reaction conditions control soil colloid facilitated phosphorus release in agricultural ultisols

11

Click here to load reader

Upload: yuji

Post on 25-Dec-2016

214 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: Reaction conditions control soil colloid facilitated phosphorus release in agricultural Ultisols

Geoderma 206 (2013) 101–111

Contents lists available at SciVerse ScienceDirect

Geoderma

j ourna l homepage: www.e lsev ie r .com/ locate /geoderma

Reaction conditions control soil colloid facilitated phosphorus release inagricultural Ultisols

Allison Rick VandeVoort a,⁎, Ken J. Livi b, Yuji Arai a

a Clemson University, School of Agricultural, Forest and Environmental Sciences, Clemson, SC 29634, USAb Johns Hopkins University, The Integrated Imaging Center/HRAEM Facility, Department of Earth and Planetary Science, Baltimore, MD 21218, USA

⁎ Corresponding author at: School of Agricultural, ForeE249 Poole Agricultural Center, Clemson, SC 29634, USA.

E-mail address: [email protected] (A.R. Vande

0016-7061/$ – see front matter © 2013 Elsevier B.V. Allhttp://dx.doi.org/10.1016/j.geoderma.2013.04.024

a b s t r a c t

a r t i c l e i n f o

Article history:Received 5 September 2012Received in revised form 18 April 2013Accepted 25 April 2013Available online 25 May 2013

Keywords:PhosphateColloid facilitated transportNanoparticlesSorptionReaction conditions

Colloid (typically b1 μm) facilitated phosphorus (P) transport in agricultural soils has received muchresearch attention in recent decades due to the possibility of a mobile solid form of P, which may increase Prunoff in some soils. The existing literature reports inconsistent results of colloidal P release as a percentageof total P in soil solution, separated at operationally defined size definitions by various centrifugation and/orfiltration methods. The amount of total soil P attributed to colloidal particles remarkably ranges from 0 to95%. The cause of colloidal P release has been difficult to assess since their experimental/environmental con-ditions (e.g., soil type, pH) vary from study to study. The case study presented here aims to focus on three con-trollable reaction conditions: ionic strength (I), pH and added P concentration, to elucidate their effect on Pcolloid release. Our case study showed that these reaction conditions do not largely alter the release of colloi-dal P in Ultisols of the southeastern United States (mildly acidic native pH, low organic C). However, the avail-ability of colloidal Pwas slightly enhanced at near neutral pH in high [P] or high I (0.1 M) solutions.We furtherdiscussed these reaction conditions in selected literature studies to better assess the occurrence of colloid fa-cilitated P transport in a wide variety of soil types and conditions.

© 2013 Elsevier B.V. All rights reserved.

1. Introduction

It has been well established that labile P originating from long-termanthropogenic agricultural amendments, has severe impacts on fresh-water and marine ecosystems (e.g., Carpenter et al., 1998; Sharpley etal., 2001; Sims et al., 1998). Primarily, eutrophication fueled by labileP has caused shortages of dissolved O2 (Lewitus et al., 2003) and exces-sive cyanobacteria contamination, creating hazardous conditions forhumans and livestock (Lawton and Codd, 1991). Nutrient (e.g., P andnitrogen)-enrichedwaters are also known to induce outbreaks of harm-ful algae (Pfiesteria spp. and/orKryptoperidinium spp.) blooms in estuar-ies and coastal waters from New York to Texas (Lewitus et al., 2002;Verity, 2010), and this has been linked to several massive fish kill inci-dents along the U.S. east coast from Delaware to the Carolinas(Burkholder et al., 1992; Rublee et al., 1999, 2001). The transport of ex-cess P from agricultural fields into surface water is undesirable for botheconomic and environmental reasons.

Several transport processes (i.e., erosion, runoff, and particulate-facilitated) are responsible for non-point P pollution. Successful predic-tion and management of mobile P is essential to protecting inland andcoastal watersheds. While traditionally all P remaining after a 0.2 μm

st and Environmental Sciences,Tel.: +1 864 656 3515 (Office).Voort).

rights reserved.

filtration (or even 0.45 μm), was considered “dissolved,” it has nowbeen well-established that colloidal- and/or nano-sized particles maybe contributing to this portion of the soil P pool (Bauer et al., 1996;Filella et al., 2006; Haygarth et al., 1997; Hens and Merckx, 2002;Kretzschmar et al., 1999). While the proportion of molybdenum reactivephosphate (MRP) made up of colloidal-sorbed P (typically b 1 μm) ishighly variable, it may contribute up to 25% (Filella et al., 2006). The im-portance of particulate-facilitated transport, in particular, has been iden-tified in recent years (Kretzschmar et al., 1999). Phosphate has the abilityto be carried on particles due to its strong sorption onto soil surfaces(Heathwaite et al., 2005; Hens and Merckx, 2001). Several case studiesshowed that colloid-facilitated transport processes were readily occur-ring under various environmental conditions (e.g., soil composition, I,pH, and [P]).

A recent study by Zang et al. (2011) observed >25% increase incolloid-facilitated P transport in soils receiving manure applicationsover those with no manure addition. Meanwhile, in soils without ma-nure additions, there was a strong correlation between colloidal Pconcentration and colloidal iron (Fe) concentration (Zang et al.,2011). Makris et al. (2006) also cited the importance of manure addi-tion in the formation of soil colloids. Both artificially increasing thenumber of water-dispersible colloids and adding manure amend-ments to soils were found to increase colloidal inorganic P in approx-imately equal amounts (Makris et al., 2006).

The proportion of P transported on colloids is also highly variablefrom month to month in the same region, as reported by Heathwaite

Page 2: Reaction conditions control soil colloid facilitated phosphorus release in agricultural Ultisols

Fig. 1. Transmission electronmicroscopic analysis of Cecil soil nanoparticles (d = 0.01 to0.2 μm). A) EDS analysis of amorphous material. Cu and C originate from the grid holderand support film; B) TEM image of amorphous Al–Si-rich material and nano-hematite;C) EDS of nano-hematite mixed with some amorphous Al–Si-rich material; and D) SAEDof B and matching of reflections with hematite.

102 A.R. VandeVoort et al. / Geoderma 206 (2013) 101–111

and Dils (2000). This is primarily ascribed to seasonal weather changes,particularly rainfall, and indicates the importance of solution compositionon colloidal-facilitated P transport processes (Heathwaite andDils, 2000).

