07.01.2015 Views

accepted manuscript

accepted manuscript

accepted manuscript

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

ÔØ ÅÒÙ×ÖÔØ<br />

Isotopic evidence for dolomite formation in soils<br />

J.L. Díaz-Hernández, A. Sánchez-Navas, E. Reyes<br />

PII:<br />

S0009-2541(13)00141-1<br />

DOI:<br />

doi: 10.1016/j.chemgeo.2013.03.018<br />

Reference: CHEMGE 16850<br />

To appear in:<br />

Chemical Geology<br />

Received date: 19 November 2012<br />

Revised date: 25 March 2013<br />

Accepted date: 28 March 2013<br />

Please cite this article as: Díaz-Hernández, J.L., Sánchez-Navas, A., Reyes, E.,<br />

Isotopic evidence for dolomite formation in soils, Chemical Geology (2013), doi:<br />

10.1016/j.chemgeo.2013.03.018<br />

This is a PDF file of an unedited <strong>manuscript</strong> that has been <strong>accepted</strong> for publication.<br />

As a service to our customers we are providing this early version of the <strong>manuscript</strong>.<br />

The <strong>manuscript</strong> will undergo copyediting, typesetting, and review of the resulting proof<br />

before it is published in its final form. Please note that during the production process<br />

errors may be discovered which could affect the content, and all legal disclaimers that<br />

apply to the journal pertain.


ACCEPTED MANUSCRIPT<br />

ISOTOPIC EVIDENCE FOR DOLOMITE FORMATION IN SOILS<br />

Díaz-Hernández, J.L. 1 , Sánchez-Navas, A. 2,3 , Reyes, E. 3<br />

1 IFAPA Camino de Purchil, Área de Recursos Naturales, Consejería de Agricultura, Pesca y Medio<br />

Ambiente, Junta de Andalucía, Apartado 2027, 18080 Granada, Spain. josel.diaz@juntadeandalucia.es<br />

2 Departamento de Mineralogía y Petrología, Universidad de Granada, Campus Fuentenueva, Granada,<br />

Spain.<br />

3 Instituto Andaluz de Ciencias de la Tierra IACT (CSIC-UGR), Avda. de las Palmeras, 18100 Armilla,<br />

Granada, Spain.<br />

Abstract<br />

Dolomite formation in soils constitute a particular challenge because of: 1) scant<br />

magnesium content in continental environments as opposed to the marine medium, 2)<br />

the kinetic problem related to the incorporation of magnesium into the carbonate, and 3)<br />

the unknown role of soil dolomites in the global carbon cycle.<br />

Pedogenic dolomite formed at deeper soil levels (subsoil) before the development of<br />

petrocalcic horizon barriers was investigated in a semiarid region of SE Spain (Guadix-<br />

Baza basin). Mineralogical characterization, textural relationships and isotopic data<br />

concerning soil dolomite, together with the results of a precipitation experiment,<br />

provided fuller knowledge of the processes and conditions governing neoformation of<br />

dolomite in these soils.<br />

In the study case, dolomite enrichment occurs beyond the limit of major biological<br />

activity, which coincides with the rooting depth of native perennial plants in the<br />

semiarid soils studied. Textural studies reveal the corrosion of inherited dolomite<br />

crystals in the upper soil horizons and the formation of dolomite in depth in relation to a<br />

ACCEPTED MANUSCRIPT<br />

clayey material, composed mainly of smectites. Stable isotope distribution in dolomites<br />

throughout the profiles indicates a fractionation with depth. This is explained by the<br />

formation of dolomites after the dissolution of the pedogenic calcite. The calcite<br />

detected in the subsoil is interpreted here as a precursor of the neoformed dolomites that<br />

transport the isotopic signal associated with biological activity of soils to deeper layers.<br />

Dolomite formation appears be favoured by the presence of clay minerals in the<br />

precipitation media. Clays retain water during evapotranspiration stages, which<br />

drastically change the transport properties of the media and promote the incorporation<br />

of Mg into the structure of the neoformed Ca,Mg-carbonate. As confirmed by<br />

laboratory experiments, diffusion-controlled crystal-growth processes lead to the<br />

1


ACCEPTED MANUSCRIPT<br />

formation a precursory “protodolomite” with disordered Ca,Mg distribution from a fluid<br />

locally supersaturated in dolomite.<br />

Keywords: Carbon-Oxygen isotopes, Dolomite problem, Inorganic carbon, Pedogenic<br />

processes, Semiarid soils, Soil carbon storage.<br />

1. INTRODUCTION<br />

The genesis of calcic and petrocalcic horizons in soils occurs as part of carbonate<br />

dissolution-precipitation cycles (Gile et al., 1966; Salomons and Mook, 1976). Detrital<br />

limestone and other calcareous materials undergoing decarbonation constitute a source<br />

for secondary carbonate accumulation in soils. Biogenic and atmogenic CO 2 are<br />

alternative sources for the carbon fixation in the soil under physico-chemical conditions<br />

stabilizing the carbonate ion. The pedogenic environment is strongly influenced by<br />

biological activity: root mats, in addition to associated lichens and microorganisms,<br />

contribute directly to mineral dissolution-precipitation in soils. In fact, pedogenesis<br />

begins with intense biological activity in the weathering zone (Wright and Tucker,<br />

1991; Wright, 1992; Cerling and Quade, 1993). Carbonic acid produced in association<br />

with biological activity (mainly roots) is the main acid involved in the dissolution of<br />

calcium carbonate. The leaching depth of the dissolved carbonate augments with<br />

increasing mean annual precipitation and temperature (Arkley, 1963; Stevenson et al.,<br />

2005). Carbonate precipitation at the deeper layers results largely from the declining<br />

P CO2 in the soil and from the higher concentrations of solutes in the soil solution due to<br />

evapotranspiration. In this sense, the carbonation process is favoured in arid-semiarid<br />

regimes. Large amounts of soil inorganic carbon, stored in the form of calcium<br />

carbonate, has been reported in cultivated mollisols under moderately cold temperatures<br />

(5-5.5ºC), and with a mean annual precipitation between 581-397 mm (Mikhailova and<br />

Post, 2006; Mikhailova et al., 2009), suggesting an increase in the carbonation process<br />

with agricultural practices.<br />

Isotopic relationships found in pedogenic carbonates in arid-semiarid ecosystems<br />

indicate that carbonate dissolution-precipitation cycles in a pedogenic environment<br />

provoke changes in the isotopic composition of the precipitated carbonate, by inducing<br />

significant inputs of biogenic carbon (Cerling, 1984). Biogenic carbon carries a unique<br />

isotopic signature, and the biological fractionation processes deplete 13 C to negative<br />

values (Salomons et al., 1977; Magaritz and Amiel, 1980; Magaritz et al., 1981).<br />

ACCEPTED MANUSCRIPT<br />

2


ACCEPTED MANUSCRIPT<br />

Therefore, the pedogenic carbonate will usually have more negative 13 C than detrital<br />

carbonates.<br />

Modern dolomites are formed only in certain environments such as meteoric, marine,<br />

hypersaline, subsurface brines, and hydrothermal (Hardie, 1987). The precipitation of<br />

dolomite and dolomitization of previous Ca-rich carbonates are among the leastunderstood<br />

problems encountered in the geochemistry of carbonates: e.g. surface<br />

seawater is strongly oversaturated in dolomite, but little evidence is available on the<br />

widespread dolomite precipitation in modern, open-marine sediments. To explain the<br />

“dolomite problem”, models have emerged ascribing microbial activity to the mediation<br />

of dolomite precipitation. Various authors have focused their attention on the role<br />

played by the bio-mediation that leads to CO 2 enrichment and raises alkalinity, and<br />

presumably drives dolomite formation in Mg-rich environments (e.g. Wright, 1999). In<br />

any case, the occurrence of dolomite in the geochemical systems of the earth’s surface<br />

is often strongly controlled by reaction kinetics (Arvidson and Mackenzie, 1999;<br />

Deelman, 2005) and/or mass-transport processes.<br />

The formation of dolomite in soils is rarely reported, its origin being related to parentmaterial<br />

inheritance, to atmospheric Mg 2+ wet deposition, and to the addition of<br />

dolomite contained in atmospheric dust (Ghebre-Egziabhier and St. Arnaud, 1983;<br />

Ming, 2002; Goddard et al., 2007; Díaz-Hernández, et al., 2011). The occurrence of<br />

neoformed dolomite in soils is usually restricted to saline environments (Shermann et.<br />

al., 1962; Kohut et al., 1995). However, the presence of high-Mg water solutions in<br />

soils derived from weathering of serpentinites and basaltic rocks may lead to the<br />

formation of authigenic dolomite in non-saline environments (Podwojewski, 1995;<br />

Capo et al., 2000). Although the formation of dolomite in soils has been discussed<br />

(Drees and Wilding, 1987; Sobecki and Karathanasis, 1987; Bui et al., 1990; Whipkey<br />

and Hayob, 2008), no mechanism explaining the precipitation of dolomite in soils has<br />

been reported. The incorporation of magnesium into the calcite structure to form high<br />

magnesian-calcites and proto-dolomites under surface conditions is broadly described in<br />

the literature (Hardie, 1987 and references therein; Deleuze and Brantley, 1997).<br />

Sedimentary and diagenetic dolomites probably precipitated as protodolomites and their<br />

isotopic composition was controlled by the protodolomite-water fractionation (Fritz and<br />

