Sunteți pe pagina 1din 46

PHOSPHATE REACTION DYNAMICS IN

SOILS AND SOIL COMPONENTS:


A MULTISCALE APPROACH
Yuji Arai1 and D. L. Sparks2
1
Department of Entomology, Soils and Plant Sciences,
Clemson University, Clemson, South Carolina 29634
2
Department of Plant and Soil Sciences,
University of Delaware, Newark, Delaware 19717

I. Introduction
II. P Chemistry
III. Phosphate Adsorption on Soil Components
A. Phosphate Adsorption on Soils (Empirical Approaches)
B. Phosphate Retention as AVected by Physicochemical
Properties of Soils
C. pH EVects on Phosphate Adsorption on Variable Charge Minerals
D. Phosphate Adsorption on Metal Oxides
E. Phosphate Adsorption on Phyllosilicate Minerals
F. Temperature EVects on P Adsorption on Soil Components
G. I EVects on P Surface Complexation
IV. Phosphate Surface Complexation on Soil Components
A. Surface Complexation‐Modeling Approaches
B. Electrophoretic Mobility Measurement Studies
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

C. Ex Situ Spectroscopic Studies


D. In Situ Spectroscopic Studies
V. Residence Time EVects on Phosphate Adsorption and Desorption in
Soils and Soil Components
A. Residence Time EVects Theory
B. Slow Adsorption
C. Slow Desorption Process and Hysteresis
D. Solid‐State, Inter‐, and Intraparticle DiVusion
E. Surface Precipitation
F. Higher Energy Binding Through Chemical Reconfiguration
VI. Future Research Needs
References

135
Advances in Agronomy, Volume 94
Copyright 2007, Elsevier Inc. All rights reserved.
0065-2113/07 $35.00
DOI: 10.1016/S0065-2113(06)94003-6
Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com
Created from qut on 2019-07-26 00:03:33.
136 Y. ARAI AND D. L. SPARKS

Macroscopic‐ to more recent molecular scale investigations have enhanced


our knowledge of soil phosphorus (P) chemistry, including the retention/
release mechanisms in soils and soil components. Phosphate uptake on metal
(oxy)hydroxide and phyllosilicate mineral surfaces and in soils generally
increases with decreasing pH. Rapid adsorption kinetics is generally ob-
served on many soil adsorbents at acidic pH. P fixation mechanisms such
as inner‐sphere complexation and intra‐ and interparticle diVusion often
result in slow P release (i.e., hysteresis and irreversible reactions), creating
challenges in remediating agricultural soils with high accumulations of P.
This chapter covers some of the historical soil P chemical research findings
via macroscopic approaches but focuses on more recent molecular scale
approaches for elucidating P retention/release mechanisms. # 2007, Elsevier Inc.

I. INTRODUCTION

Phosphorus (P) is an indispensable element for all living protoplasmic


organisms because of its genetic role in ribonucleic acid and as an essential
nutrient, along with N, for growth. More than 90% of the total P in the soil–
plant–animal system is in soils and less than 10% is in the remaining biological
systems (Ozanne, 1980). The P content of the lithosphere is 1200 ppm while
it is 200–5000 ppm (an average of 600 ppm) in soils (Lindsay, 1979). In the
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

hydrosphere, typical concentrations of total P in domestic wastewater, agricul-


tural drainage, and lake surface waters are 3–15, 0.05–1, and 0.01–0.04 ppm,
respectively (Snoeyink and Jenkins, 1980).
In the last several decades, excess P has been recognized as point and
nonpoint source pollutant throughout the world due to long‐term anthropo-
genic inputs (i.e., municipal and industrial eZuents, and synthetic and animal‐
based fertilizers) (Parry, 1998; Ryden et al., 1973; Sharpley et al., 2000;
Vaithiyanathan and Correll, 1992; Withers, 1996). While point source
pollution has been eVectively reduced since the late 1960s, many diYculties
still remain in controlling nonpoint P pollution which impacts freshwater and
seawater biogeochemical cycles. Eutrophication causes problems in recreational,
industrial, and drinking water supplies due to overgrowth of algae and
cyanobacteria and their decomposition which leads to a dissolved oxygen
shortage (Kotak et al., 1993; Sharpley and Rekolainen, 1997). Consumption
of such water, containing cyanobacteria, can be a serious health hazard to
livestock and humans due to its neuro‐ and hepatotoxic eVects (Lawton and
Codd, 1991). P mobility in agricultural settings, due to surface runoV and
leaching, also contributes to a wide range of water‐related problems, including