The influences of dispersion and aggregation imparted by I havean effect on the number and size of soil colloids present in soil solu-tion. A dramatic example of this was shown by Koopmans and co-workers in their 2005 paper. When I was low (b1 mM), MRP wasabundant in the colloidal fraction of soil solution (filtered at0.45 μm), but when I was increased to > 0.5 M prior to filtration,MRP in the colloidal portion decreased by over 90% (Koopmans etal., 2005). This suggests that the high I condition promoted floccula-tion, producing particles > 0.45 μm in diameter; and that the soil col-loids present in the dispersed (low I) sample contributed primarily tothe MRP value. Work by Heckrath et al. (1995) showed that phos-phate can easily be mobilized when Olsen-extractable P, includingdissolved and colloidal P, is >60 mg/kg, while it was effectively im-mobile at lower P concentrations. This relationship between P con-centration and mobility/desorption was further investigated by Araiet al. (2005). They reported that the addition of P to soil solution en-hanced the release of P (i.e., dissolved P plus colloidal P via inductive-ly coupled plasma-atomic emission spectroscopy (ICP-AES) analysis)beyond the addition concentration in P rich Ultisols (Arai et al., 2005),which suggests the effects of the P ligand on total P release from soils.Ilg et al. (2008) hypothesized that excessive P inputs to soil systemsmay induce colloid-facilitated P transport processes (Ilg et al., 2008).

While there is clear evidence of soil colloid-facilitated P transportprocesses, few recent studies indicated the absence of colloid facilitat-ed P transport in soils. In our previous work, the occurrence of colloi-dal P was insignificant in weakly acidic agricultural soils (Rick andArai, 2011). Another recent field-scale study by Fuchs et al. (2009)also observed a lack of colloid-facilitated P transport through bothpreferential and non-preferential flow paths in a natural settingunder mildly acidic conditions (Fuchs et al., 2009).

The discrepancy in these studies leads us to question whether thecolloid facilitated transport process is important in all soil systems.While differences in the methodology of particle separation couldhave resulted in different findings, we believe that the occurrence ofcolloid (e.g., >200 nm)-facilitated P transport processes is highly de-pendent on dispersibility and coagulation of soil nanoparticles (NPs)(e.g., b200 nm) that are influenced by environmental conditions.The greater the NP coagulation, the greater the formation of colloids,resulting in colloid facilitated transport. For example, at high I, a com-pressed electrical double layer facilitates the coagulation of NPs intocolloidal-sized particles. Anderson et al. (1985) reported that phos-phate induced goethite coagulation in aqueous system (Anderson etal., 1985). Unfortunately, the influence of multiple variables (e.g., I,[P]) on soil NP stability has been rarely tested together in a singlecase study. Therefore, it is difficult to validate the effect of reactionconditions on the occurrence of colloid facilitated P transport in het-erogeneous systems like soils.

In this study, we hypothesize that colloid facilitated P transport iscontrolled by coagulation of NPs influenced by reaction conditions inagricultural soils. Based on the studies discussed above, the followingreaction conditions: I, pH, and P concentration, were tested togetherin the same P rich agricultural soil. To test the hypothesis, experimen-tal geochemistry (batch desorption experiments) was carried out inconjunction with zeta potential measurements, dynamic light scatter-ing, and transmission electron microscopy.

2. Materials and methods

2.1. Materials

A long-term poultry litter amended topsoil (0–20 cm), Cecil (CC)sandy loam (fine, kaolinitic, thermic Typic Kanhapludult), was collectedin South Carolina. Physicochemical properties of soils are characterized

and described in the previous work by Rick and Arai (2011). Operation-ally defined soil colloidal (200–400 nm) particleswere collected using acombination of filtration and centrifugation method as described in theP desorption section. Crystalline kaolinite (KL) (KGa-1, ClayMineral So-ciety, Chantilly, Virginia), nano-hematite (HM) (99.5% purity, 30 nmdi-ameter, US Research Nanomaterials, Inc. Houston, Texas), muscovite(MS) (Paris Hill, Maine), and lab synthesized goethite (GT) were usedfor zeta potential measurements and particle size distribution analysis.Characterization and chemical treatments (removal of carbonates andoxides) of KL are described in previous work by this research group(Arai, 2010). Muscovite flakes were ground using a dish and puckmill, and passed through a 2 μm sieve. Goethite was synthesizedaccording to the ferric nitrate method described by Schwertmann andCornell (1991). These aluminosilicate and Fe oxyhydroxide mineralsrepresent model minerals found in soil NPs via transmission electronmicroscopy (TEM) analysis, described later.

Page 3: Reaction conditions control soil colloid facilitated phosphorus release in agricultural Ultisols

103A.R. VandeVoort et al. / Geoderma 206 (2013) 101–111

2.2. Transmission electron microscopy analysis

Transmission electronmicroscopy analyseswere conducted on a CM300 FEG operating at 297 kV. Soil NP suspensions were deposited on aCu TEMgrid. Particles of interest were chosen for conventional TEM, se-lected area electron diffraction (SAED), and fast Fourier transforms(FFT) of high-resolution images to determine the crystalline nature ofthe phases. X-rays were collected using a light element Oxford energydispersive spectrometer (EDS) and processed by the software ESVision4. Imageswere processed (FFT) using theGatanDigitalMicrograph soft-ware package.

2.3. Phosphate desorption experiments

Phosphate desorption from CC soil was measured via batch methodusing air-dried whole soil b 2 mm. For each batch method sample,1.5 g CC soil was measured into 35 mL polycarbonate Oakridge centri-fuge tubes, along with 30 mL solution with 0.01 or 0.1 M I, adjustedwith NaNO3, and appropriate bicarbonate concentration for the desired

Fig. 2. Transmission electronmicroscopic analysis of Cecil soil nanoparticles (d = 0.01 to 0.2 μmB) EDS of nano-hematite cluster.

final pH, as determined through a solubility calculation from VisualMINTEQ version 3.0 (Gustafsson, 2010). The samples were allowed tohydrate for approximately 24 h, after which the pH was adjusted tothe desired level (3 to 8) using NaOH and/or HNO3. The samples wereallowed to hydrate for an additional 24 h, after which 0, 0.15 mL, or1.5 mL of a 20 mM Na2HPO4 solution was added, resulting in a final Pconcentration of 0, 0.1, or 1 mM as P, respectively (referred to hereafteras “initial P” or Pi). After a 48 hour reaction time on an end-over-endshaker, all samples were centrifuged at 2300 ×g for 3 min, resulting inan aliquot containing only particles smaller than 0.41 μm in diameter.Of this aliquot, a portion was reserved for elemental analysis viaICP-AES after acidification at the Clemson University Agricultural Ser-vices Laboratory (“total elements” or PT, FeT, AlT, or CaT hereafter). Therest of the samplewas filtered at 0.2 μm through polyvinylidene fluoride(PVDF) filters; elemental analysis was again conducted with ICP-AESafter acidification (“filtered elements” or Pf, Fef, Alf, or Caf hereafter).

