You are on page 1of 8

Published May, 2012

Journal of Environmental Quality SPECIAL SECTION


TECHNICAL REPORTS
EMERGING TECHNOLOGIES TO REMOVE NONPOINT P SOURCES FROM SURFACE WATER AND GROUNDWATER

Factors Controlling Phosphate Interaction with Iron Oxides


Liping Weng,* Willem H. Van Riemsdijk, and Tjisse Hiemstra

E
nhanced primary production in surface water (i.e.,
Factors such as pH, solution ion composition, and the presence
eutrophication) has become a worldwide problem that
of natural organic matter (NOM) play a crucial role in the
effectiveness of phosphorous adsorption by iron oxides. The results in algal blooms, the development of harmful
interplay between these factors shows a complicated pattern cyanobacteria, fish kills, smell nuisance, and other detrimen-
and can sometimes lead to controversial results. With the help tal consequences (Carpenter et al., 1998). In the past decades,
of mechanistic modeling and adsorption experiments, the net considerable reduction of P loss has been achieved thanks to
macroscopic effect of single and combined factors can be better
measures to control industrial release and sewage water treat-
understood and predicted. In the present work, the relative
importance of the above-mentioned factors in the adsorption of ment. As a result, the relative importance of P release from
phosphate was analyzed using modeling and comparison between agricultural fields has increased in recent years (Barberis et al.,
the model prediction and experimental data. The results show 1996; Chardon and Koopmans, 2005; Koopmans et al., 2007;
that, under normal soil conditions, pH, concentration of Ca, and Sharpley et al., 1994).
the presence of NOM are the most important factors that control
Various strategies and measures have been proposed, tested,
adsorption of phosphate to iron oxides. The presence of Ca not
only enhances the amount of phosphate adsorbed but also changes and implemented to reduce release of P from agricultural fields,
the pH dependency of the adsorption. An increase of dissolved including input reduction and leaching prevention (Kronvang
organic carbon from 0.5 to 50 mg L−1 can lead to a >50% decrease et al., 2005; Sharpley et al., 1994; Withers et al., 2000). In
in the amount of phosphate adsorbed. Silicic acid may decrease agriculture fields, to guarantee sufficient nutrient supply for
phosphate adsorption, but this effect is only important at a very
crop production, a solution concentration of above 3 mg L−1
low phosphate concentration, in particular at high pH.
is required, and P concentrations as high as 30 mg L−1 can be
found in soil runoff (Withers et al., 2000). The P concentra-
tion in leachate from agricultural fields is much higher than the
target value for P in surface water (<0.1 mg L−1). One attrac-
tive approach to reduce leaching is to use P-fixing materials
that are directly mixed with soil or used to build a reactive
barrier (Codling et al., 2000; McDowell et al., 2008; Stout
et al., 2000). Iron-rich materials are good candidates to build
P-removing filters (McDowell et al., 2008; Zeng et al., 2004;
Zhang et al., 2009). Such materials can be metal iron (e.g.,
zerovalent iron) (McDowell et al., 2008; Sakadevan and Bavor,
1998) and iron oxides in various forms (Boujelben et al., 2008;
Pan et al., 2010; Zeng et al., 2004).
Phosphorus in soil drainage water consists of orthophos-
phate, organic P, and P associated with mineral colloidal particles
(adsorbed or precipitated). A permeable reactive filter/barrier
containing iron oxides can, in principle, remove all three forms
of P. The mineral colloidal P can be intercepted by filtration
effects, and orthophosphate and organic P can be immobilized
by adsorption reactions. The efficiency and effective lifespan
of a P barrier depends on various factors, including the basic
Copyright © 2012 by the American Society of Agronomy, Crop Science Society characteristic of the adsorptive material itself (e.g., the porosity,
of America, and Soil Science Society of America. All rights reserved. No part of reactive surface area, P-binding affinity, and binding capacity)
this periodical may be reproduced or transmitted in any form or by any means, and environmental factors (e.g., pH, ion composition, the pres-
electronic or mechanical, including photocopying, recording, or any information
storage and retrieval system, without permission in writing from the publisher. ence of other compounds, and water flow). The reactive groups
present on the surface of iron oxides can bind various cations
J. Environ. Qual. 41:628–635 (2012)
doi:10.2134/jeq2011.0250
Posted online 13 Jan. 2012. Dep. of Soil Quality, Wageningen Univ., P.O. Box 47, 6700 AA, Wageningen, The
Received 14 July 2011. Netherlands. Assigned to Associate Editor Ray Bryant.
*Corresponding author (liping.weng@wur.nl).
© ASA, CSSA, SSSA Abbreviations: DOC, dissolved organic carbon; FA, fulvic acid; LCD, ligand and
5585 Guilford Rd., Madison, WI 53711 USA charge distribution; NOM, natural organic matter.

