and function of
silica in plants. Biol. Rev. 58:179–207
Righi D, Bravard S, Chauvel A, Ranger J,
Robert M. 1990. In situ study of soil pro-
cesses in an oxisol-spodosol sequence of
Amazonia (Brazil). Soil Sci. 150(1):438–45
Robert M, Berthelin J. 1986. Role of biologi-
cal and biochemical factors in soil mineral
weathering. In Interactions of Soil Minerals
withNatural Organics andMicrobes, ed. PM
Huang, M Schnitzer, pp. 453–96. Madison,
WI: Soil Sci. Soc. Am.
Robert M, Chenu C. 1992. Interactions be-
tween soil minerals and microorganisms. In
Soil Biochemistry, ed. G Stotzky, JM Bollag,
7:307–404. New York: Marcel Dekker
Robert M, Razzaghe MK, Vicente MA, Veneau
G. 1979. Rˆole du facteur biochimique dans
l’alt´eration des min´eraux silicat´es. Sci. Sol
2-3:153–74
Roche MA. 1990. Hydrologie et ´erosion de
l’´ecosyst`eme forestier guyanais. A. Etude
du milieu naturel. In Mise en Valeur de
l’Ecosyst`
eme Forestier Guyanais: Op´
era-
tion ECEREX, ed. JM Sarrailh, pp. 47–64.
Paris: Inst. Natl. Rech. Agron.
Rose AW, Kato T, Machesky ML. 1993. The
significance of biogenic element cycling in
ancient tropical soils. Chem. Geol. 107:401–
3
Rozanski K, Araguas-Araguas L, Plata Bedmar
A, Franken W, Tancredi AC, Tundis Vital A.
1991. Water movement in the Amazon soils
traced by the means of hydrogen isotopes.
In Proc. Int. Symp. Stable Isot. Plant Nutr.,
Soil Fertil. Environ. Stud., IAEA-SM-313,
pp. 147–48. Vienna: Int. At. Energy Agency
Schwartz D. 1988. Some podzols on Bateke
sands and their origins, People’s Republic of
Congo. Geoderma 43:229–47
Sieffermann G, Millot G. 1969. Equatorial and
tropical weathering of recent basalts from
Cameroon: allophanes, halloysite, metahal-
loysite, kaolinite and gibbsite. In Proc. Int.
Clay Conf., Tokyo, pp. 417–30. Jerusalem:
Israel Univ. Press
Sieffermann GR. 1990. Origin of iron-
carbonatelayers inTertiarycoastalsediments
of central Kalimantan Province (Borneo), In-
donesia. Spec. Pub. Int. Assoc. Sediment.
11:139–46
Sioli H, ed. 1984. The Amazon: Limnology
and Landscape Ecology of a Mighty Tropi-
cal River and its Basin: Monogaphiae Bio-
logica. Dordrecht, Neth: Junk
Soler JM, Lasaga AC. 1996. A mass transfer
model of bauxite formation. Geochim. Cos-
mochim. Acta 60:4913–31
Sombroek WG. 1966. Amazon Soils: A Re-
connaissance of the Soils of the Brazilian
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
P1: FXZ/LRV P2: FXZ/FXB QC: FXZ
March 17, 2001 13:23 Annual Reviews AR125-06
THE ROLE OF PLANTS IN WEATHERING 163
Amazon Region. Wageningen, The Nether-
lands: PUDOC. 300 pp.
StallardRF.1985.River chemistry,geology,ge-
omorphology and soils in the Amazon and
Orinoco basins. In The Chemistry of Weath-
ering. NATC ASI Ser. 149, ed. JI Drever, pp.
293-316. Dordrecht, Holl: Reidel
Stallard RF, Edmond JM. 1983. Geochemistry
of the Amazon. 2. The influence of geology
andweathering environmenton thedissolved
load. J. Geophys. Res. 88:9671–88
Stevenson FJ, Vance GF. 1989. Naturally oc-
curringaluminum-organic complexes.InThe
Environmental Chemistry of Aluminum, ed.
GSposito, pp.117–45. BocaRaton, FL: CRC
Tan KH. 1991. Differences in decomposition of
carbonate and silicate minerals as influenced
by humic acids. In Diversity of Environmen-
talBiogeochemistry, ed. J Berthelin, 341–42.
