Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Chemical Geology 347 (2013) 20–33

Contents lists available at SciVerse ScienceDirect

Chemical Geology
journal homepage: www.elsevier.com/locate/chemgeo

Isotopic evidence for dolomite formation in soils


J.L. Díaz-Hernández a,⁎, A. Sánchez-Navas b, c, E. Reyes c
a
IFAPA Camino de Purchil, Área de Recursos Naturales, Consejería de Agricultura, Pesca y Medio Ambiente, Junta de Andalucía, Apartado 2027, 18080 Granada, Spain
b
Departamento de Mineralogía y Petrología, Universidad de Granada, Campus Fuentenueva, Granada, Spain
c
Instituto Andaluz de Ciencias de la Tierra IACT (CSIC–UGR), Avda. de las Palmeras, 18100 Armilla, Granada, Spain

a r t i c l e i n f o a b s t r a c t

Article history: Dolomite formation in soils constitute a particular challenge because of: 1) scant magnesium content in
Received 19 November 2012 continental environments as opposed to the marine medium, 2) the kinetic problem related to the incorporation
Received in revised form 25 March 2013 of magnesium into the carbonate, and 3) the unknown role of soil dolomites in the global carbon cycle.
Accepted 28 March 2013
Pedogenic dolomite formed at deeper soil levels (subsoil) before the development of petrocalcic horizon
Available online 6 April 2013
barriers was investigated in a semiarid region of SE Spain (Guadix–Baza basin). Mineralogical characterization,
Editor: U. Brand textural relationships and isotopic data concerning soil dolomite, together with the results of a precipitation
experiment, provided fuller knowledge of the processes and conditions governing neoformation of dolomite in
Keywords: these soils.
Carbon–oxygen isotopes In the study case, dolomite enrichment occurs beyond the limit of major biological activity, which coincides
Dolomite problem with the rooting depth of native perennial plants in the semiarid soils studied. Textural studies reveal the
Inorganic carbon corrosion of inherited dolomite crystals in the upper soil horizons and the formation of dolomite in depth
Pedogenic processes in relation to a clayey material, composed mainly of smectites. Stable isotope distribution in dolomites
Semiarid soils
throughout the profiles indicates a fractionation with depth. This is explained by the formation of dolomites
Soil carbon storage
after the dissolution of the pedogenic calcite. The calcite detected in the subsoil is interpreted here as a
precursor of the neoformed dolomites that transport the isotopic signal associated with biological activity
of soils to deeper layers. Dolomite formation appears to be favoured by the presence of clay minerals in
the precipitation media. Clays retain water during evapotranspiration stages, which drastically change the
transport properties of the media and promote the incorporation of Mg into the structure of the neoformed
Ca,Mg-carbonate. As confirmed by laboratory experiments, diffusion-controlled crystal-growth processes
lead to the formation a precursory “protodolomite” with disordered Ca,Mg distribution from a fluid locally
supersaturated in dolomite.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction dissolution–precipitation in soils. In fact, pedogenesis begins with


intense biological activity in the weathering zone (Wright and Tucker,
The genesis of calcic and petrocalcic horizons in soils occurs as 1991; Wright, 1992; Cerling and Quade, 1993). Carbonic acid produced
part of carbonate dissolution–precipitation cycles (Gile et al., 1966; in association with biological activity (mainly roots) is the main acid
Salomons and Mook, 1976). Detrital limestone and other calcareous involved in the dissolution of calcium carbonate. The leaching depth
materials undergoing decarbonation constitute a source for secondary of the dissolved carbonate augments with increasing mean annual
carbonate accumulation in soils. Biogenic and atmogenic CO2 are alter- precipitation and temperature (Arkley, 1963; Stevenson et al., 2005).
native sources for the carbon fixation in the soil under physico-chemical Carbonate precipitation at the deeper layers results largely from the
conditions stabilizing the carbonate ion. The pedogenic environment is declining PCO2 in the soil and from the higher concentrations of solutes
strongly influenced by biological activity: root mats, in addition to in the soil solution due to evapotranspiration. In this sense, the carbon-
associated lichens and microorganisms, contribute directly to mineral ation process is favoured in arid–semiarid regimes. Large amounts of
soil inorganic carbon, stored in the form of calcium carbonate, have
been reported in cultivated mollisols under moderately cold tempera-
Abbreviations: a.s.l., above sea level; AEM, analytical electron microscopy; apfu, tures (5–5.5 °C), and with a mean annual precipitation between 581
atoms per formula unit; BSE, back scattered electrons; DIC, dissolved inorganic carbon; and 397 mm (Mikhailova and Post, 2006; Mikhailova et al., 2009),
EDX, energy-dispersion X-ray; FESEM, field-emission scanning electron microscope; suggesting an increase in the carbonation process with agricultural
MCL, Meteoric Calcite Line; MDL, Meteoric Dolomite Line; OA, oriented aggregated;
TEM, transmission electron microscopy; XRD, X-ray diffraction.
practices.
⁎ Corresponding author. Isotopic relationships found in pedogenic carbonates in arid–semiarid
E-mail address: josel.diaz@juntadeandalucia.es (J.L. Díaz-Hernández). ecosystems indicate that carbonate dissolution–precipitation cycles in a

0009-2541/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.chemgeo.2013.03.018
J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33 21