The colloidal contribution (PC, FeC, AlC, or CaC hereafter) to PT, FeT,AlT, and CaT was determined by subtracting the filtered element con-centration value for each sample from its respective total element

). A) HRTEM image of nano-hematite particles ranging from 7 to 15 nm in diameter; and

Page 4: Reaction conditions control soil colloid facilitated phosphorus release in agricultural Ultisols

104 A.R. VandeVoort et al. / Geoderma 206 (2013) 101–111

concentration value (PC = PT − PF, etc.). Colloids are operationallydefined in this study as soil particles between 0.41 and 0.2 μm indiameter.

Fig. 3. Transmission electron microscopic analysis of Cecil soil nanoparticles (d = 0.01 to 0ciated nano-hematite cluster; C) EDS of phyllosilicate; and D) SAED of phyllosilicate and m

The P sorption onto soil surfaces was determined on a soil massbasis for both the total and filtered portions. The contribution ofthe colloidal particles towards PT sorption was calculated for each

.2 μm). A) EDS of nano-hematite cluster; C) TEM image of phyllosilicate with an asso-atching of measured reflections with kaolinite.

Page 5: Reaction conditions control soil colloid facilitated phosphorus release in agricultural Ultisols

Fig. 4. pH dependent release of soil colloids (d: 0.2–0.4 μm) associated with phosphorus (A), iron (B), aluminum (C), and calcium (D) as a function of ionic strength (0.01 and 0.1 MNaNO3). All experiments were conducted in the presence of 0.1 mM NaH2PO4.

105A.R. VandeVoort et al. / Geoderma 206 (2013) 101–111

data point by subtracting the Pf sorption value from the PT sorptionvalue. The same technique was also followed for Fe, Al, and Caanalysis.

Fig. 5. Average particle diameter of goethite and two soil components (hematite andkaolinite) as determined by dynamic light scattering measurements at pH 7 in 0.01or 0.1 M NaNO3 solutions.

2.4. Zeta potential and particle size characterization

Model mineral solutions were analyzed for zeta potential and par-ticle size (via dynamic light scattering, DLS) using a Brookhaven In-struments particle separation system (Brookhaven Instruments,Holtsville, New York). Model minerals were prepared by hydratingfor several days at pH 5.5 ± 0.2 in 0.01 M NaNO3. One day beforezeta potential and DLS measurements were performed, some sampleswere adjusted for ionic strength, up to 0.1 M NaNO3. All samples wereadjusted for pH, and achieved pH 5 or pH 7 with the use of HNO3

and/or NaOH. These samples were within 0.2 pH units of the targetlevel. Finally, phosphate solution was added to select samples toachieve 0.1 to 1 mM [P].

The pH range (pH = 5 and 7)was chosen tomimic the optimumpHrange of soil NP release in preliminary experiments. Because of our in-terest in the effect of phosphate ligand, all samples were keptsuspended in distilled deionized water (18 mΩ) prior to the additionof orthophosphate stock solution ([P] = 1.005 mM). Zeta potentialand DLS measurements were taken on model mineral samples withno added P, and again after the addition of 0.01 mL P stock solution tothe 2 mL samples, resulting in a [P] of 0.1 mM. Zeta potential and parti-cle size were measured for a third time after the further addition of0.99 mL of P stock solution, resulting in the final [P] or 1 mM. Suspen-sion density of model minerals is approximately 0.17 g/L for KL,0.10 g/L for GT, and 0.21 g/L for HM. Baseline data were establishedusing a reference of 92 nm latex spheres. Baseline and all sampleswere read continuously for 50 s. Data were analyzed using BrookhavenInstruments 90Plus Particle Sizing Software, ver. 4.14.

3. Results and discussion

3.1. Mineralogy and chemical properties of soil nanoparticles

High resolution (HR) TEM equipped with EDS revealed both mor-phological and elemental data of soil NPs. There was evidence of Al,Si, and Fe in these natural NPs (~5–150 nm), which contained onlyminor amounts of P, Ca, Ti, Cl, S, and Mg. Based on a number of TEMobservations (several tens of micrographs), we identified the follow-ing three major types NPs: amorphous aluminosilicate materials, HM,and KL/MS.

Soil NPs are primarily composed of amorphous materials, whichrich in Al and Si, and variable in composition (Fig. 1A and B). SAEDpatterns of this material contained broad rings indicative of an

Page 6: Reaction conditions control soil colloid facilitated phosphorus release in agricultural Ultisols

106 A.R. VandeVoort et al. / Geoderma 206 (2013) 101–111

amorphous structure mixed with variable amounts of sharp reflec-tions indexed as an iron oxide (Fig. 1D).

Hematite was found in small (~10 nm) electron dense particlescontaining high Fe concentrations. Figs. 1B and 2A have clusters ofHM nanocrystals (~7 to 15 nm in diameter). EDS analyses of the FeNPs always contained Al, Si, P and Ca (Figs. 1C, 2B). SAED patternsof these NPs index as HM, clearly indicating that the NPs are HMand not GT (Fig. 1D). Fast Fourier transforms processing of severalHRTEM images yielded reflections for HM. Often, these Fe NPs weremixed with amorphous Al–Si material (Fig. 1B).

Several EDS spectra (e.g., Fig. 3C) showed near equal intensity of Siand Al, inferring the presence of 1:1 or 2:1 layer silicate phase. In ad-dition, the SAED pattern in Fig. 3D can be indexed as the hk0 net of aphyllosilicate mineral such as KL. However, these reflections are notdiagnostic of any one phyllosilicate mineral. The ratio of Al to Si isclose to 1:1, and K and Na are present. Yet, these alkali elements arepresent everywhere on the grid, including the carbon support film.If K is structural, then the phyllosilicate would be a 2:1 layered silicatesuch as illite or MS, but if K is a contaminant, then the phase is KL.Both KL and illite are present is the soil so neither phase can beruled out.

Because of the common occurrence of apatite-based P fertilizer inagricultural soils, the presence of hydroxyapatite was also investigatedvia EDS analysis. Phosphorus was usually associated with both the

Fig. 6. pH dependent release of soil colloids (d: 0.2–0.4 μm) associated with iron (A), alu-minum (B), and calcium (C) as a function of [P]: 0, 0.1, and 1 mM NaH2PO4.

amorphous phases and HM. In these analyses, the intensity of Ca in50% of the spot analyses was half the intensity of P. A Ca–P rich particlewas found during the particle survey, suggesting that, althoughapatite-like NPs are present, they are not a primary contributor to thecolloid facilitated P transport process. Overall, the results of HRTEManalyses suggested that P in soil solutions was predominantly associat-ed with amorphous Al and Si rich NPs, along with nano-HM.

We also compared the zeta potential values of GT, MS, and soil col-loid components in 0.01 NaNO3 solutions at pH 7. MS, KL, and HM, allshowed negative potentials (−25.71 (±1.17) mV, −26.76 (±2.42)mV, −16.74 (±3.03) mV, respectively), supporting the findings ofTEM that soil NPs consists of these minerals components. However,GT showed a slightly positive zeta potential value, 3.97(±2.84) mV.To further investigate the mobility of P associated coagulated soil NPs,we studied the effect I and P ligand concentration on the formation ofsoil colloids using zeta potential measurements, DLS, and batch desorp-tion experiments.