628
and anions. The chemical reactions at the surface become more The goethite material was prepared based on the procedure
difficult to predict the more the ions and compounds interact. described by Hiemstra et al. (1989). The specific surface area
The complete set of chemical reactions that may take place in the of the goethite material was 94 m2 g−1, as determined using
field situation leads to a very complex system. the BET-N2 adsorption method (Weng et al., 2005). Fulvic
In adsorption reactions, when equilibrium is maintained, acid (FA) was used to simulate NOM, and the FA material
the relation between solution concentration of phosphate (C, was purified from a peat soil (Strichen Soil Assoc., Scotland)
mol L−1) and the corresponding loading on the surface (Γ, mol following the International Humic Substance Society proce-
kg−1) can be expressed using the Langmuir model: dures. The purified FA material contains 43% C, and its aver-
age molar mass is 0.68 kDa, which was measured using size
K 'C
Γ = Γ 'max [1] exclusion chromatography.
1 + K 'C This study included three series of treatments: (i) no FA,
where Γ′max is the apparent maximum binding capacity (mol (ii) 150 mg L−1 FA (64.5 mg C L−1), and (iii) 300 mg L−1 FA
kg−1) and K′ the conditional affinity constant. For a reactive (129 mg C L−1). In all series, the suspension contained final
barrier, the increase of the P loading equals the net amount of total concentrations of 2.7 g L−1 goethite, 0.01 CaCl2, and 0.48
P added per unit adsorptive material; this is equal to V(C0 – mmol L−1 NaH2PO4. The solutions were added immediately
C), where V (L kg−1) is the volume of solute drained expressed after each other in the order CaCl2, goethite, NaH2PO4, and
per kg of adsorptive material, and C0 and C are the input and FA. For all series, acid or base solution (0.01 mol L−1 NaOH
output P concentrations, respectively. If the latter is small, the and 0.05 mol L−1 HCl) was added to the suspension to adjust
average loading of the reactive material can be approached by the pH to selected values in the range of pH 3 to 7. During
VC0. Under this condition, Eq. [1] can be written as the preparation, the gas-tight vessels used in the experiments
were flushed with N2 gas to minimize the influence of CO2.
K 'C
Γ = VC 0 = Γ 'max [2] The prepared suspensions were shaken at 20°C for 4 d and
1 + K 'C centrifuged at 18,000 g for 30 min. The final pH in the super-
Rearrangement of Eq. [2] results in natant was measured using a pH meter. The P concentration
in the supernatant was measured with ICP–MS. The Ca con-
VC 0 centration was measured with ICP–AES. The concentration of
C= [3]
(Γ 'max −VC0 ) K ' dissolved organic carbon (DOC) was measured with a TOC
analyzer. The amounts of P, Ca, and FA adsorbed were derived
To maintain the C under a certain target value, the maxi- from the difference between the treatments and the blank (no
mum amount of drainage that can be treated depends strongly goethite). The small contribution of organic P (P in FA) to the
on both Γ′max and K′. In a mechanistic model, the maximum total soluble P measured is corrected using P content in FA
binding capacity depends on the amount and type of iron measured in a FA solution (0.16 μg mg−1) and DOC. The final
oxides in the material, the particle size (surface area), and acces- DOC concentration ranged from 15 to 75 mg L−1.
sibility of the surface, i.e., Γmax is characterized by the mate-
rial. When this type of Langmuir model is derived empirically, Modeling
the apparent Γ′max also depends on factors that may promote
adsorption (like Ca2+) or decrease the adsorption due to com- CD-MUSIC Model
petition with natural organic matter (NOM). The conditional Iron (hydr)oxides can occur in nature in various crystal forms,
affinity constant K′ depends on the type of reactive groups such as goethite (α-FeOOH), hydrous ferric oxide, or ferrihy-
present on the surface (a character of the material) and is also drite, hematite (α-Fe2O3), and lepidocrocite (γ-FeOOH). All
strongly affected by the solution chemistry (e.g., pH and pres- these solid phases have a strong affinity for phosphate. From
ence of other adsorbing compounds). a more fundamental point of view, much more is known with
The objective of this study was to apply the most recent respect to the adsorption of phosphate to goethite, whereas
developments in the understanding of ion and NOM adsorp- much less information is available for the other phases men-
tion to iron oxides to gain insight into the most relevant factors tioned. Goethite with varying crystal morphology shows similar
that influence the performance of P-filters under field condi- binding properties for phosphate (Torrent et al., 1990). It has
tions. The relative importance of pH, major cations (Ca2+ and been shown (Hiemstra et al., 2010a) that phosphate desorption
Mg2+), anions (HCO3− and SO42−), neutral species (H4SiO40), data for various soils can be described well by assuming that the
and NOM in phosphate adsorption were analyzed using mech- reactive metal oxide phase in soils can be interpreted based on
anistic modeling and the model prediction were compared the well calibrated adsorption model for goethite, although the
with data of an adsorption experiment. We show that modern process may involve several metal oxide phases and clay miner-
surface complexation models such as the CD-MUSIC model als (Devau et al., 2009, 2011).
(Hiemstra and Van Riemsdijk, 2006) can be applied to explain The surface of the iron oxides consist of a few different types
and predict the net effects of various factors on the quality and surface oxygen groups that differ in their degree of coordina-
capacitance of a barrier. tion with the iron in the solid. These oxygens have a variable
number of protons associated with them. Due to breaking
Materials and Methods of bonds, surface groups with a coordination number that is
Batch experiments were performed to study the effects less than the maximum can occur. The formal charge of the
of pH, Ca, and NOM on phosphate adsorption to goethite. various surface groups can be derived using a simple Pauling

www.agronomy.org • www.crops.org • www.soils.org 629


bond valence analysis or bond valence theory. In the Pauling ≡FeOH−0.5 + H+ = ≡FeOH2+0.5 log KH,1 [4]
approach, one assigns a formal charge of 0.5 (3/6) per Fe bond
to the oxygen. This approach is used in the MUSIC (Multi Site ≡Fe3O−0.5 + H+ = ≡Fe3OH+0.5 log KH,2 [5]
Complexation) model (Hiemstra et al., 1989), which, in com- Depending on environmental conditions such as pH, metal (hydr)
bination with the CD (Charge Distribution) model (Hiemstra oxides can be positively charged, negatively charged, or uncharged.
and Van Riemsdijk, 1996), forms the basis of ion adsorption Based on the structure analysis, the overall effective site
models for oxides. The CD-MUSIC model is based as much as density of goethite is set at 3.45 site nm−2 for the singly coor-
possible on available molecular level information with respect dinated oxygen and 2.7 site nm−2 for the triply coordinated
to the surface speciation. oxygen (Table 1). The (refined) MUSIC model also allows for
The application of the MUSIC model to various minerals an a priori estimate of the logK values of the various proton-
shows that sites that have a very high or very low value of the ation reactions for various minerals. In practice, the model for
formal charge are not present at the surface to any significant goethite can be simplified by assuming that the proton affin-
extent. In other words, such surface species are thermodynami- ity constants for the two types of reactive sites are equal. The
cally not very stable. For the singly coordinated surface oxygens intrinsic proton affinities (logKH,1 and logKH,2) follow in that
on goethite, the amount of the ≡FeO−1.5 site is insignificant, case directly from the point of zero charge of goethite (Table 1).
and the major surface species are thus the ≡FeOH−0.5 and the In the CD model, the charge of an adsorbing ion is dis-
≡FeOH2+0.5 species, whereas for the triply coordinated oxygen tributed over two electrostatic planes. The advantage of this
the relevant species are ≡Fe3OH+0.5 and ≡Fe3O−0.5. The doubly approach is that the electrostatic energy component to the
coordinated oxygens have a high affinity for protons, which binding is now a combination of two terms, one for each
leads to the formation of a stable uncharged species (≡Fe2OH0). plane. The CD approach is a practical way to account for the
In the MUSIC model, the basic charging of the mineral goe- fact that there is a gradient of the electrostatic potential over
thite is therefore assumed to be caused by the protonation and the adsorbed ion. It is also an advantage that the value of the
deprotonation of the singly (≡FeOH−0.5) and triply (≡Fe3O−0.5) CD parameter is linked to the structure of the surface com-
coordinated surface oxygens: plex. The simplest electrostatic model that takes the structure
of the inner-sphere surfaces into account is a CD approach

