Vous êtes sur la page 1sur 16

This article was downloaded by: [IVIC Instituto Venezolano de] On: 27 June 2012, At: 12:24 Publisher:

Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Communications in Soil Science and Plant Analysis


Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/lcss20

Nitrate Sorption in a Mexican Allophanic Andisol using Intact and Packed Columns
Blanca Prado , Cline Duwig , Mauricio Escudey & Michel Esteves
a a a b c

Institut de Recherche pour le Dveloppment, c/o Colegio de Postgraduados, Laboratorio de Fertilidad de Suelo, Montecillo, Mexico
b

Departamento de Qumica de los Materiales, Facultad de Qumica y Biologa, Universidad de Santiago de Chile, Santiago, Chile
c

Institut de Recherche pour le Dveloppment, Laboratoire d'tude des Transferts en Hydrologie et Environnement, Grenoble, France Version of record first published: 31 Oct 2011

To cite this article: Blanca Prado, Cline Duwig, Mauricio Escudey & Michel Esteves (2006): Nitrate Sorption in a Mexican Allophanic Andisol using Intact and Packed Columns, Communications in Soil Science and Plant Analysis, 37:15-20, 2911-2925 To link to this article: http://dx.doi.org/10.1080/00103620600833017

PLEASE SCROLL DOWN FOR ARTICLE Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

Communications in Soil Science and Plant Analysis, 37: 29112925, 2006 Copyright # Taylor & Francis Group, LLC ISSN 0010-3624 print/1532-2416 online DOI: 10.1080/00103620600833017

POSTER PAPER

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

Nitrate Sorption in a Mexican Allophanic Andisol using Intact and Packed Columns
line Duwig Blanca Prado and Ce
veloppment, c/o Colegio de Institut de Recherche pour le De Postgraduados, Laboratorio de Fertilidad de Suelo, Montecillo, Mexico

Mauricio Escudey
mica de los Materiales, Facultad de Qu mica y Departamento de Qu a, Universidad de Santiago de Chile, Santiago, Chile Biolog

Michel Esteves
veloppment, Laboratoire de tude des Institut de Recherche pour le De Transferts en Hydrologie et Environnement, Grenoble, France

Abstract: Contamination of groundwater by nitrate is a worldwide environmental issue. A better knowledge of nitrate sorption characteristics by soils contributes to efcient fertilizer use and prevents aquifer contamination. In volcanic soils, nitrate sorption is induced by variable charges due to the presence of amorphous materials and aluminum (Al) and iron (Fe) oxides. Anion transport in packed and intact columns was investigated in a Mexican Allophanic Andisol, under different 2 permanent ow regimes in unsaturated conditions and several NO2 3 -N and Br input 2 concentrations. In the packed columns, the NO3 -N adsorption in the soil was nonlinear. In the intact columns, the retardation coefcient variation was directly correlated to the increase of amorphous material with depth. The presence of preferential ow in the intact columns signicantly increased the mobility and velocity of nitrate moving through the columns, whereas in the packed columns, NO2 3 -N fate was only affected by soil chemical composition and mineralogy. Keywords: Allophane, physical nonequilibrium, retardation, volcanic soil

Received 28 January 2005, Accepted 7 May 2005 line Duwig, Institut de Recherche pour le De veloppAddress correspondence to Ce ment, c/o Colegio de Postgraduados, Laboratorio de Fertilidad de Suelo, CP 56230, Montecillo, Mexico. E-mail: celine.duwig@hmg.inpg.fr 2911