Amsterdam: Elsevier
Tardy Y, Nahon D. 1985. Geochemistry of lat-
erites, stability of Al-goethite, Al-hematite,
and Fe3+kaolinites in bauxites and ferri-
cretes: an approach to the mechanism of
concretion formation. Am. J. Sci. 285:865–
903
Tardy Y, Novikoff A. 1988. Activit´e de l’eau
et d´eplacement des ´equilibres gibbsite-
kaolinite dans les profils lat´eritiques. C. R.
Acad. Sci. Paris (II)306:39–44
Taylor RM. 1988. Proposed mechanism for the
formation of soluble Si-Al and Fe(III)-Al
hydroxy complexes in soils. Geoderma
42:65–77
Turenne JF. 1977. Mode d’Humification
et Diff´
erenciation Podzolique dans deux
Topos´
equences Guyanaises: M´
em. No 84.
Paris: ORSTOM. 173 pp.
Veillon L. 1990. Sols Ferrallitiques et Podzols
en Guyane Septentrionale: Relations entre
Syst`
emes de Transformations P´
edologiques
et ´
Evolution Historique d’un Milieu Tropical
Humide et Forestier. PhD thesis. Univ. Paris
VI, 191 pp.
Velbel MA. 1985. Geochemical mass balances
and weathering rates in forested watersheds
of the southern Blue Ridge. Am. J. Sci.
285:904–30
Volkoff B. 1985. Organisations r´egionales de la
couverture p´edologique du Br´esil. Chronolo-
gie des diff´erenciations. Cah. ORSTOM S´
er.
P´
edol. 21(4):225–36
Watteau F, Villemin G, Burtin G, Ghanbaja J,
Toutain F. 2001. Biogenic silica tracing from
plant to soil in a forest ecosystem. In Phy-
toliths: Applications in Earth Science and
Human History, ed. JD Meunier, F Colin.
Rotterdam, Neth: Balkema. In press
Webb LJ, Tracey JG, Williams WT, Lance GN.
1969. The pattern of mineral return in leaf
litter of three subtropical Australian forests.
Aust. For. 93:99–110
Wright VP, Platt NH, Mariott SB, Beck VH.
1995. A classification of rhizogenic (root-
formed) calcretes, with examples from the
Upper Jurassic-Lower Cretaceous of Spain
and Upper Cretaceous of southern France.
Sediment. Geol. 100:143–58
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
Annual Review of Earth and Planetary Science
Volume 29, 2001
CONTENTS
BREAKTHROUGHS IN OUR KNOWLEDGE AND
UNDERSTANDING OF THE EARTH AND PLANETS, G Schubert 1
HUMAN IMPACTS ON ATMOSPHERIC CHEMISTY, PJ Crutzen, J
Lelieveld 17
INNER-CORE ANISOTROPY AND ROTATION, Jeroen Tromp 47
PARTIAL MELTING EXPERIMENTS ON PERIDOTITE AND
ORIGIN OF MID-OCEAN RIDGE BASALT, Ikuo Kushiro 71
TECTONIC EVOLUTION OF THE JAPANESE ISLAND ARC
SYSTEM, Asahiko Taira 109
THE ROLE OF PLANTS IN CONTROLLING RATES AND
PRODUCTS OF WEATHERING: Importance of Biological Pumping, Y
L
ucas 135
RUSTY RELICS OF EARTH HISTORY: Iron(III) Oxides, Isotopes, and
Surficial Environments, Crayton Yapp 165
USING SPRINGS TO STUDY GROUNDWATER FLOW AND
ACTIVE GEOLOGIC PROCESSES, Michael Manga 201
GROUND PENETRATING RADAR FOR ENVIRONMENTAL
APPLICATIONS, Rosemary Knight 229
DATING MODERN DELTAS: Progress, Problems, and Prognostics,
Jean-Daniel Stanley 257
RHEOLOGICAL PROPERTIES OF WATER ICE--APPLICATIONS TO
SATELLITES OF THE OUTER PLANETS, WB Durham, LA Stern 295
THE LATE ORDOVICIAN MASS EXTINCTION, Peter M Sheehan 331
HYDROGEN IN THE DEEP EARTH, Quentin Williams, Russell J.