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 137

summer fish kills. For example, the microscopic organism Pfiesteria (Burkholder
and Glasgow, 1997) produces a toxin, killing various lower food chain
marine organisms (fish and crustaceans). The toxin may also be harmful to
humans because it is known to cause symptoms such as nausea, migraine
headaches, skin sores, acute loss of learning, and memory problems in labo-
ratory rats (Levin et al., 1997). Water quality problems as well as P‐related
environmental impacts remain despite many attempts to better understand
abiotic P chemical processes in soils and sediments. This chapter provides an
overview of P research on soils and soil mineral components, which has been
conducted at the macroscopic and, more recently, the molecular level scale.

II. P CHEMISTRY

P belongs to the Group VA in the periodic table with an electronic configu-


ration of ([Ne] 4s2 4p3). It is stable in the pentavalent state to form an ortho-
phosphate anion (i.e., phosphate) that retains a near tetrahedral complex
surrounded by four oxygen atoms. In most soil/water environments, H2 PO 4
and HPO2 4 are the thermodynamically favorable species (pKa1 ¼ 2.1, pKa2 ¼
7.2, and pKa3 ¼ 12.3). There are several other forms of P‐containing com-
pounds (polyphosphates, metaphosphates, and organic P). The condensed
forms of inorganic P (i.e., polyphosphate and metaphosphates) are formed
with two or more orthophosphate groups. Whereas polyphosphates are
linear O‐P‐O linkages, the metaphosphates are cyclic.
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

The organic P percentage of total soil P can range from 20 to 80% (Dalal,
1977). Several forms of organic P have been identified in soils. They are inositol
phosphate, nucleic acids, and phospholipids. Inositol phosphate (phytic acid)
makes up more than 50% of the total organic P due to its high stability in soils
whereas the phospholipid content comprises as little as 0.5–7% of total organic P
(Dalal, 1977). Nucleic acid, which originates from the decomposition of
microbes, plants, and animal remains, is the smallest (less than 3%) fraction
of the total organic P (Dalal, 1977). P forms complex minerals with a wide
variety of elements. About 150 P minerals are known (Cathcart, 1980). Accord-
ing to Povarennykh’s structural classification, P minerals can be placed into
four groups. They are framework, insular, chain, and layer minerals. A majority
of the P minerals belong to insular minerals, including the apatite group
(Ca2Ca3(PO4)3(OH, F)) and wavellite (Al3(PO4)2(OH)3 5H2O). P solubility
products are generally controlled by pH, concentration of P, and divalent
(e.g., Ca2þ, Mg2þ, and Fe2þ) and trivalent cations (e.g., Al3þ and Fe3þ) in
bulk solutions. Where the latter cations are less available, P solubility is
strongly controlled by adsorption onto clays and clay minerals in subsurface
horizons. In reduced soil environments, P solubility can be increased due to the

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
138 Y. ARAI AND D. L. SPARKS

reductive dissolution of iron‐containing adsorbents like iron oxides and ferric‐P


minerals (Diaz et al., 1993; Gale et al., 1992; Shapiro, 1957).

III. PHOSPHATE ADSORPTION ON


SOIL COMPONENTS

Surface species (inner sphere or outer sphere) and the P‐bonding environ-
ment (e.g., monodentate mononuclear and bidentate binuclear) can be identi-
fied and studied using macroscopic, microscopic, and spectroscopic techniques.
Electrophoretic mobility (EM) measurements, ionic strength (I ) eVects on the
adsorption envelope, and sorption‐proton balance data can be used to indi-
rectly distinguish predominant surface complexes at mineral–water inter-
faces. X‐ray absorption spectroscopy (XAS) and attenuated total
reflectance Fourier transform infrared (ATR‐FTIR) spectroscopy are two
powerful techniques that can be employed to directly identify not only
the types of surface complexes (inner‐sphere or outer‐sphere) but also the
bonding mode at mineral–water interfaces under in situ conditions.