2912

B. Prado et al.

INTRODUCTION Ion-exchange reactions in soils are processes dependant on soil properties, which are inuenced by local conditions such as ionic strength, soil solution concentration, pH, and water content of soil. These processes govern the fate of dissolved nutrients and pesticides applied at the soil surface in crop elds. The study of anions such as nitrate, which is both a nutrient and a potential pollutant, and their transport in soil has intensied in recent years because of increasing nitrate concentrations in drinking water resources all over the world (Goodrich, Lykins, and Clark 1991; Spalding and Exner 1993). Nitrogen (N) is an important plant nutrient, and NO2 3 -N is the most plant-available form. Nitrate naturally present in the soil or applied as fertilizer is considered inert in most aerobics soils and thus subject to leaching through the vadose zone. In variable-charge volcanic soils, anion transport is retarded because of electrostatic adsorption (Ishiguro, Song, and Yuita 1992), a fact that could decrease risk of subsoil nitrate contamination. However, Katou, Clothier, and Green (1996) found that nitrate-leaching depth in an Andisol increased in the presence of Cl2, because of competitive adsorption. Similarly, Ishiguro and Shoji (2002) studied the B horizon of a volcanic soil, showing that nitrate retention time was reduced when sulfate was previously adsorbed on the soil. Schalscha, Pratt, and Domecq (1973) demonstrated that nitrate adsorption on Chilean volcanic soils decreased when the pH of soil solution increased; additionally, the adsorption rate was lower in surface horizons than subsoil. These results can be explained by the increased amorphous material and the decreased organic matter with increased depth. Kinjo and Pratt (1971a), in their study on nitrate adsorption in acid soil from Mexico and South America, found a signicant correlation between the amount of nitrate adsorbed and the soil content of amorphous material. At acid soil pH and low anion input concentrations, Qafoku, Summer, and Radcliffe (2000) reported that nitrate adsorption was directly related to nitrate concentration in the soil solution. A better understanding of the adsorption processes during ion displacement in soil will contribute to more efcient management of fertilizer applications to crops, thus reducing loss of fertilizers in groundwater and increasing economic and environmental benet (Kinjo and Pratt 1971a). The use of miscible-displacement or ow-through techniques in soil columns allows the analysis of processes occurring between the liquid phase and soil matrix. Batch technique is commonly used in adsorption studies. However, the ow-through technique represents conditions closer to eld, including the soilsolution ratio, the immobility of the solid phase, and the contact time between the soil and the solution (Miller, Summer, and Miller 1989). Considering the debate about the representativeness of packed versus intact columns (Cassel, Genuchten, and Wierenga 1975; Camobreco et al. 1996) to a natural system, results obtained from both types of columns were compared.

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

Nitrate Sorption in a Mexican Allophanic Andisol

2913

The objective of this work was to study the transfer of reactive anions through a Mexican volcanic soil. The retardation was estimated at different soil depths, using intact and packed soil columns. The effects of the input anion concentration and the ow rate were examined.

MATERIALS AND METHODS Study Site and Soils Sampled

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

Soils were sampled from a small hydrologic catchment (53 ha), La Loma (198 160 N and 998 59 W; 2600 3000 m above sea level; in Amanalco de Becerra, State of Mexico, Mexico), instrumented for water runoff, erosion, and fertilizer transport monitoring. The climate at the site is tropical and high altitude, with average annual precipitation of 1300 mm and annual mean temperature of 10.78C. The soil in the catchment is a Typic Hapludand (Soil Survey Staff 1999); indurated volcanic tuff (Tepetate) soils are also present. The study was based on the rst soil type, where the agricultural practices and use of fertilizers are more intensive. Nitrogen fertilizer is applied at an average rate of about 150 kg N/ha to maize, which represents 35% of the land in the catchment. A soil prole was excavated under maize and each horizon was sampled for physical, chemical, and mineralogical characteristics. Bulk density was estimated from samples of known volume taken with a 5-cm diameter steel core. The pH was measured in water suspension using a 1:2.5 (w/v) soil solution ratio. Soil texture was determined by laser granulometry. Total organic carbon content (TOC) was analyzed by dry combustion (TOC-5050A Shimadzu). Phosphate retention was determined according to Blackmore, Searle, and Daly (1997). Cation and anion exchange capacity were assessed according to Zelazny, He, and Vanwormhoudt (1996) at soil pH. Selective dissolutions of iron (Fe), aluminium (Al), and silica (Si) were carried out to characterize the poorly ordered amorphous materials. Ammonium oxalate extraction at pH 3.0 was conducted to extract the poorly crystalline active Fe, Al, and Si compounds (Feo, Alo, and Sio respectively) (Blackmore, Searle, and Daly 1997). The soils were extracted with sodium pyrophosphate to get Fe and Al bound to organic matter (Fep and Alp) (Blackmore, Searle, and Daly 1997). A dithionite-citrate dissolution procedure was used to determine free Al, Fe, and Mn oxides (Ald, Fed, and Mnd) (Mehra and Jackson 1960). Concentrations of Fe, Al, Si, and Mn in solutions were determined using an atomic absorption spectrophotometer (SpectrAA 220 Varian). The allophane and imogolite were calculated with the relation Sio 6 (Partt 1990), the ferrihydrite concentration with the relation Feo 1.7 (Partt 1998), and the crystalline Fe oxides with the relation Fed Feo (Partt 1998).