Table 1. CD-MUSIC model parameters for goethite.†


PZC‡ Site density Capacitance
≡FeOH−0.5 ≡Fe3O−0.5 C1 C2
9.0
Surface species 3.45 nm−2 2.7 nm−2 0.93 F m−2 0.75 F m−2
Sites Charge distribution§
Ions logK
≡ FeOH−0.5 ≡ Fe3O−0.5 Δz0 Δz1 Δz2
≡FeOH2+0.5 1 0 1 0 0 1H+ 9.00
≡Fe3OH+0.5 0 1 1 0 0 1H+ 9.00
≡FeOH…Na+0.5 1 0 0 1 0 1Na+ −0.60
≡Fe3O…Na+0.5 0 1 0 1 0 1Na+ −0.60
≡FeOH…K+0.5 1 0 0 1 0 1K+ −1.61
≡Fe3O…K+0.5 0 1 0 1 0 1K+ −1.61
≡FeOHCa+1.5 1 0 0.31 1.69 0 1Ca2+ 3.23
≡FeOHCaOH+0.5 1 0 0.31 0.69 0 −1H+, 1Ca2+ −6.42
≡FeOH…Ca+1.5 1 0 0 2 0 1Ca2+ 1.8
≡Fe3O…Ca+1.5 1 0 0 2 0 1Ca2+ 1.8
≡(FeOH)2Mg+1.0 2 0 0.72 1.28 0 1Mg2+ 4.52
≡(FeOH)2MgOH0 2 0 0.72 0.28 0 −1H+, 1Mg2+ −6.78
≡FeOH2…Cl−0.5 1 0 1 −1 0 1H+, 1Cl− 8.55
≡Fe3OH…Cl−0.5 0 1 1 −1 0 1H+, 1Cl− 8.55
≡(FeO)2CO−1.0 2 0 0.62 −0.62 0 2H+, 1CO32− 22.01
≡(FeO)2CO…Na0 2 0 0.62 0.38 0 2H+, 1Na+,1CO32− 22.03
≡FeOH2…CO3−1.5 1 0 1 −2 0 1H+, 1CO32− 10.22
≡Fe3OH…CO3−1.5 0 1 1 −2 0 1H+, 1CO32− 10.22
≡(FeO)2Si(OH)2.−1.0 2 0 0.29 −0.29 0 1H4SiO40 5.85
≡(FeO)2SiOHOSi3O2(OH)7.−1.0 2 0 0.29 −0.29 0 1H4SiO40 13.98
≡(FeO)2SiOHOSi3O3(OH)6.−2.0 2 0 0.29 −1.29 0 −1H+, 1H4SiO40 7.47
≡(FeO)2PO2−2.0 2 0 0.46 −1.46 0 2H+, 1PO43− 29.72
≡FeOPO2OH−1.5 1 0 0.28 −1.28 0 2H+, 1PO43− 27.63
† Adapted from Hiemstra et al., 2007, 2010a.
‡ Point of zero charge.
§ Δz0, charge attributed to 0-plane; Δz1, charge attributed to 1-plane; Δz2, charge attributed to 2-plane.