Smith, 1970). In any case, the incorporation of magnesium to the rhombohedral<br />

carbonate structure is not easy and appears to be controlled by kinetic factors (Mucci<br />

and Morse, 1983; Putnis et al., 1995; Fernández-Díaz et al., 1996).<br />

ACCEPTED MANUSCRIPT<br />

3


ACCEPTED MANUSCRIPT<br />

2. SCOPE OF THE INVESTIGATION<br />

In this paper, we study the occurrence of pedogenic dolomite in soils of an arid/semiarid<br />

region. Calcite is by far the most abundant and best-studied pedogenic carbonate. On<br />

the other hand, dolomite is not usually evaluated in soils and its role in the pedogenic<br />

processes is scarcely known. Here, we describe the formation of dolomite at soil depths<br />

below indurated layers delimitated by petrocalcic horizons or caliches. Deeper layers<br />

(subsoil) in these types of soils are scarcely studied because caliches are hard to<br />

excavate and their features at these depths are irrelevant for soil taxonomy or for<br />

judging the soil as a support for plants.<br />

Several reasons underlie this approach: 1) Knowledge of the distribution of carbon in<br />

the deep-soil-layer system in semiarid areas is essential to assess soil carbonates (calcite<br />

and dolomite) as a sink for atmospheric carbon. However, the carbon distribution in<br />

deeper layers (subsoil) has rarely been determined (Jobággy and Jackson, 2000; Díaz-<br />

Hernández et al., 2003). 2) Well-ordered dolomite, together with protodolomites, has<br />

recently been documented in soils (Whipkey and Hayob, 2008). 3) The soils of the<br />

study area are particularly rich in inherited dolomites and, therefore, could be suitable<br />

for the development of pedogenic dolomites in addition to calcites. 4) Periodic moisture<br />

movement to depths below solum horizons serves as a regulating factor in maintaining<br />

soluble Mg at levels favourable to the precipitation of low-magnesium calcites (St.<br />

Arnaud and Herbillon, 1973), and occasionally also leads to the formation of dolomites.<br />

5) The few studies available concerning pedogenic dolomites (St. Arnaud, 1979; Botha<br />

and Hughes, 1992; Capo et al., 2000) point to the occurrence of pedogenic dolomites in<br />

depth.<br />

The main objective of this research is to demonstrate the formation of dolomite in soils.<br />

For this, we firstly studied compositional, mineralogical, and textural variations<br />

throughout the soil profiles, finding that dolomite was dissolved in superficial horizons<br />

and precipitated at greater depths than was the pedogenic calcite. Also, we analysed the<br />

isotopes in calcite and dolomite of the soil samples. This second stage revealed that the<br />

biogenic isotopic signature of the pedogenic calcite was transferred to a neoformed<br />

dolomite through the dissolution-precipitation cycles that affected the soil carbonates.<br />

Finally, we discuss the physico-chemical constraints on the dolomite formation in<br />

subsoil. According to the model proposed for the precipitation of dolomite: 1) inherited<br />

dolomite constitutes the main source for the Mg 2+ ; 2) hydrodynamics in the soil studied<br />

ACCEPTED MANUSCRIPT<br />

4


ACCEPTED MANUSCRIPT<br />

controls moisture and solute transport properties in the subsoil; 3) the incorporation of<br />

magnesium to the carbonate structure in Mg-rich confined porous media to form<br />

dolomite is explained from the standpoint of kinetics.<br />

3. MATERIAL AND METHODS<br />

3.1. Geological setting and climate<br />

The Guadix-Baza basin is located in the Betic cordillera (SE Spain, Fig. 1A). The area<br />

studied was developed as an endorheic depression from the Late Neogene to the Late<br />

Pleistocene (Vera, 1970; Peña, 1985; García-Aguilar and Martín, 2000; Gibert et al.,<br />

2007, among others), and occupies about 3500km 2 . During this period, the basin was<br />

filled with alluvial and lacustrine materials for which the composition varies depending<br />

on the nature of source materials. Those from the Subbetic ranges (North border) are<br />

rich in limestone, while those from the Betic ranges (South border) are rich in quartzite,<br />

micaschist, and dolomite (IGME, 1973-1999). The Guadix sub-basin is comprised of<br />

fluvial sediments (silts, sands and conglomerates) and the Baza sub-basin is formed<br />

mainly by lacustrine deposits (limestones, marly limestones, and gypsum). The rocky<br />

relief surrounding the Guadix-Baza depression coincides partially with the Atlantic-<br />

Mediterranean divide. The depression is characterized by having elevations ranging<br />

from 1500m a.s.l. (above sea level) on the edges to 650m a.s.l. at the Guadiana Menor<br />

River, the main drain of the area.<br />

The capture of the internal drainage network of the depression by the Guadiana Menor<br />

River (a tributary of the Guadalquivir River) modelled the landscape into two main<br />

geomorphological groups (Díaz-Hernández and Juliá, 2006):<br />

1) Old surfaces (350,000-205,000 yr): S1, S2, S3, consisting of dissected piedmont<br />

plains gently sloping towards the centre of the basin (Fig. 1B), capped by caliche soils<br />

with thick petrocalcic horizons (Petric Calcisols and Chromic Luvisols (FAO-<br />

UNESCO, 1988) (Calciorthids, Paleorthids, Paleargids, Haploxeralfs, Rodoxeralfs and<br />

Xerochrepts, following the Soil Taxonomy (SSS, 1999)). Dolomite enrichment<br />

observed under the petrocalcic horizon (see below) suggests that the formation of<br />

dolomite began within this broad time range.<br />

2) The youngest surfaces (115,000-48,000 yr): S4 (badlands, with Gypsic, Calcaric and<br />

Eutric Regosols (FAO-UNESCO, 1988) (Xerortents, SSS, 1999) and S5 (alluvials, with<br />

Calcaric Fluvisols (FAO-UNESCO, 1988) (Xerofluvents, SSS, 1999).<br />

ACCEPTED MANUSCRIPT<br />

5


ACCEPTED MANUSCRIPT<br />

The climate of the basin is semiarid, hot in summer, with frequent frosts in winter. The<br />

mean annual temperature is 15ºC. Mean annual precipitation is about 250-300mm in the<br />

central part and about 400mm in the higher fans located along the mountain fronts<br />

(MAPA, 1989). Rainfall occurs mainly during the autumn-winter months; in summer,<br />

rain is scarce, falling as infrequent torrential storms.<br />

3.2. Sampling<br />

Soil profiles were studied at 81 sites in the Guadix-Baza basin (Fig. 1A), where the<br />

depth and features of the caliches could be accurately determined. Natural or artificial<br />

pits reached an average depth of about 200cm (in several cases down to 400cm) and the<br />

sampling sites were distributed as evenly as possible, in 10 × 10km cells to simulate a<br />

probabilistic sampling, with one or two sites selected per cell. Data from the deepest<br />

layers were not considered here because of the statistical results from the few data<br />

(N


ACCEPTED MANUSCRIPT<br />

The fine fraction (


ACCEPTED MANUSCRIPT<br />

3.6. Electron microscopy<br />

Two samples corresponding to the fine fraction of soil material located below the<br />

petrocalcic horizon at 120 and 220cm depths (profile 46) were selected on the basis of<br />

their dolomite content. They were consolidated and prepared as polished sections and<br />

later carbon coated for the textural and mineralogical study by analytical scanning<br />

electron microscopy. Back-scattered electron (BSE) images and energy-dispersion X-<br />

ray (EDX) spot analyses (200nm of diameter) were performed with a high-resolution<br />

field-emission scanning electron microscope (FESEM) Auriga (Carl Zeiss) equipped<br />

with a LinK INCA 200 (Oxford Instruments) microanalysis system operated at an<br />

acceleration voltage of 10-20kV.<br />

For transmission electron microscopy (TEM), the same


ACCEPTED MANUSCRIPT<br />

4. RESULTS<br />

4.1. Compositional, mineralogical and particle-size variations in the soil profiles<br />

The distribution of organic and inorganic carbon with depth varied considerably in the<br />

soils of the study region (Figs. 2A-C; see also Fig. 1 in Díaz-Hernández, 2010),<br />

reflecting the heterogeneity of the area, in which several soil types are represented<br />

(Díaz-Hernández et al., 2003). Profiles of organic and inorganic carbon (calcite and<br />

dolomite) distribution with depth for alluvial soils, badlands, and piedmont units are<br />

depicted in Figs. 2A-C. Organic carbon decreases with depth in all soil types, and it is<br />

particularly marked in the case of piedmonts (Fig. 2A). Petrocalcic horizon, represented<br />

by the enrichment in calcite content in depth, is detected mostly in piedmonts. This<br />

horizon appears between 10cm and 90cm depth, with 30-130cm of thickness (Fig. 2B,<br />

see also Díaz-Hernández et al., 2003). An increase of the dolomite content with depth is<br />

also better appreciated in piedmont profiles, usually below the petrocalcic horizon<br />

(compare Fig. 2B with 2C). In the soils studied, dolomite enrichment can occur at great<br />

depth (below 200cm), as is the case of the profile PC-43 of the Fig. 2C. This can be<br />

explained by the presence of Mg-rich substrates, constituted by altered ophitic rocks in<br />

the subsoil. The development of the petrocalcic horizons and the occurrence of<br />

associated dolomite enrichment in depth are better observed in the case of piedmonts.<br />