2914

B. Prado et al.

Column Experiments Seven soil columns were studied. Four were excavated on the eld at different depths (C1 to C4 at 0.05, 0.3, 0.55, and 0.8 m respectively) by inserting a 0.25-m long and 0.05-m inner diameter PVC tube into the soil. Intact soil columns were stored at 48C until use. A soil sample was taken at the same site between 0.2 and 0.6 m deep with a screw-type auger. This soil was used to pack three columns (Cd1 to Cd3). The soil sample was sieved moist at 2 mm before being packed into a 0.025-mdiameter glass tube. The same procedure was followed for each of the three packed columns to achieve a uniform and identical bulk density. The eld soil solution was analyzed and contained Ca2, Mg2, and K1 at concentrations of 0.307, 0.024, and 0.0486 mM respectively. A background solution containing the same concentrations of these cations with Cl as accompanying cation was prepared. Columns were oriented vertically, and a constant ow rate of the background solution under unsaturated condition was passed through the columns using a peristaltic pump. This solution was leached through the column for a minimum of two pore volumes until reaching a steady state for the output ow rate and electrical conductivity (EC). Half of a pore volume of KNO3, KBr, and 18O was 2 injected at concentrations of 6, 8, or 9 mM for NO2 and 34.1 3 -N and Br 18 (% of atomic excess) for O and pushed through by the background solution. The 18O was used as a water tracer to calibrate the soil hydrodynamic parameters of a numerical model. EC, pH, and temperature of the leachate were monitored continuously. Samples were automatically collected for chemical analysis. The experiments were terminated when the leachates EC reached a level close to the background solutions EC. Water contents were determined by weighting the column at the beginning and at the end of each run and by knowing the initial water content. The pore water velocity was calculated as the input solution ux over the volumetric water content. The leachate fractions were weighed to 2 check the ux at the column exit and were analyzed for NO2 3 -N, Br , 2 18 and Cl by capillary electrophoresis and for O by mass spectrometry. The runs were conducted the same way in both packed and intact columns. Two different ow rates were used, and the input concentration was varied only into the packed columns. Table 1 presents the bulk density, nal water content (u), mean pore water velocity (n), pore volume (Vo) and input concentration (Co) for each column.

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

Transport Parameters Determination


18

Convection dispersion equation transport parameters were determined for the 2 O, NO2 3 -N, and Br breakthrough curves (BTC) data using the CXTFIT 2.1 code (Toride, Leij, and van Genuchten 1999). The deterministic linear

Nitrate Sorption in a Mexican Allophanic Andisol

2915

Table 1. Column parameters: bulk density (r) input concentration (Co), nal water 2 content (u), mean pore water velocity (n), and pore volume (Vo) for NO2 3 -N and Br for a columns of a Typic Hapludand from Amanalco, Mexico Columns Depth (cm)

r (g cm23)
0.72 0.61 0.62 0.63 0.77 0.75 0.77

Co (mM)

u (cm3 cm23)
0.71 0.69 0.80 0.72 0.66 0.70 0.71

n (cm min21)

Vo (cm3)

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

Intact soil columns C1 5 30 C2 30 55 C3 55 80 C4 80 105 Packed soil columns Cd1 20 60 Cd2 20 60 Cd3 20 60

8 8 8 8 6 6 9

0.04 0.05 0.04 0.04 0.05 0.033 0.036

347 338 391 352 81 86 87

equilibrium adsorption model was used in inverse mode for ux concentration in packed columns and the surface intact column to t the dispersion coefcient D and the retardation R. This procedure assumed that all the water was mobile and therefore at physical equilibrium. However, this hypothesis proved incorrect on the intact columns, where the 18O BTC peak appeared before one pore volume. The two-region nonequilibrium model for ux concentration was tted on the 18O BTC using R 1 to obtain D, the mobile water fraction (um/u), and the dimensionless mobileimmobile region exchange term (v). D and um/u were then kept constant, and the same two-region 2 model was tted on the NO2 3 -N and Br BTCs to obtain R, b (dimensionless partitioning coefcient), and v. The use of CXTFIT assumes that adsorption is linear.

RESULTS AND DISCUSSION Soil Characteristics and Evidence for Anion Sorption Main physical and chemical properties for each soil layer are given in Table 2. Clay content, 27%, is dominated by allophane and imogolite. The average content of allophane and imogolite varied from 18 to 28% of the clay fraction depending on soil depth. The soil exhibits a high AEC, consistent with previous studies on volcanic ash soils (Wong, Hughes, and Rowell 1990; Kinjo and Pratt 1971a). The TOC content is about 5.3% and did not vary greatly with depth, as did the other soil characteristics except the allophane and imogolite contents and the bulk density. As in other Andisols, the soil is variably charged, a characteristic associated with the presence of noncrystalline materials and high organic matter content (Nanzyo, Dahlgren, and Shoji 1993).