630 Journal of Environmental Quality


with two electrostatic planes, a CD-Basic Stern model. The 2010a, 2010b; Weng et al., 2008). For modeling the adsorption
electrostatic approach has recently been further refined by con- of NOM and the interaction between phosphate and NOM at
sidering the fact that ions adsorbed as outer-sphere complexes the surface of iron oxides, a simplified approach has been pro-
can have different sizes, which means the electrostatic energy posed and used (Hiemstra et al., 2010b), in which the formation
component for different size ions is also different (Hiemstra reaction of a virtual NOM surface complex is defined as:
and Van Riemsdijk, 2004; Hiemstra et al., 2004; Rahnemaie
et al., 2006). In this case, an Extended Stern model is required ≡FeOH−0.5 + HNOM−1.0 ⇔ ≡FeNOM−1.5 + H2O [8]
(Hiemstra and Van Riemsdijk, 2006).
Phosphate ions can form surface complexes by reacting chemi- The virtual NOM species can be used to model the appar-
cally with one or two singly coordinated surface groups (≡FeOH− ent effective NOM loading active in the phosphate compe-
0.5
). The group that forms the bond is an oxygen ion in both cases tition. In the CD-MUSIC modeling, the overall charge z =
(Fig. 1). Originally, the adsorption behavior of phosphate was −1.5 of ≡FeNOM−1.5 is distributed in the interface, and this
described in the CD-MUSIC model using three surface species: a charge contributes strongly to the competition with phosphate
monodentate, a bidentate, and a protonated bidentate. Recently, (Hiemstra et al., 2010b). As shown in Hiemstra et al. (2010a),
Rahnemaie et al. (2007) reinterpreted the results using an exten- for most soils studied, the calculated ≡FeNOM−1.5 loading is
sive data set and concluded that P adsorption can be described between about 0.5 and 1.5 μmol m−2.
using two inner-sphere complexes: the bidentate and protonated A more advanced approach to account for the presence of
monodentate. The formation reaction for the dominant bidentate NOM coating is the ligand and charge distribution (LCD)
surface phosphate species can be written as: model (Weng et al., 2007, 2008, 2006b). It is assumed that
the interaction between NOM and phosphate at the surface
2≡FeOH−0.5 + PO43− + 2H+ = (≡FeO)2PO22− + 2H2O [6] of oxides can be simulated in the LCD model using a layer of
adsorbed FA, which is present and equally distributed between
and formation reaction for the less important monodentate the first and second Stern layer (Weng et al., 2008). Previous
species can be written as: study has shown that the FA particles can fit into the com-
pact part of the electrostatic double layer, whereas a large part
≡FeOH−0.5 + PO43− + 2H+ = ≡FeOPO2OH−1.5 + 2H2O [7] of the adsorbed humic acid is located beyond the electrostatic
double layer, in the diffuse double layer (Weng et al., 2008).
The mechanistic nature of the CD-MUSIC model allows pre- It was assumed that the carboxylic groups (RCOO−) of NOM
diction of effects of pH and the synergistic and competitive in the first compact layer can form innersphere complexes
effects of cations and other anions. Model parameters for ions (≡Fe1OOCR−0.5) with the singly coordinated surface sites on
that are most relevant in a soil environment are available in the goethite (Weng et al., 2006b). The corresponding charge is
literature. The parameter set used in the following calculation distributed between the 0-plane and the 1-plane. The other
was adopted from Hiemstra et al. (2010b), except those for carboxylic and phenolic groups on the NOM can bind H+
silicic acid adsorption, which came from Hiemstra et al. (2007) and other cations like Ca2+. The reactions between the NOM
(Table 1). ligands and surface sites, protons, and other cations were cal-
culated in the LCD with the NICA model. It was assumed that
Modeling Natural Organic Matter Effects the NICA model parameters remain the same as for FA in the
In natural systems, NOM is usually present. These organic solution (Milne et al., 2001, 2003).
molecules also interact with iron oxides (Weng et al., 2006a)
and other soil minerals by forming an organic “coating” on these Effects of Solution Chemistry
mineral surfaces. It has been demonstrated that adsorption of
NOM strongly influences phosphate adsorption to goethite on Phosphorus Adsorption
(Bhatti et al., 1998; Borggaard et al., 1990; Hiemstra et al., The interaction of major cations and anions with goethite
(FeOOH) has been quantified systematically during the last
decade, resulting in a consistent set of parameters (Table 1)
that can now be used to quantify the major factors acting on
the phosphate binding in chemical barriers. These interactions
are explored for P-filter systems with and without the presence
of NOM.

In the Absence of Natural Organic Matter


The inorganic chemistry of natural aqueous solution is
dominated by a selective number of light elements. In non-
acid and oxidized systems, the main cations are Na+, K+, Ca2+,
and Mg2+, and the main inorganic anions are HCO3−, Cl−, and
SO42−. In addition, neutral silicic acid H4SiO40 is present at sig-
nificant concentrations. These major elements affect the bind-
ing of minor elements (e.g., heavy metals) and oxyanions (e.g.,
Fig. 1. Representation of the structure of the protonated monodentate phosphate). The interaction of adsorbed ions is mainly of an
(upper) and bidentate (lower) phosphate surface species on goethite
www.agronomy.org • www.crops.org • www.soils.org 631
electrostatic nature (repulsion or attraction) and includes the is less than 0.1 μmol m−2. The carbonate loading ΓCO3 decreases
pH-dependent proton adsorption by the surface groups. with the increase of the phosphate concentration mainly due
The interactions of cations and anions in multi-component to electrostatic effects that change outside the surface in the
systems with iron oxides can be described with the CD-MUSIC 1-plane due to the adsorption of both ions.
model, as explained above. Figure 2a gives the calculated sur- Adsorption of Mg2+ increases with pH, and the highest Mg
face composition (Γ, in μmol m−2) of goethite in equilibrium loading is for pH 8 (ΓMg < 0.1 μmol m−2). High pH is preferred
with a representative natural solution with [Na+] = 0.1 mmol by cations because at high pH the particle is less positively or
L−1, [K+] = 0.1 mmol L−1, [Ca2+] = 1 mmol L−1, [Mg2+] = 0.2 more negatively charged. In the presence of phosphate, the
mmol L−1, [Cl−] = 2.5 mmol L−1, [H4SiO40] = 0.1 mmol L−1, oxide particles as a whole are often negatively charged due to
and [HCO3−] in equilibrium with atmospheric CO2 (4 × 10−4 the adsorption of the PO43− ions. This negative charge is attrac-
bar) for pH 4, 6, and 8, respectively, as a function of the total tive for Mg2+. However, despite the cooperative effect of phos-
phosphate concentration in solution. The calculated loading of phate, the Mg2+ loading starts to decrease at higher phosphate
specifically bound Na+, K+, and Cl− was not given in the figure concentrations in solution. This is due to the formation of dis-
due to their insignificance as a result of a very low affinity. solved MgHPO40 (aq), reducing the Mg2+ activity.
The relative importance of solution species in their effect The behavior of Mg2+ contrasts with that of Ca2+ ions.
on phosphate adsorption can be derived from their adsorbed The loading of Ca2+ (ΓCa) increases with the phosphate con-
amount relative to that of phosphate (Fig. 2a). In Fig. 2a, both centration. The cooperative adsorption is due to electrostatic
scales are log scale. The calculation (Fig. 2a) shows that HCO3−/ interaction (attraction) between adsorbed phosphate (anion)
CO32− and Mg2+ are relatively unimportant at the chosen ref- and calcium (cation). The adsorption of Ca2+ increases as pH
erence composition of the equilibrium solution. Carbonate increases. Figure 2a shows that adsorption of Ca2+ becomes
adsorption increases as pH increases and decreases as phos- significant at a relatively high pH (> pH 5) and above a phos-
phate concentration increases. The highest carbonate loading phate concentration of 1 μmol L−1. Under these conditions, an
(ΓCO3) is found in the system with the highest concentration of
HCO3− (pH 8) and lowest phosphate concentration, and ΓCO3