A total of 81 profiles were analysed, 8 corresponding to alluvial, 11 to badlands, and 62<br />

to piedmonts, in order to gain statistical information concerning carbon distribution in<br />

depth for the soils studied. Table 1 and Fig. 2D show the depth distribution of soil<br />

organic carbon content (from organic matter) and calcite and dolomite content (soil<br />

inorganic carbon), expressed as an average percentage. The average organic carbon<br />

decreased rapidly with depth, tending to near-zero values below 120cm (N = 81<br />

profiles), and continued to decline gradually to a depth of 220cm (N = 27 profiles). The<br />

calcite content initially increased with depth, showing minor amounts near the surface<br />

and a maximum between 40-60cm depth (43% in the bulge). Below this depth the<br />

calcite diminished gradually to 200cm in depth (around 30%). A second bulge at a<br />

depth of 200-280cm appeared from the profiles of the units of type S1. Polycyclic<br />

profiles showing this second maximum of calcite accumulation within this depth range<br />

were found for S1 unit (not shown). These minor maxima correspond to hard substrates<br />

or paleosols in those units, probably caused by regional sedimentological changes, and<br />

are not considered in this work. The average depth distribution of dolomite began with a<br />

low content (9% in the superficial horizons) which progressively increased from 40cm,<br />

ACCEPTED MANUSCRIPT<br />

9


ACCEPTED MANUSCRIPT<br />

reaching 17% at 140cm depth. However, the depth distribution of the dolomite and<br />

calcite contents proved somewhat irregular below 180cm due to the low number of<br />

samples for these depths (Table 1) Average particle-size distributions throughout the<br />

soil profiles studied show high contents of sand and gravels with respect clays and silts<br />

(Fig. 2E).<br />

The XRD study evidenced that the major minerals of soil samples were quartz, calcite,<br />

and dolomite, with minor amounts of clays, feldspars, and micas. For dolomite-rich<br />

samples, superstructure reflection (015) (the main reflection of well-ordered dolomites;<br />

Goldsmith and Graf, 1958; Reeder and Sheppard, 1984) was present. The depth<br />

distribution of quartz, calcite, and dolomite was shown in the XRD mapping of Fig. 3,<br />

corresponding to a depth sequence of diffractogram patterns for profile 46 (piedmont).<br />

This image shows a decrease in quartz with depth on the basis of the change in the<br />

intensity of the (101) reflection, and an interruption of the signal coinciding with the<br />

maximum of calcite enrichment (100cm depth). By contrast, the dolomite (104 peak)<br />

showed an inverse distribution to that of quartz (in-depth enrichment), with the same indepth<br />

interruption and lack in superficial horizon. Calcite enrichment between 60-<br />

130cm of depth, as indicated by the increase in the intensity of the (104) peak,<br />

coincided with the development of a petrocalcic horizon of 70cm thick in this profile.<br />

4.2. Isotopic data<br />

The frequency distribution of<br />

13 C and<br />

18 O of the calcite in the study samples are<br />

shown in Figs. 4A and B, respectively. These values were consistently negative, and<br />

13 C ranged from -13 to -1‰ and<br />

18 O from -10 to -2‰, with a normal distribution and<br />

ACCEPTED MANUSCRIPT<br />

modes around -7‰ in both cases. These negative values indicate a contribution of<br />

isotopically light carbonate. The histograms for the<br />

13 C and<br />

18 O of the dolomite<br />

showed bimodal distributions (Figs. 4C and D): the two modal values for<br />

13 C were -5<br />

and -1‰, and -6 and -3‰ for<br />

18 O. The scatter diagram relating both values for the<br />

dolomite samples also showed these two different tendencies in the histograms, and two<br />

clusters may be identified within the diagram (Fig. 4E). The average isotopic<br />

composition of the inherited dolomites from the coarse fraction of the studied soils is<br />

also represented in the bivariate plot (star in Fig. 4E). It plotted close to the analysed<br />

dolomite cluster with less negative and even positive values (cluster I in Fig. 4E).<br />

Samples of deeper layers (>70cm) defined the cluster with more negative<br />

13 C and<br />

18 O<br />

10


ACCEPTED MANUSCRIPT<br />

values (cluster II in Fig. 4E) that corresponded to the smallest left-hand maximum<br />

observed in the histograms (Figs. 4C and D). However, none of single profiles studied<br />

here showed a clear isotopic fractionation with the dolomite enrichment in depth.<br />

4.3. SEM and TEM study<br />

Electron microscopy confirmed the XRD observations of Fig. 3 regarding dolomite<br />

dissolution and calcite precipitation in relation to the formation of petrocalcic horizon.<br />

BSE images of Figs. 5A and 5B correspond to a sample below the petrocalcic horizon<br />

(120cm depth), where the corrosion affecting dolomite and quartz crystals and the<br />

precipitation of calcite are visible. In Fig. 5A the voids, after quartz dissolution, appear<br />

to be filled mainly with calcite. Frequently, calcification is widespread and calcite<br />

nodules occur with dolomite relicts within the concretions (Fig. 5B). The calcified<br />

nodules frequently had a highly porous texture. The dissolution process of inherited<br />

dolomites sometimes began along crystallographic planes and advanced from inner to<br />

outer regions with progressive corrosion, forming hollow dolomite crystals (Fig. 5C). In<br />

addition, corroded dolomite grains sometimes showed central voids (Fig. 5D). Calcite<br />

rhyzocretions also occurred close to the petrocalcic horizon (Fig. 5D).<br />

Samples from 220cm in depth showed corroded calcite grains surrounded by dolomite<br />

and clay minerals (Figs. 5E-F). Fig. 5F, corresponding to an inset of Fig. 5E, shows<br />

dolomite+clays secondary products located in a corrosion gulf which includes<br />

fragments of partially altered calcite relics. A detailed observation of precursory calcite<br />

evidenced the presence of a poorly crystalline material, basically clay minerals, close to<br />

irregular corrosion boundaries at the surface of calcite grains, in addition to inherited<br />

chlorites. Clay minerals also surrounded sub-idiomorphic dolomite crystals with sizes<br />

around 2 m in diameter, and sometimes localized preferentially between corroded<br />

calcite grains and rhombohedral dolomite crystals (Figs. 6A and 6B). Poorly defined<br />

crystalline material surrounding dolomite consisted of smectites plus Si-rich amorphous<br />

substance (Figs. 6C, 7 and Table 2). According to AEM analyses of Table 2 the<br />

smectite is intermediate between beidellite and montmorillonite, although several<br />

analyses correspond to mixtures of clay minerals and Si-rich amorphous substances. Si<br />

excess cannot be incorporated in the smectite structure and must therefore derive from<br />

adjacent non-crystalline material. Bubbles occurring in relation to clay minerals and<br />

amorphous substance in TEM image (Fig. 6C) resulted from the gradual release of<br />

volatile substances after pervasive electron bombardment. Clay minerals were also<br />

ACCEPTED MANUSCRIPT<br />

11


ACCEPTED MANUSCRIPT<br />

studied by XRD from separate fractions of the sample at 220cm depth (Fig. 8). This<br />

study confirms the occurrence of the following clay minerals: 1) An inherited chlorite is<br />

evidenced by (002) and (004) peaks (Fig. 8A; see also BSE image and qualitative<br />

microanalysis of Fig. 6B); 2) a detrital illite with peaks at 1.0 and 0.5nm d-spacing<br />

corresponding to (001) and (002) reflections; 3) abundant smectites and minor illitesmectite<br />

mixed-layer phases (probably R3-ordered I/S). The latter is evidenced by the<br />

broad basal reflection of the 1.0-1.2nm zone in the untreated sample (Fig. 8A). The<br />

smectite had a basal spacing of around 1.4nm in the air-dried state, which expands up to<br />

1.8nm upon ethylene glycol saturation (Fig. 8B).<br />

5. DISCUSSION<br />

5.1. Origin of compositional and mineralogical variations along the profiles<br />

Organic carbon shows the expected distribution (Figs. 2A and D) for these types of soils<br />

with a usual decrease with depth due to the more intense biogenic activity in the upper<br />

horizons. The combination of water from rainfall, Ca 2+ ions from diverse sources and<br />

CO 2 generated mostly from organic-matter decomposition and root respiration (Gile et<br />

al., 1966; Deines, 1980; Rech et al., 2003), the transient decrease in P CO2 and the<br />

increase in the concentration of solutes due to evapotranspiration lead to the<br />

accumulation in depth of pedogenic calcium carbonate in arid and semiarid regions. The<br />

formation of the petrocalcic horizon explains the occurrence of the bulge at 40-60cm<br />

depth in the calcite average regional profile (Fig. 2D), and the calcite enrichment found<br />

in the individual profiles shown in this work (Figs. 2B and 3). Although the dolomite<br />

content in the individual profiles was somewhat irregular (Fig. 2C), and in particular for<br />

depths below 180cm, it was evident that a regional decrease occurred near the surface<br />

with a marked enrichment in depth (Fig. 2D, Table1). Moreover, a smooth increase<br />

began at 40cm and continued to 140cm depth. Below 180cm depth, the values in the<br />

profile may have resulted from the interference of authigenic processes with the effects<br />

of the substratum on the mineralogy. Although the maximum in the case of the dolomite<br />

was less marked than in the calcite, the overall variation in dolomite (around 10%) was<br />

quite similar to the percentage change observed in calcite (around 12%). As occurred<br />

for calcite, dolomite dissolution/precipitation may explain the occurrence of the smooth<br />

maximum, just below the petrocalcic horizon, preceded by the change of sign in the<br />

curvature of the dolomite pattern in the Fig. 2D. Therefore, at the regional scale,<br />

dolomite underwent superficial dissolution and its precipitation occurred at greater<br />