Hemley 365
PHYSICS OF PARTIALLY SATURATED POROUS MEDIA: Residual
Saturation and Seismic-Wave Propagation, Xun Li, Lirong Zhong, Laura
J Pyrak-Nolte 419
RESPONSE OF LATE CARBONIFEROUS AND EARLY PERMIAN
PLANT COMMUNITIES TO CLIMATE CHANGE, William A
DiMichele, Hermann W Pfefferkorn, Robert A Gastaldo 461
GIANT DIKE SWARMS: Earth, Venus, and Mars, RE Ernst, EB
Grosfils, D Mège 489
THE CARBON BUDGET IN SOILS, Ronald Amundson 535
CONTINUOUS FREE OSCILLATIONS: Atmosphere-Solid Earth
Coupling, Toshiro Tanimoto 563
Annu. Rev. Earth Planet. Sci. 2001.29:135-163. Downloaded from arjournals.annualreviews.org
by University of British Columbia Library on 03/03/10. For personal use only.
... By these and other activities, roots and associated soil biota also strongly affect soil aggregate formation and turnover (Six et al., 2004), producing legacy effects on soil biota and plant communities. Roots also affect mineral weathering, mobilize and redistribute plant, nutrients in the soil profile (Jobbágy and Jackson, 2001;Lucas, 2001). These phenomena likely affect not only the plants that cause them but also generations of plants that follow. ...
... This in interaction with climate and topography affects the intensity of soil erosion (Yair, 1995;Le Bayon and Binet, 2001;Jouquet et al., 2006;Román-Sánchez et al., 2019) and deposition, which again alter soil properties. On an even longer temporal and large spatial scale, soil-forming processes, affected by vegetation and soil biota (Jobbágy and Jackson, 2001;Lucas, 2001), affect the balance between the acquisition of new nutrients by weathering and their loss from the system. This affects nutrient availability, which leads to changes in vegetation, plant species traits, nutrient turnover, and how organic matter is stored in soil (Vitousek, 2004;Kuneš et al., 2011;Vindušková et al., 2019). ...
... For example, dissolved silicon can improve SOC stability by inhibiting the hydrolysis and polymerisation of Fe 3+ and reducing the transformation of low-crystalline minerals into the crystalline Fe 3+ phase (Hiemstra et al., 2007;Jones et al., 2009;Pokrovski et al., 2003), moreover. Silicon and aluminium in soil solution can stabilise SOC by combining with Si-Al secondary clay minerals and their nano-, micro-, and macroaggregates (Lucas, 2001;Song et al., 2018). Additionally, investigations conducted on farmland soil have demonstrated that the utilisation of silicon fertiliser can enhance the activity of the soil bacterial community and the concentrations of microbial organic carbon by regulating soil pH . ...
... Notably, soil inorganic and organic environments of forests are not independent . Plants typically accumulate large quantities of potassium (K) in their tissues and K is primarily cycled via mineral weathering, but also by biological pumps during leaf litter decomposition (Lucas 2001;Castro and Fernandez-Nu ñez 2014). However, K is easily leached from leaves and plant tissues (Aber and Melillo 2001;Castro and Fernandez-Nu ñez 2014). ...
Article
Full-text available
Context. Studies of afforestation have traditionally neglected the influences of plant microhabitats on the growth and carbon sink capacities of planted forests. Aims. We investigated the potential mechanisms related to the relationship of afforestation elevation to soil organic carbon density (SOCD). Methods. The carbon density of three plantation ecosystems and barren land soils were evaluated at two elevations in the Donglingshan Mountains of Beijing, with structural equation modelling and variation partitioning analyses used to identify the environmental factors that influenced the carbon densities of plantation ecosystems. Key results. Afforestation elevation was related to the vegetation phenology of plantation forests. Specifically, growth periods at higher elevations were delayed relative to those at lower elevations, while different growth periods affected growth rate of diameter at breast height (R DBH), in addition to the carbon and nitrogen contents of ground surface litters. Consequently, lower elevation afforestation reduced the carbon sink capacity of coniferous plantation ecosystems in the study area. Lower plantation elevations were associated with significantly reduced R DBH values of Pinus tabuliformis. Further, biomass carbon density (BCD) and SOCD of Larix principis-rupprechtii plantations were significantly lower due to decreased elevations. Soil nitrogen concentrations, litter nitrogen density (LND), soil phosphorus concentrations, and BCD were the primary drivers of plantation SOCD. Conclusions. Overall, different plantation elevations were associated with different vegetation phenologies and R DBH values, which further affected LND and BCD, thereby ultimately affecting variation of SOCD. Implications. This study provides important insights into the selection of afforestation plots to maximise plantation carbon sequestration capacities.