Fig. 2. The surface composition of goethite in equilibrium with K+, Na+, Ca2+, Mg2+, H4SiO40, Cl−, and HCO3− at concentrations representative for
natural soil solutions (see text) as a function of the phosphate concentration in solution for pH 4, 6, and 8 at PCO2 = 4 x 10−4 bar. Compositions
have been calculated with the CD-MUSIC model using the parameters of Table 1 (Hiemstra et al., 2007, 2010a). In the systems of Fig. 2a and 2b, no
natural organic matter is present, whereas in Fig. 2c and 2d, the natural organic matter loading is 1.66 μmol m−2. Figure 2a and 2b refer to the same
calculation, but the loading is given at the logarithmic and linear scales, respectively. Figure 2c and 2d refer to the same solution composition,
except a 10-fold higher H4SiO4 concentration (1.0 mmol L−1) was used in Fig. 2d than in 2c.

632 Journal of Environmental Quality


increase in Ca concentration promotes phosphate adsorption Similar to the case when no NOM was present (Fig. 2b),
to iron oxides. silicic acid is dominantly adsorbed at low phosphate concentra-
Adsorption of silicic acid increases as pH increases. At a tion in solution and at high pH value (pH 8) (Fig. 2c), as may
given pH, the competition between phosphate and silicic acid be found in surface waters. The high loading of Si is more pro-
for the adsorption leads to a strong decrease of silicic acid nounced at a 10 times higher Si concentration (1.0 mmol L−1)
adsorbed when phosphate concentration increases (Fig. 2a). (Fig. 2d). This situation can be typical for aquifers. At an ele-
Adsorbed silicic acid becomes significant at a low phosphate vated silicic acid concentration (1.0 mmol L−1) (Fig. 2d), com-
concentration (<0.1 μmol L−1). petition between silicic acid and phosphate plays an important
In summary, the calculation suggests that, in many natural role at relatively high pH and at a phosphate concentration of
systems, the surface of iron oxides is dominated by adsorbed <10 μmol L−1. In fertile top soils, the phosphate concentration
PO43−, H4SiO4, and Ca2+. The influence of CO32− and Mg2+ on is usually several micromolar or higher. At these concentrations,
phosphate adsorption is predicted to be insignificant. Silicic phosphate is the most important oxyanion at the oxide surfaces
acid (H4SiO40) may decrease phosphate adsorption, but this in acid and close to neutral soils (pH 4–6).
effect is only important at a very low phosphate concentration, The role of SO42− in the surface speciation is not discussed
in particular at high pH, as can be found in surface water. in detail here. Sulfate adsorption to oxides can be large at low
In Fig. 2b, the surface composition is given at the linear pH (pH < 5). The adsorption is significant in simple model
scale for the same solution conditions used for Fig. 2a. Because systems with goethite, sulfate, and a background electrolyte
the binding of Mg2+ and carbonate are very low (≤0.1 μmol (Geelhoed et al., 1997). However, in the presence of phosphate,
m−2), they were not plotted in Fig. 2b. At high phosphate con- the anion competition determines the sulfate behavior. Because
centrations, only Ca2+ and phosphate are relevant. The pH and sulfate is relatively weakly bound in comparison to phosphate,
Ca concentration can be considered as most relevant in affect- phosphate adsorption removes adsorbed sulfate from the oxide
ing the phosphate adsorption to iron oxides. Only at very low surface, while the phosphate adsorption remains largely unaf-
phosphate levels does the binding of Si become relevant. fected (Geelhoed et al., 1997). The presence of adsorbed NOM
further reduces the sulfate loading. Only at very acidic condi-
In the Presence of Natural Organic Matter tions (pH < ?3–4), sulfate may be a relevant ion to be incor-
Natural organic matter has a significant influence on porated in surface complexation modeling of soil systems, such
phosphate adsorption to iron oxides and may strongly influ- as acid sulfate soils and acid mine tailings.
ence the performance of the P filter. Using the CD-MUSIC Another point that requires consideration in practice is the
model (Table 1) and the virtual NOM approach (Eq. [8]), we contribution of organic P to the P release in the field. Organic
calculated the change of surface composition of goethite as a P adsorbs to iron oxides as well. Therefore, the adsorption of
function of the solution concentration of PO4 in the presence NOM can diminish inorganic P adsorption via competition,
of NOM. This calculation was performed for a representative but it can also enhance P adsorption due to adsorption of
soil solution composition as used in the calculation in Fig. 2a organic P.
and 2b: [Na+] = 0.1 mmol L−1, [K+] = 0.1 mmol L−1, [Ca2+] =
1 mmol L−1, [Mg2+] = 0.2 mmol L−1, [Cl−] = 2.5 mmol L−1, Comparing Modeling and Experimental Adsorption Data
[H4SiO40] = 0.1 mmol L−1, and [HCO3−] in equilibrium with The above model simulation indicates that pH, Ca, and
atmospheric CO2 (4 × 10−4 bar) for pH 4, 6, and 8. The results NOM are the most important factors influencing phosphate
are shown in Fig. 2c. The difference between Fig. 2c and 2b is adsorption under the most common conditions in field. To
the presence of adsorbed NOM species (1 virtual NOM nm−2 test the reliability of the model prediction for the effects of
or 1.66 μmol NOM m−2). A second calculation is given in Fig. these three factors, a series of batch adsorption experiments was
2d for the same conditions but at a 10 times higher H4SiO4 performed (see Materials and Methods). The results (Fig. 3,
concentration (1.0 mmol L−1). symbols) can be compared with the model prediction (Fig. 3,
To reduce the complexity of the figure, the adsorption of lines). The effect of the presence of adsorbed FA on phosphate
carbonate and magnesium is not given. Comparing Fig. 2c adsorption was calculated in this modeling using the LCD
(with NOM) with Fig. 2b (no NOM), one can see that the approach. We can see from this figure that the CD-MUSIC
presence of NOM decreased phosphate and silicic acid adsorp- and LCD model predictions are in good agreement with the
tion and promoted Ca adsorption. The presence of NOM data.