ACCEPTED MANUSCRIPT<br />

12


ACCEPTED MANUSCRIPT<br />

depths than in the case of calcite. In short, the calcite-and dolomite-distribution patterns<br />

found here clearly indicate that they were affected by similar processes, but, in the case<br />

of dolomite, different dynamics seem to have occurred within the soil (see below).<br />

Dissolution of the dolomite in the superficial horizon of the studied profiles (e.g. Fig. 3)<br />

in relation to the development of petrocalcic horizon was also evidenced by Figs. 5A-D,<br />

where detrital dolomite appeared partially or almost completely dissolved and calcite<br />

precipitates in voids and formed rhyzocretions. A general decline in the quartz content<br />

was noted in the profile in the Fig. 3, contradicting the assumption that quartz content<br />

remains constant throughout the soil profiles (Lelong, 1969; Souchier, 1971; Sohet et<br />

al., 1988). The appearance of a physical barrier represented by the petrocalcic horizon<br />

prevents the redistribution of the upward-added quartz from diverse sources, mainly<br />

aeolian (Díaz-Hernández et al., 2011), and is therefore responsible for the observed<br />

decrease of quartz below it. However, the quartz content may decrease below the<br />

petrocalcic horizon as a result of dissolution processes, as evidenced by Fig. 5A, where<br />

the precipitation of calcite after quartz dissolution is evident. Calcite concretions after<br />

the dissolution dolomite and quartz had a marked porous texture (Figs. 5A and 5B)<br />

which may be interpreted in relation to bio-mediated precipitation processes.<br />

As mentioned above, dolomite enrichment below 40cm in depth was observed in the<br />

depth-wise variation of dolomite at the regional scale (Fig. 2D) and in the representative<br />

profiles of the Figs. 2C and 3. One possible explanation would imply no dolomite<br />

precipitation, but only dissolution; thus, the increment in-depth would be due to more<br />

intense dissolution for superficial horizons, the solutes being leached out of the system.<br />

On the other hand, in dissolution/precipitation mechanism proposed here, dolomite<br />

dissolved above 40cm is potentially precipitated below this depth. Figs. 5E-F and 6,<br />

corresponding to a sample situated at 220cm in depth in the studied profile, show the<br />

occurrence of dolomite and clay overgrowth in relation to calcite. Textural relations<br />

indicate the dissolution of calcite and the precipitation of dolomite in relation to clays.<br />

The specific mechanisms explaining the observed spatial association of calcite,<br />

dolomite and clays in SEM and TEM images are discussed below.<br />

ACCEPTED MANUSCRIPT<br />

5.2. Pedogenic isotopic fractionation and neoformed dolomites<br />

The proposed dolomite dissolution/precipitation model is consistent with the<br />

heterogeneity observed in the isotopic data from the dolomite samples analysed, which<br />

showed two different signals (Figs. 4C, 4D, and 4E). Cluster I in Fig. 4E, with higher<br />

13


ACCEPTED MANUSCRIPT<br />

isotopic values, lies close to the parent material that supplies the dolomite to the system<br />

(star in Fig. 4E). The precipitation of a neoformed dolomite below the petrocalcic (Figs.<br />

5E, 5F and 6) as a result of dissolution of inherited dolomite and pedogenic calcite of<br />

upper horizons explains the high dispersion observed in the δ 13 C values of the samples<br />

of deeper layers (cluster II in Fig. 4E and smallest left-hand maximum in the histograms<br />

of Figs. 4C and D). These values define the MDL (Meteoric Dolomite Line), a term<br />

similar to that proposed by Lohmann (1988) for the Meteoric Calcite Line (MCL). The<br />

C dispersion is explained by a mixing of inherited carbon from dolomite-rich parent<br />

material, with positive values, and pedogenic carbon from neoformed dolomite, with<br />

very negative C values. In addition, O values close to 6‰ (V-PDB) observed for<br />

pedogenic dolomites (cluster II in Fig. 4E) are consistent with the O measured in<br />

surface waters, values ranging from -7.5 and -9‰ (V-SMOW), and temperatures<br />

between ~17 and 24ºC.<br />

The isotopic composition of the dissolved inorganic carbon (DIC) and oxygen in soil<br />

waters below the petrocalcic horizon is in equilibrium with the isotopic composition of<br />

neoformed dolomites. Carbon and oxygen are distributed among three chemical species<br />

with a pH-dependent relative abundance (CO 2 , HCO - 3 and CO 2- 3 ). Most of the carbonic<br />

acid derives from the vegetal cover and prompts an increase of negative C in DIC<br />

(Cerling, 1984, 1991; Usdowski and Hoefs, 1993; Reyes et al., 1998). In the pedogenic<br />

medium studied, this isotopic signature is transported towards deeper layers through<br />

carbonate dissolution (mostly calcite) above the petrocalcic horizon. The negative<br />

carbon isotope composition of the neoformed dolomite is determined mainly by the<br />

isotopic pedogenic calcite because pore fluids generally have low carbon contents so<br />

ACCEPTED MANUSCRIPT<br />

that the 13 C of the precursor is generally retained (Hoefs, 2004). Low<br />

13 C and<br />

values are expected in biomediated calcite and dolomite precipitates (Jones and<br />

Donnelly, 2004; Wacey et al., 2007). Multiple cyclic dissolution/precipitation processes<br />

that affected the carbonates in the soils explain not only the formation of an isotopically<br />

light calcite, related to pedogenic-biogenic processes above the petrocalcic horizon<br />

(Magaritz and Amiel, 1980; Magaritz et al., 1981), but also the formation of an<br />

isotopically light dolomite, occurring below the petrocalcic horizon, in which the<br />

isotopic composition differs from that observed for the inherited dolomites (Fig. 4). The<br />

isotopic values of the pedogenic calcite co-existing with neoformed dolomite may be<br />

assigned to the maximum observed in the histograms of Figures 4A-B. Calcite isotopic<br />

18 O<br />

14


ACCEPTED MANUSCRIPT<br />

values are more negative than those of dolomite, as expected, because most of the<br />

calcite in soil is pedogenic in origin. The pedogenic dolomite precipitated in depth has a<br />

marked isotopic contribution from inherited dolomite, and therefore lower negative<br />

values than those measured for the pedogenic calcite are expected.<br />

Dolomite isotopic composition is not usually determined in soils, and much less in deep<br />

layers (“subsoil”). Until now, the occurrence of a pedogenic 13 C (and 18 O) signal in soils<br />

is associated exclusively with the calcite genesis. In the case studied, the neoformed<br />

dolomite, characterized by relatively negatives values, forms after partial dissolution of<br />

the calcite, with a partial “transference” of the biogenic isotopic signature. Therefore,<br />

calcite precipitation is interpreted here as an intermediate stage for dolomite<br />

precipitation, below the petrocalcic horizon, because of the carbonate dissolutionprecipitation<br />

cycles forming pedogenic calcite are also responsible of the formation of<br />

the pedogenic dolomites.<br />

Kinetic effects, such as diffusion transport, may also induce a certain fractionation in<br />

soils. The carbon isotope fractionation in soils caused by diffusion processes has been<br />

estimated to be around 4‰ (Cerling, 1984; Hesterberg and Siegenthaler, 1991).<br />

According to these authors 13 CO 2 diffuses into the atmosphere at lower rates than the<br />

lighter 12 CO 2 molecule. However this effect is clearly counteracted by the major 12 CO 2<br />

production related to root respiration and organic-matter decomposition in the upper<br />

horizons of soil, as evidenced clearly by the lower<br />

13 C values recorded for the<br />

pedogenic calcite of soil in relation to the pedogenic dolomite occurring in the subsoil.<br />

The kinetic effect discussed above resulting in 13 C enrichment would partly explain the<br />

shift from the very negative<br />

13 C at the C3 vegetal cover (~ -27‰ vs. V-PDB, Deines,<br />

ACCEPTED MANUSCRIPT<br />

1980) due to CO 2 from organic matter decomposition, toward values around -12‰ in<br />

the soil (Cerling, 1984; 1991; Reyes et al., 1998). Other kinetic processes such as<br />

intracrystalline diffusion of C and O within the dolomite structure might explain isotope<br />

exchange between the solid and the pore fluid in soils, and the observed fractionation in<br />

Fig. 4, without the need for an explanation involving the dolomite dissolutionprecipitation<br />

process. However, alternative explanations of these isotopic data based on<br />

intracrystalline diffusion processes should be discarded in our case because of: 1)<br />

According to the Arrhenius relationship between the diffusion rate and the temperature,<br />

intracrystalline diffusion is negligible at low temperature (Putnis, 2003). 2) Volume<br />

diffusion of C within the structure of the rhombohedral carbonates occurs by diffusion<br />

of carbonate anions, whereas the lattice diffusion of O is due to vacancy migration<br />

15


ACCEPTED MANUSCRIPT<br />

(Labotka et al., 2000); this explains why the diffusivity of carbon is 4 orders of<br />

magnitude smaller than the diffusivity of oxygen in calcite (Ferry et al., 2010). The<br />

latter point disagrees with our observations, where a broader range of variation is found<br />

for<br />

13 C compared to that of<br />

dolomite located at deeper levels.<br />

18 O for the cluster II (Fig. 4E) corresponding to the<br />

5.3. Physico-chemical constraints on the dolomite formation in the subsoil<br />

As shown, dissolution of inherited dolomite above petrocalcic constitutes the main<br />

source for the Mg 2+ enrichment in the soil solution necessary for dolomite<br />

neoformation. An exception to this is provided by the profiles with altered ophitic rocks<br />

in their subsoil, where minor dolomite was formed at great depth (below 200cm, Fig.<br />