pedogenic environment provoke changes in the isotopic composition of with protodolomites, has recently been documented in soils (Whipkey
the precipitated carbonate, by inducing significant inputs of biogenic and Hayob, 2008). 3) The soils of the study area are particularly rich in
carbon (Cerling, 1984). Biogenic carbon carries a unique isotopic signa- inherited dolomites and, therefore, could be suitable for the develop-
ture, and the biological fractionation processes deplete 13C to negative ment of pedogenic dolomites in addition to calcites. 4) Periodic mois-
values (Salomons et al., 1977; Magaritz and Amiel, 1980; Magaritz ture movement to depths below solum horizons serves as a regulating
et al., 1981). Therefore, the pedogenic carbonate will usually have more factor in maintaining soluble Mg at levels favourable to the precipitation
negative 13C than detrital carbonates. of low-magnesium calcites (St. Arnaud and Herbillon, 1973), and occa-
Modern dolomites are formed only in certain environments such as sionally also leads to the formation of dolomites. 5) The few studies
meteoric, marine, hypersaline, subsurface brines, and hydrothermal available concerning pedogenic dolomites (St. Arnaud, 1979; Botha
(Hardie, 1987). The precipitation of dolomite and dolomitization of and Hughes, 1992; Capo et al., 2000) point to the occurrence of pedo-
previous Ca-rich carbonates are among the least-understood problems genic dolomites in depth.
encountered in the geochemistry of carbonates: e.g. surface seawater The main objective of this research is to demonstrate the forma-
is strongly oversaturated in dolomite, but little evidence is available tion of dolomite in soils. For this, we firstly studied compositional,
on the widespread dolomite precipitation in modern, open-marine mineralogical, and textural variations throughout the soil profiles,
sediments. To explain the “dolomite problem”, models have emerged finding that dolomite was dissolved in superficial horizons and
ascribing microbial activity to the mediation of dolomite precipitation. precipitated at greater depths than was the pedogenic calcite. Also,
Various authors have focused their attention on the role played by the we analysed the isotopes in calcite and dolomite of the soil samples.
bio-mediation that leads to CO2 enrichment and raises alkalinity, This second stage revealed that the biogenic isotopic signature of
and presumably drives dolomite formation in Mg-rich environments the pedogenic calcite was transferred to a neoformed dolomite
(e.g. Wright, 1999). In any case, the occurrence of dolomite in the through the dissolution–precipitation cycles that affected the soil
geochemical systems of the earth's surface is often strongly controlled carbonates. Finally, we discuss the physico-chemical constraints on
by reaction kinetics (Arvidson and Mackenzie, 1999; Deelman, 2005) the dolomite formation in subsoil. According to the model proposed
and/or mass-transport processes. for the precipitation of dolomite: 1) inherited dolomite constitutes
The formation of dolomite in soils is rarely reported, its origin being the main source for the Mg 2+; 2) hydrodynamics in the soil studied
related to parent-material inheritance, to atmospheric Mg2+ wet controls moisture and solute transport properties in the subsoil; and
deposition, and to the addition of dolomite contained in atmospheric 3) the incorporation of magnesium to the carbonate structure in
dust (Ghebre-Egziabhier and St. Arnaud, 1983; Ming, 2002; Goddard Mg-rich confined porous media to form dolomite is explained from
et al., 2007; Díaz-Hernández et al., 2011). The occurrence of neoformed the standpoint of kinetics.
dolomite in soils is usually restricted to saline environments (Shermann
et al., 1962; Kohut et al., 1995). However, the presence of high-Mg 3. Material and methods
water solutions in soils derived from weathering of serpentinites and
basaltic rocks may lead to the formation of authigenic dolomite in 3.1. Geological setting and climate
non-saline environments (Podwojewski, 1995; Capo et al., 2000).
Although the formation of dolomite in soils has been discussed (Drees The Guadix–Baza basin is located in the Betic cordillera (SE Spain,
and Wilding, 1987; Sobecki and Karathanasis, 1987; Bui et al., 1990; Fig. 1A). The area studied was developed as an endorheic depression
Whipkey and Hayob, 2008), no mechanism explaining the precipitation from the Late Neogene to the Late Pleistocene (Vera, 1970; Peña,
of dolomite in soils has been reported. The incorporation of magnesium 1985; García-Aguilar and Martín, 2000; Gibert et al., 2007, among
into the calcite structure to form high magnesian-calcites and proto- others), and occupies about 3500 km 2. During this period, the basin
dolomites under surface conditions is broadly described in the was filled with alluvial and lacustrine materials for which the compo-
literature (Hardie, 1987 and references therein; Deleuze and Brantley, sition varies depending on the nature of source materials. Those from
1997). Sedimentary and diagenetic dolomites probably precipitated as the Subbetic ranges (North border) are rich in limestone, while those
protodolomites and their isotopic composition was controlled by the from the Betic ranges (South border) are rich in quartzite, micaschist,
protodolomite–water fractionation (Fritz and Smith, 1970). In any and dolomite (IGME, 1973–1999). The Guadix sub-basin is comprised
case, the incorporation of magnesium to the rhombohedral carbonate of fluvial sediments (silts, sands and conglomerates) and the Baza
structure is not easy and appears to be controlled by kinetic factors sub-basin is formed mainly by lacustrine deposits (limestones, marly
(Mucci and Morse, 1983; Putnis et al., 1995; Fernández-Díaz et al., limestones, and gypsum). The rocky relief surrounding the Guadix–
1996). Baza depression coincides partially with the Atlantic–Mediterranean
divide. The depression is characterized by having elevations ranging
2. Scope of the investigation from 1500 m a.s.l. (above sea level) on the edges to 650 m a.s.l. at the
Guadiana Menor River, the main drain of the area.
In this paper, we study the occurrence of pedogenic dolomite in The capture of the internal drainage network of the depression by
soils of an arid/semiarid region. Calcite is by far the most abundant the Guadiana Menor River (a tributary of the Guadalquivir River)
and best-studied pedogenic carbonate. On the other hand, dolomite modelled the landscape into two main geomorphological groups
is not usually evaluated in soils and its role in the pedogenic processes (Díaz-Hernández and Juliá, 2006):
is scarcely known. Here, we describe the formation of dolomite at soil
depths below indurated layers delimitated by petrocalcic horizons or 1) Old surfaces (350,000–205,000 yr): S1, S2, S3, consisting of
caliches. Deeper layers (subsoil) in these types of soils are scarcely dissected piedmont plains gently sloping towards the centre of the
studied because caliches are hard to excavate and their features at basin (Fig. 1B), capped by caliche soils with thick petrocalcic hori-
these depths are irrelevant for soil taxonomy or for judging the soil zons (Petric Calcisols and Chromic Luvisols, FAO-UNESCO, 1988)
as a support for plants. (Calciorthids, Paleorthids, Paleargids, Haploxeralfs, Rodoxeralfs and
Several reasons underlie this approach: 1) Knowledge of the dis- Xerochrepts, following the Soil Taxonomy, Soil Survey Staff, 1999).
tribution of carbon in the deep-soil-layer system in semiarid areas is Dolomite enrichment observed under the petrocalcic horizon (see
essential to assess soil carbonates (calcite and dolomite) as a sink below) suggests that the formation of dolomite began within this
for atmospheric carbon. However, the carbon distribution in deeper broad time range.
layers (subsoil) has rarely been determined (Jobággy and Jackson, 2) The youngest surfaces (115,000–48,000 yr): S4 (badlands, with
2000; Díaz-Hernández et al., 2003). 2) Well-ordered dolomite, together Gypsic, Calcaric and Eutric Regosols, FAO-UNESCO, 1988) (Xerortents,
22 J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33

Fig. 1. A) Location of the study area showing the position of 81 profiles sampled. B) Cross section indicated in panel A showing the spatial relationship between the geomorphological
units of the area.

Soil Survey Staff, 1999) and S5 (alluvials, with Calcaric Fluvisols, 3.2. Sampling
FAO-UNESCO, 1988) (Xerofluvents, Soil Survey Staff, 1999).
Soil profiles were studied at 81 sites in the Guadix–Baza basin
The climate of the basin is semiarid, hot in summer, with frequent (Fig. 1A), where the depth and features of the caliches could be accu-
frosts in winter. The mean annual temperature is 15 °C. Mean rately determined. Natural or artificial pits reached an average depth
annual precipitation is about 250–300 mm in the central part and of about 200 cm (in several cases down to 400 cm) and the sampling
about 400 mm in the higher fans located along the mountain fronts sites were distributed as evenly as possible, in 10 × 10 km cells to
(MAPA, 1989). Rainfall occurs mainly during the autumn–winter simulate a probabilistic sampling, with one or two sites selected per
months; in summer, rain is scarce, falling as infrequent torrential cell. Data from the deepest layers were not considered here because
storms. of the statistical results from the few data (N b 4) are of doubtful
J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33 23

reliability. Samples were taken at intervals of 20 cm deep after 3.6. Electron microscopy
cleaning the sections to expose fresh material. From the 81 profiles
here studied, 43 correspond to rainfed cultivated areas, 33 to Two samples corresponding to the fine fraction of soil material
Mediterranean bush, and 5 to Mediterranean forest. located below the petrocalcic horizon at 120 and 220 cm depths
(profile 46) were selected on the basis of their dolomite content.
3.3. Carbon soil analysis and particle-size measures They were consolidated and prepared as polished sections and later
carbon coated for the textural and mineralogical study by analytical
Organic and inorganic carbon were analysed in sieved samples scanning electron microscopy. Back-scattered electron (BSE) images
(b2 mm) using standard techniques (USDA-NRCS, 1996). Organic and energy-dispersion X-ray (EDX) spot analyses (200 nm of diameter)
carbon was determined by the Walkley and Black method, the CaCO3 were performed with a high-resolution field-emission scanning electron
equivalent was measured by a manometric method (Loeppert and microscope (FESEM) Auriga (Carl Zeiss) equipped with a LinK INCA 200
Suárez, 1996), and the calcite and dolomite contents were inferred from (Oxford Instruments) microanalysis system operated at an acceleration
the CaCO3 equivalent together with the peak-height ratios detected in voltage of 10–20 kV.
the X-ray diffraction (XRD) powder diagrams. Soil-particle-size analysis For transmission electron microscopy (TEM), the same b 2 μm frac-
was performed by the Robinson pipette method (USDA-NRCS, 1996). A tion of the mudstone samples studied by powder-XRD was embedded
total of 892 samples were analysed. in an epoxy resin, reoriented, and sectioned by diamond-knife ultrami-
crotome. Diverse thin-sections below 100 nm were deposited on a
copper grid and coated with carbon for the TEM study. Analytical elec-
3.4. XRD methods tron microscopy (AEM) and TEM images were obtained with a Philips
CM20 instrument operated at 200 kV and equipped with an EDX
The specimens were ground to particle sizes b20 μm and examined model EDAX. Analysis of smectite and amorphous material tightly
using a Philips PW 1710/00 diffractometer, with a graphite monochro- intergrown with clay particles was performed in STEM mode using a
mator, automatic slit, Cu–Kα radiation and on-line computer connec- beam 5 nm in diameter and a scanning area of 20 × 100 nm. For quan-
tion. Data were collected for 0.4 s integration time in 0.05° 2θ steps at titative micro-analyses, EDX data were corrected by the thin-film meth-
40 kV and 40 mA, in a 2θ interval between 3 and 70°. od of Lorimer and Cliff (1976). The k-factors were determined using
The fine fraction (b2 mm) of the sample at 220 cm depth was muscovite, albite, biotite, spessartine, olivine, and titanite standards.
prepared by dispersing the sample with an ultrasonic probe and The quantitative analysis corresponding to the smectite packet was
pipetting the suspension to obtain the b 2 μm fraction (fraction 1). selected according to the following crystal-chemistry constraints:
Carbonates were then removed after attack with diluted acetic acid. Si: 3.5–4 apfu (atoms per formula unit); sum of octahedral cations
This fraction was centrifuged at 3000 rpm, and a second suspension (∑VI): 1.95–2.2 apfu; and K + Na + 2Ca (∑XII): b1 apfu.
was obtained from the supernatant (fraction 2). The supernatant
liquid was deposited onto a circular glass slide for the oriented clay- 3.7. Experimental precipitation of magnesian calcites
mineral aggregates (OA). The occurrence of smectite in the sample
was investigated by treating the OAs with ethylene glycol. X-ray Solution growth of magnesian calcites was carried out by bulk
diffraction diagrams of the OAs were recorded using a PANalytical X'Pert crystallisation (free-drift experiments) where two solutions of rela-
Pro diffractometer (Cu–Kα radiation, 45 kV, 40 mA) equipped with an tively easily dissolved salts (0.1 M Na2CO3 and 50 mM CaCl2·6H2O–
X'Celerator solid-state lineal detector. The diffraction patterns were 50 mM MgCl2·6H2O) were mixed, and the resulting solution became
made by a continuous scan between 3 and 50° 2θ, with a 0.01° 2θ step supersaturated with respect to the less soluble (Ca,Mg)CO3 substance.
size and 20 s time step. In one of the experiments, a low-viscous medium was prepared by
The data were processed using the XPowder® programme in order adding 0.1% of agar to water, boiling the solution, and cooling to
to determine the qualitative mineral composition and XRD map 25 °C. Precipitates were later isolated and studied by XRD using the
(Martín-Ramos, 2004). XPowder® programme.