decreased phosphate adsorption by 37 to 97% (compare Fig. For comparison, the dissolved phosphate concentration in
2c and Fig. 2b). The effect of NOM on Ca and phosphate the absence of Ca and NOM calculated using the CD-MUSIC
adsorption is stronger at a lower P concentration. In the pres- model (for an electrolyte background of 0.01 mol L−1 of NaCl)
ence of NOM, the surface loading of Ca is usually higher than is also given in Fig. 3. Compared with the situation when nei-
that of phosphate for soils of pH 6 to 8 under the Ca concen- ther Ca nor NOM was present, adsorption of Ca promoted
tration used (Fig. 2c, 2d). The Ca2+ adsorption is stimulated by phosphate adsorption (and therefore led to a lower soluble
adsorbed NOM, whereas the phosphate loading is suppressed phosphate), and this effect becomes significant when the pH
by NOM, showing, respectively, the synergistic and the com- is not too low. In addition to the promotion of phosphate
petitive action of NOM on the adsorption of inorganic ions. adsorption, the presence of Ca changed the pH dependency
At pH > 5 and in the presence of NOM, phosphate adsorption of phosphate adsorption as well. Due to a stronger adsorption
is more sensitive to the change of Ca concentration in the solu- of Ca at increased pH, phosphate adsorption increases with
tion compared with that when NOM is absent. increased pH between pH about 4 to 7, whereas the trend
www.agronomy.org • www.crops.org • www.soils.org 633
is the opposite in a background of Na or K. The presence of Langmuir adsorption isotherm (Eq. [1]). A difference is the
FA decreased phosphate adsorption and led to strong increase presence of an exponent n at the concentration term. Modeling
of phosphate in solution. The pH dependency of phosphate shows that this term reflects the effect of electrostatics on the
adsorption in the presence of Ca and NOM shows a negative shape of phosphate adsorption. The equation has been param-
relation in the range pH 3 to 7. eterized in this study using a large set of soil data (Hiemstra
et al., 2010a), for which the effective reactive surface area has
Implications been determined, leading to K′ = 140 and n = 0.3 in the case
The above analysis shows that major factors determining of Γ′max = 3 μmol m−2. A good description can be found (R2 =
the phosphate adsorption to iron oxides are the phosphate con- 0.69) for the data set of soils.
centration, the binding of NOM, the presence of Ca, and the The practical equation (Eq. [9]) can be used for instance
pH. This can be compared with the results found for a series to evaluate the possible influence of NOM on the adsorption
of soils. For these soils, the phosphate loading was measured capacitance of a chemical barrier. In Fig. 4, the calculated effect
and the equilibrium concentration in 0.01 mol L−1 CaCl2 was (using Eq. 9) of DOC on the PO4 adsorption in 0.01 mol L−1
established, and pH and DOC concentration were measured CaCl2 is given for 0.5, 5, and 50 mg L−1 DOC. The variation in
(Hiemstra et al., 2010b). Without scaling, no relation between DOC has a significant effect on amount of phosphate adsorbed.
the PO4 loading on iron oxides (Γ) and the concentration of In the micromolar range of phosphate concentration in solution,
dissolved phosphate (CPO4) is found. The reason is that the natu- an increase of DOC from 0.5 to 50 mg L−1 reduces amount of
ral materials strongly differ in reactive surface area. Therefore, phosphate adsorbed by about 50% (Fig. 4). This result shows
scaling per unit reactive surface area is required. This surface that the binding capacity of a reactive barrier depends on the
area can be approached experimentally using a new soil dilution combination of CPO4 and CDOC in the solute. In addition, a reli-
method in 0.5 mol L−1 NaHCO3 (Hiemstra et al., 2010a). After able estimate of the reactive surface area is essential as a param-
scaling the PO4 loading (Γ) to mol m−2, the loading is related to eter (Hiemstra et al., 2010a) because the isotherms are expressed
the logarithm of solution parameters. The strongest correlation per unit surface area, not per unit mass.
is with DOC (R2 = 0.66). The correlation is negative, point-
ing to the competitive character of DOC for phosphate. An Conclusions
increase in NOM loading leads to a decrease in PO4 loading, In summary, the above discussion shows that solution pH,
and vice versa (Hiemstra et al., 2010b). Considering the DOC– concentration of Ca, and the loading of NOM are the major
PO4 interaction as an exchange process, the competition can be factors that may significantly influence the adsorption of phos-
described with a nonelectrostatic exchange equation: phate to iron oxide materials in natural soil solutions at pH
4 to 8. Silicic acid may decrease phosphate adsorption, but
K '(C PO4 / C DOC )n
Γ = Γ ' max [9] this effect is only important at a very low phosphate concen-
1 + K '(C PO4 / C DOC )n trations, in particular at high pH, as can be found in surface
water. Natural organic matter and PO43− are strong competi-
where CPO4 is in mol L−1 and CDOC is in mg L−1. At a constant tors, and adsorbed Ca2+ acts cooperatively on the binding of
DOC concentration, the isotherm has similarities with the these anions. The binding of NOM cannot be ignored when
attempting to understand the behavior of phosphate and other

Fig. 3. Effect of the presence of fulvic acid (FA) (0, 150, and 300 mg L−1)
on the concentration of phosphate in a 0.01 mol L−1 CaCl2 solution
in a system with goethite (2.7 g L−1) having initially 0.48 mmol L−1
PO4 (experimental data points). The effects of FA on phosphate Fig. 4. Effect of dissolved organic carbon (DOC) (0.5–50 mg L−1) on
adsorption were calculated using the ligand and charge distribution the phosphate loading of the natural oxides in 0.01 mol L−1 CaCl2.
approach (lines) applying the CD-MUSIC model parameters of Table The isotherms have been calculated using Eq. 9, which has been cali-
1. For comparison, the calculated soluble phosphate concentration brated on a large set of top soils with variable phosphate equilibrium
is given (thin line) for a system with 0.01 mol L−1 NaCl as background concentrations (~1–30 μmol L−1), DOC (~5–30 mg L−1), and pH (~4–7)
(no Ca and FA). in 0.01 mol L−1 CaCl2.