2C). The montmorillonitic component of the smectites localized preferentially between<br />

corroded calcite grains and rhombohedral dolomite crystals (Table 2 and Figs. 6A-B)<br />

may constitute an additional source for Mg 2+ . Despite the occurrence of high Mg:Ca<br />

mole ratio, Mg rich calcite form only at very high salinity (Stanley and Hardie, 1998).<br />

However, for medium salinities, a low Mg content was expected in the calcite formed<br />

from solutions in soils. This is because Mg 2+ ions tend to remain in solution during<br />

carbonate precipitation, where they are thought to be heavily solvated (de Leeuw and<br />

Parker, 2001), due to strongly attractive Mg-water interactions.<br />

The formation of the petrocalcic horizon prevents rain-water infiltration into the subsoil.<br />

The observed dolomite enrichment in the subsoil probably occurred during the initial<br />

stages of the formation of the petrocalcic horizon (soft caliches), and before its<br />

hardening that led to the consolidation of the impermeable barrier. The high content of<br />

gravels and sands throughout the profile (Fig. 2E) favoured rainwater infiltration into<br />

the subsoil. Under these physical conditions, evapotranspiration processes affected not<br />

only the soil, but also deeper levels during the less rainy climatic episodes. The<br />

decreased water content in the subsoil diminished bulk flow of the grain-boundary fluid<br />

containing dissolved solutes. Under this regime, the fluid film is restricted to a<br />

monolayer of molecules adsorbed onto the smectite particles. To understand the role of<br />

the water molecules adsorbed onto smectites in the incorporation of Mg to the structure<br />

of the calcite, we dealt here with the transport of solute in a viscous solid media (Putnis<br />

et al., 1995; Fernández-Díaz et al., 1996; Sánchez-Navas et al., 2009). It can be deduced<br />

from Table 3 that the quantity of Mg incorporated into the structure of the calcite<br />

increased in viscous medium and that the incorporation of Mg into magnesian-calcite<br />

ACCEPTED MANUSCRIPT<br />

16


ACCEPTED MANUSCRIPT<br />

produced a lattice contraction of the rhombohedral carbonate, as well as a reduction of<br />

crystallinity. In addition, metastable phases also appeared when a viscous medium was<br />

used for the precipitation experiment (Fig. 9). Micro-textural data (Fig. 6) indicate that<br />

the precipitation of neoformed dolomite in depth after partial dissolution of pedogenic<br />

calcite was mediated by the smectites of the soils studied (Figs. 7 and 8). Clay minerals<br />

surrounding neoformed dolomite crystals acted as a catalytic material in the process of<br />

incorporation of Mg into the structure of the rhombohedral carbonates. This was<br />

because the transport of reacting material from dissolution sites to those of growth was<br />

the rate-determining process. Therefore, as suggested by our laboratory experiments, it<br />

is necessary to consider the constraints that transport properties may make on the<br />

incorporation of an element during crystal growth. Fig. 10A shows the concentration<br />

profile of Mg at the crystal-water interface during interface-controlled calcite growth.<br />

As suggested at the beginning of this section, the partition coefficient between the<br />

concentration of magnesium in carbonate crystal and the concentration of magnesium in<br />

bulk medium, [Mg] carb /[Mg] m , must be clearly less than 1. For a diffusion-controlled<br />

growth of dolomite, the concentration of Mg rises above [Mg] m at the crystal-medium<br />

interface ([Mg]i m ). If the partition coefficient remains constant, [Mg] carb increases with<br />

decreasing diffusion rate until, in the limit, [Mg] carb =[Mg] m . As a result the effective<br />

partition coefficient equals 1, and therefore most of the Mg of the medium is then<br />

incorporated into the structure of the rhombohedral carbonate (Fig. 10B).<br />

Previous experimental data and the occurrence in soils of high-magnesium calcite<br />

together with well-ordered dolomite (Whipkey and Hayob, 2008) suggest that our<br />

neoformed dolomite that precipitated in relation to smectites in the “subsoil” should be,<br />

at least initially, a “protodolomite” or “Mg-kutnohorite” with essentially disordered Ca-<br />

Mg distribution.<br />

ACCEPTED MANUSCRIPT<br />

6. CONCLUSIONS<br />

The present study detected the presence of a clear enrichment of dolomite beyond the<br />

lower limit of biological activity in semiarid soils. This limit coincides with the rooting<br />

depth of native perennial plants and, in the case study, is defined by the occurrence of<br />

the petrocalcic horizon.<br />

Isotopic data establish a bimodal distribution of<br />

13 C and<br />

18 O for dolomites, indicating<br />

the occurrence of an inherited and a neoformed dolomite. Isotopically light dolomite<br />

formed below the petrocalcic horizon after partial dissolution of the pedogenic calcite,<br />

17


ACCEPTED MANUSCRIPT<br />

which transferred the pedogenic (biological) 13 C and 18 O signature from the soil (s. str.)<br />

to subsoil.<br />

Before the formation of nonpermeable petrocalcic horizon, the high contents in gravel<br />

and sand fractions in the studied soils favoured infiltration and evapotranspiration<br />

processes in the subsoil. The moisture in the subsoil was a determining factor<br />

controlling the precipitation of Ca and Mg carbonates, the precursors of dolomite. The<br />

precipitation of the neoformed dolomite below the petrocalcic horizon was related to the<br />

presence of smectites. Smectite retained soil water and acted as a catalyst for the<br />

incorporation of Mg into the carbonate.<br />

The geochemical model proposed here for the incorporation of Mg to the carbonate in<br />

slow diffusive transport medium, such as a clayey-rich and water deficient<br />

environments, would also explain the close association of Mg-carbonates with clays<br />

observed on the surface of Mars (Ehlmann et al., 2008).<br />

This study documents the occurrence of changes in carbonate mineralogy in depth.<br />

Levels below the petrocalcic are traditionally considered as zones not affected by the<br />

pedogenic processes.<br />

In the soils studied, the dolomite content reached 43 kg m -2 (Díaz-Hernández, 2010),<br />

and therefore dolomite should be considered in the quantification of carbon stored in<br />

arid-semiarid soils. Pedogenic processes occurring in the subsoil, such as the dolomite<br />

formation, should be included in the carbon-soil balances.<br />

ACKNOWLEDGMENTS<br />

We acknowledge support from grant P11-RNM-7067 of the Junta de Andalucía<br />

(C.E.I.C.-S.G.U.I.T.) and the projects CGL2007-65572-C02-01/BTE, CGL-2009-<br />

09249, CGL2010-21257-C02-01 and CGL2012-32169 of the Ministerio de Ciencia e<br />

Innovación (MICINN, Spain). We would like to thank Drs. D. Martín-Ramos and F.<br />

Nieto for help with the XRD analyses, A. González, M.M. Abad and M.J. Guerrero for<br />

guidance with SEM and TEM studies and Dr. A. Delgado for his assistance with the<br />

stable isotope data. We thank the anonymous reviewers for their suggestions, which<br />

contributed to improve the <strong>manuscript</strong>. Mr. David Nesbitt revised the English<br />

<strong>manuscript</strong>.<br />

ACCEPTED MANUSCRIPT<br />

REFERENCES<br />

Al-Aasm, I.S., Taylor, B.E., South, B., 1990. Stable isotope analysis of multiple<br />

18


ACCEPTED MANUSCRIPT<br />

carbonate samples using selective acid extraction. Chemical Geology 80, 119-125.<br />

Arkley, R.J., 1963. Calculation of carbonate and water movement in soil from climatic<br />

data. Soil Science 96, 239-248.<br />

Arvidson, R.S., Mackenzie, F.T., 1999. The dolomite problem: control of precipitation<br />

kinetics by temperature and saturation state. American Journal of Science 299,<br />

257-288.<br />

Botha, G.A., Hughes, J.C., 1992. Pedogenic palygorskite and dolomite in a late<br />

Neogene sedimentary succession, nothwestern Transvaal, South Africa. Geoderma<br />

53, 139-154.<br />

Bui, E.N., Loeppert, R.H., Wilding L.P., 1990. Carbonate phases in calcareous soils of<br />

the western United States. Soil Science Society of America Journal 54, 520-529.<br />

Capo, R.C., Whipkey, C.E., Blachère, J.R., Chadwick, O.A., 2000. Pedogenic origin of<br />

dolomite in a basaltic weathering profile, Kohala peninsula, Hawaii. Geology 28,<br />

271-274.<br />

Cerling, T.E., 1984. The estable isotopic composition of modern soil carbonates and its<br />

relationships to climate. Earth and Planetary Science Letters 71, 229-240.<br />

Cerling, T.E., 1991. Carbon dioxide in the atmosphere: evidence from cenozoic and<br />

mesozoic paleosols. American Journal of Science 291, 377-400.<br />

Cerling, T.E., Quade, J., 1993. Stable carbon and oxygen isotopes in soil carbonates. In:<br />

Swart, P.K., Lohmann, K.C., Mckenzie, J., Savin, S. (Eds.), Climate Change in<br />