3.5. Isotopic analysis 4. Results

The soil samples used in this study, selected from among the 81 4.1. Compositional, mineralogical and particle-size variations in the soil
profiles on the basis of an even distribution by geomorphic units, profiles
were ground to b 200 mesh to be later treated with 100% phosphoric
acid for 12 h in a thermostatic bath at 25 °C (McCrea, 1950). The sam- The distribution of organic and inorganic carbon with depth varied
ples containing calcite and dolomite carbonates were also treated considerably in the soils of the study region (Fig. 2A–C; see also Fig. 1
according to the method of Al-Aasm et al. (1990). This method con- in Díaz-Hernández, 2010), reflecting the heterogeneity of the area,
sists essentially of three steps: 1) obtaining CO2 after a 2-h reaction in which several soil types are represented (Díaz-Hernández et al.,
with phosphoric acid at 25 °C (labelled as calcite); 2) after of 24 h 2003). Profiles of organic and inorganic carbon (calcite and dolomite)
at 25 °C elimination of CO2 from residual calcite and a small quantity distribution with depth for alluvial soils, badlands, and piedmont units
of dolomite, and 3) obtaining CO2 from the continuation of the reaction are depicted in Fig. 2A–C. Organic carbon decreases with depth in all
at 50 °C for 24 h that was labelled as dolomite. The carbon dioxide soil types, and it is particularly marked in the case of piedmonts
resulting from both methods was analysed using a Finnigan Mat 251 (Fig. 2A). Petrocalcic horizon, represented by the enrichment in calcite
mass dual inlet mass spectrometer. The experimental error was less content in depth, is detected mostly in piedmonts. This horizon appears
than ±0.1‰ for δ 13C and δ 18O. The in-house standards Carrara and between 10 cm and 90 cm depth, with 30–130 cm of thickness (Fig. 2B,
EEZ-1 were previously verified against NBS-18 and NBS-19. All the see also Díaz-Hernández et al., 2003). An increase of the dolomite
samples were compared to a reference carbon dioxide obtained from content with depth is also better appreciated in piedmont profiles, usu-
a calcite standard prepared at the same time. The oxygen isotope ratios ally below the petrocalcic horizon (compare Fig. 2B with C). In the soils
for dolomite were calculated taking into account the fractionation factor studied, dolomite enrichment can occur at great depth (below 200 cm),
for acid decomposition at 50 °C, 1.01057 for dolomite (Rosenbaum and as is the case of the profile PC-43 of the Fig. 2C. This can be explained by
Sheppard, 1986) and at 25 °C, 1.01044 for calcite (Kim and O'Neil, the presence of Mg-rich substrates, constituted by altered ophitic rocks
1997). in the subsoil. The development of the petrocalcic horizons and the
24 J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33
J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33 25

Table 1
Arithmetic mean percentage (mean in %) of the organic and inorganic (calcite and dolomite) carbon content and of the diverse particle-size fractions for the specified depths in the
studied profiles (number of samples, N, for each depth is also indicated).

Depth (cm) Organic carbon Inorganic carbon Particle-size (b2 mm) Particle-size (>2 mm)

Calcite Dolomite Clay Silt Sand Gravels

N Mean (%) Mean (%) Mean (%) Mean (%) N Mean (%) N

0 81 1.82 29.35 9.19 19.48 31.26 42.26 69 34.44 81


20 81 1.13 37.60 8.42 19.82 29.59 50.60 55 34.83 68
40 81 0.67 42.59 10.34 19.38 25.58 55.04 43 37.85 60
60 81 0.46 42.23 12.95 15.51 28.27 56.22 43 43.14 57
80 81 0.34 39.35 14.68 13.62 26.45 59.93 46 39.89 65
100 81 0.25 35.90 15.24 12.04 23.43 64.53 43 37.73 65
120 81 0.19 32.64 16.16 11.88 23.96 64.16 50 39.05 67
140 75 0.16 31.41 16.83 11.46 21.49 67.05 42 35.55 65
160 64 0.15 29.78 15.81 11.08 19.66 69.26 37 41.63 59
180 55 0.12 31.25 13.28 10.94 22.82 66.23 35 41.14 48
200 44 0.13 28.09 16.97 12.72 20.46 66.82 31 45.99 40
220 27 0.11 33.63 18.33 12.12 17.99 69.89 17 53.06 24
240 19 0.12 38.25 13.57 13.76 25.17 61.08 14 52.15 17
260 12 0.14 37.31 16.26 10.18 15.63 74.19 6 51.91 9
280 9 0.15 28.81 12.73 14.57 17.92 67.51 5 56.14 7
300 8 0.18 26.78 12.77 18.89 19.18 61.94 4 59.46 6
320 8 0.22 25.03 13.51 11.72 29.10 59.18 4 55.71 6
340 4 0.22 31.14 14.00 9.20 15.66 75.14 2 58.12 3
Total 892 546 747

occurrence of associated dolomite enrichment in depth are better ob- mapping of Fig. 3, corresponding to a depth sequence of diffractogram
served in the case of piedmonts. patterns for profile 46 (piedmont). This image shows a decrease in
A total of 81 profiles were analysed, 8 corresponding to alluvial, quartz with depth on the basis of the change in the intensity of the
11 to badlands, and 62 to piedmonts, in order to gain statistical in- (101) reflection, and an interruption of the signal coinciding with
formation concerning carbon distribution in depth for the soils the maximum of calcite enrichment (100 cm depth). By contrast,
studied. Table 1 and Fig. 2D show the depth distribution of soil or- the dolomite (104 peak) showed an inverse distribution to that of
ganic carbon content (from organic matter) and calcite and dolo- quartz (in-depth enrichment), with the same in-depth interruption
mite content (soil inorganic carbon), expressed as an average and lack in superficial horizon. Calcite enrichment between 60 and
percentage. The average organic carbon decreased rapidly with 130 cm of depth, as indicated by the increase in the intensity of the
depth, tending to near-zero values below 120 cm (N = 81 pro- (104) peak, coincided with the development of a petrocalcic horizon
files), and continued to decline gradually to a depth of 220 cm of 70 cm thick in this profile.
(N = 27 profiles). The calcite content initially increased with
depth, showing minor amounts near the surface and a maximum
between 40 and 60 cm depth (43% in the bulge). Below this 4.2. Isotopic data
depth the calcite diminished gradually to 200 cm in depth (around
30%). A second bulge at a depth of 200–280 cm appeared from the The frequency distribution of δ 13C and δ 18O of the calcite in the
profiles of the units of type S1. Polycyclic profiles showing this sec- study samples are shown in Fig. 4A and B, respectively. These values
ond maximum of calcite accumulation within this depth range were consistently negative, and δ 13C ranged from −13 to −1‰ and
were found for S1 unit (not shown). These minor maxima δ18O from −10 to −2‰, with a normal distribution and modes around
correspond to hard substrates or paleosols in those units, probably −7‰ in both cases. These negative values indicate a contribution of
caused by regional sedimentological changes, and are not consid- isotopically light carbonate. The histograms for the δ13C and δ 18O of
ered in this work. The average depth distribution of dolomite the dolomite showed bimodal distributions (Fig. 4C and D): the two
began with a low content (9% in the superficial horizons) which modal values for δ 13C were −5 and −1‰, and −6 and −3‰ for
progressively increased from 40 cm, reaching 17% at 140 cm δ18O. The scatter diagram relating both δ values for the dolomite sam-
depth. However, the depth distribution of the dolomite and calcite ples also showed these two different tendencies in the histograms,
contents proved somewhat irregular below 180 cm due to the low and two clusters may be identified within the diagram (Fig. 4E). The
number of samples for these depths (Table 1). Average particle-size average isotopic composition of the inherited dolomites from the coarse
distributions throughout the soil profiles studied show high con- fraction of the studied soils is also represented in the bivariate plot (star
tents of sand and gravels with respect to clays and silts (Fig. 2E). in Fig. 4E). It plotted close to the analysed dolomite cluster with less
The XRD study evidenced that the major minerals of soil samples negative and even positive δ values (cluster I in Fig. 4E). Samples of
were quartz, calcite, and dolomite, with minor amounts of clays, feld- deeper layers (>70 cm) defined the cluster with more negative δ13C
spars, and micas. For dolomite-rich samples, superstructure reflection and δ18O values (cluster II in Fig. 4E) that corresponded to the smallest
(015) (the main reflection of well-ordered dolomites; Goldsmith and left-hand maximum observed in the histograms (Fig. 4C and D).
Graf, 1958; Reeder and Sheppard, 1984) was present. The depth However, none of single profiles studied here showed a clear isotopic
distribution of quartz, calcite, and dolomite was shown in the XRD fractionation with the dolomite enrichment in depth.