634 Journal of Environmental Quality


oxyanions (e.g., B, As, Se, Cr, Mo, etc.) in natural systems. droxides. J. Colloid Interface Sci. 301:1–18. doi:10.1016/j.jcis.2006.05.008
The effects of NOM are stronger at a low phosphate concen- Hiemstra, T., W.H. Van Riemsdijk, and G.H. Bolt. 1989. Multisite pro-
ton adsorption modeling at the solid-solution interface of (hydr)
tration. However, even at a micromolar concentration level of oxides: A new approach. 1. Model description and evaluation of in-
phosphate, a 100 times increase in DOC can lead to a 50% trinsic reaction constants. J. Colloid Interface Sci. 133:91–104.
decrease in phosphate adsorption. Consequently, the soluble doi:10.1016/0021-9797(89)90284-1
Koopmans, G.F., W.J. Chardon, and R.W. McDowell. 2007. Phosphorus move-
phosphate concentration may increase by several orders of ment and speciation in a sandy soil profile after long-term animal manure
magnitude because soluble phosphate is only a small fraction applications. J. Environ. Qual. 36:305–315. doi:10.2134/jeq2006.0131
of total phosphate. The mechanistic CD-MUSIC and LCD Kronvang, B., M. Bechmann, H. Lundekvam, H. Behrendt, G.H. Rubaek,
O.F. Schoumans, N. Syversen, H.E. Andersen, and C.C. Hoffmann.
models are useful tools in the understanding and prediction of 2005. Phosphorus losses from agricultural areas in river basins: Effects
the net effects of combined factors. and uncertainties of targeted mitigation measures. J. Environ. Qual.
34:2129–2144. doi:10.2134/jeq2004.0439
References McDowell, R.W., A.N. Sharpley, and W. Bourke. 2008. Treatment of drainage wa-
ter with industrial by-products to prevent phosphorus loss from tile-drained
Barberis, E., F.A. Marsan, R. Scalenghe, A. Lammers, U. Schwertmann, A.C. land. J. Environ. Qual. 37:1575–1582. doi:10.2134/jeq2007.0454
Edwards, R. Maguire, M.J. Wilson, A. Delgado, and J. Torrent. 1996. Milne, C.J., D.G. Kinniburgh, and E. Tipping. 2001. Generic NICA-Donnan
European soils overfertilized with phosphorus: 1. Basic properties. Fert. model parameters for proton binding by humic substances. Environ. Sci.
Res. 45:199–207. doi:10.1007/BF00748590 Technol. 35:2049–2059. doi:10.1021/es000123j
Bhatti, J.S., N.B. Comerford, and C.T. Johnston. 1998. Influence of oxalate Milne, C.J., D.G. Kinniburgh, W.H. Van Riemsdijk, and E. Tipping. 2003. Ge-
and soil organic matter on sorption and desorption of phosphate onto neric NICA-Donnan model parameters for metal-ion binding by humic
a spodic horizon. Soil Sci. Soc. Am. J. 62:1089–1095. doi:10.2136/ substances. Environ. Sci. Technol. 37:958–971. doi:10.1021/es0258879
sssaj1998.03615995006200040033x Pan, G., L. Li, D.Y. Zhao, and H. Chen. 2010. Immobilization of non-point
Borggaard, O.K., S.S. Jorgensen, J.P. Moberg, and B. Rabenlange. 1990. phosphorus using stabilized magnetite nanoparticles with enhanced
Influence of organic-matter on phosphate adsorption by alu- transportability and reactivity in soils. Environ. Pollut. 158:35–40.
minum and iron-oxides in sandy soils. J. Soil Sci. 41:443–449. doi:10.1016/j.envpol.2009.08.003
doi:10.1111/j.1365-2389.1990.tb00078.x Rahnemaie, R., T. Hiemstra, and W.H. Van Riemsdijk. 2006. A new structural
Boujelben, N., J. Bouzid, Z. Elouear, A. Feki, F. Jamoussi, and A. Montiel. approach for outersphere complexation tracing the location of electrolyte
2008. Phosphorus removal from aqueous solution using iron coated ions. J. Colloid Interface Sci. 293:312–321. doi:10.1016/j.jcis.2005.06.089
natural and engineered sorbents. J. Hazard. Mater. 151:103–110. Rahnemaie, R., T. Hiemstra, and W.H. Van Riemsdijk. 2007. Geometry,
doi:10.1016/j.jhazmat.2007.05.057 charge distribution, and surface speciation of phosphate on goethite.
Carpenter, S.R., N.F. Caraco, D.L. Correll, R.W. Howarth, A.N. Sharp- Langmuir 23:3680–3689. doi:10.1021/la062965n
ley, and V.H. Smith. 1998. Nonpoint pollution of surface wa- Sakadevan, K., and H.J. Bavor. 1998. Phosphate adsorption characteristics of
ters with phosphorus and nitrogen. Ecol. Appl. 8:559–568. soils, slags and zeolite to be used as substrates in constructed wetland
doi:10.1890/1051-0761(1998)008[0559:NPOSWW]2.0.CO;2 systems. Water Res. 32:393–399. doi:10.1016/S0043-1354(97)00271-6
Chardon, W.J., and G.F. Koopmans. 2005. Phosphorus workshop. J. Environ. Sharpley, A.N., S.C. Chapra, R. Wedepohl, J.T. Sims, T.C. Daniel, and K.R.
Qual. 34:2091–2092. doi:10.2134/jeq2005.0001in Reddy. 1994. Managing agricultural phosphorus for protection of surface
Codling, E.E., R.L. Chaney, and C.L. Mulchi. 2000. Use of aluminum- waters: Issues and options. J. Environ. Qual. 23:437–451. doi:10.2134/