3.2. Ionic strength (I) effect on colloid release associated with P, Fe,and Al

Fig. 4 shows the effect of I on colloid (200–410 nm) release atpH 3.5–7.5. Because of common association of P with Al that was foundin the TEM analysis, the release of PC, FeC, and AlC are each individuallyshown in Fig. 4A–C. To assure the uncommonoccurrence of Ca associatedNP/colloids, we reported the release of CaC in Fig. 4D. We conducted ex-periments under constant [P] = 0.1 mM because this condition yieldsnegligible effects on the release of FeC and AlC at I = 0.01 M (Fig. 4Aand B), allowing the effect of I to be studied. It is interesting that samplesat the lower I consistently showed little release of colloids for all ele-ments. Ionic strength is known to affect diffused double layer thickness,resulting in coagulation and dispersion of NPs (Hiemenez, 1986). HigherI (0.1 M) promoted increased colloidal fractions over low I (0.01 M). Al-though this amount P is low (0.5–2 mg/L), it is high enough to trigger eu-trophication in waterways. Eutrophication can be induced when labile Pis as low as 0.04 mg/L in natural environments (Sims et al., 1996).

Increasing the I of the soil solution reduces the thickness of thediffused double layer, which should promote coagulation of soilNPs (b 200 nm). In our previous investigation, Fe-rich aluminosili-cate NPs were readily available from soils of the same series (Rickand Arai, 2011). It is likely that the stability of dispersed NPs was af-fected by I. This coagulation was especially pronounced at pH ≥ 6.5

Fig. 7. Zeta potential of soil components muscovite (MS), kaolinite (KL), and hematite(HM); and goethite (GT) at ionic strength = 0.01 M NaNO3. The reaction conditionwas varied at pH 5 or 7 and total [P] = 0, 0.1 or 1 mM.

Page 7: Reaction conditions control soil colloid facilitated phosphorus release in agricultural Ultisols

107A.R. VandeVoort et al. / Geoderma 206 (2013) 101–111

for PC, FeC, and AlC (Fig. 4A–C). It is interesting that the release of PC,FeC, and AlC follows the same pH trend (Fig. 4), suggesting thatcolloid-facilitated transport of P is associated with Al- and Fe-richparticles, such as amorphous aluminosilicates and nano-HM. Thistrend drastically increases with increasing pH from 3.5 to 6.5. Onthe contrary, the release of CaC was not pH dependent (Fig. 4D). Be-cause of weakly acidic soil pH, Ca-based colloids were not expecteddue to its high solubility at this pH (Fig. 4D).

To observe the effect of I on particle size distribution, DLS measure-ments of soil components (HM and KL) and GT were compared at bothlow (0.01 M) and high (0.1 M) I levels at pH 7 (Fig. 5). No data werecollected for MS, due to its large average particle size (>100 μm),which was beyond the capabilities of the DLS instrument. For all sam-ples, the average particle size at high I was approximately twice thatof the particle size at low I. For GT, HM, and KL, it was 2.72, 2.21, and2.38 times as high, respectively. Within each component, the differencein particle size between high and low I was > 1 standard deviation. Insummary, DLS data of NP components support I-dependent soil colloidrelease (Fig. 4) that increasing ionic strength promotes the coagulationof soil nanoparticles strongly.

3.3. Effects of phosphate on colloid release associated with Fe, Al, and Ca

Fig. 6 shows the effect of [PT] on colloid release from the soil at theconstant background electrolyte solution, 0.01 MNaNO3. The total con-centration of FeC, AlC, and CaC is individually presented in Fig. 6A, B, andC, respectively. Varying P concentration greatly affected the release ofFeC and AlC (Fig. 6A and B). When soils were suspended in “P-free” so-lution, moderate colloid release (5–70 mg/L for Fe, 12–115 mg/L forAl) was observed across the pH spectrum (approx. pH 4–9) (Fig. 6A).Interestingly, the addition of P (total P = 0.1 mM) reduced the releaseof colloid to near negligible levels. A small quantity of CaC was observedin the low P concentration solution ([PT] = 0.1 mM) at pH>5 (Fig. 6C).However, the amount of Ca-P colloids was two orders of magnitude lessthan those of AlC and FeC. Colloidal Ca release was not significantly af-fected by [P]. This was not a surprising result since the TEM analysis

Fig. 8. Particle size as determined by dynamic light scattering of goethite and soil nanoparticand pH 5 and 7: A) goethite (GT) at pH 5; B) GT at pH 7; C) hematite (HM) at pH 5; D) Hoperationally defined soil colloid definition of 0.2 to 0.41 μm (=200–410 nm). Several linessamples) due to all particles in these samples being >9900 nm in diameter.

indicated only rare occurrences of CaC in the soil extracts.Whenwe fur-ther increased the P addition to 1 mM, the release of soil colloids wasenhanced. The release of colloids was much greater than that of the0.1 mM P system, especially at near neutral pH values (160 mg/L forFeC and 350 mg/L for AlC).

To explain the negligible PC portion at PT = 0.1 mM, it is likely thatspecific adsorption of P on NPs facilitated the dispersion of colloidsand/or the coagulation of NPs to > 0.41 μm. Several theories can bepostulated to explain the coagulation. First, addition of 1 mMP resultedin greater I in the system comparing to the system with lower [P](0–0.1 mM). The increased I resulted in a thinner diffuse double layer,resulting in turn, in the formation of colloids. Second, the potential ofP sorbed NPs was near equal to the ISP at given soil pH, which could re-sult in the coagulation of NPs. Third, it is known that the specific adsorp-tion of PO4 on goethite can bridge goethite particles, enhancing thecoagulation when P was added (Anderson et al., 1985). To test the the-ory above, wemeasured 1) the change in surface charge of soil NP com-ponents, goethite, and muscovite via zeta potential measurements, and2) changes in particle size distribution during P sorption via DLS mea-surements for soil NP components and goethite.

Zeta potential, or electrophoretic mobility (EM), measurements are auseful microscopic approach for determining surface charge propertiesof NPs and colloids. Inner-sphere complexes cause shifts in both EM andisoelectric point (IEP) (lowering zeta potential) due to specific ion adsorp-tion inside the shear plane (Hunter, 1981). The shear plane is at the outeredge of the inner part of the double layer and near the outer Helmholtzplane or the Stern layer, depending on the models to describe the inter-face (Hunter, 1981). Fig. 7 shows the zeta potential values of soil compo-nents HMandKL, alongwith phyllosilicateMS and Fe oxyhydroxideGT at0.01 M I at 0, 0.1, and 1 mM P addition at both pH 5 and pH 7.