Fig. 2. A–C) Vertical distribution of organic and inorganic (calcite and dolomite) carbon in alluvial soils (PA, Fluvisols), badlands (PB, Regosols) and piedmonts (PC, Calcisols).
Geomorphological units are also indicated in brackets. Arrows in panel C mark dolomite enrichment in depth for the diverse type of soils. D) Depth-wise (y-axis) variation of
the organic carbon, calcite, and dolomite (inorganic carbon) content, expressed as an average percentage, in the study profiles of Guadix–Baza basin, drawn from the data of
Table 1. E) Distribution with depth of diverse granulometric fractions of soils; only mean values calculated from the studied profiles have been represented (gravel + fraction
b2 mm = 100% and contents of the fractions b2 mm (clay + silt + sand) are, in turn, recalculated to 100%). Panels D–E show the limits of the layers referred to in the text
(soil, petrocalcic, and subsoil). Superficial horizon (epipedon) and endopedon in the subsuperficial horizon are also indicated in these figures.
26 J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33

Fig. 3. False colour map drawn from XRD analyses of powdered samples (profile 46, piedmont S2). The y-axis represents the depth of the samples at intervals of 20 cm, with soil
surface on the top. Colour corresponds to the relative intensities of XRD peaks (26–32° 2θ interval only is represented). The approximate limits of petrocalcic horizon are indicated.

4.3. SEM and TEM study sometimes showed central voids (Fig. 5D). Calcite rhyzocretions also
occurred close to the petrocalcic horizon (Fig. 5D).
Electron microscopy confirmed the XRD observations of Fig. 3 Samples from 220 cm in depth showed corroded calcite grains
regarding dolomite dissolution and calcite precipitation in relation surrounded by dolomite and clay minerals (Fig. 5E–F). Fig. 5F, corre-
to the formation of petrocalcic horizon. BSE images of Fig. 5A and B sponding to an inset of Fig. 5E, shows dolomite + clays secondary
correspond to a sample below the petrocalcic horizon (120 cm products located in a corrosion gulf which includes fragments of
depth), where the corrosion affecting dolomite and quartz crystals partially altered calcite relics. A detailed observation of precursory
and the precipitation of calcite are visible. In Fig. 5A the voids, after calcite evidenced the presence of a poorly crystalline material, basically
quartz dissolution, appear to be filled mainly with calcite. Frequently, clay minerals, close to irregular corrosion boundaries at the surface of
calcification is widespread and calcite nodules occur with dolomite calcite grains, in addition to inherited chlorites. Clay minerals also
relicts within the concretions (Fig. 5B). The calcified nodules frequently surrounded sub-idiomorphic dolomite crystals with sizes around
had a highly porous texture. The dissolution process of inherited dolo- 2 μm in diameter, and sometimes localized preferentially between
mites sometimes began along crystallographic planes and advanced corroded calcite grains and rhombohedral dolomite crystals (Fig. 6A
from inner to outer regions with progressive corrosion, forming hollow and B). Poorly defined crystalline material surrounding dolomite
dolomite crystals (Fig. 5C). In addition, corroded dolomite grains consisted of smectites plus Si-rich amorphous substance (Figs. 6C, 7

Fig. 4. A–D) Histograms of the δ13C ‰ (V-PDB) and δ18O ‰ (V-PDB) of calcites (A and B) and dolomites (C and D). E) Scatter diagram showing the isotopic values (δ13C vs. δ18O) of
the dolomites contained in the fraction b2 mm coming from selected profiles of the studied area. Number data = 72. Note that dolomite samples below 70 cm in depth (full
circles) are highly fractionated. Star corresponds to the average isotopic composition of the inherited dolomites that constitute the fine pebble fraction of the studied soils. The
Meteoric Dolomite Line (MDL) was calculated using the equation for the dolomite-water system (Irwin et al., 1977) with the range of local meteoric water (−8 to −9‰ vs.
V-SMOW) and surface temperatures from 17 to 24 °C.
J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33 27

Fig. 5. A–B) Back-scattered electron images showing precipitation of pedogenic calcite at 120 cm depth (profile 46): Grain of quartz partly corroded and filled with calcite and probably
siderite (A). Calcite nodule including dolomite relicts (B). Note the porous texture of the calcite precipitates in these images. C) Back-scattered electron micrograph of a large idiomorphic
crystal of dolomite in an advanced stage of corrosion located at 120 cm in depth. Note that this corrosion occurs in preferential plains, and progresses from the centre to outer edges.
D) Small dolomite crystals with central voids; in addition, there are pedogenic calcites as rhyzocretions. E) Back-scattered electron image of a partly corroded large calcite grain
surrounded by a mixture of clays and carbonates in a sample from 220 cm in depth (profile 46). F) Inset of E in which we can see some dolomite grains surrounded by clays and corroded
calcite grains.

and Table 2). According to AEM analyses of Table 2 the smectite is inter- 5. Discussion
mediate between beidellite and montmorillonite, although several
analyses correspond to mixtures of clay minerals and Si-rich amor- 5.1. Origin of compositional and mineralogical variations along the
phous substances. Si excess cannot be incorporated in the smectite profiles
structure and must therefore derive from adjacent non-crystalline
material. Bubbles occurring in relation to clay minerals and amorphous Organic carbon shows the expected distribution (Fig. 2A and D) for
substance in TEM image (Fig. 6C) resulted from the gradual release of these types of soils with a usual decrease with depth due to the more in-
volatile substances after pervasive electron bombardment. Clay min- tense biogenic activity in the upper horizons. The combination of water
erals were also studied by XRD from separate fractions of the sample from rainfall, Ca 2+ ions from diverse sources and CO2 generated mostly
at 220 cm depth (Fig. 8). This study confirms the occurrence of the from organic-matter decomposition and root respiration (Gile et al.,
following clay minerals: 1) An inherited chlorite is evidenced by (002) 1966; Deines, 1980; Rech et al., 2003), the transient decrease in PCO2
and (004) peaks (Fig. 8A; see also BSE image and qualitative microanal- and the increase in the concentration of solutes due to evapotranspira-
ysis of Fig. 6B); 2) a detrital illite with peaks at 1.0 and 0.5 nm d-spacing tion lead to the accumulation in depth of pedogenic calcium carbonate
corresponding to (001) and (002) reflections; 3) abundant smectites in arid and semiarid regions. The formation of the petrocalcic horizon
and minor illite–smectite mixed-layer phases (probably R3-ordered explains the occurrence of the bulge at 40–60 cm depth in the calcite
I/S). The latter is evidenced by the broad basal reflection of the 1.0– average regional profile (Fig. 2D), and the calcite enrichment found in
1.2 nm zone in the untreated sample (Fig. 8A). The smectite had a the individual profiles shown in this work (Figs. 2B and 3). Although
basal spacing of around 1.4 nm in the air-dried state, which expands the dolomite content in the individual profiles was somewhat irregular
up to 1.8 nm upon ethylene glycol saturation (Fig. 8B). (Fig. 2C), and in particular for depths below 180 cm, it was evident that
28 J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33

Fig. 7. A) General TEM image of clays and poorly crystalline material surrounding
dolomite crystals in a sample from 220 cm depth (profile 46). In addition to well-
defined smectite packets (S), with a composition shown in Table 2, amorphous
substance (A) with a mottled appearance and marked dark contrast also occurs in
the poorly crystalline material located around the neoformed dolomite crystal (Dol).
Other areas in the image lacking fringe contrast and of mottled aspect may correspond
to smectites with a slight disorientation of the layers (Sd) or to a mixture of poorly
crystalline smectites and amorphous substance (Table 2). B) Structural features of
the small smectite packets, some of them composed of a few basal layers. The packets
appear to be part of a (gel-like) matrix formed by disoriented smectite-like flakes and a
substance of relatively low atomic number lacking fringe contrast.

a regional decrease occurred near the surface with a marked enrich-


ment in depth (Fig. 2D, Table 1). Moreover, a smooth increase began
at 40 cm and continued to 140 cm depth. Below 180 cm depth, the
values in the profile may have resulted from the interference of

Table 2
Analytical electron microscopy chemical analyses, expressed in atoms per formula unit,
of smectite packets and mixture of smectites and amorphous substance. Normalisation
to 22 charges.