A. PHOSPHATE ADSORPTION ON SOILS (EMPIRICAL APPROACHES)

The retention of P in soil–water environments has received much atten-


tion in soil chemical studies due to its important biogeochemical role in the
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

environment. It is often found that P adsorption in soils increases with


decreasing pH (Sanchez and Uehara, 1980). P adsorption reactions on acidic
soils are typically biphasic, characterized by an initial rapid reaction fol-
lowed by a much slower reaction (Barrow, 1985; Parfitt and Smart, 1978).
The slow and continuous P adsorption on soils and soil components
has been reported by many researchers (Barrow and Shaw, 1975; Munns
and Fox, 1976; van der Zee and van Riemsdijk, 1988; van Riemsdijk and
de Haan, 1981). Researchers have attempted to understand P adsorption
phenomena using macroscopic approaches (e.g., adsorption isotherms and
envelopes) coupled with empirical and surface complexation models.
In the past, P adsorption isotherms were extensively fitted to the Langmuir
equation to describe diVerent adsorption ‘‘sites.’’ These sites were defined by
multiple linear portions of the Langmuir plot (Fried and Shapiro, 1956;
Muljadi et al, 1966; Olsen and Watanabe, 1957; Ryden et al., 1977b; Yao
and Millero, 1996). Some investigators used a two‐site Langmuir equation
to describe the low‐ and high‐energy sites involved in the adsorption reaction.
A P adsorption reaction at pH  7 on 19 Vertisols (clay rich soils) was
described using a two‐site Langmuir equation. The maximum adsorption

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 139

values from the two‐site model were highly correlated with the calcite content
for high‐energy sites and dithionite‐citrate‐bicarbonate (DCB) extractable
iron for low‐energy sites (López‐Pineiro and Navarro, 1997). Yao and Millero
(1996) used a Langmuir adsorption isotherm equation to understand P
adsorption on birnessite surfaces (Yao and Millero, 1996). The data exhibited
three linear segments, which were related to a variety of adsorption‐binding
sites. Veith and Sposito (1977) showed that the Langmuir equation can
equally well describe both adsorption and precipitation reactions (Veith and
Sposito, 1977). Such empirical equations should not be used to interpret
any particular adsorption mechanism or even if adsorption, as opposed to
precipitation, actually has occurred (Sposito, 1989).

B. PHOSPHATE RETENTION AS AFFECTED BY PHYSICOCHEMICAL


PROPERTIES OF SOILS

The retention of P on soils is highly dependent on the physicochemical


properties of the soils, for example crystalline and amorphous iron and alumi-
num oxides, and organic matter (OM), clay and calcium contents. Many
researchers have shown a correlation between these parameters and adsorption
and chemical extraction data.
Long‐term (256 days) P adsorption on 12 Oxisols (well‐weathered soils) was
investigated by Barrón and Torrent (1995). In their experiments, the adsorption
maximum from the Langmuir equation suggested that slow adsorption was
highly correlated with the ratio of the OM content and the specific surface area
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

of the solids. High P retention in an OM (i.e., an aluminum‐substituted woody


peat) was also observed (Bloom, 1981; López‐Hernandez and Burnhan, 1974).
Several researchers have reported that P retention in various soils (e.g.,
Entisols, Oxisols, and soil clays) are highly associated with ammonium
oxalate extractable iron (Feox) and/or aluminum (Alox), DCB extractable
iron and/or aluminum, and Fe oxides detected by Mössbauer spectroscopy
(Arai et al., 2005; Bloom, 1981; Lookman et al., 1995; López‐Hernandez and
Burnhan, 1974; Maguire et al., 2000; Sei et al., 2002).
In the study of Arai et al. (2005), results of sequential inorganic P fraction-
ation indicated that ammonium oxalate extractable P/Al/Fe fractions were
predominant in long‐term poultry litter‐amended Southern Delaware agricul-
tural soils, followed by crystalline iron phosphate‐associated fractions, crystal-
line aluminum phosphate‐associated fractions, soluble P fractions, and calcium
phosphate‐associated fractions (Arai et al., 2005). These Al/Fe‐P associations
were also supported by results from electron microprobe analyses.
High P retention in calcium rich soils and soil components at pH > 7 has
also been reported. Kuo and Lotse (1972) observed strong P retention on
calcite and Ca‐kaolinite, and a second‐order kinetic equation was fitted to