In nearly all cases, with the exception of MS at pH 7, the additionof 0.1 mM P caused a dramatic decrease in the zeta potential from the0 mM P level, indicating more negatively charged particles. This de-crease was more than one standard deviation below the 0 mM Pvalue for all soil component samples, again, except MS at pH 7. Thechange in zeta potential was most dramatic for GT, which decreased

le components at ionic strength = 0.01 M NaNO3 as a function of [P] = 0, 0.1 or 1 mMM at pH 7; E) kaolinite (KL) at pH 5; and F) KL at pH 7. Shaded regions represent thenot displayed (i.e., HM 0.1 mM P at pH 5 and 7, KL 0.1 mM P at pH 5, and all muscovite

Page 8: Reaction conditions control soil colloid facilitated phosphorus release in agricultural Ultisols

Fig. 9. Negative phosphate sorption (P release) envelope on Cecil sandy loamas a functionof [P]: A) 0, B) 0.1, and C)1 mM NaH2PO4. All experiments were conducted in 0.1 MNaNO3. Light gray shaded areas indicate dissolved P concentration, while dark gray shad-ed area indicate colloidal P concentration.

108 A.R. VandeVoort et al. / Geoderma 206 (2013) 101–111

by 49.5 mV at pH 5 and 50.7 mV at pH 7with the addition of 0.1 mMP,followed by HM, which decreased by 20.3 and 11.5 mV at pH 5 and 7,respectively. This supports the hypotheses of inner-sphere P adsorptionby GT, KL, and HM. A further addition of P to 1 mM concentrationresulted in a much less pronounced decrease in zeta potential. Goethiteand HM were again the components with the largest decrease in zetapotential from 0.1 to 1 mM [P]. For GT, the zeta potential dropped by10.9 and 7.3 mV at pH 5 and 7, respectively. For HM, the zeta potentialdropped by 13.9 and 5.0 mV at pH 5 and 7, respectively. While all othersamples had a decrease in zeta potential by > 1 standard deviation be-tween 0.1 and 1 mMP, KL at pH 7 actually saw a slight increase or nearequal in zeta potential. Our findings of P specific adsorption on soilmineral colloids agree with several studies. Phosphate sorption onferrihydrite, HM, GT, and boehmite (γ-AlOOH) is known to forminner-sphere surface complexes due to charge reversal and a lowerIEP with increasing P loading level (Anderson and Malotky, 1979; Araiand Sparks, 2001; Bleam et al., 1991; Elzinga and Sparks, 2007;Hansmann and Anderson, 1985; Tejedor-Tejedor and Anderson, 1990;Van Emmerik et al., 2007).

Next, we examined the effects of P specific sorption on particlesize distribution in GT, HM, and KL (Fig. 8). Shaded regions in Fig. 8indicate the particle size range of colloids (200–410 nm) defined inthis study. As stated previously, the release of soil colloids fluctuatedwith changes in [P] (Fig. 6). While colloids were not detectable at[P] = 0.1 mM at pH 5 to 7, they do became available when [P] iszero or 1 mM. We find that the effect of P on coagulation is mineralspecific. When [P] is zero, some HM particles are available at the200 to 410 nm range, but KL is only available at pH 7. KL is not withinthe operationally defined colloid size at pH 5. Interestingly, GT parti-cles are not within the range at either pH value. With increasing [P] to0.1 mM, GT particles became dispersed to be within the defined col-loid size range at both pH values. When we further increased [P] to1 mM, HM particles are only within this size range at pH 7. In thecase of KL, the availability of colloids within the range increasedwith decreasing pH. Although we were unable to demonstrate DLSmeasurements on MS and amorphous aluminosilicate NPs, the re-lease of soil colloids can be postulated by mineral specific coagulationproperties of soil colloid components.

3.4. Effects of phosphate on colloidal phosphorus release

As discussed in the previous section, changing the P ligand concen-tration could contribute to the coagulation of NPs. The concentrationof dissolved P in the field can be varied depending on timing of agricul-tural practices. While high [P] in runoff is expected in manure amendedsoils during storm events (e.g., Addiscott and Thomas, 2000; Hollmannet al., 2008; Makris et al., 2006), low [P] in soil solutions is a likely sce-nario after the growing season (e.g., Mayer and Jarrell, 1995). Therefore,it is important to study the effect of P ligand concentration on colloid fa-cilitated P release. The PC contribution over various [PT]was analyzed bycomparingMRP (b0.2 μm) to PC over the pH spectrum (Fig. 9). In Fig. 9,negative sorption corresponds to the release of PT. The concentration ofP associated colloid andMRP is shown in the dark shaded area and lightshaded area, respectively. The following sections discuss: 1) PC releaseas a function of [P] and 2) a comparison between PC and MRP.

Similar to overall colloid concentration, as determined by FeC andAlC (Fig. 6A–B), the PC fraction (dark shaded area in Fig. 9A) wasgreatest at 1 mM added P, followed by 0 mM added P, and lastly by0.1 mM added P. When soils were suspended in “P free” solutions, thePC fraction was largest (6.24 mg/kg) at pH = 4.67, while with0.1 mM added P, PC only occurred at approximately 3.12 mg/kg atpH 6.2 (Fig. 9A–B). Increasing the P addition to 1 mM caused PC to belargest (36.25 mg/kg at pH 4.23, and 18.31 mg/kg at pH 7.07), with adecrease in the mildly acidic pH range.

In our system, the contribution of PC is very small comparing to theMRP at respective pH values in each P treatment. At the largest PC release

in each system, the % ratio of PC: MRP is 6.85% when [P] = 0 mM, 2.65%when [P] = 0.1 mM, and 5.31% when [P] = 1 mM. This finding is con-sistent is our previous investigation that PC release in this Ultisol isnear negligible (Rick and Arai, 2011). Overall, there are substantial quan-tities ofMRP release in each system. The [MRP] increaseswith increasingthe addition of P. In the 0, 0.1, and 1 mM [P] systems, [MRP] was as highas ~250 mg/L at pH 7.2 (Fig. 9A), ~240 mg/L at pH 3.5 (Fig. 9B), and~720 mg/L at pH 7.4 (Fig. 9C), respectively. Overall, the MRP fractionshowed a similar trend across the P additions, with the lowest MRP re-leased in mildly acidic pH conditions, between pH 5.5 and 6.5 (Fig. 9).This decrease is probably attributed to the formation of Ca/Mgphosphatesolubility products. Overall, P release for both 0.1 mM and 1 mMwas ingreat excess of the amount added (Fig. 9B–C), especially below pH 4 andabove pH 7 (Fig. 9A and B). Considering the total P level in this P rich soil,it was likely that P sorption was already at maximum sorption prior tothe experiments. Our agronomic Mehlich-1 test suggested the level ofP = 222.5 mg/kg, which is rated excessively high.