Smectite analyses Mixed analyses

Si 3.79 3.88 3.83 3.78 4.16 4.17 4.12 4.41


AlIV 0.21 0.12 0.17 0.22 −0.16 −0.17 −0.12 −0.41
AlVI 1.55 0.84 1.57 1.69 1.55 1.28 1.43 1.53
Fe(3+) 0.24 0.62 0.17 0.14 0.05 0.28 0.24 0.16
Mg 0.34 0.53 0.22 0.14 0.31 0.48 0.26 0.21
Fig. 6. Electron images of a sample from 220 cm depth (profile 46). A–B) Back-scattered VI
2.14 2.00 1.97 1.97 1.91 2.03 1.93 1.90
electron images of rhombohedral dolomite crystals adjacent to corroded calcite grains and
Ca 0.03 0.03 0.00 0.02 0.05 0.05 0.05 0.03
surrounded by clays. Detrital clays (chlorites) also occur in the area shown in panel B.
K 0.07 0.36 0.34 0.31 0.28 0.16 0.22 0.21
C) Transmission electron microscopy image of a small subhedral dolomite crystal XII
0.10 0.40 0.34 0.33 0.33 0.21 0.28 0.24
surrounded by smectite plus Si-rich amorphous substance.
J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33 29

Fig. 8. A) Powder X-ray diagram obtained from oriented aggregated (fraction 2) of the sample at 220 cm depth (profile 46), with an indication of d values corresponding to the
reflections of the observed mineral phases. (B) Comparison of air-dried and ethylene-glycol-treated diagrams, where the basal spacing of smectites is displaced from 1.46 nm to
1.76 nm after glycolation.

authigenic processes with the effects of the substratum on the mineral- which may be interpreted in relation to bio-mediated precipitation
ogy. Although the maximum in the case of the dolomite was less marked processes.
than in the calcite, the overall variation in dolomite (around 10%) was As mentioned above, dolomite enrichment below 40 cm in depth
quite similar to the percentage change observed in calcite (around was observed in the depth-wise variation of dolomite at the regional
12%). As occurred for calcite, dolomite dissolution/precipitation may scale (Fig. 2D) and in the representative profiles of the Figs. 2C and 3.
explain the occurrence of the smooth maximum, just below the One possible explanation would imply no dolomite precipitation, but
petrocalcic horizon, preceded by the change of sign in the curvature of only dissolution; thus, the increment in-depth would be due to more
the dolomite pattern in the Fig. 2D. Therefore, at the regional scale, do- intense dissolution for superficial horizons, the solutes being leached
lomite underwent superficial dissolution and its precipitation occurred out of the system. On the other hand, in dissolution/precipitation
at greater depths than in the case of calcite. In short, the calcite- and mechanism proposed here, dolomite dissolved above 40 cm is poten-
dolomite-distribution patterns found here clearly indicate that they tially precipitated below this depth. Figs. 5E–F and 6, corresponding
were affected by similar processes, but, in the case of dolomite, different to a sample situated at 220 cm in depth in the studied profile, show
dynamics seem to have occurred within the soil (see below). the occurrence of dolomite and clay overgrowth in relation to calcite.
Dissolution of the dolomite in the superficial horizon of the studied Textural relations indicate the dissolution of calcite and the precipita-
profiles (e.g. Fig. 3) in relation to the development of petrocalcic horizon tion of dolomite in relation to clays. The specific mechanisms explaining
was also evidenced by Fig. 5A–D, where detrital dolomite appeared the observed spatial association of calcite, dolomite and clays in SEM
partially or almost completely dissolved and calcite precipitates in and TEM images are discussed below.
voids and formed rhyzocretions. A general decline in the quartz content
was noted in the profile in Fig. 3, contradicting the assumption that 5.2. Pedogenic isotopic fractionation and neoformed dolomites
quartz content remains constant throughout the soil profiles (Lelong,
1969; Souchier, 1971; Sohet et al., 1988). The appearance of a physical The proposed dolomite dissolution/precipitation model is consistent
barrier represented by the petrocalcic horizon prevents the redistribu- with the heterogeneity observed in the isotopic data from the dolomite
tion of the upward-added quartz from diverse sources, mainly aeolian samples analysed, which showed two different signals (Fig. 4C, D, and
(Díaz-Hernández et al., 2011), and is therefore responsible for the E). Cluster I in Fig. 4E, with higher isotopic values, lies close to the parent
observed decrease of quartz below it. However, the quartz content material that supplies the dolomite to the system (star in Fig. 4E). The
may decrease below the petrocalcic horizon as a result of dissolution precipitation of a neoformed dolomite below the petrocalcic (Figs. 5E,
processes, as evidenced by Fig. 5A, where the precipitation of calcite F and 6) as a result of dissolution of inherited dolomite and pedogenic
after quartz dissolution is evident. Calcite concretions after the dissolu- calcite of upper horizons explains the high dispersion observed in the
tion dolomite and quartz had a marked porous texture (Fig. 5A and B) δ13C values of the samples of deeper layers (cluster II in Fig. 4E and
30 J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33