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
140 Y. ARAI AND D. L. SPARKS

obtain the rate coeYcient. The coeYcient was 30,000 times higher for calcite
than for Ca‐kaolinite. A similar correlation was also observed for P adsorption
on calcite‐rich Vertisols (López‐Pineiro and Navarro, 1997).
While many macroscopic studies have shown a strong correlation
between P retention and specific soil properties, they have not provided
any information on the spatial distribution of P in heterogeneous systems.
Several microscopic techniques have been applied to investigate P retention
phenomena in heterogeneous systems.
Phosphate fixation in P fertilizer‐amended soils (pH  7) was investigated
using scanning transmission electron microscopy (STEM) along with energy
dispersive X‐ray spectroscopy (EDX) (Pierzynski et al., 1990b). In the
separated clay fractions, it was observed that discrete P‐rich particles were
highly associated with Al and Ca in the highest density‐separated fraction
(<2.2 Mg m–3). Furthermore, the EDX model calculations, using elemental
ratios (cations/P), also showed an Al and Ca enrichment associated with P.
These results were later supported by a P solubility equilibrium study which
predicted the formation of varicsite (aluminum phosphate) like solids at pH < 6.8
and calcium phosphate like solids at pH > 6.8 (Pierzynski et al., 1990a).
The P mineralogy of Florida soils derived from phosphoric deposits was
investigated using X‐ray diVraction (XRD), thermal analysis, and selective
dissolution (Wang et al., 1991). The XRD data showed the presence of
carbonate‐fluroapatite (Ca10(PO4, CO3)6F2–3), wavellite (Al3(PO4)(OH)3 
5H2O), and crandallite (CaAl3(PO4)2(OH)5 H2O). The oxalate extractable
P (Pox) was associated with Alox in the surface horizons. Endothermic diVer-
ential scanning calorimetry peaks (90–100  C), which originated from dehy-
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

droxylation of amorphous Al‐ and Fe‐P minerals, were eliminated by oxalate


extraction. This suggests that P might associate with amorphous Al and Fe
minerals.
A similar study was conducted on manure‐derived surface soils and sedi-
ments (pH 6.9–9.5) (Harris et al., 1994). Amorphous apatite and ferrous‐P
minerals (vivianite) were detected by XRD in stream sediment samples. The
lack of crystalline Ca‐P minerals suggests that the manure components might
inhibit the crystallization of Ca‐P minerals.

C. pH EFFECTS ON PHOSPHATE ADSORPTION ON VARIABLE


CHARGE MINERALS

Phosphate adsorption on soil minerals is greatly influenced by the pH of


the bulk solution. Metal oxides and phyllosilicate minerals in soils contain
surface functional groups (unsatisfied bonds with respect to the repeated
bonding of the unit cells). Examples include: (1) the siloxane surface asso-
ciated with the plane of oxygen atoms bound to the silica tetrahedral layer of

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 141

phyllosilicate minerals, and (2) hydroxyl groups that are associated with the
edges of inorganic minerals such as phyllosilicates, metal oxides, oxyhydr-
oxides, and hydroxides (Sparks, 1995a). When the mineral is hydrated, the
metal ion (Lewis acid) sites are occupied with water molecules and/or are
associated with Lewis base sites (hydroxylated surface) because of the disso-
ciative chemisorption of the water molecules. The surface sites can be proto-
nated and deprotonated depending on the pH of the bulk fluid. Therefore,
they are often called variable charge (pH‐dependent charge) mineral surfaces.
The speciation of oxyanions is also influenced by pH. The dissociation
constant (equilibrium constant), Ka, for the oxyanions refers to the reaction
in which an acid donates a proton to water. The larger the Ka value, the
higher the tendency to donate protons to water molecules.
Therefore, oxyanion adsorption on soil components is a function of both the
net surface charge density of the adsorbent and the chemical speciation of
the adsorbate which in turn are dependent on the pH of the bulk fluid (pHb).
In general, P adsorption on inorganic minerals increases with decreasing pHb
due to: (1) the negatively charged chemical species (i.e., H2 PO 2
4 and HPO4 )
and (2) the positively charged mineral surfaces, when pHb < PZC, of the
solids.