Continental Isotopic Records. Geophysical Monograph, vol. 78. American<br />

Geophysical Union, Washington, pp. 217-231.<br />

de Leeuw, N.H., Parker, S.C., 2001. Surface-water interactions in the dolomite problem.<br />

Physical Chemistry and Chemical Physics 3, 3217-3221.<br />

Deelman, J.C. 2005. Low temperature formation of dolomite and magnesite. Geology<br />

Series, version 2.1. Eindhoven: Compact Disc Publication.<br />

Deines, P., 1980. The isotopic composition of reduced carbon. In: A. Fritz and P. Fontes<br />

(Eds.). The terrestrial environment. Handbook of environmental isotope<br />

geochemistry: Elsevier Scientific Press, 329-434.<br />

Deleuze, M., Brantley, S.L., 1997. Inhibition of calcite crystal growth by Mg 2+ at 100°C<br />

and 100 bars: Influence of growth regime. Geochimica et Cosmochimica Acta 61,<br />

1475-1485.<br />

Díaz-Hernández, J.L. 2010. Is carbon storage in soils understimated. Chemosphere 80,<br />

346–349.<br />

ACCEPTED MANUSCRIPT<br />

19


ACCEPTED MANUSCRIPT<br />

Díaz-Hernández, J.L., Barahona-Fernández, E., Linares-González, J., 2003. Organic and<br />

Inorganic Carbon in soils of semiarid regions: a case study from the Guadix-Baza<br />

basin (Southeast Spain). Geoderma 114, 65-80.<br />

Díaz-Hernández, J.L., Juliá, R., 2006. Geochronological position of badlands and<br />

geomorphological patterns in the Guadix-Baza basin (SE Spain). Quaternary<br />

Research 65, 467-477.<br />

Díaz-Hernández, J.L., Martín-Ramos, J.D., López-Galindo, A., 2011. Quantitative<br />

analysis of mineral phases in atmospheric dust deposited in the south-eastern<br />

Iberian Peninsula. Atmospheric Environment 45, 3015-3024.<br />

Drees L.R., Wilding, L.P., 1987. Micromorphic record and interpretations of carbonate<br />

forms in the rolling plains of Texas. Geoderma 40, 157-175.<br />

Ehlmann, B.L., Mustard, J.F., Murchie, S.L., Poulet, F., Bishop, J.L., Brown, A.J.,<br />

Calvin, W.M., Clark, R.N., Marais, D.J., Milliken, R.E., Roach, L.H., Roush, T.L.,<br />

Swayze, G.A., Wray, J.J., 2008. Orbital Identification of Carbonate-Bearing<br />

Rocks on Mars. Science 322, 1828-1832.<br />

FAO-UNESCO, 1988. Soil Map of the World. Revised Legend. FAO, Roma.<br />

Fernández-Díaz, L., Putnis, A., Prieto, M., Putnis, C.V., 1996. The role of magnesium<br />

in the crystallization of calcite and aragonite in a porous medium. Journal of<br />

Sedimentary Research 66, 482-491.<br />

Ferry, J.M., Ushikubo, T., Kita, N.T., Valley, J.W., 2010. Assessment of grain-scale<br />

homogeneity and equilibration of carbon and oxygen isotope compositions of<br />

minerals in carbonate-bearing metamorphic rocks by ion microprobe. Geochimica<br />

et Cosmochimica Acta 74, 6517-6540.<br />

Fritz, P., Smith D.G.W., 1970. The isotopic composition of secondary dolomites:<br />

Geochimica et Cosmochimica Acta 34, 1161-1173.<br />

García-Aguilar, J.M., Martín J.M., 2000. Late Neogene to recent continental history and<br />

evolution of the Guadix-Baza basin (SE Spain). Revista de la Sociedad Geológica<br />

de España 13, 65-77.<br />

Ghebre-Egziabhier, K., St. Arnaud, R.J., 1983. Carbonate mineralogy of lake sediments<br />

and surrounding soils. 2. The Qu’Appelle Lakes. Canadian Journal of Soil Science<br />

63, 259-268.<br />

Gibert, L., Ortí, F., Rosell, L., 2007. Plio-Pleistocene lacustrine evaporates of the Baza<br />

Basin (Betic Chain, SE Spain). Sedimentary Geology 200, 89-116.<br />

Gile, L.H., Peterson, F.F., Grossman, R.B., 1966. Morphological and genetic sequences<br />

ACCEPTED MANUSCRIPT<br />

20


ACCEPTED MANUSCRIPT<br />

of carbonate accumulation in desert soils. Soil Science 1015, 347-360.<br />

Goddard, M.A., E.A. Mikhailova, C.J. Post, and M.A. Schlautman. 2007. Atmospheric<br />

Mg 2+ wet deposition within the continental United States and implications for soil<br />

inorganic carbon sequestration. Tellus B 59, 50-56.<br />

Goldsmith, J.R., Graf, D.L., 1958. Structural and compositional variations in some<br />

natural dolomites. Journal of Geology 66, 678-693.<br />

Hardie, L.A., 1987. Perspectives dolomitization: a critical view of some current views.<br />

Journal of Sedimentary Petrology 57, 166-183.<br />

Henderson, P., 1986. Inorganic geochemistry. Pergamon Press, Oxford, 356 pp.<br />

Hesterberg, R., Siegenthaler, U., 1991. Production and stable isotopic composition of<br />

CO 2 in a soil near Bern. Switzerland. Tellus 43B, 197-205.<br />

Hoefs, J., 2004. Stable isotope geochemistry, Springer-Verlag, Berlin Heidelberg, 244<br />

pp.<br />

IGME (1973-1999) Mapa Geológico de España, escala 1:50.000, Hojas 929, 930, 950,<br />

951, 972, 973, 993, 994, 1011 and 1012.<br />

Irwin, H., Curtis, C., Coleman, M., 1977. Isotopic evidence for source of diagenetic<br />

carbonates formed during burial of organic-rich sediments. Nature. 269, 209-213.<br />

Jobággy, E.G., Jackson, R.B., 2000. The vertical distribution of soil organic carbon and<br />

its relation to climate and vegetation. Ecological Applications 10, 423-436.<br />

Jones, M.B., Donnelly, A., 2004. Carbon sequestration in temperate grassland<br />

ecosystems and the influence of management, climate and elevated CO 2 . New<br />

Phytologist 164, 423-439.<br />

Kim, S.T., O´Neil, J.R., 1997. Equilibrium and nonequilibrium oxygen isotope effects<br />

in synthetic carbonates. Geochimica et Cosmochimica Acta 61, 3461-3475.<br />

Kohut, C., Dudas, M.J., Muehlenbachs, K., 1995. Authigenic Dolomite in a Saline Soil<br />

in Alberta, Canada. Soil Science Society of American Journal 59, 1499-1504.<br />

Labotka, T.C., Cole, D.R., Riciputi, L.R., 2000. Diffusion of C and O in calcite at 100<br />

MPa. American Mineralogist 85, 488-494.<br />

Lelong, F., 1969. Nature et gènèse des produits d'altération de roches cristallines sous<br />

climat tropical humide (Guyane Française). Mèmoires Science de la Terre, Nancy,<br />

14, 187 pp.<br />

Loeppert, R.H., Suárez, D.L., 1996. Carbonate and gypsum. In: Sparks, D.L. (Ed),<br />

Methods of Soil Analysis, part 3 Chemical Analysis, SSSA Book series, No 5,<br />

ASA & SSSA, Madison, WI, 437-474.<br />

ACCEPTED MANUSCRIPT<br />

21


ACCEPTED MANUSCRIPT<br />

Lohmann, K.Z., 1988. Geochemical patterns of meteoric diagenetic systems and their<br />

application to studies of paleokarst. In: James N.P., Choquette P.W. (Eds.),<br />

Paleokarst: Springer-Verlag, 58-80, Berlin.<br />

Lorimer, G.W., Cliff, G., 1976. Analytical electron microscopy of minerals. In: Electron<br />

microscopy in Mineralogy (Ed. H.R. Wenk), pp. 506-519. Springer-Verlag,<br />

Berlin.<br />

Magaritz, M., Amiel, A.J., 1980. Calcium carbonate in a calcareous soil from the Jordan<br />

Valley, Israel: Its origin as revealed by the stable carbon isotope method. Soil<br />

Science Society of American Journal 44, 1059-1062.<br />

Magaritz, M., Kaufman, A., Yaalon, D.H., 1981. Calcium carbonate nodules in soils:<br />

18 O/ 16 O and 13 C/ 12 C ratios and 14 C contents. Geoderma 25, 157-172.<br />

MAPA, 1989. Caracterización agroclimática de la Provincia de Granada. Ed. Secretaría<br />

General Técnica del MAPA. 197 pp.<br />

Martín-Ramos, J.D., 2004. Using XPowder®, a software package for powder X-Ray<br />

diffraction analysis. D.L.GR-1001/04. Spain. ISBN: 84-609-1497-6.<br />

Mikhailova, E.A., Post, C.J., 2006. Effects of land use on soil inorganic carbon stocks in<br />

the russian Chernozem. Journal of Environmental Quality 35, 1384-1388.<br />

Mikhailova, E.A., Post, C.J., Cihacek, L., Ulmer, M., 2009. Soil inorganic carbon<br />

sequestration as a result of cultivation in the mollisols, in Carbon sequestration<br />

and its role in the global carbon cycle, Geophysical Monograph Series 183, edited<br />

by B.J. McPherson and E.T. Sundquist, 129-133, A.G.U., Washington, D.C.<br />

Ming, D.W. 2002. In: Encyclopedia of Soil Science (eds R. Lal). Marcel Dekker, New<br />