smallest left-hand maximum in the histograms of Fig. 4C and D). These diffusion of C and O within the dolomite structure might explain isotope
values define the MDL (Meteoric Dolomite Line), a term similar to that exchange between the solid and the pore fluid in soils, and the observed
proposed by Lohmann (1988) for the Meteoric Calcite Line (MCL). The fractionation in Fig. 4, without the need for an explanation involving
δ13C dispersion is explained by a mixing of inherited carbon from the dolomite dissolution–precipitation process. However, alternative
dolomite-rich parent material, with positive values, and pedogenic explanations of these isotopic data based on intracrystalline diffusion
carbon from neoformed dolomite, with very negative δ13C values. In processes should be discarded in our case because of: 1) According to
addition, δ 18O values close to 6‰ (V-PDB) observed for pedogenic dolo- the Arrhenius relationship between the diffusion rate and the tempera-
mites (cluster II in Fig. 4E) are consistent with the δ 18O measured in ture, intracrystalline diffusion is negligible at low temperature (Putnis,
surface waters, values ranging from −7.5 and −9‰ (V-SMOW), and 2003). 2) Volume diffusion of C within the structure of the rhombohe-
temperatures between ~17 and 24 °C. dral carbonates occurs by diffusion of carbonate anions, whereas the
The isotopic composition of the dissolved inorganic carbon (DIC) lattice diffusion of O is due to vacancy migration (Labotka et al.,
and oxygen in soil waters below the petrocalcic horizon is in equilib- 2000); this explains why the diffusivity of carbon is 4 orders of magni-
rium with the isotopic composition of neoformed dolomites. Carbon tude smaller than the diffusivity of oxygen in calcite (Ferry et al., 2010).
and oxygen are distributed among three chemical species with a The latter point disagrees with our observations, where a broader range
pH-dependent relative abundance (CO2, HCO3− and CO32−). Most of of variation is found for δ 13C compared to that of δ 18O for the cluster II
the carbonic acid derives from the vegetal cover and prompts an (Fig. 4E) corresponding to the dolomite located at deeper levels.
increase of negative δ 13C in DIC (Cerling, 1984, 1991; Usdowski and
Hoefs, 1993; Reyes et al., 1998). In the pedogenic medium studied, 5.3. Physico-chemical constraints on the dolomite formation in the
this isotopic signature is transported towards deeper layers through subsoil
carbonate dissolution (mostly calcite) above the petrocalcic horizon.
The negative carbon isotope composition of the neoformed dolomite As shown, dissolution of inherited dolomite above petrocalcic con-
is determined mainly by the isotopic pedogenic calcite because pore stitutes the main source for the Mg 2+ enrichment in the soil solution
fluids generally have low carbon contents so that the 13C of the pre- necessary for dolomite neoformation. An exception to this is provided
cursor is generally retained (Hoefs, 2004). Low δ 13C and δ 18O values by the profiles with altered ophitic rocks in their subsoil, where minor
are expected in biomediated calcite and dolomite precipitates (Jones dolomite was formed at great depth (below 200 cm, Fig. 2C). The
and Donnelly, 2004; Wacey et al., 2007). Multiple cyclic dissolution/ montmorillonitic component of the smectites localized preferentially
precipitation processes that affected the carbonates in the soils between corroded calcite grains and rhombohedral dolomite crystals
explain not only the formation of an isotopically light calcite, related (Table 2 and Fig. 6A–B) may constitute an additional source for Mg 2+.
to pedogenic–biogenic processes above the petrocalcic horizon Despite the occurrence of high Mg:Ca mole ratio, Mg rich calcite form
(Magaritz and Amiel, 1980; Magaritz et al., 1981), but also the forma- only at very high salinity (Stanley and Hardie, 1998). However, for
tion of an isotopically light dolomite, occurring below the petrocalcic medium salinities, a low Mg content was expected in the calcite formed
horizon, in which the isotopic composition differs from that observed from solutions in soils. This is because Mg2+ ions tend to remain in
for the inherited dolomites (Fig. 4). The isotopic values of the pedogenic solution during carbonate precipitation, where they are thought to be
calcite co-existing with neoformed dolomite may be assigned to the heavily solvated (de Leeuw and Parker, 2001), due to strongly attractive
maximum observed in the histograms of Fig. 4A–B. Calcite isotopic Mg–water interactions.
values are more negative than those of dolomite, as expected, because The formation of the petrocalcic horizon prevents rain-water infil-
most of the calcite in soil is pedogenic in origin. The pedogenic dolomite tration into the subsoil. The observed dolomite enrichment in the
precipitated in depth has a marked isotopic contribution from inherited subsoil probably occurred during the initial stages of the formation
dolomite, and therefore lower negative values than those measured for of the petrocalcic horizon (soft caliches), and before its hardening
the pedogenic calcite are expected. that led to the consolidation of the impermeable barrier. The high
Dolomite isotopic composition is not usually determined in soils, content of gravels and sands throughout the profile (Fig. 2E) favoured
and much less in deep layers (“subsoil”). Until now, the occurrence rainwater infiltration into the subsoil. Under these physical conditions,
of a pedogenic 13C (and 18O) signal in soils is associated exclusively evapotranspiration processes affected not only the soil, but also deeper
with the calcite genesis. In the case studied, the neoformed dolomite, levels during the less rainy climatic episodes. The decreased water
characterized by relatively negative values, forms after partial dissolu- content in the subsoil diminished bulk flow of the grain-boundary
tion of the calcite, with a partial “transference” of the biogenic isotopic fluid containing dissolved solutes. Under this regime, the fluid film is
signature. Therefore, calcite precipitation is interpreted here as an inter- restricted to a monolayer of molecules adsorbed onto the smectite
mediate stage for dolomite precipitation, below the petrocalcic horizon, particles. To understand the role of the water molecules adsorbed
because of the carbonate dissolution–precipitation cycles forming pedo- onto smectites in the incorporation of Mg to the structure of the calcite,
genic calcite are also responsible of the formation of the pedogenic we dealt here with the transport of solute in a viscous solid media
dolomites. (Putnis et al., 1995; Fernández-Díaz et al., 1996; Sánchez-Navas et al.,
Kinetic effects, such as diffusion transport, may also induce a 2009). It can be deduced from Table 3 that the quantity of Mg incorpo-
certain fractionation in soils. The carbon isotope fractionation in soils rated into the structure of the calcite increased in viscous medium and
caused by diffusion processes has been estimated to be around 4‰ that the incorporation of Mg into magnesian-calcite produced a lattice
(Cerling, 1984; Hesterberg and Siegenthaler, 1991). According to contraction of the rhombohedral carbonate, as well as a reduction of
these authors 13CO2 diffuses into the atmosphere at lower rates than
the lighter 12CO2 molecule. However this effect is clearly counteracted Table 3
by the major 12CO2 production related to root respiration and organic- X-ray diffraction results for magnesian calcites obtained in batch precipitation experi-
matter decomposition in the upper horizons of soil, as evidenced clearly ments by using two types of media. In both cases the concentrations of the mixed soluble
salts are the same: 50 mM Na2CO3, 50 mM CaCl2·6H2O, and 0.1 M MgCl2·6H2O (starting
by the lower δ13C values recorded for the pedogenic calcite of soil in
pH 7.4).
relation to the pedogenic dolomite occurring in the subsoil. The kinetic
effect discussed above resulting in 13C enrichment would partly explain Type of medium d104 (Å) FWHM (° 2θ)a % Mg
the shift from the very negative δ 13C at the C3 vegetal cover (~−27‰ Liquid 3.01 0.3 10
vs. V-PDB, Deines, 1980) due to CO2 from organic matter decomposi- Viscous 2.95 1.26 25
tion, towards values around −12‰ in the soil (Cerling, 1984, 1991; a
FWHM: full width at half maximum (peak broadening measured for 104 reflection
Reyes et al., 1998). Other kinetic processes such as intracrystalline in the XRD diagrams of Fig. 10).
J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33 31

crystallinity. In addition, metastable phases also appeared when a


viscous medium was used for the precipitation experiment (Fig. 9).
Micro-textural data (Fig. 6) indicate that the precipitation of neoformed
dolomite in depth after partial dissolution of pedogenic calcite was
mediated by the smectites of the soils studied (Figs. 7 and 8). Clay min-
erals surrounding neoformed dolomite crystals acted as a catalytic
material in the process of incorporation of Mg into the structure of the
rhombohedral carbonates. This was because the transport of reacting
material from dissolution sites to those of growth was the rate-
determining process. Therefore, as suggested by our laboratory experi-
ments, it is necessary to consider the constraints that transport proper-
ties may make on the incorporation of an element during crystal
growth. Fig. 10A shows the concentration profile of Mg at the crystal–
water interface during interface-controlled calcite growth. As suggested
at the beginning of this section, the partition coefficient between the
concentration of magnesium in carbonate crystal and the concentration
of magnesium in bulk medium, [Mg]carb/[Mg]m, must be clearly less
than 1. For a diffusion-controlled growth of dolomite, the concentration
of Mg rises above [Mg]m at the crystal–medium interface ([Mg]im). If Fig. 10. Concentration profiles of Mg at the crystal–medium interface. A) Concentration
the partition coefficient remains constant, [Mg]carb increases with distribution at an infinitely fast bulk-diffusion rate. B) Steady-state distribution at very
decreasing diffusion rate until, in the limit, [Mg]carb = [Mg]m. As a slow surface-diffusion rate along the grain boundaries of clay minerals. ([Mg]carb,
concentration of magnesium in carbonate crystal; [Mg]m, concentration of magnesium
result the effective partition coefficient equals 1, and therefore most of
in bulk medium; [Mg]im, concentration of magnesium at the carbonate crystal–medium
the Mg of the medium is then incorporated into the structure of the interface; Δ, thickness of the diffusion boundary layer). Based on Henderson (1986).
rhombohedral carbonate (Fig. 10B).
Previous experimental data and the occurrence in soils of high-
magnesium calcite together with well-ordered dolomite (Whipkey Before the formation of nonpermeable petrocalcic horizon, the
and Hayob, 2008) suggest that our neoformed dolomite that precipitated high contents in gravel and sand fractions in the studied soils
in relation to smectites in the “subsoil” should be, at least initially, a favoured infiltration and evapotranspiration processes in the subsoil.
“protodolomite” or “Mg-kutnohorite” with essentially disordered Ca– The moisture in the subsoil was a determining factor controlling the
Mg distribution. precipitation of Ca and Mg carbonates, the precursors of dolomite.
The precipitation of the neoformed dolomite below the petrocalcic
6. Conclusions horizon was related to the presence of smectites. Smectite retained
soil water and acted as a catalyst for the incorporation of Mg into
The present study detected the presence of a clear enrichment of the carbonate.
dolomite beyond the lower limit of biological activity in semiarid The geochemical model proposed here for the incorporation of Mg
soils. This limit coincides with the rooting depth of native perennial to the carbonate in slow diffusive transport medium, such as a clayey-
plants and, in the case study, is defined by the occurrence of the rich and water deficient environments, would also explain the close
petrocalcic horizon. association of Mg-carbonates with clays observed on the surface of
Isotopic data establish a bimodal distribution of δ 13C and δ 18O for Mars (Ehlmann et al., 2008).
dolomites, indicating the occurrence of an inherited and a neoformed This study documents the occurrence of changes in carbonate
dolomite. Isotopically light dolomite formed below the petrocalcic mineralogy in depth. Levels below the petrocalcic are traditionally
horizon after partial dissolution of the pedogenic calcite, which trans- considered as zones not affected by the pedogenic processes.
ferred the pedogenic (biological) 13C and 18O signature from the soil In the soils studied, the dolomite content reached 43 kg m−2
(s. str.) to subsoil. (Díaz-Hernández, 2010), and therefore dolomite should be considered