D. PHOSPHATE ADSORPTION ON METAL OXIDES

The first dissociation constant of P is 2.2, this is followed by constants


of 7 and 12.8. At most environmental pH values (4–8), the species are
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

predominantly in deprotonated forms (negatively charged species), and the


charge properties of metal oxides are positive due to the PZC of the solids
[i.e., 6.5–8.5 for iron oxides, 8.2–9.1 for aluminum oxides; an exception is
manganese oxides (e.g., birnessite) 2.8]. Therefore, P is expected to adsorb
on metal oxide surfaces strongly via electrostatic interaction when pHb‐PZC
is less then zero and to predominantly adsorb via ligand exchange when
pHb‐PZC is greater than zero.
Yao and Millero (1996) studied P adsorption on birnessite surfaces in 0.7‐M
NaCl solutions and seawaters. In both media, P adsorption was maximized at
pH  3 and gradually decreased with increasing pH (Yao and Millero, 1996).
Hingston et al. (1967) reported that P adsorption on goethite increased
with decreasing pH. The P adsorption envelope showed inflection points
near the pK values of phosphoric acid (Hingston et al., 1967).
A similar pH‐dependent adsorption behavior has been observed for P on
akaganeite, ferrihydrite, hematite, goethite, boehmite, and amorphous
Al(OH)3 (Bleam et al., 1991; Chen et al., 1973; Chitrakar et al., 2006;
Shang et al., 1992). The results of our recent investigation agree with the
previous research findings. Phosphate adsorption envelope kinetics at the

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
142 Y. ARAI AND D. L. SPARKS

100

PZSE ∼ 8
80
24 h
12 h
P sorbed (%)

60 4h
2h
1h
40 30 min
10 min
20 5 min

0
3.5 4.5 5.5 6.5 7.5 8.5 9.5
pH

Figure 1 Phosphate adsorption kinetics (suspension density, 1.25 g liter1; I, 0.1‐M NaCl;
and initial P concentration, 1 mM) at the ferrihydrite–water interface as a function of pH.

ferrihydrite–water interface [five‐point Brunauer‐Emmett‐Teller (BET) sur-


face of two‐line ferrihydrite, 260 m2 g1; suspension density, 1.25 g liter1;
I, 0.1‐M NaCl; and initial P concentration, 1 mM] at pH 4.0–9.5 are shown in
Fig. 1. Phosphate adsorption increases with decreasing pH from 9.5 to 4
(Fig. 1). Whereas, 98.2% adsorption (2.70 mmol m2) at pH 4 was achieved
after 24 h, slow adsorption continued at pH > 4 after 24 h (Fig. 1). Only 58%
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

(1.59 mmol m2) of total P uptake was achieved at pH 9.5 after 24 h and the
slow P uptake continued after 24 h (Fig. 1). The surface charge density of
ferrihydrite becomes more negatively charged with increasing pH from 4.0 to
the PZSE of the solid (pHPZSE  8). The mineral surface at pH < 8 is more
positively charged than at pH > 8, and the protonated mineral surfaces at
pH 4 strongly attract negatively charged P solution species (e.g., H2 PO 4 ). It is
interesting that the extent of P adsorption kinetics rapidly drops after pH  8.
This is probably due to changes in net surface charge density of ferrihydrite
from positively to negatively charged surfaces that repel HPO2 4 species.

E. PHOSPHATE ADSORPTION ON PHYLLOSILICATE MINERALS

While the metal oxides exhibit a strong aYnity for P at acidic pH values,
the phyllosilicate minerals show a diVerent adsorption capacity. In general, the
PZC of clays is lower than that of iron and aluminum (oxyhdr)oxides (i.e., 4.6
for kaolinite and 2.5 for montmorillonite) (Sparks, 1995a). Therefore, the

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 143

pH‐dependent edge sites of phyllosilicate minerals are generally negatively


charged at most environmental pHs (4–8). If there is oxyanion adsorption at
the higher pH range, the oxyanions usually specifically adsorb onto edge sites
via ligand exchange. Phosphate adsorption on illite and kaolinite gradually
increases from pH 3 to 5, and then the adsorption decreases with increasing pH
(Chen et al., 1973; Edzwald et al., 1976).