2916

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

Table 2.

Selected physical and chemical parameters of a Typic Hapludand from Amanalco, Mexico

Depth (cm) 0 15 15 20 20 45 45 65 65 85 85 110 105 140


a b

Horizon Ap A 2A 2B 2Bw 2Bw2

ra (g cm23)
0.71 0.66 0.49 0.51 0.49 0.50

OCb (%) 5.4 5.3 5.6 5.3 4.7 5.1

AECc (cmolc kg21) 12.3 11.6 10.3 11.2

CECd (cmolc kg21) 20.1 23.4 17.9 21.1

pH H2O 5.5 6.1 6.2 6.3 6.3 6.5

Allophane and imogolite (%) 18.5 23.1 22.5 25.5 26.0 27.7

Fe oxides (%) 2.2 3.0 2.7 2.0 3.0 2.3

Ferrihydrite (%) 3.2 2.0 1.8 2.6 2.7 2.1

r: bulk density. OC: organic carbon content. c AEC: anion exchange capacity. d CEC: cation exchange capacity.
B. Prado et al.

Nitrate Sorption in a Mexican Allophanic Andisol

2917

Packed Columns Table 1 presents the bulk density, nal water content (u), mean pore water velocity (n), pore volume (Vo), and input concentration (Co) for each packed column. Figure 1 shows the Br2 and NO2 3 -N BTCs for one packed column (Cd3). As indicated before, curves are symmetrical. An equilibrium convection dispersion equation satisfactorily describes the experimental results. Model parameters are presented in Table 3. The maximum C/Co peak occurred in the column with the lower pore water velocity (Cd2). A lower injection ux leads to a lower dispersion, which results in a narrower curve with a higher peak. Because of the presence of variable charges in the soil, the peak of NO2 3 -N BTCs appeared after 1.25 pore volume, which would be the position of the BTC peak for an inert tracer. The retardation coefcient R varies with the concentration of input solution (Cd2 and Cd3, Table 3), indicating a nonlinear adsorption of NO2 3 -N in the soil columns. This parameter decreases when the input concentration increases following a Freundlich-type isotherm, similar to results obtained by Qafoku, Sumner, and Redcliffe (2000). The authors reported that at low concentration (, 5 mM ), more NO2 3 -N is adsorbed by the soil than in solution. Thus, the maximum R value is obtained at low concentrations and decreases at higher concentrations. There is a controversy concerning the mechanism of NO2 3 -N adsorption at low concentration in soil: Toner, Sparks, and Carski (1989) proposed a totally reversible adsorption that is in fact a simple electrostatic retention, whereas according to Singh and Kanehiro (1969), the adsorption is the result of van der Waals interactions. Lately, Qafoku, Sumner, and Radcliffe (2000) explained the adsorption as an overlapping or interpenetration of double layers around the positively charged Al polymers and negatively charged silicate minerals. Each of these suggestions could imply a different effect on retardation.

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

Figure 1. Measured (circle and cross) and simulated (line) BTCs for NO2 3 -N and Br2 for a packed soil column (Cd3) of a Typic Hapludand, Amanalco, Mexico. The rectangle indicates the pulse.