... Bauxitization is favoured by abundant vegetation because root systems favour effective water percolation, organic matter creates acidic conditions favouring the dissolution and transport of Fe and Al, and plants also remove silica from soils (Bárdossy & Aleva, 1990;Lucas, 2001;Petersen, 1971;Power & Loh, 2010). Organic acids lower environmental pH and accelerate mineral dissolution rates (Drever & Stillings, 1997). ...
Article
Lithology plays a fundamental role in rock weathering and erosion, and in landscape evolution. When weathering‐ and erosion‐prone lithologies are protected from erosion by more resilient rock types (e.g., quartzites and banded iron formations) unusual weathering products result. At the Southern Espinhaço Range, Minas Gerais, Brazil, bauxitic weathering profiles are found in a unique geomorphological–lithological–climatic setting. Resistant quartzites acted as a barrier against erosion of interbedded hematite‐phyllite lenses, channelling solution flows and facilitating the formation of deep weathering profiles. The long‐term exposure of the hematite‐phyllites under alternating wet and dry tropical climates favoured widespread bauxitization. Here we investigate the geochemical, mineralogical, geochronological and micromorphological signatures of scaffolded bauxites in order to reconstruct their evolutionary history. Our data suggest that recurrent aluminium and iron mobilization within the profiles were mainly driven by mineral dissolution‐reprecipitation mediated by bioturbation and the influx of vegetation‐derived organic species. (U–Th)/He geochronology of Al‐goethite reveals that bauxitization started at least since the Lower Miocene, with important intensification of weathering in the Upper Miocene and Lower Pleistocene. The adjacent resilient quartzites acted as scaffolds for bauxitization and supported the preservation of more erodible weathering profiles developed over phyllites. Surface waters that could not infiltrate into the impermeable adjacent quartzites preferentially infiltrated into the more weathereable phyllites, enhancing their porosity and permeability, further enhancing weathering. The evolutionary history of Southern Espinhaço Range bauxites suggests a new model of bauxitization in ancient land surfaces evolution underlain by quartzites, where erosion‐prone lithologies are scaffolded by resilient quartzites and survive long‐term weathering with minimum erosion, producing bauxites.
... The multiplier by which plants increase CO 2 drawdown as a global average has proven challenging to measure; there are many feedbacks, and plants may inhibit weathering in some settings (Volk, 1987;Pagani et al., 2009;Beerling et al., 2012;Brantley et al., 2012;Doughty et al., 2014;Lawrence et al., 2014;Quirk et al., 2014;D'Antonio et al., 2020). However, results from field, glasshouse, and modeling studies indicate that plants can multiply weathering rates by 2-109 or more, especially in wet and warm climates, on acidic mafic and ultramafic rocks, and in areas with high topographic relief (Schwartzman & Volk, 1989;Drever, 1994;Cochran & Berner, 1996;Moulton et al., 2000;Lucas, 2001;Berner, 2004;Taylor et al., 2012;Johnson et al., 2014;Lenton et al., 2018;Ibarra et al., 2019;Perez-Fodich & Derry, 2019). ...
Article
Full-text available
Many tree genera in the Malesian uplands have Southern Hemisphere origins, often supported by austral fossil records. Weathering the vast bedrock exposures in the everwet Malesian tropics may have consumed sufficient atmospheric CO2 to contribute significantly to global cooling over the past 15 Myr. However, there has been no discussion of how the distinctive regional tree assemblages may have enhanced weathering and contributed to this process. We postulate that Gondwanan‐sourced tree lineages that can dominate higher‐elevation forests played an overlooked role in the Neogene CO2 drawdown that led to the Ice Ages and the current, now‐precarious climate state. Moreover, several historically abundant conifers in Araucariaceae and Podocarpaceae are likely to have made an outsized contribution through soil acidification that increases weathering. If the widespread destruction of Malesian lowland forests continues to spread into the uplands, the losses will threaten unique austral plant assemblages and, if our hypothesis is correct, a carbon sequestration engine that could contribute to cooler planetary conditions far into the future. Immediate effects include the spread of heat islands, significant losses of biomass carbon and forest‐dependent biodiversity, erosion of watershed values, and the destruction of tens of millions of years of evolutionary history.