F. TEMPERATURE EFFECTS ON P ADSORPTION ON SOIL COMPONENTS

Temperature may have two distinct eVects on a chemical reaction. These


are: (1) the rate of reaction and (2) the equilibrium end point (Barrow, 1987).
Reaction at high temperature results in an increase in the reaction rate and a
decrease in the subsequent desorption if the reaction involves activated
complexes (intermediate‐ or high‐energy states) (Barrow, 1979b).
The eVect of temperature on the chemical reaction can be explained using
transition state theory (Eyring reaction rate theory). The schematic reaction
flow is:
A þ B $ ABz ! Products ð1Þ

where A and B are the reactants and ABz is the activated complex.
Using transition state theory, the elementary reaction process can be
expressed as: (1) the total molecular partition functions per unit volume
(qi) for the reactant species and for the activated complexes (qz) and (2) the
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

diVerence in zero‐point potential energies between the activated complex


and reactants (E0):
    
ðkB TÞ qz E0
k¼ exp  ð2Þ
h qA qB ðkTÞ

where k is the elementary rate constant, kB is the Boltzmann’s constant, T is


the absolute temperature, and h is the Planck’s constant (Stumm and Morgan,
1995b).
The previous equation can also be thermodynamically reexpressed as
  
kB T z gA gB
k¼ K ð3Þ
h gz
h  z
i
where Kz is the thermodynamic formation constant exp  DG RT and g is
the activity coeYcient. Since the activation complexes are diYcult to mea-
sure using the above equations during temperature‐dependent reactions, the
linearized Arrhenius empirical rate law is commonly used.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
144 Y. ARAI AND D. L. SPARKS
 
Ea
ln k ¼ ln A þ  ð4Þ
RT

where A is the the pre‐exponential factor, Ea is the activation energy, R is the ideal
gas constant, and T is the temperature. Combining the transition state theory
and the above thermodynamic expression gives A ¼ [(kBT)/h]exp(DSz /R)
and Ea ¼ DHz þ RT, assuming the activity coeYcients are all unity. If an
endothermic adsorption reaction (positive enthalpy) is observed, one can
estimate the apparent activation energy of the reaction. A linear plot between
ln k versus 1/T allows one to obtain the value of the slope (Ea/R). The
reaction is generally diVusion limited when Ea < 42 kJ mol1, whereas a
chemically controlled reaction is suggested when Ea > 42 kJ mol1 (Sparks,
1995b).
Various results have been reported on temperature‐dependent P adsorp-
tion and desorption on soils and soil components. In acid soils, P adsorption
increases with increasing temperature and desorption is subsequently
reduced, suggesting that the reaction involves activated complexes. (Barrow,
1979b; Barrow and Shaw, 1977; Chien et al., 1982; Sheppard and Racz,
1984). van Riemsdijk and Lyklema (1980) reported that P adsorption on
gibbsite at pH 5 increased with increasing temperature (2, 12, 22, and 45  C).
An estimated activation energy of 63  4 kJ mol1 was determined, suggest-
ing that the reaction was controlled more by chemical than by physical
processes (van Riemsdijk and Lyklema, 1980). Yao and Millero (1996)
reported that P adsorption kinetics on birnessite surfaces were temperature
dependent at pH 8. P uptake increased with increasing temperature from 5 to
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

35  C during the 50 h of the kinetic experiments. Assuming a pseudo first‐


order reaction during the initial 5 min of reaction, they reported an apparent
activation energy of 9 kJ mol1. With goethite, however, P adsorption
increased with decreasing temperature from 68 to 25  C. This indicated
an exothermic temperature‐dependent reaction and nonactivated complex
formation (Madrid and Posner, 1979).
Temperature‐dependent P desorption kinetics from desert soils were studied
using an anion‐exchange resin (Evans and Jurinak, 1976). The energy of
activation was estimated from the data based on 4 h of desorption. An activa-
tion energy of less than 42 kJ mol1 in all soils was determined, suggesting that
P release might depend on diVusion processes. Conversely, Barrow (1979b)
found that the activation energy for the P adsorption (forward) reaction was
similar to that of the desorption (backward) reaction. The values for both steps
were 80 kJ mol1, suggesting that the rate‐limiting steps for both reactions
were not diVusion‐controlled. Bar‐Yosef and Kafkafi (1978) estimated activa-
tion energies for initial rapid and slow P desorption processes for kaolinite of
67.8 and 20.8 kJ mol1, respectively. This suggested that chemically controlled
desorption was followed by diVusion‐controlled desorption.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 145