2918 Table 3. Columns C1 C4 Cd1 Cd2 Cd3 Model tted parameters D (cm min21)
2

B. Prado et al.

l (cm21)
0.77 5.5 0.81 0.82 0.77

RNO2 3 -N 1.5 2.2 1.54 1.56 1.5

KdNO2 3 -N (cm3 kg21) 0.47 1.37 0.50 0.52 0.46

RBr

KdBr2 (cm3 kg21) 0.57 1.49 0.56 0.56 0.46

0.03 0.23 0.04 0.03 0.03

1.6 2.3 1.6 1.6 1.5

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

Bromide adsorption was also studied, an anion not naturally present in soil system and not subject to microbial degradation. A slightly higher retardation coefcient was found for Br2 compared to NO2 3 -N (not shown). This result can either be due to a slight preference of the soil for Br2 over 2 NO2 3 -N or to some bacterial degradation of the NO3 -N anion during the experiment duration. Degradation can be excluded because the mass balances for both anions at conclusion were found to be close to 100%. Additionally, the Darcy velocity ux is greater than the one indicated by Corey, Nielsen, and Kirkham (1967) as being too rapid to allow microorganism degradation of NO2 3 -N during its transport through the soil column. The dispersivity parameter l was calculated using the equation l D/n, assuming that the molecular diffusion was negligible compared to the convective dispersion, because in this study the pore volume velocity was high lek and Nielsen 1994). The dispersion and also the dispersivity (n did (Kut not vary greatly) were identical in all packed columns, which indicates that the packing method was effective to obtain uniformity. According to Magesan et al. (2003), the dispersivity is an indicator of how variable solute mobility is in a soil. Flux velocity or timescale can inuence the transport and reactions of anions in soils. If sorption process is truly at equilibrium, sorption and transport parameters are invariant with ux velocity (Jardine et al. 1998). As mentioned earlier, R and D values obtained in columns at different ux velocity (Cd1 and Cd2) did not vary. It can be concluded that in the packed columns, within the range of velocities and concentrations studied, the sorption and transport processes are at equilibrium.

Intact Soil Columns Table 1 presents the bulk density, nal water content (u), mean pore water velocity (n), pore volume (Vo), and input concentration (Co) for the intact columns.

Nitrate Sorption in a Mexican Allophanic Andisol

2919

Competitive Adsorption with Cl2 Figure 2a shows the variations of dimensionless Cl2 and NO2 3 -N measured concentrations at the bottom of column C1. As indicated in the method section, Cl2 is the accompanying anion for all cations in the recomposed soil solution. This solution is also used to prepare the pulse solution of Br2 2 and NO2 3 -N. This means that Cl is constantly injected in the column at a concentration of 1.5 mM. At the time of the pulse, Cl2 is in an equilibrium state, the concentration in the leachate being equal to that in the input 2 solution. When the NO2 pulse is injected in the soil column, 3 -N and Br 2 Cl is supposed to be released into the leachate in higher proportion (however, early samples were not analyzed), meaning that the Cl2 adsorption level decreases. After the pulse, Cl2 concentration in the leachate decreases (Figure 2), showing that in the absence of the two other anions, Cl2 is adsorbed. Kinjo and Pratt (1971b) reported a slight preference for Cl2 over Br2 and NO2 3 -N in soils containing amorphous materials, whereas Harsh, Chorover, and Nizeyimana (2002) pointed out that sorption of Cl2, Br2 and 2 2 NO2 3 -N on allophane is not selective. In this study, NO3 -N and Br

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

2 Figure 2. Measured NO2 3 -N and Cl BTCs for the intact soil columns: (a) in the top horizon (5 30 cm, column C1) and (b) at 80 105 cm deep (column C4).

2920

B. Prado et al.

concentrations were 5.3 times higher than Cl2 concentration, meaning that the latter cannot compete for the adsorption sites in the presence of the two other concentrated anions. Figure 2b shows the same BTCs for column C4. A similar reduction in Cl2 concentration can be noted but to a lesser extent, which may be explained by the effect of dispersion being seven times higher in C4 than in C1 (see Table 3). In Figure 2, it can also be observed that the NO2 3 -N BTC peak does not reach the level in C1 because of the dispersion.

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

Physical Nonequilibrium Figure 2 shows the variation of dimensionless measured concentrations (C/Co) for Cl2 and NO2 3 -N with pore volumes, at the bottom of the intact soil columns C1 (a) and C4 (b). The C1 NO2 3 -N BTC is symmetrical (Figure 2a), which indicates that NO2 3 -N movement is under physical and chemical equilibrium. On the other hand, the C4 NO2 3 -N BTC is asymmetrical (Figure 2b), with an early peak and tailing at the end of the curve, indicative of the presence of physical or chemical nonequilibrium. Figure 3 shows the dimensionless measured and simulated NO2 3 -N BTCs of two intact columns, in the top horizon (5 30 cm, column C1) (a) and at 0.8 m deep (column C4) (b). The 18O (nonreactive tracer) BTC for C1 is symmetrical, indicating that solute transport is under physical equilibrium. However, in Figure 3b, the 18O BTC with an advanced peak and pronounced tail indicates the presence of physical nonequilibrium between advectiondominated pore domains and the soil matrix pores within the aggregates (Jardine et al. 1998). Preferential ux or the presence of pores that do not participate to the overall ux reduces the contact time between solute and soil matrix and provokes an anticipated exit of the solute, inducing increasing risks of groundwater contamination (Beven and Germann 1982). The structure of C1 layer is the result of tillage, as this column was sampled between 5 and 30 cm deep in a maize plot. Bejat et al. (2000) indicates that tillage destroys the natural pore structure of surface soils, disrupting macropore continuity and reducing the extent of bypass ow. In contrast to C1, C4 was sampled at the bottom of the root zone between 0.8 and 1.1 m, where the soil has never been disrupted and keeps its natural structure. Adsorption and Relation with Soil Characteristics On both graphs in Figure 3, a shift toward the right of the N-NO2 3 BTC compared to the 18O one can be observed, indicative of nitrate retardation in the soil. As in packed columns, Br2 was slightly more retarded than NO2 3 -N in intact soil columns (Table 3). The retardation coefcient increases signicantly with depth (Table 3), whereas AEC slightly decreases (Table 2). This nding is contradictory to results obtained by Qafoku, Sumner, and Radcliffe (2000), who found that on packed columns of variable charge