... The formation of secondary minerals such as clay minerals and aluminum hydroxide is among other factors controlled by biogenic activity since organic acids and an acidity increase by elevated organic-derived CO 2 contents accelerate dissolution rates of primary minerals (see e.g., Lucas, 2001;Lawrence et al., 2014). This effect needs to be considered for the organic-rich and acidic subsurface of NA (see Bernhard et al., 2018). ...
Article
Full-text available
Subsurface fluid pathways and the climate-dependent infiltration of fluids into the subsurface jointly control the intensity and depth of mineral weathering reactions. The products of these weathering reactions (secondary minerals), such as Fe(III) oxyhydroxides and clay minerals, in turn exert a control on the subsurface fluid flow and hence on the development of weathering profiles. We explored the dependence of mineral transformations on climate during the weathering of granitic rocks in two 6 m deep weathering profiles in Mediterranean and humid climate zones along the Chilean Coastal Cordillera. We used geochemical and mineralogical methods such as (micro-) X-ray fluorescence (μ-XRF and XRF), oxalate and dithionite extractions, X-ray diffraction (XRD), and electron microprobe (EMP) mapping to elucidate the transformations involved during weathering. In the profile of the Mediterranean climate zone, we found a low weathering intensity affecting the profile down to 6 m depth. In the profile of the humid climate zone, we found a high weathering intensity. Based on our results, we propose mechanisms that can intensify the progression of weathering to depth. The most important is weathering-induced fracturing (WIF) by Fe(II) oxidation in biotite and precipitation of Fe(III) oxyhydroxides and by the swelling of interstratified smectitic clay minerals that promotes the formation of fluid pathways. We also propose mechanisms that mitigate the development of a deep weathering zone, like the precipitation of secondary minerals (e.g., clay minerals) and amorphous phases that can impede the subsurface fluid flow. We conclude that the depth and intensity of primary mineral weathering in the profile of the Mediterranean climate zone is significantly controlled by WIF. It generates a surface–subsurface connectivity that allows fluid infiltration to great depth and hence promotes a deep weathering zone. Moreover, the water supply to the subsurface is limited in the Mediterranean climate, and thus, most of the weathering profile is generally characterized by a low weathering intensity. The depth and intensity of weathering processes in the profile of the humid climate zone, on the other hand, are controlled by an intense formation of secondary minerals in the upper section of the weathering profile. This intense formation arises from pronounced dissolution of primary minerals due to the high water infiltration (high precipitation rate) into the subsurface. The secondary minerals, in turn, impede the infiltration of fluids to great depth and thus mitigate the intensity of primary mineral weathering at depth. These two settings illustrate that the depth and intensity of primary mineral weathering in the upper regolith are controlled by positive and negative feedbacks between the formation of secondary minerals and the infiltration of fluids.
Article
Full-text available
Two major barriers hinder the holistic understanding of subsurface critical zone (CZ) evolution and its impacts: (a) an inability to measure, define, and share information and (b) a societal structure that inhibits inclusivity and creativity. In contrast to the aboveground portion of the CZ, which is visible and measurable, the bottom boundary is difficult to access and quantify. In the context of these barriers, we aim to expand the spatial reach of the CZ by highlighting existing and effective tools for research as well as the “human reach” of CZ science by expanding who performs such science and who it benefits. We do so by exploring the diversity of vocabularies and techniques used in relevant disciplines, defining terminology, and prioritizing research questions that can be addressed. Specifically, we explore geochemical, geomorphological, geophysical, and ecological measurements and modeling tools to estimate CZ base and thickness. We also outline the importance of and approaches to developing a diverse CZ workforce that looks like and harnesses the creativity of the society it serves, addressing historical legacies of exclusion. Looking forward, we suggest that to grow CZ science, we must broaden the physical spaces studied and their relationships with inhabitants, measure the “deep” CZ and make data accessible, and address the bottlenecks of scaling and data‐model integration. What is needed—and what we have tried to outline—are common and fundamental structures that can be applied anywhere and used by the diversity of researchers involved in investigating and recording CZ processes from a myriad of perspectives.