3.5. Significance of colloid facilitated phosphorus transport in soils

This study, as well as previous work by our research group (Rickand Arai, 2011), showed very little contribution of PC in the samesoil series. Several other studies also recorded little to no observationof PC (Fuchs et al., 2009; Liang et al., 2010; Siemens et al., 2008). Thiswas particularly apparent in the study by Fuchs et al. (2009). While PCwas not the primary P form in these studies, experiments conductedunder different setting and environmental condition produced a

Page 9: Reaction conditions control soil colloid facilitated phosphorus release in agricultural Ultisols

Table 1Summary of selected colloidal P literature. Studies arranged in descending order of observance of colloidal P.

Soil type pH Org. C (%) Total P (mg/kg) Settings of study I (M) Part. size definition Coll.-P as % of total P Notes Ref.

Heavy clay fromWaardenburg,The Netherlands

6.7 to 7.7 5.4% at surface 747 at surface Batch 0.03 M using CaCl2,adjusted to 0.51 Mwith NaCl

>0.45 μm Refiltered sampleshad avg. 94% morecolloidal P

0.45 μm low-I filtrates given0.51 M NaCl, then refiltered

Koopmanset al. (2005)

Sandy loam nearRøgen, Denmark

6.42 to 8.16 1.5% 734 Column study 0.0088 M 0.24 to 2 μm 75% Downward particle transportin soil columns structureimportant

de Jongeet al. (2004)

Fine sandy soilsnear Münster,Germany

5.5 to 6.2 atsurface, 7.1to 8.2 below4 m

1.1 to 2.5% at surface,decreasing w/ depth

610 to 884 at surface,decreasing w/ depth

Column study No added electrolytes b0.01 to 0.45 μm Highly variable,from approx. 0 to 95%

Artificially increasing [P]may promote colloidal Pformation, esp. at low I

Siemenset al. (2004)

Peaty podzol fromnortheastScotland

5.0 High, due to peat present 3356, additionalfrom P mineralfertilizers

Batch 0.03 M w/ CaCl2,adjusted furtherto 0.1 M w/ NaCl

Multi-filtration at1.6, 1.2, 0.45,and 0.22 μm

MRP: 55% >1.2 μm,19% >0.45, 4%>0.22 μm;OP: 34% >1.2 μm,39% >0.45 μm, 7%>0.22 μm

MRP assoc. w/ large soilcolloids, while OP assoc.w/ small colloids in highOM soils

Shand et al.(2000)

Several UK soilsranging fromsilt loam to clay

Ranges from5.3 to 8.1

Ranges from 0.13to 0.37%

Ranges from 461 to1373

Both batchand field studies

Varies from 0 addedto 0.03 M w/ NaCland CaCl2

Size cutoffs at 2, 0.8,0.45, 0.22,and 0.001 μm

Highly variablebetween size fractionsand P additions

Colloidal P increasedw/ fertilizer P additions,esp. on colloids >0.45 μmin size

Heathwaiteet al. (2005)

Sand and sandyloam fromNorthwestGermany

Top soil 6.1to 7.2, below30 cm 4.7–6.8

Not given, but inc. w/addn of org. fertilizer

Available P 47 to 140.Various P fertilizerused

Column study 0 added and 0.01 M KCl Approx. 0.01to 1.2 μm

Around 28% for top30 cm, up to 94% below30 cm depth

Colloidal P release fac.by low I

Ilg et al.(2005)

Silty loam nearHang-zhou, China

Soil pH 6.0,adjusted tovariable levelsfrom 1.4 to 9.9expt.

1.74% 1720 Batch No added electrolytes Approx. 0.01to 1 μm

b5% between pH 4 and7; 25.5% at pH 7.6

Colloidal P greatest belowpH 4 and above pH 7

Liang et al.(2010)

Sandy loam soilfrom upstateS. Carolina

5.32 to 5. 69 b0.5% 262 to 900 Batch 0.065 M adjustedw/ NaNO3

0.01 to 0.2 μm Neg., b15% P not assoc. w/ smallcolloids in acidic, sandy,low OM soils

Rick andArai (2011)

Gravelly loam nearTahlequah,Oklahoma

Mildly acidiccharacteristicof region

Low (0.5–1%)characteristic of region

33 in top 15 cm, nofertilizer for severalyears

100 mg/L P injectedconst. trench sites infield

No added electrolytes Filt. at 0.45 μm;colloid vs. dissolveddetermined throughanalytical methods

Neg. Essent. no colloidal Punder pref. or non-pref.flow paths

Fuchs et al.(2009)

Georgetown, S.Carolina sed.,95% quartz

5.2 inground-water

0.1% 0 to 120 mg/L Padded

Column study 0.0004 M in ground-water >0.03 μm Not meas. Most coll. rel: Addn of0.5 mM P + 0.5 mMascorbic acid (by turbidity)

Swartz andGschwend(1998)

109A.R.V

andeVoort

etal./

Geoderm

a206

(2013)101

–111

Page 10: Reaction conditions control soil colloid facilitated phosphorus release in agricultural Ultisols

110 A.R. VandeVoort et al. / Geoderma 206 (2013) 101–111

wide range of results. We believe that the lack of colloid-facilitatedtransport, including the case study information provided here, canbe attributed to the following physicochemical parameters: 1) totalP level in native soils, 2) % organic matter (OM) content in soils,3) soil pH, and 4) I of soil solutions. To facilitate the evaluation ofthese physicochemical parameters, the following colloid facilitated Ptransport studies were selected and the results were carefullysummarized in Table 1. The studies are arranged by degree of PCobservation, decreasing down the column.

A low level of native P (33 mg/kg) in the soil, especially apparent inthe Fuchs et al. (2009) study, may lead to overall P sorption in bulk soilminerals, instead of P release with particulates. Because of the strongchemisorption of P in soils, the input (100 mg/L) of P ligands may notcontribute sufficiently to surface charge properties of NPs or solubilityproducts. If P saturation has not yet been reached in a soil, the soil sur-faces themselves may be providing a sink for overall P sorption.

Our investigations including this manuscript and the previous work(Rick andArai, 2011) have been conducted in a soil with lowOMcontent(b0.5%). This is typically the case for other soils lacking PC contributionsas well (Fuchs et al., 2009; Siemens et al., 2008). In the literature sur-veyed, studies conducted in high OM soils had much higher rates of PCcontent (de Jonge et al., 2004; Koopmans et al., 2005; Shand et al.,2000; Siemens et al., 2004). Taking this a step further, artificially enhanc-ing the OM of the soil (e.g., adding manure fertilizer) has been shown toincrease the colloidal fraction of P even further. Similar research findings(i.e., OM enhanced colloidal P release) were also reported in severalother studies not included in Table 1 (Makris et al., 2006; Zang et al.,2011). Organically bound P and/or organic P (e.g., inositol phosphate)in humic substances might be more susceptible to solute movementthan orthophosphate bound to inorganic mineral fractions in soils.