Fig. 9. X-ray diffraction diagrams for magnesian calcites obtained in batch precipitation experiments (see Section 3.7 and Table 3) by using viscous (A) and liquid media (B). d104
spacing of calcite varies from 0.301 nm for the liquid medium to 0.295 nm for the viscous media. In the viscous medium monohydrocalcite also precipitates together with the
magnesian calcite.
32 J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33

in the quantification of carbon stored in arid–semiarid soils. Pedogenic García-Aguilar, J.M., Martín, J.M., 2000. Late Neogene to recent continental history and
evolution of the Guadix–Baza basin (SE Spain). Revista de la Sociedad Geológica de
processes occurring in the subsoil, such as the dolomite formation, España 13, 65–77.
should be included in the carbon-soil balances. Ghebre-Egziabhier, K., St. Arnaud, R.J., 1983. Carbonate mineralogy of lake sediments
and surrounding soils. 2. The Qu'Appelle Lakes. Canadian Journal of Soil Science
63, 259–268.
Acknowledgements Gibert, L., Ortí, F., Rosell, L., 2007. Plio-Pleistocene lacustrine evaporates of the Baza
Basin (Betic Chain, SE Spain). Sedimentary Geology 200, 89–116.
Gile, L.H., Peterson, F.F., Grossman, R.B., 1966. Morphological and genetic sequences of
We acknowledge support from grant P11-RNM-7067 of the Junta de carbonate accumulation in desert soils. Soil Science 1015, 347–360.
Andalucía (C.E.I.C.—S.G.U.I.T.) and the projects CGL2007-65572-C02-01/ Goddard, M.A., Mikhailova, E.A., Post, C.J., Schlautman, M.A., 2007. Atmospheric Mg2+
wet deposition within the continental United States and implications for soil
BTE, CGL-2009-09249, CGL2010-21257-C02-01 and CGL2012-32169 of inorganic carbon sequestration. Tellus Series B 59, 50–56.
the Ministerio de Ciencia e Innovación (MICINN, Spain). We would like Goldsmith, J.R., Graf, D.L., 1958. Structural and compositional variations in some
to thank Drs. D. Martín-Ramos and F. Nieto for their help with the XRD natural dolomites. Journal of Geology 66, 678–693.
Hardie, L.A., 1987. Perspectives dolomitization: a critical view of some current views.
analyses, A. González, M.M. Abad and M.J. Guerrero for their guidance
Journal of Sedimentary Petrology 57, 166–183.
with SEM and TEM studies and Dr. A. Delgado for his assistance with Henderson, P., 1986. Inorganic Geochemistry. Pergamon Press, Oxford (356 pp.).
the stable isotope data. We thank the anonymous reviewers for their sug- Hesterberg, R., Siegenthaler, U., 1991. Production and stable isotopic composition of
gestions, which contributed to improvement the manuscript. Mr. David CO2 in a soil near Bern Switzerland. Tellus 43B, 197–205.
Hoefs, J., 2004. Stable Isotope Geochemistry. Springer-Verlag, Berlin Heidelberg (244
Nesbitt revised the English manuscript. pp.).
IGME (1973–1999) Mapa Geológico de España, escala 1:50.000, Hojas 929, 930, 950,
951, 972, 973, 993, 994, 1011 and 1012.
References Irwin, H., Curtis, C., Coleman, M., 1977. Isotopic evidence for source of diagenetic
carbonates formed during burial of organic-rich sediments. Nature 269, 209–213.
Al-Aasm, I.S., Taylor, B.E., South, B., 1990. Stable isotope analysis of multiple carbonate Jobággy, E.G., Jackson, R.B., 2000. The vertical distribution of soil organic carbon and its
samples using selective acid extraction. Chemical Geology 80, 119–125. relation to climate and vegetation. Ecological Applications 10, 423–436.
Arkley, R.J., 1963. Calculation of carbonate and water movement in soil from climatic Jones, M.B., Donnelly, A., 2004. Carbon sequestration in temperate grassland ecosystems
data. Soil Science 96, 239–248. and the influence of management, climate and elevated CO2. New Phytologist 164,
Arvidson, R.S., Mackenzie, F.T., 1999. The dolomite problem: control of precipitation 423–439.
kinetics by temperature and saturation state. American Journal of Science 299, Kim, S.T., O'Neil, J.R., 1997. Equilibrium and nonequilibrium oxygen isotope effects in
257–288. synthetic carbonates. Geochimica et Cosmochimica Acta 61, 3461–3475.
Botha, G.A., Hughes, J.C., 1992. Pedogenic palygorskite and dolomite in a Late Neogene Kohut, C., Dudas, M.J., Muehlenbachs, K., 1995. Authigenic dolomite in a saline soil in
sedimentary succession, nothwestern Transvaal, South Africa. Geoderma 53, Alberta, Canada. Soil Science Society of America Journal 59, 1499–1504.
139–154. Labotka, T.C., Cole, D.R., Riciputi, L.R., 2000. Diffusion of C and O in calcite at 100 MPa.
Bui, E.N., Loeppert, R.H., Wilding, L.P., 1990. Carbonate phases in calcareous soils of the American Mineralogist 85, 488–494.
western United States. Soil Science Society of America Journal 54, 520–529. Lelong, F., 1969. Nature et gènèse des produits d'altération de roches cristallines sous
Capo, R.C., Whipkey, C.E., Blachère, J.R., Chadwick, O.A., 2000. Pedogenic origin of climat tropical humide (Guyane Française). Mèmoires Science de la Terre, Nancy,
dolomite in a basaltic weathering profile, Kohala peninsula, Hawaii. Geology 28, 14, 187 pp.
271–274. Loeppert, R.H., Suárez, D.L., 1996. Carbonate and gypsum. In: Sparks, D.L. (Ed.),
Cerling, T.E., 1984. The estable isotopic composition of modern soil carbonates and its Methods of Soil Analysis, Part 3 Chemical Analysis. SSSA Book series, No 5. ASA &
relationships to climate. Earth and Planetary Science Letters 71, 229–240. SSSA, Madison, WI, pp. 437–474.
Cerling, T.E., 1991. Carbon dioxide in the atmosphere: evidence from cenozoic and Lohmann, K.Z., 1988. Geochemical Patterns of Meteoric Diagenetic Systems and their
mesozoic paleosols. American Journal of Science 291, 377–400. Application to Studies of Paleokarst. In: James, N.P., Choquette, P.W. (Eds.),
Cerling, T.E., Quade, J., 1993. Stable carbon and oxygen isotopes in soil carbonates. In: Paleokarst. Springer-Verlag, Berlin, pp. 58–80.
Swart, P.K., Lohmann, K.C., Mckenzie, J., Savin, S. (Eds.), Climate Change in Continental Lorimer, G.W., Cliff, G., 1976. Analytical electron microscopy of minerals. In: Wenk, H.R.
Isotopic Records. Geophysical Monograph, vol. 78. American Geophysical Union, (Ed.), Electron Microscopy in Mineralogy. Springer-Verlag, Berlin, pp. 506–519.
Washington, pp. 217–231. Magaritz, M., Amiel, A.J., 1980. Calcium carbonate in a calcareous soil from the Jordan
de Leeuw, N.H., Parker, S.C., 2001. Surface–water interactions in the dolomite problem. Valley, Israel: its origin as revealed by the stable carbon isotope method.
Physical Chemistry Chemical Physics 3, 3217–3221. Soil Science Society of America Journal 44, 1059–1062.
Deelman, J.C., 2005. Low Temperature Formation of Dolomite and Magnesite. Geology Magaritz, M., Kaufman, A., Yaalon, D.H., 1981. Calcium carbonate nodules in soils: 18O/16O
Series, Version 2.1. Compact Disc Publication, Eindhoven. and 13C/12C ratios and 14C contents. Geoderma 25, 157–172.
Deines, P., 1980. The isotopic composition of reduced carbon. In: Fritz, A., Fontes, P. MAPA, 1989. Caracterización agroclimática de la Provincia de Granada. Ed. Secretaría
(Eds.), The Terrestrial Environment. Handbook of Environmental Isotope General Técnica del MAPA. 197 pp.
Geochemistry. Elsevier Scientific Press, pp. 329–434. Martín-Ramos, J.D., 2004. Using XPowder®, A Software Package for Powder X-ray
Deleuze, M., Brantley, S.L., 1997. Inhibition of calcite crystal growth by Mg2+ at 100 °C Diffraction Analysis. D.L.GR-1001/0484-609-1497-6 (Spain).
and 100 bars: influence of growth regime. Geochimica et Cosmochimica Acta 61, McCrea, J.M., 1950. On the isotopic chemistry of carbonates and a paleotemperature
1475–1485. scale. Journal of Chemical Physics 18, 849–857.
Díaz-Hernández, J.L., 2010. Is carbon storage in soils underestimated? Chemosphere Mikhailova, E.A., Post, C.J., 2006. Effects of land use on soil inorganic carbon stocks in
80, 346–349. the Russian Chernozem. Journal of Environmental Quality 35, 1384–1388.
Díaz-Hernández, J.L., Juliá, R., 2006. Geochronological position of badlands and Mikhailova, E.A., Post, C.J., Cihacek, L., Ulmer, M., 2009. Soil inorganic carbon sequestration
geomorphological patterns in the Guadix–Baza basin (SE Spain). Quaternary as a result of cultivation in the mollisols. In: McPherson, B.J., Sundquist, E.T. (Eds.),
Research 65, 467–477. Carbon Sequestration and Its Role in the Global Carbon Cycle, Geophysical Monograph
Díaz-Hernández, J.L., Barahona-Fernández, E., Linares-González, J., 2003. Organic and Series 183. A.G.U., Washington, D.C., pp. 129–133.
inorganic carbon in soils of semiarid regions: a case study from the Guadix–Baza Ming, D.W., 2002. In: Lal, R. (Ed.), Encyclopedia of Soil Science. Marcel Dekker,
basin (Southeast Spain). Geoderma 114, 65–80. New York, pp. 139–141.
Díaz-Hernández, J.L., Martín-Ramos, J.D., López-Galindo, A., 2011. Quantitative analysis Mucci, A., Morse, J.W., 1983. The incorporation of Mg2+ and Sr2+ into calcite
of mineral phases in atmospheric dust deposited in the south-eastern Iberian overgrowths: influences of growth rate and solution composition. Geochimica
Peninsula. Atmospheric Environment 45, 3015–3024. et Cosmochimica Acta 47, 217–233.
Drees, L.R., Wilding, L.P., 1987. Micromorphic record and interpretations of carbonate Peña, J.A., 1985. La Depresión de Guadix–Baza. Estudios Geológicos 41, 33–46.
forms in the rolling plains of Texas. Geoderma 40, 157–175. Podwojewski, P., 1995. The occurrence and interpretation of carbonate and sulphate
Ehlmann, B.L., Mustard, J.F., Murchie, S.L., Poulet, F., Bishop, J.L., Brown, A.J., Calvin, minerals in a sequence of Vertisols in New Caledonia. Geoderma 65, 223–248.
W.M., Clark, R.N., Marais, D.J., Milliken, R.E., Roach, L.H., Roush, T.L., Swayze, G.A., Putnis, A., 2003. Introduction to Mineral Sciences. Cambridge University Press (457 pp.).
Wray, J.J., 2008. Orbital identification of carbonate-bearing rocks on mars. Science Putnis, A., Prieto, M., Fernández-Díaz, L., 1995. Fluid supersaturation and crystallization
322, 1828–1832. in porous media. Geological Magazine 132, 1–13.
FAO-UNESCO, 1988. Soil Map of the World. Revised Legend. FAO, Roma. Rech, J.A., Quade, J., Hart, W.S., 2003. Isotopic evidence for the source of Ca and S in soil
Fernández-Díaz, L., Putnis, A., Prieto, M., Putnis, C.V., 1996. The role of magnesium in gypsum, anhydrite and calcite in the Atacama Desert, Chile. Geochimica et
the crystallization of calcite and aragonite in a porous medium. Journal of Cosmochimica Acta 67, 575–586.
Sedimentary Research 66, 482–491. Reeder, R.J., Sheppard, C.E., 1984. Variation in lattice parameters in some sedimentary
Ferry, J.M., Ushikubo, T., Kita, N.T., Valley, J.W., 2010. Assessment of grain-scale homo- dolomites. American Mineralogist 69, 520–527.
geneity and equilibration of carbon and oxygen isotope compositions of minerals Reyes, E., Pérez del Villar, L., Delgado, A., Cortecci, G., Núñez, R., Pelayo, M., Cózar, J.,
in carbonate-bearing metamorphic rocks by ion microprobe. Geochimica et 1998. Carbonatation processes at the El Berrocal analogue granitic system
Cosmochimica Acta 74, 6517–6540. (Spain): mineralogical and isotopic study. Chemical Geology 150, 293–315.
Fritz, P., Smith, D.G.W., 1970. The isotopic composition of secondary dolomites. Rosenbaum, J., Sheppard, S.M.F., 1986. An isotopic study of siderites, dolomites and an-
Geochimica et Cosmochimica Acta 34, 1161–1173. kerites at high temperatures. Geochimica et Cosmochimica Acta 50, 1147–1150.
J.L. Díaz-Hernández et al. / Chemical Geology 347 (2013) 20–33 33