and iron-rich residues to immobilize phosphorus in poultry litter and jeq1994.00472425002300030006x
litter-amended soils. J. Environ. Qual. 29:1924–1931. doi:10.2134/ Stout, W.L., A.N. Sharpley, and J. Landa. 2000. Effectiveness of coal combus-
jeq2000.00472425002900060027x tion by-products in controlling phosphorus export from soils. J. Environ.
Devau, N., E. Le Cadre, P. Hinsinger, B. Jaillard, and F. Gerard. 2009. Soil Qual. 29:1239–1244. doi:10.2134/jeq2000.00472425002900040030x
pH controls the environmental availability of phosphorus: Experimental Torrent, J., V. Barron, and U. Schwertmann. 1990. Phosphate adsorption and
and mechanistic modelling approaches. Appl. Geochem. 24:2163–2174. desorption by goethites differing in crystal morphology. Soil Sci. Soc. Am.
doi:10.1016/j.apgeochem.2009.09.020 J. 54:1007–1012. doi:10.2136/sssaj1990.03615995005400040012x
Devau, N., P. Hinsinger, E. Le Cadre, B. Colomb, and F. Gerard. 2011. Weng, L.P., W.H. Van Riemsdijk, and T. Hiemstra. 2007. Adsorption of hu-
Fertilization and pH effects on processes and mechanisms controlling mic acids onto goethite: Effects of molar mass, pH and ionic strength.
dissolved inorganic phosphorus in soils. Geochim. Cosmochim. Acta J. Colloid Interface Sci. 314:107–118. doi:10.1016/j.jcis.2007.05.039
75:2980–2996. doi:10.1016/j.gca.2011.02.034
Weng, L.P., W.H. Van Riemsdijk, and T. Hiemstra. 2008. Humic nanoparticles
Geelhoed, J.S., T. Hiemstra, and W.H. Van Riemsdijk. 1997. Phosphate at the oxide–water interface: Interactions with phosphate ion adsorption.
and sulfate adsorption on goethite: Single anion and competitive ad- Environ. Sci. Technol. 42:8747–8752. doi:10.1021/es801631d
sorption. Geochim. Cosmochim. Acta 61:2389–2396. doi:10.1016/
Weng, L.P., W.H. Van Riemsdijk, L.K. Koopal, and T. Hiemstra. 2006a.
S0016-7037(97)00096-3
Adsorption of humic substances on goethite: Comparison between
Hiemstra, T., J. Antelo, R. Rahnemaie, and W.H. Van Riemsdijk. 2010a. humic acids and fulvic acids. Environ. Sci. Technol. 40:7494–7500.
Nanoparticles in natural systems I: The effective reactive surface area of doi:10.1021/es060777d
the natural oxide fraction in field samples. Geochim. Cosmochim. Acta
Weng, L.P., W.H. Van Riemsdijk, L.K. Koopal, and T. Hiemstra. 2006b. Li-
74:41–58. doi:10.1016/j.gca.2009.10.018
gand and charge distribution (LCD) model for the description of ful-
Hiemstra, T., J. Antelo, A.M.D. Van Rotterdam, and W.H. Van Riemsdijk. vic acid adsorption to goethite. J. Colloid Interface Sci. 302:442–457.
2010b. Nanoparticles in natural systems II: The natural oxide fraction at doi:10.1016/j.jcis.2006.07.005
interaction with natural organic matter and phosphate. Geochim. Cos-
Weng, L.P., L.K. Koopal, T. Hiemstra, J.C.L. Meeussen, and W.H. Van Riemsdijk.
mochim. Acta 74:59–69. doi:10.1016/j.gca.2009.10.019
2005. Interactions of calcium and fulvic acid at the goethite-water interface.
Hiemstra, T., M.O. Barnett, and W.H. Van Riemsdijk. 2007. Interaction of si- Geochim. Cosmochim. Acta 69:325–339. doi:10.1016/j.gca.2004.07.002
licic acid with goethite. J. Colloid Interface Sci. 310:8–17. doi:10.1016/j.
Withers, P.J.A., I.A. Davidson, and R.H. Foy. 2000. Prospects for controlling
jcis.2007.01.065
nonpoint phosphorus loss to water: A UK perspective. J. Environ. Qual.
Hiemstra, T., R. Rahnemaie, and W.H. Van Riemsdijk. 2004. Interpretation 29:167–175. doi:10.2134/jeq2000.00472425002900010021x
of surface species from spectroscopy and CD modelling. Geochim. Cos-
Zeng, L., X.M. Li, and J.D. Liu. 2004. Adsorptive removal of phosphate from
mochim. Acta 68:A159.
aqueous solutions using iron oxide tailings. Water Res. 38:1318–1326
Hiemstra, T., and W.H. Van Riemsdijk. 1996. A surface structural approach to 10.1016/j.watres.2003.12.009. doi:10.1016/j.watres.2003.12.009
ion adsorption: The charge distribution (CD) model. J. Colloid Interface
Zhang, G.S., H.J. Liu, R.P. Liu, and J.H. Qu. 2009. Removal of phosphate
Sci. 179:488–508. doi:10.1006/jcis.1996.0242
from water by a Fe-Mn binary oxide adsorbent. J. Colloid Interface Sci.
Hiemstra, T., and W.H. Van Riemsdijk. 2004. Interfacial distribution of 335:168–174. doi:10.1016/j.jcis.2009.03.019
charge: Relationship between microand macroscopic ion adsorption
phenomena. Abs. Papers Am. Chem. Soc. 227:U1215.
Hiemstra, T., and W.H. Van Riemsdijk. 2006. On the relationship between charge
distribution, surface hydration, and the structure of the interface of metal hy-

www.agronomy.org • www.crops.org • www.soils.org 635

You might also like