When the results of studies that were reviewed in Table 1were com-pared, we found the importance of soil pH on PC release, especially in theabsence of high levels of OM. Mildly acidic pH appeared to impede P as-sociation with colloids in all studies considered (Fuchs et al., 2009; Lianget al., 2010; Rick and Arai, 2011; Siemens et al., 2008). It seems, as in sev-eral cases, that high OM levels can promote PC formation even in unfa-vorable pH conditions (Ilg et al., 2005; Mayer and Jarrell, 1995; Shandet al., 2000; Siemens et al., 2004). Likewise, it also appears that high pHcan promote colloidal P formation even in the absence of large amountsof OM (Heathwaite et al., 2005; Turner et al., 2004). The increase of PC atneutral to high pH values could be due, in part, to the charge character-istics of soil minerals. Many aluminum and iron (oxy)hydroxides experi-ence their point of zero charge (PZC) between pH 6.5 and 8.5 (Stummand Morgan, 1996), promoting flocculation and coagulation of particleswhen pH is near or equal to PZC, likely in the colloidal size range. Thissudden increase in available colloidal-sized particles above pH 6.5could facilitate the PC transport process.

Finally, I also appears to play a role in soil colloid formation and Passociation. The trend is not as clear as that of native [P], OM, or pH.Several studies reviewed in Table 1 exhibit a large amount of PC ob-servation after the addition of excessive amounts of mineral fertil-izers (Heathwaite et al., 2005; Ilg et al., 2005; Shand et al., 2000;Siemens et al., 2004). While electrolytes may not have been addedto these studies with the intention of increasing I, the excessive addi-tion of fertilizer will cause the same effect. Increasing I along with theaddition of P causes a compression of the diffuse double layer, leadingsoil particles to coagulate. This is likely the mechanism by which PCformed in these case studies.

In summary, it is important to note that the current general percep-tions in the field of geo/soil chemistry may overstate the importance ofcolloid-facilitated transport of P. While PC is clearly the predominantform of P in some studies conducted in high native [P] soils, high pH,high OM, and/or high fertilization/I conditions (de Jonge et al., 2004;Heathwaite et al., 2005; Ilg et al., 2005; Koopmans et al., 2005; Shand etal., 2000; Siemens et al., 2004; Turner et al., 2004), it is not the predomi-nant form of P under other conditions (Fuchs et al., 2009; Liang et al.,

2010; Mayer and Jarrell, 1995; Rick and Arai, 2011; Siemens et al.,2008). While there has been much evidence to support colloidal-facilitated P transport in the last few decades, the contribution of PCwith respect to “non-point pollution” should be carefully evaluatedbased on environmental conditions in assessing the environmental riskfrom P rich agricultural lands.

References

Addiscott, T.M., Thomas, D., 2000. Tillage, mineralization and leaching: phosphate. Soiland Tillage Research 53 (3–4), 255–273.

Anderson,M.A., Malotky, D.T., 1979. The adsorption of protolyzable anions on hydrous ox-ides at the isoelectric pH. Journal of Colloid and Interface Science 72 (3), 413–427.

Anderson, M.A., Tejedor-Tejedor, M.I., Stanforth, R.R., 1985. Influence of aggregation onthe uptake kinetics of phosphate by goethite. Environmental Science and Technol-ogy 19 (7), 632–637.

Arai, Y., 2010. Effects of dissolved calcium on arsenate sorption at the kaolinite–waterinterface. Soil Science 175 (5), 207–213.

Arai, Y., Sparks, D.L., 2001. ATR-FTIR spectroscopic investigation on phosphate adsorp-tion mechanisms at the ferrihydrite–water interface. Journal of Colloid and Inter-face Science 241, 317–326.

Arai, Y., Livi, K.J.T., Sparks, D.L., 2005. Phosphate reactivity in long-term poultry litter-amended southern Delaware sandy soils. Soil Science Society of America Journal69, 616–629.

Bauer, J.E., Ruttenberg, K.C., Wolgast, D.M., Monaghan, E., Schrope, M.K., 1996. Cross-flow filtration of dissolved and colloidal nitrogen and phosphorus in seawater: re-sults from an intercomparison study. Marine Chemistry 55 (1–2), 33–52.

Bleam, W.F., Pfeffer, P.E., Goldberg, S., Taylor, R.W., Dudley, R., 1991. A 31P solid-state nu-clearmagnetic resonance study of phosphate adsorption at the boehmite/aqueous so-lution interface. Langmuir 7 (8), 1702–1712.

Burkholder, J.M., Noga, E.J., Hobbs, C.H., Glasgow, H.B., 1992. New ‘phantom’ dinoflagellateis the causative agent of major estuarine fish kills. Nature 358 (6385), 407–410.

Carpenter, S.R., et al., 1998. Nonpoint pollution of surface waters with phosphorus andnitrogen. Ecological Applications 8 (3), 559–568.

de Jonge, L.W., Moldrup, P., Rubaek, G.H., Schelde, K., Djurhuus, J., 2004. Particle leachingand particle-facilitated transport of phosphorus at field scale. Vadose Zone Journal 3(2), 462–470.

Elzinga, E.J., Sparks, D.L., 2007. Phosphate adsorption onto hematite: an in situ ATR-FTIR investigation of the effects of pH and loading level on the mode of phosphatesurface complexation. Journal of Colloid and Interface Science 308 (1), 53–70.

Filella, M., Deville, C., Chanudet, V., Vignati, D., 2006. Variability of the colloidal molyb-date reactive phosphorous concentrations in freshwaters. Water Research 40 (17),3185–3192.

Fuchs, J.W., Fox, G.A., Storm, D.E., Penn, C.J., Brown, G.O., 2009. Subsurface transport ofphosphorus in riparian floodplains: influence of preferential flow paths. Journal ofEnvironmental Quality 38 (2), 473–484.

Gustafsson, J.P., 2010. Visual MINTEQ. KTH, Department of Land and Water ResourcesEngineering, Stockholm, Sweden.

Hansmann, D.D., Anderson, M.A., 1985. Using electrophoresis in modeling sulfate, sel-enite, and phosphate adsorption onto goethite. Environmental Science and Tech-nology 19 (6), 544–551.

Haygarth, P.M., Warwick, M.S., House, W.A., 1997. Size distribution of colloidal molybdatereactive phosphorus in river waters and soil solution.Water Research 31 (3), 439–448.

Heathwaite, A.L., Dils, R.M., 2000. Characterising phosphorus loss in surface and sub-surface hydrological pathways. Science of the Total Environment 251, 523–538.

Heathwaite, L., Haygarth, P., Matthews, R., Preedy, N., Butler, P., 2005. Evaluating colloi-dal phosphorus delivery to surface waters from diffuse agricultural sources. Journalof Environmental Quality 34 (1), 287–298.