York, 139-141.<br />

McCrea, J.M., 1950. On the isotopic chemistry of carbonates and a paleotemperature<br />

scale. Journal of Chemical Physics 18, 849-857.<br />

Mucci, A., Morse, J.W., 1983. The incorporation of Mg 2+ and Sr 2+ into calcite<br />

overgrowths: influences of growth rate and solution composition Geochimica et<br />

Cosmochimica Acta 47, 217-233.<br />

Peña, J.A., 1985. La Depresión de Guadix-Baza. Estudios Geológicos 41, 33-46.<br />

Podwojewski, P., 1995. The occurrence and interpretation of carbonate and sulphate<br />

minerals in a sequence of vertisols in New Caledonia. Geoderma 65, 223-248.<br />

Putnis, A., 2003. Introduction to Mineral Sciences. Cambridge University Press, 457 pp.<br />

Putnis, A., Prieto, M., Fernández-Díaz, L., 1995. Fluid supersaturation and<br />

crystallization in porous media. Geological Magazine 132, 1-13.<br />

ACCEPTED MANUSCRIPT<br />

22


ACCEPTED MANUSCRIPT<br />

Reeder, R.J., Sheppard, C.E., 1984. Variation in lattice parameters in some sedimentary<br />

dolomites. American Mineralogist 69, 520-527.<br />

Rech, J.A., Quade, J., Hart, W.S., 2003. Isotopic evidence for the source of Ca and S in<br />

soil gypsum, anhydrite and calcite in the Atacama Desert, Chile. Geochimica et<br />

Cosmochimica Acta 67, 575-586.<br />

Reyes, E., Pérez del Villar, L., Delgado, A., Cortecci, G., Núñez, R., Pelayo, M., Cózar, J.,<br />

1998. Carbonatation processes at the El Berrocal analogue granitic system (Spain):<br />

mineralogical and isotopic study. Chemical Geology 150, 293-315.<br />

Rosenbaum, J., Sheppard, S.M.F., 1986. An isotopic study of siderites, dolomites and<br />

ankerites at high temperatures. Geochimica et Cosmochimica Acta 50, 1147-1150.<br />

Salomons, W., Mook, W.G., 1976. Isotope geochemistry of carbonate dissolution and<br />

reprecipitation in soils. Soil Science 122, 15-24.<br />

Salomons, W., Goudie, A., Mook, W.G., 1977. Isotopic composition of calcrete<br />

deposits from Europe, Africa and India. Earth Surface Processes and Landforms 3,<br />

43-57.<br />

Sánchez-Navas, A., Martín-Algarra, A., Rivadeneyra, M.A., Melchor, S., Martín-Ramos<br />

J.D., 2009. Crystal-growth behaviour in Ca-Mg carbonate bacterial spherulites.<br />

Crystal Growth and Design 9, 2690-2699.<br />

Shermann, D.G., Shultz, F., Always, F.J., 1962. Dolomitization of soils of the Red<br />

River Valley, Minnesota. Soil Science 94, 304-313.<br />

Sobecki, T.M., Karathanasis, A.D., 1987. Quantification and compositional<br />

characterization of pedogenic calcite and dolomite in calcic horizons of selected<br />

Aquolls. Soil Science Society of American Journal 51, 683-690.<br />

Sohet, K., Herbauts, J., Gruber, W., 1988. Changes caused by Norway spruce in<br />

ochreous brown earth, assessed by isoquartz method. Journal of Soil Science 39,<br />

549-561.<br />

Soil Survey Staff, 1999. Soil taxonomy, a basic system of soil classification for making<br />

and interpreting soil surveys. U.S. Department of Agriculture, Soil Conservation<br />

Service, Agriculture Handbook no. 436, Washington, D.C., 754 pp.<br />

Souchier, B., 1971. Evolution des sols sur roches cristallines à l'étage montagnard<br />

(Vosges). Mémoires du Service de la Carte Géologique d’Alsace et de Lorraine<br />

33, 134 pp.<br />

St. Arnaud, R.J., 1979. Nature and distribution of secondary soil carbonates within<br />

landscapes in relation to soluble Mg 2+ /Ca 2+ ratios. Canadian Journal of Soil<br />

ACCEPTED MANUSCRIPT<br />

23


ACCEPTED MANUSCRIPT<br />

Science 59, 87-98.<br />

St. Arnaud, R.J., Herbillon, A.J., 1973. Occurrence and genesis of magnesium-bearing<br />

calcites in soils. Geoderma 9, 279-298.<br />

Stanley, S.M., Hardie, L.A., 1998. Secular oscillations in the carbonate mineralogy of<br />

reef-building and sediment-producing organisms driven by tectonically forced<br />

shifts in seawater chemistry. Palaeogeography, Palaeoclimatology, Palaeoecology<br />

144, 3-19.<br />

Stevenson, B.A., Kelly, E.F., McDonald, E.V., Busacca, A.J., 2005. The stable carbon<br />

isotope composition of soil organic carbon and pedogenic carbonates along a<br />

bioclimatic gradient in the Palouse region, Washington State, USA. Geoderma<br />

124, 37-47.<br />

USDA-NRCS, 1996. Soil Survey Laboratory Methods Manual. S.S. Investigation<br />

Report No. 42, Version 3.0, 716 pp.<br />

Usdowski, E., Hoefs, J., 1993. Oxygen isotope exchange between carbonic acid,<br />

bicarbonate, carbonate, and water: A re-examination of the data of McCrea (1950)<br />

and an expression for the overall partitioning of oxygen isotopes between the<br />

carbonate species and water. Geochimica et Cosmochimica Acta 57, 3815-3818.<br />

Vera, J.A., 1970. Estudio estratigráfico de la Depresión de Guadix-Baza. Boletín<br />

Geológico y Minero 81, 429-462.<br />

Wacey, D., Wright, D.T., Boyce, A.J., 2007. A stable isotope study of microbial<br />

dolomite formation in the Coorong region, South Australia. Chemical Geology,<br />

244, 155-174.<br />

Whipkey, Ch.E., Hayob J.L., 2008. Textural and compositional evidence for the<br />

evolution of pedogenic calcite and dolomite in a weathering profile on the Kohala<br />

Peninsula, Hawaii. Carbonate Evaporite 23, 104-112.<br />

Wright, D.T., 1999. The role of sulphate-reducing bacteria and cyanobacteria in<br />

dolomite formation in distal ephemeral lakes of the Coorong region, South<br />

Australia. Sedimentary Geology 126, 147-157.<br />

Wright, V.P., 1992. Paleosol recognition: a guide to early diagenesis in terrestrial<br />

settings. In: Wolf K.H. and Chilingarian G.V. (Eds.), Diagenesis, III: Amsterdam<br />

(Elsevier), Developments in Sedimentol. 47, 591-619.<br />

Wright, V.P., Tucker, M.E., 1991. Calcretes. Ed. Blackwell, London, 352 pp.<br />

ACCEPTED MANUSCRIPT<br />

24


ACCEPTED MANUSCRIPT<br />

FIGURE CAPTIONS<br />

Fig. 1. A) Location of the study area showing the position of 81 profiles sampled. B)<br />

Cross section indicated in A showing the spatial relationship between the<br />

geomorphological units of the area.<br />

Fig. 2. A-C) Vertical distribution of organic and inorganic (calcite and dolomite) carbon<br />

in alluvial soils (PA, Fluvisols), badlands (PB, Regosols) and piedmonts (PC,<br />

Calcisols). Geomorphological units are also indicated in brackets. Arrows in Fig. C<br />

mark dolomite enrichment in depth for the diverse type of soils. D) Depth-wise (y-axis)<br />

variation of the organic carbon, calcite, and dolomite (inorganic carbon) content,<br />

expressed as an average percentage, in the study profiles of Guadix-Baza basin, drawn<br />

from the data of Table 1. E) Distribution with depth of diverse granulometric fractions<br />

of soils; only mean values calculated from the studied profiles has been represented<br />

(gravel + fraction


ACCEPTED MANUSCRIPT<br />

the dolomite-water system (Irwin et al., 1977) with the range of local meteoric water (-8<br />

to -9‰ vs. V-SMOW) and surface temperatures from 17 to 24ºC.<br />

Fig. 5. A-B) Back-scattered electron images showing precipitation of pedogenic calcite<br />

at 120cm depth (profile 46): Grain of quartz partly corroded and filled with calcite and<br />

probably siderite (A). Calcite nodule including dolomite relicts (B). Note the porous<br />

texture of the calcite precipitates in these images. C) Back-scattered electron<br />

micrograph of a large idiomorphic crystal of dolomite in an advanced stage of corrosion<br />

located at 120cm in depth. Note that this corrosion occurs in preferential plains, and<br />

progresses from the centre to outer edges. D) Small dolomite crystals with central voids;<br />

in addition, there are pedogenic calcites as rhyzocretions. E) Back-scattered electron<br />

image of a partly corroded large calcite grain surrounded by a mixture of clays and<br />

carbonates in a sample from 220cm in depth (profile 46). F) Inset of E in which we can<br />

see some dolomite grains surrounded by clays and corroded calcite grains.<br />

Fig 6. Electron images of a sample from 220cm depth (profile 46). A-B) Back-scattered<br />

electron images of rhombohedral dolomite crystals adjacent to corroded calcite grains<br />

and surrounded by clays. Detrital clays (chlorites) also occur in the area shown in Fig.<br />