Nitrate Sorption in a Mexican Allophanic Andisol

2921

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

Figure 3. Measured (circle and triangle) and simulated (line) BTCs for NO2 3 -N and 18 O in the intact soil columns: (a) in the top horizon (5 30 cm, column C1) and (b) at 80 105 cm deep (column C4).

subsoils, the retardation coefcient was positively correlated to the AEC. The AEC determined through classical batch method [e.g., Zelazny, He, and Vanwormhoudt (1996) in this study] is more related to the number of sites where exchanges can occur and gives no information about the type of anion interaction on these exchange sites. Duwig et al. (2003) determined the AEC with both static and dynamic methods using the same anion and obtained different values. They explained the discrepancy by differences in the experimental conditions that can affect the adsorption. In this study, the AEC was estimated through batch experiments with the anion Cl2. Cl2 is considered to be a good tracer for NO2 3 -N (McMahon and Thomas 1974). However, the experimental conditions between batch and columns are not the same; mostly the way the anion and the soil particles are in contact differs. In addition to the difference in the anion and experimental conditions used, the role of soil physical parameters is important when the retardation coefcient through displacement experiments is determined, in comparison to static methods.

2922

B. Prado et al.

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

The increase of R with depth can directly be related to the increase of soil allophane and imogolite content, as reported by Kinjo and Pratt (1971a) in their study of NO2 3 -N adsorption in acidic soils from Mexico and South America. In Table 1, it can be noted that the pH increases with depth as the carbon (C) content decreases, whereas the retardation increases. As found in many studies on anion transport in variable charge soils where soil conditions are modied (Qafoku, Sumner, and Radcliffe 2000; Magesan et al. 2003), retardation is negatively correlated with pH. However, while studying different soil layers under natural conditions, this relationship is not so evident, because other factors affect anion adsorption such as amorphous materials. Dispersion The dispersion was found to increase with depth. This phenomenon is linked to the difference in soil structure from one soil layer to another. The soil has a loamy structure; however the tillage, the C content, the amorphous material content, and the water content modify the structure. It varies from granular of loose consistency on the surface, to granular with subangular blocky moderately thick of friable consistency at depth. The dispersivity coefcient was calculated as for packed columns, and the values obtained are in the range of those reported by Magesan et al. (2003). These authors studied solute movement through intact columns of an allophanic soil under unsaturated ux.

Comparison between Packed and Intact Columns The discrepancy found between the two columns type are linked to difference in soil structure. The ne and loose structure of the top intact columns (C1) is homogenous, due to tillage practices, and comparable to the structure of packed columns. It is conrmed by the fact that the dispersion of packed columns (Cd1 to Cd3) where the soil was sieved and compacted to the same bulk density was found to be similar to column C1. At a greater depth, the soil is more aggregated and the deeper intact core dispersivity increases. Whereas solute transport in packed columns was found to be uniform and under physical equilibrium, deep intact soil core exhibited preferential ux, leading to accelerated solute movement. Retardation coefcient obtained in packed columns was equal to those obtained in intact soil cores from the surface horizon. This result can be explained by two factors: similar soil characteristics because the soils were sampled at the same depth and similar soil structure. In intact soil cores, retardation increased with depth, due to variation in soil amorphous material contents. Although the Br2 and NO2 3 -N anions are more retarded at depth compared to the water tracer, it must be noted that they will move through this deep layer more quickly than an inert tracer in the same soil without preferential ux (see Figure 3b). Consequently, nitrate present at this depth [which is below the root zone (90 cm)] cannot be used by plants and will quickly move to the groundwater.