Heckrath, G., Brookes, P.C., Poulton, P.R., Goulding, K.W.T., 1995. Phosphorus leachingfrom soils containing different phosphorus concentrations in the Broadbalk Exper-iment. Journal of Environmental Quality 24 (5), 904–910.

Hens, M., Merckx, R., 2001. Functional characterization of colloidal phosphorus species inthe soil solution of sandy soils. Environmental Science andTechnology35 (3), 493–500.

Hens, M., Merckx, R., 2002. The role of colloidal particles in the speciation and analysisof “dissolved” phosphorus. Water Research 36 (6), 1483–1492.

Hiemenez, P.C., 1986. Principles of Colloid and Surface Chemistry. Marcel Dekker, Inc.,New York.

Hollmann, M., Knowlton, K.F., Brosius, M.R., McGilliard, M.L., Mullins, G.L., 2008. Phos-phorus runoff potential of varying sources of manure applied to fescue pastures inVirginia. Soil Science 173 (10), 721–735.

Hunter, R.J., 1981. Zeta Potential in Colloid Science. Academic Press, London.Ilg, K., Siemens, J., Kaupenjohann, M., 2005. Colloidal and dissolved phosphorus in

sandy soils as affected by phosphorus saturation. Journal of Environmental Quality34 (3), 926–935.

Ilg, K., Dominik, P., Kaupenjohann,M., Siemens, J., 2008. Phosphorus-inducedmobilizationof colloids:model systems and soils. European Journal of Soil Science 59 (2), 233–246.

Koopmans, G.F., Chardon, W.J., van der Salm, C., 2005. Disturbance of water-extractablephosphorus determination by colloidal particles in a heavy clay soil from TheNetherlands. Journal of Environmental Quality 34 (4), 1446–1450.

Kretzschmar, R., Borkovec, M., Grolimund, D., Elimelech, M., 1999. Mobile subsurface col-loids and their role in contaminant transport. Advances in Agronomy 66, 121–193.

Lawton, L.A., Codd, G.A., 1991. Cyanobacterial (blue-green algal) toxins and their sig-nificance in UK and European waters. Journal of the Institution of Water and Envi-ronment Management 5 (4), 460–465.

Page 11: Reaction conditions control soil colloid facilitated phosphorus release in agricultural Ultisols

111A.R. VandeVoort et al. / Geoderma 206 (2013) 101–111

Lewitus, A.J., et al., 2002. Low abundance of the dinoflagellates, Pfiesteria piscicida,P. shumwayae, and Cryptoperidiniopsis spp., in South Carolina tidal creeks andopen estuaries. Estuaries 25 (4A), 586–597.

Lewitus, A.J., et al., 2003. Harmful algal blooms in South Carolina residential and golfcourse ponds. Population and Environment 24 (5), 387–413.

Liang, X.Q., et al., 2010. Effect of pH on the release of soil colloidal phosphorus. Journalof Soils and Sediments 10 (8), 1548–1556.

Makris, K.C., Grove, J.H., Matocha, C.J., 2006. Colloid-mediated vertical phosphorustransport in a waste-amended soil. Geoderma 136 (1–2), 174–183.

Mayer, T.D., Jarrell, W.M., 1995. Assessing colloidal forms of phosphorus and iron in theTualatin River Basin. Journal of Environmental Quality 24 (6), 1117–1124.

Rick, A.R., Arai, Y., 2011. Role of natural nanoparticles in phosphorus transport process-es in ultisols. Soil Science Society of America Journal 75 (2), 335–347.

Rublee, P.A., et al., 1999. PCR and FISH detection extends the range of Pfiesteria piscicidain estuarine waters. Virginia Journal of Science 50, 325–336.

Rublee, P.A., et al., 2001. Use of molecular probes to assess geographic distribution ofPfiesteria species. Environmental Health Perspectives 109, 765–767.

Schwertmann, U., Cornell, R.M., 1991. Iron Oxides in the Laboratory. Wiley, Weinheim,Germany.

Shand, C.A., Smith, S., Edwards, A.C., Fraser, A.R., 2000. Distribution of phosphorus inparticulate, colloidal and molecular-sized fractions of soil solution. Water Research34 (4), 1278–1284.

Sharpley, A.N., McDowell, R.W., Kleinman, P.J.A., 2001. Phosphorus loss from land towater: integrating agricultural and environmental management. Plant and Soil237 (2), 287–307.

Siemens, J., Ilg, K., Lang, F., Kaupenjohann, M., 2004. Adsorption controls mobilization ofcolloids and leaching of dissolved phosphorus. European Journal of Soil Science 55(2), 253–263.

Siemens, J., Ilg, K., Pagel, H., Kaupenjohann, M., 2008. Is colloid-facilitated phosphorusleaching triggered by phosphorus accumulation in sandy soils? Journal of Environ-mental Quality 37 (6), 2100–2107.

Sims, J.T., et al., 1996. Assessing the Impact of Agricultural Drainage on Ground andSurface Water Quality in Delaware: Development of Best Management Practicesfor Water Quality Protection. Delaware Department of Natural Resources and Envi-ronmental Control, Dover, Delaware.

Sims, J.T., Simard, R.R., Joern, B.C., 1998. Phosphorus loss in agricultural drainage: historicalperspective and current research. Journal of Environmental Quality 27 (2), 277–293.

Stumm, W., Morgan, J.J., 1996. Aquatic Chemistry. John Wiley & Sons, Inc., New York.Swartz, C.H., Gschwend, P.M., 1998. Mechanisms controlling release of colloids to

groundwater in a Southeastern coastal plain aquifer sand. Environmental Science& Technology 12 (32), 1779–1785.

Tejedor-Tejedor,M.I., Anderson,M.A., 1990. Protonation of phosphate on the surface of goe-thite as studied by CIR-FTIR and electrophoretic mobility. Langmuir 6 (3), 602–611.

Turner, B.L., Kay, M.A., Westermann, D.T., 2004. Colloidal phosphorus in surface runoffand water extracts from semiarid soils of the western United States. Journal of En-vironmental Quality 33 (4), 1464–1472.

Van Emmerik, T.J., Sandstrom, D.E., Antzutkin, O.N., Angove, M.J., Johnson, B.B., 2007.31P solid-state nuclear magnetic resonance study of the sorption of phosphateonto gibbsite and kaolinite. Langmuir 23 (6), 3205–3213.

Verity, P.G., 2010. Expansion of potentially harmful algal taxa in a Georgia estuary(USA). Harmful Algae 9 (2), 144–152.

Zang, L., Tian, G., Liang, X., Liu, J., Peng, G., 2011. Effect of water-dispersible colloids inmanure on the transport of dissolved and colloidal phosphorus through soil column.African Journal of Agricultural Research 6 (30), 6369–6376.