Salomons, W., Mook, W.G., 1976. Isotope geochemistry of carbonate dissolution and shifts in seawater chemistry. Palaeogeography Palaeoclimatology Palaeoecology
reprecipitation in soils. Soil Science 122, 15–24. 144, 3–19.
Salomons, W., Goudie, A., Mook, W.G., 1977. Isotopic composition of calcrete deposits Stevenson, B.A., Kelly, E.F., McDonald, E.V., Busacca, A.J., 2005. The stable carbon
from Europe, Africa and India. Earth Surface Processes and Landforms 3, 43–57. isotope composition of soil organic carbon and pedogenic carbonates along a
Sánchez-Navas, A., Martín-Algarra, A., Rivadeneyra, M.A., Melchor, S., Martín-Ramos, bioclimatic gradient in the Palouse region, Washington State, USA. Geoderma
J.D., 2009. Crystal-growth behaviour in Ca–Mg carbonate bacterial spherulites. 124, 37–47.
Crystal Growth & Design 9, 2690–2699. USDA-NRCS, 1996. Soil survey laboratory methods manual. S.S. Investigation Report
Shermann, D.G., Shultz, F., Always, F.J., 1962. Dolomitization of soils of the Red River No. 42, Version 3.0 (716 pp.).
Valley, Minnesota. Soil Science 94, 304–313. Usdowski, E., Hoefs, J., 1993. Oxygen isotope exchange between carbonic acid, bicarbonate,
Sobecki, T.M., Karathanasis, A.D., 1987. Quantification and compositional characterization carbonate, and water: a re-examination of the data of McCrea (1950) and an expres-
of pedogenic calcite and dolomite in calcic horizons of selected Aquolls. Soil Science sion for the overall partitioning of oxygen isotopes between the carbonate species
Society of America Journal 51, 683–690. and water. Geochimica et Cosmochimica Acta 57, 3815–3818.
Sohet, K., Herbauts, J., Gruber, W., 1988. Changes caused by Norway spruce in ochreous Vera, J.A., 1970. Estudio estratigráfico de la Depresión de Guadix–Baza. Boletín
brown earth, assessed by isoquartz method. Journal of Soil Science 39, 549–561. Geológico y Minero 81, 429–462.
Soil Survey Staff, 1999. Soil taxonomy, a basic system of soil classification for making Wacey, D., Wright, D.T., Boyce, A.J., 2007. A stable isotope study of microbial dolomite
and interpreting soil surveys. U.S. Department of Agriculture, Soil Conservation formation in the Coorong region, South Australia. Chemical Geology 244, 155–174.
Service, Agriculture Handbook No. 436, Washington, D.C. (754 pp.). Whipkey, Ch.E., Hayob, J.L., 2008. Textural and compositional evidence for the evolution of
Souchier, B., 1971. Evolution des sols sur roches cristallines à l'étage montagnard pedogenic calcite and dolomite in a weathering profile on the Kohala Peninsula,
(Vosges). Mémoires du Service de la Carte Géologique d'Alsace et de Lorraine 33, Hawaii. Carbonate Evaporite 23, 104–112.
134 pp. Wright, V.P., 1992. Paleosol recognition: a guide to early diagenesis in terrestrial
St. Arnaud, R.J., 1979. Nature and distribution of secondary soil carbonates within settings. In: Wolf, K.H., Chilingarian, G.V. (Eds.), Diagenesis, III: Amsterdam
landscapes in relation to soluble Mg2+/Ca2+ ratios. Canadian Journal of Soil Science (Elsevier): Developments in Sedimentol, 47, pp. 591–619.
59, 87–98. Wright, D.T., 1999. The role of sulphate-reducing bacteria and cyanobacteria in dolomite
St. Arnaud, R.J., Herbillon, A.J., 1973. Occurrence and genesis of magnesium-bearing formation in distal ephemeral lakes of the Coorong region, South Australia. Sedimen-
calcites in soils. Geoderma 9, 279–298. tary Geology 126, 147–157.
Stanley, S.M., Hardie, L.A., 1998. Secular oscillations in the carbonate mineralogy of Wright, V.P., Tucker, M.E., 1991. Calcretes. Ed. Blackwell, London (352 pp.).
reef-building and sediment-producing organisms driven by tectonically forced

You might also like