Nitrate Sorption in a Mexican Allophanic Andisol

2923

CONCLUSIONS The Andisol (Typic Hapludand) from the Mexican catchment la Loma shows a NO2 3 -N adsorption capacity that could retard its transfer toward the aquifer. However, in deeper layers, the preferential ow is the dominant process during the displacement experiments, a factor that accelerates nitrate transfer toward groundwater resources. In the packed columns, where the concentration and ux density effect on the reactive anions was studied, the retardation decreased when the input concentration increased, indicating that the N-NO2 3 adsorption in the soil is nonlinear. The dispersivity was similar in all packed columns, showing that the packing method was effective to recreate comparable structure in each column. In the intact columns where the variation of anion adsorption with depth was analyzed, the retardation increased with depth. The variation of retardation coefcient was directly correlated to the increase of amorphous materials with depth but not with the anion exchange capacity. The dispersivity also increased with depth, due to the change in soil structure between soil surface and the horizon at 1-m depth. In both type of columns, Br2 was slightly more retarded than NO2 3 -N. The packed columns are useful to study the effect of the soil properties on the anion adsorption; nevertheless, the intact columns describe better the solute transport toward the aquifer as they parallel more closely the soil structure and pore geometry found under eld conditions. Results indicate that extremely thoughtful consideration must be taken before extrapolating fertilizer movement study results obtained from packed soils to the eld situation. Andisols, although having a ne and poorly developed structure with mostly micropores, can display preferential ux and, as noted by Jarvis (1998), this process must be considered more like a rule rather than an exception. ACKNOWLEDGMENTS Blanca Prado is indebted to Consejo Nacional de Ciencia y Tecnologia te Franc (Mexico) and Socie aise d0 Exportation des Resources Educatives (France) for her PhD grant. The authors thank the Laboratorio de Fertilidad de Suelo of Colegio de Postgraduados, Montecillo for the soil analyses. veloppeThe research was funded by the Institut de Rercherche pour le De ment (IRD), France. REFERENCES
Bejat, L., Perfect, E., Quisenberry, V.L., and Haszler, G.R. (2000) Solute transport as related to soil structure in unsaturated intact soil blocks. Soil Science Society of America Journal, 64: 818826. Beven, K.J. and Germann, P.F. (1982) Macropores and water ows in soils. Water Resources Research, 18: 1311 1325.

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

2924

B. Prado et al.

Blakemore, L.C., Searle, P.L., and Daly, B.K. (1997) Procedures for Soil Analysis (Technical Paper No. 9). International Soil Reference and Information Centre: Wageningen, The Netherlands. Camobreco, V.J., Richards, B.K., Steenhuis, T.S., Peverly, J.H., and McBride, M.B. (1996) Movement of heavy metals through undisturbed and homogenized soil columns. Soil Science, 161: 740 750. Cassel, D.K., Genuchten, M.T.V., and Wierenga, P.J. (1975) Predicting anion movement in disturbed and undisturbed soils. Soil Sci. Soc. Am. P, 39: 1015 1919. Corey, J.C., Nielsen, D.R., and Kirkham, D. (1967) Miscible displacement of nitrate through soil columns. Soil Science Society of America Journal, 31: 497 501. Duwig, C., Becquer, T., Charlet, L., and Clothier, B.E. (2003) Estimation of nitrate retention by a Ferrasol by a transient-ow method. Eur. J. Soil Sci., 54: 505 515. Goodrich, J.A., Lykins, B.W., and Clark, R.M. (1991) Drinking water from agriculturally contaminated groundwater. Journal of Environmental Quality, 20: 707 717. Harsh, J., Chorover, J., and Nizeyimana, E. (2002) Allophane and imogolite. In Soil Mineralogy with Environmental Applications; Dixon, J.B. and Schulze, D.G. (eds.); SSSA: Madison, Wisconsin, 291 322. Ishiguro, M., Song, K.C., and Yuita, K. (1992) Ion transport in an allophanic Andisol under the inuence of variable charge. Soil Science Society of America Journal, 56: 1789 1793. Ishiguro, M. and Shoji, S. (2002) Nitrate and sulphate transport in volcanic ash soil of A and B horizons. In 17th World Congress of Soil Science; IUSS: Bangkok, Thailand; Vol. V, 1629. Jardine, P.M., OBrien, R., Wilson, G.V., and Gwo, J.P. (1998) Experimental techniques for conrming and quantifying physical nonequilibrium processes in soils. In Physical Nonequilibrium in Soils: Modeling and Application; Selim, H.M. and Ma, L. (eds.); Ann Arbor Press: Chelsea, Michigan, 243 271. Jarvis, N. (1998) Modeling the impact of preferential ow on nonpoint source pollution. In Physical Nonequilibrium in Soils: Modeling and Application; Selim, H.M. and Ma, L. (eds.); Ann Arbor Press: Chelsea, Michigan, 195 221. Katou, H., Clothier, B.E., and Green, S.R. (1996) Anion transport involving competitive adsorption during transient water ow in an Andisol. Soil Sciience Society of America Journal, 60: 1368 1375. Kinjo, T. and Pratt, P.F. (1971a) Nitrate adsorption, I: In some acid soils of Mexico and South America. Soil Science Society of America Proceedings, 35: 722 725. Kinjo, T. and Pratt, P.F. (1971b) Nitrate adsorption, II: In competition with chloride, sulfate, and phosphate. Soil Sci. Soc. Am. Proc., 35: 725 728. lek, M. and Nielsen, D.R. (1994) Soil Hydrology; Catena Verlag: CremlingenKut Destedt, Germany. Magesan, G.N., Vogeler, I., Clothier, B.E., Green, S.R., and Lee, R. (2003) Solute movement through an allophanic soil. Journal of Environmental Quality, 32: 2325 2333. McMahon, M.A. and Thomas, G.W. (1974) Chloride and tritiated water ow in disturbed and undisturbed soil cores. Soil Science Society of America Proceedings, 38: 727 732. Mehra, O.P. and Jackson, M.L. (1960) Iron oxide removal from soil and clays by a dithionite citrate system buffered with sodium bicarbonate. Clays and Clay Minerals, 7: 317 327. Miller, D.M., Summer, M.E., and Miller, W.P. (1989) A comparison of batch and ow generated anion adsorption isotherms. Soil Science Society of America Journal, 53: 373 380.