Article
Full-text available
Terrestrial cosmogenic nuclides (TCN) are widely employed to infer denudation rates in mountainous landscapes. The calculation of an inferred denudation rate (Dinf) from TCN concentrations is typically performed under the assumptions that denudation rates were steady during TCN accumulation and that soil chemical weathering negligibly impacted soil mineral abundances. In many landscapes, however, denudation rates were not steady and soil composition was significantly impacted by chemical weathering, which complicates interpretation of TCN concentrations. We present a landscape evolution model that computes transient changes in topography, soil thickness, soil mineralogy, and soil TCN concentrations. We used this model to investigate TCN responses in transient landscapes by imposing idealized perturbations in tectonically (rock uplift rate) and climatically sensitive parameters (soil production efficiency, hillslope transport efficiency, and mineral dissolution rate) on initially steady‐state landscapes. These experiments revealed key insights about TCN responses in transient landscapes. (a) Accounting for soil chemical erosion is necessary to accurately calculate Dinf. (b) Responses of Dinf to tectonic perturbations differ from those to climatic perturbations, suggesting that spatial and temporal patterns in Dinf are signatures of perturbation type and magnitude. (c) If soil chemical erosion is accounted for, basin‐averaged Dinf inferred from TCN in stream sediment closely tracks actual basin‐averaged denudation rate, showing that Dinf is a reasonable proxy for actual denudation rate, even in many transient landscapes. (d) Response times of Dinf to perturbations increase with hillslope length, implying that response times should be sensitive to the climatic, biological, and lithologic processes that control hillslope length.
Article
Biochemical decomposition of primary minerals was investigated by shaking calcite, hornblende and microcline with fulvic (FA) or humic acid (HA) at pH 7.0 and 2.5 for 0 to 885 hrs. Silicon, Al, Fe, K and Ca released were determined in the extracts by atomic absorption spectroscopy. Changes in mineral structure were studied by X-ray diffraction analysis, petrographic and scanning electron microscopy. The results indicated that the dissolution of Si and Al from the silicate minerals yielded a parabolic curve. The dissolution of Ca and Fe from hornblende, and K from microcline was also curvilinear in regression. -from Author
Article
Membranes define the outer boundary of the living protoplast and the internal compartmentation of plant cells. From a structural point of view, membranes consist of a lipid bilayer and proteins essential for functions such as solute transport, signal transduction, and numerous metabolic reactions. While membranes can represent a significant barrier to the free movement of many solutes, those with sufficient lipid solubility may move across membranes by dissolving into the lipid bilayer. However, selective membrane transport is generally observed for hydrophilic solutes such as mineral nutrients and cell metabolites. Such selective transport requires an input of metabolic energy, and in plants this occurs via the production of proton electrochemical gradients across the membrane by substrate- (ex. ATP, PPi) driven proton pumps. Selective solute transport is then mediated by membrane-associated secondary transport systems which utilize the proton electrochemical gradient to drive the transport process. This review of membrane structure and transport system function provides a background for a further examination of herbicide interactions with plant membranes.
Article
Relief, geology, climate, vegetation allow dividing Brazil into 4 main regions. In each region, areas can be characterized by distinct pedologic types (latosols, lateritic or red-yellow podzolic soils, fersiallitic soils, cambisols). The latosol areas are related to 2 pediplaned surfaces with old pedologic covers. They quite differ from other areas that show a young topography and recent pedologic cover. -from English summary
Article
Cet article de synthèse sur les cycles biogéochimiques des éléments minéraux dans les écosystèmes forestiers tempérés tente de faire le point des travaux de recherche effectués, en essayant tout particulièrement de saisir l'incidence des cycles biogéochimiques sur le mode de dissolution des minéraux des sols, les préconcentrations géochimiques et les formes de transfert des éléments minéraux dans les rivières.
Chapter
Interest in the hydrological cycle of the Amazon Basin has recently increased due to the possibility of change in the water balance as a result of intensive deforestation. In this chapter we will demonstrate our work on water and nutrient cycling in selected catchment areas in the central Amazon Basin near Manaus. This research work began in 1976 with the installation of the first catchment area in the forest reserve Ducke, a watershed called ‘Barro Branco’, and continued in 1978 with a much larger undisturbed catchment area in the forest reserve Km 60 of the INPA, the watershed called ‘Bacia Modelo’. In 1980 two other experimental watersheds were installed in the neighbourhood of the ‘Bacia Modelo’; one of them is used as a buffer watershed and the other as an experimental watershed mainly for forestry.