B. C) Transmission electron microscopy image of a small subhedral dolomite crystal<br />

surrounded by smectite plus Si-rich amorphous substance.<br />

Fig. 7. A) General TEM image of clays and poorly crystalline material surrounding<br />

dolomite crystals in a sample from 220cm depth (profile 46). In addition to well-defined<br />

smectite packets (S), with a composition shown in Table 2, amorphous substance (A)<br />

with a mottled appearance and marked dark contrast also occurs in the poorly crystalline<br />

material located around the neoformed dolomite crystal (Dol). Other areas in the image<br />

lacking fringe contrast and of mottled aspect may correspond to smectites with a slight<br />

disorientation of the layers (Sd) or to a mixture of poorly crystalline smectites and<br />

amorphous substance (Table 2). B) Structural features of the small smectite packets,<br />

some of them composed of a few basal layers. The packets appear to be part of a (gellike)<br />

matrix formed by disoriented smectite-like flakes and a substance of relatively low<br />

atomic number lacking fringe contrast.<br />

ACCEPTED MANUSCRIPT<br />

26


ACCEPTED MANUSCRIPT<br />

Fig. 8 A) Powder X-ray diagram obtained from oriented aggregated (fraction 2) of the<br />

sample at 220cm depth (profile 46), with an indication of d values corresponding to the<br />

reflections of the observed mineral phases. (B) Comparison of air-dried and ethyleneglycol-treated<br />

diagrams, where the basal spacing of smectites is displaced from 1.46nm<br />

to 1.76nm after glycolation.<br />

Fig. 9. X-ray diffraction diagrams for magnesian calcites obtained in batch precipitation<br />

experiments (see Section 3.7 and Table 3) by using viscous (A) and liquid media (B).<br />

d 104 spacing of calcite varies from 0.301nm for the liquid medium to 0.295nm for the<br />

viscous media. In the viscous medium monohydrocalcite also precipitates together with<br />

the magnesian calcite.<br />

Fig. 10. Concentration profiles of Mg at the crystal-medium interface. A) Concentration<br />

distribution at an infinitely fast bulk-diffusion rate. B) Steady-state distribution at very<br />

slow surface-diffusion rate along the grain boundaries of clay minerals. ([Mg] carb ,<br />

concentration of magnesium in carbonate crystal; [Mg] m , concentration of magnesium<br />

in bulk medium; [Mg] i m, concentration of magnesium at the carbonate crystal-medium<br />

interface; thickness of the diffusion boundary layer). Based on Henderson (1986).<br />

TABLE CAPTIONS<br />

Table 1: Arithmetic mean percentage (Mean in %) of the organic and inorganic (calcite<br />

and dolomite) carbon content and of the diverse particle-size fractions for the specified<br />

depths in the studied profiles (number of samples, N, for each depth is also indicated).<br />

ACCEPTED MANUSCRIPT<br />

Table 2. Analytical electron microscopy chemical analyses, expressed in atoms per<br />

formula unit, of smectite packets and mixture of smectites and amorphous substance.<br />

Normalization to 22 charges.<br />

Table 3. X-ray diffraction results for magnesian calcites obtained in batch precipitation<br />

experiments by using two types of media. In both cases the concentrations of the mixed<br />

soluble salts are the same: 50mM Na 2 CO 3 , 50 mM CaCl 2·6H 2 O, and 0.1 M<br />

MgCl 2·6H 2 O (starting pH 7.4).<br />

27


ACCEPTED MANUSCRIPT<br />

SPECIFIC ABBREVIATIONS<br />

a.s.l.: Above Sea Level<br />

AEM: Analytical Electron Microscopy<br />

apfu: Atoms Per Formula Unit<br />

BSE: Back Scattered Electrons<br />

DIC: Dissolved Inorganic Carbon<br />

EDX: Energy-Dispersion X-ray<br />

FESEM: Field-Emission Scanning Electron Microscope<br />

MCL: Meteoric Calcite Line<br />

MDL: Meteoric Dolomite Line<br />

OA: Oriented Aggregated<br />

TEM: Transmission Electron Microscopy<br />

XRD: X-ray diffraction<br />

ACCEPTED MANUSCRIPT<br />

28


ACCEPTED MANUSCRIPT<br />

Table 1<br />

Organic Carbon Inorganic Carbon Particle-Size (2mm)<br />

Depth Calcite Dolomite Clay Silt Sand Gravels<br />

(cm) N Mean (%) Mean (%) Mean (%) Mean (%) N Mean (%) N<br />

0 81 1.82 29.35 9.19 19.48 31.26 42.26 69 34.44 81<br />

20 81 1.13 37.60 8.42 19.82 29.59 50.60 55 34.83 68<br />

40 81 0.67 42.59 10.34 19.38 25.58 55.04 43 37.85 60<br />

60 81 0.46 42.23 12.95 15.51 28.27 56.22 43 43.14 57<br />

80 81 0.34 39.35 14.68 13.62 26.45 59.93 46 39.89 65<br />

100 81 0.25 35.90 15.24 12.04 23.43 64.53 43 37.73 65<br />

120 81 0.19 32.64 16.16 11.88 23.96 64.16 50 39.05 67<br />

140 75 0.16 31.41 16.83 11.46 21.49 67.05 42 35.55 65<br />

160 64 0.15 29.78 15.81 11.08 19.66 69.26 37 41.63 59<br />

180 55 0.12 31.25 13.28 10.94 22.82 66.23 35 41.14 48<br />

200 44 0.13 28.09 16.97 12.72 20.46 66.82 31 45.99 40<br />

220 27 0.11 33.63 18.33 12.12 17.99 69.89 17 53.06 24<br />

240 19 0.12 38.25 13.57 13.76 25.17 61.08 14 52.15 17<br />

260 12 0.14 37.31 16.26 10.18 15.63 74.19 6 51.91 9<br />

280 9 0.15 28.81 12.73 14.57 17.92 67.51 5 56.14 7<br />

300 8 0.18 26.78 12.77 18.89 19.18 61.94 4 59.46 6<br />

320 8 0.22 25.03 13.51 11.72 29.10 59.18 4 55.71 6<br />

340 4 0.22 31.14 14.00 9.20 15.66 75.14 2 58.12 3<br />

Total 892 546 747<br />

Table 2<br />

Smectite analyses<br />

Mixed analyses<br />

Si 3.79 3.88 3.83 3.78 4.16 4.17 4.12 4.41<br />

Al IV 0.21 0.12 0.17 0.22 -0.16 -0.17 -0.12 -0.41<br />

ACCEPTED MANUSCRIPT<br />

Al VI 1.55 0.84 1.57 1.69 1.55 1.28 1.43 1.53<br />

Fe (3+) 0.24 0.62 0.17 0.14 0.05 0.28 0.24 0.16<br />

Mg 0.34 0.53 0.22 0.14 0.31 0.48 0.26 0.21<br />

Ʃ VI 2.14 2.00 1.97 1.97 1.91 2.03 1.93 1.90<br />

Ca 0.03 0.03 0.00 0.02 0.05 0.05 0.05 0.03<br />

K 0.07 0.36 0.34 0.31 0.28 0.16 0.22 0.21<br />

Ʃ XII 0.10 0.40 0.34 0.33 0.33 0.21 0.28 0.24<br />

29


ACCEPTED MANUSCRIPT<br />

Table 3<br />

Type of medium d 104 (Å) FWHM (º2θ)* % Mg<br />

Liquid 3.01 0.3 10<br />

Viscous 2.95 1.26 25<br />

*FWHM: full width at half maximum (peak broadening measured for 104 reflection in the XRD<br />

diagrams of the Fig. 10 ).<br />

ACCEPTED MANUSCRIPT<br />

30


ACCEPTED MANUSCRIPT<br />

Fig. 1<br />

ACCEPTED MANUSCRIPT<br />

31


ACCEPTED MANUSCRIPT<br />

ACCEPTED MANUSCRIPT<br />

Fig. 2<br />

32


ACCEPTED MANUSCRIPT<br />

Fig. 3<br />

ACCEPTED MANUSCRIPT<br />

33


ACCEPTED MANUSCRIPT<br />

Fig. 4<br />

ACCEPTED MANUSCRIPT<br />

34


ACCEPTED MANUSCRIPT<br />

Fig. 5<br />

ACCEPTED MANUSCRIPT<br />

35


ACCEPTED MANUSCRIPT<br />

ACCEPTED MANUSCRIPT<br />

Fig. 6<br />

36


ACCEPTED MANUSCRIPT<br />

Fig. 7<br />

ACCEPTED MANUSCRIPT<br />

37


ACCEPTED MANUSCRIPT<br />

Fig. 8<br />

ACCEPTED MANUSCRIPT<br />

38


ACCEPTED MANUSCRIPT<br />

Fig. 9<br />

ACCEPTED MANUSCRIPT<br />

39


ACCEPTED MANUSCRIPT<br />

Fig. 10<br />

ACCEPTED MANUSCRIPT<br />

40


ACCEPTED MANUSCRIPT<br />

RESEARCH HIGHLIGHTS<br />

1. Neoformed dolomite appears in semiarid soils below the petrocalcic horizon<br />

2. An isotopically light dolomite forms after dissolution of pedogenic calcite<br />

3. Smectite acts as a catalyst for dolomite/protodolomite precipitation<br />

4. Pedogenic dolomite must be taken into account in carbon-soil balances<br />

ACCEPTED MANUSCRIPT<br />

41

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!