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

Nitrate Sorption in a Mexican Allophanic Andisol

2925

Downloaded by [IVIC Instituto Venezolano de] at 12:24 27 June 2012

Nanzyo, M., Dahlgren, R., and Shoji, S. (1993) Chemical characteristics of volcanic ash soils. In Volcanic Ash Soils, Genesis, Properties and Utilization; Shoji, S., Nanzyo, M., and Dahlgren, R. (eds.); Elsevier: Amsterdam, The Netherlands, 145 187. Partt, R.L. (1990) Allophane in New Zealanda review. Australian Journal of Soil Research, 28: 343 360. Partt, R.L. and Childs, C.W. (1998) Estimation of forms of Fe and Al: A review and analysis of constrating soils by dissolution and Moessbauer methods. Aust. J. Soil Res., 26: 121 144. Qafoku, N.P., Sumner, M.E., and Radcliffe, D.E. (2000) Anion transport in columns of variable charge subsoils: Nitrate and chloride. Journal of Environmental Quality, 29: 484 493. Schalscha, E.B., Pratt, P.F., and Domecq, T.C. (1973) Nitrate adsorption by some volcanic-ash soils of southern Chile. Soil Science Society of America Journal, 117: 44 45. Singh, B.P. and Kanehiro, Y. (1969) Adsorption of nitrate in amorphous and kaolinitic Hawaiian soils. Soil Science Society of America Journal, 33: 681683. Soil Survey Staff (1999) Soil Taxonomy. A Basic System of Soil Classication for Making and Interpreting Soil Surveys, 2nd Ed. (Agriculture Handbook 436), Department of Agriculture, Natural Resources Conservation Service: Washington, D.C. Spalding, R.F. and Exner, M.E. (1993) Occurrence of nitrate in groundwatera review. Journal of Environmental Quality, 22: 392 402. Toner, C.V., Sparks, D.L., and Carski, T.H. (1989) Anion exchange chemistry of middle Atlantic soils: Charge properties and nitrate retention kinetics. Soil Science Society of America Journal, 53: 1061 1067. Toride, N., Leij, F.J., and van Genuchten, M.T. (1999) The CXTFIT code for estimating transport parameters from laboratory or eld tracer experiments, version 2.1. Research Report 137, U.S. Salinity Laboratory: Riverside, California. Wong, M.T.F., Hughes, R., and Rowell, D.L. (1990) Retarded leaching of nitrate in acid soils from the tropics. Measurement of the effective anion exchange capacity. Journal of Soil Science, 41: 655 663. Zelazny, L.W., He, L., and Vanwormhoudt, A. (1996) Charge analysis of soils and anion exchange. In Methods of Soil Analysis, Part 3: Chemical Methods, 3rd edn.; Sparks, D.L. (ed.), SSSA: Madison, Wisconsin, 1231 1253.

Vous aimerez peut-être aussi