G. I EFFECTS ON P SURFACE COMPLEXATION

I can have an influence on both the rate of the elementary reaction and the
type of surface complexation (inner‐sphere and/or outer‐sphere complexa-
tion) (Hayes et al., 1988; Stumm and Morgan, 1995b). Oxyanions adsorb onto
variable charge mineral surfaces by both inner‐sphere (via ligand exchange)
and outer‐sphere complexation (via electrostatic interaction) (McBride,
1989). These surface species can be indirectly inferred by studying the I eVects
on the degree of adsorption (Hayes et al., 1988). Inner‐sphere complexes (i.e.,
selenite) are insensitive to changes in I due to ligand exchange adsorption
mechanisms, while outer‐sphere complexes (e.g., selenate) are sensitive to
changes in I because of competition from the counteranions of the indiVerent
electrolytes.
Many macroscopic studies have investigated I‐dependent oxyanion ad-
sorption behavior on clays and clay minerals. Ryden et al. (1977b) reported
that P adsorption isotherm on New Zealand loamy soils was not significantly
aVected by I eVects ([NaCl] ¼ 0.001–1 M) at <6‐mM equilibrium concentra-
tions (Ceq). Arai and Sparks (2001) showed no I‐dependent P adsorption on
amorphous iron oxyhydroxide at pH 4–7.5. Similarly, P adsorption on bir-
nessite at pH 2–8.5 was not aVected by changes in salinity (5–35 ppt) (Yao and
Millero, 1996). On the basis of the theory above, these results might indicate
the formation of inner‐sphere surface species on the adsorbents.
However, there is an exception to this theory. Phosphate adsorption often
increases with increasing I. Several researchers have documented the reac-
tion on soils (New Zealand loamy soils) and soil components (kaolinite,
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

montmorillonite, illite, and goethite) (Barrow, 1980; Edzwald et al., 1976;


Helyar et al., 1976a,b; Ryden et al., 1977b). Two theories have been sug-
gested to explain this unique adsorption behavior. They are: (1) the quadru-
ple layer model theory and (2) the diVuse double layer theory. Barrow’s data
were successfully fitted using a quadruple layer adsorption model (Bowden
et al., 1980). Such complex model applications are generally not subject to
direct experimental confirmation because they employ several fitting para-
meters that cannot be analytically measured (McBride, 1997). Decreased
double layer thickness (DLT), due to increased I, allows the oxyanions to
closely approach the negatively charged surface, and then they adsorb via
ligand exchange. Inner‐sphere adsorption, however, should not be influ-
enced by either the DLT or the repulsion forces because the specific adsorp-
tion is in direct coordination with discrete surface metal cations (McBride,
1997). Therefore, the quadruple model and DLT are inappropriate to
describe the above P adsorption behavior.
McBride used the simple mass action principle to explain adsorption
phenomena. This approach is similar to the constant capacitance model
(CCM) described by Goldberg and Sposito (1984b). It ignores, however, the

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
146 Y. ARAI AND D. L. SPARKS

surface electrical potential correction term, c, by assuming that any inner‐


sphere and outer‐sphere counterions are adsorbed to balance the surface
charge created in the process. In the case of high I, the negatively charged
surfaces created by inner‐sphere P adsorption are likely to be neutralized by
coadsorption of cations (i.e., Naþ) from indiVerent electrolytes maintaining
the charge balance. This mass action principle favors such reactions when a
higher concentration of indiVerent electrolyte is present. No spectroscopic
evidence, however, is available to support this theory.

IV. PHOSPHATE SURFACE COMPLEXATION ON


SOIL COMPONENTS

While numerous macroscopic studies have investigated adsorption behav-


ior with isotherms and envelopes, they have not provided any information on
adsorption mechanisms (i.e., surface complexation) at the molecular scale.
Recent empirical and modeling approaches [proton balance measurements
and surface complex models (SCM)], and microscopic [electrophoretic mo-
bility (EM) measurements] and spectroscopic (FTIR spectroscopy and XAS)
studies have provided better insight on oxyanion adsorption mechanisms at
the molecular scale.

A. SURFACE COMPLEXATION‐MODELING APPROACHES


Copyright © 2007. Elsevier Science & Technology. All rights reserved.

The lack of mechanistic information obtained from previous empirical


approaches (e.g., Langmuir and Freundlich adsorption isotherm equations)
has led to the use of SCM to describe adsorption phenomena. SCM are chemical
models that are based on the molecular description of the electric double layer
using equilibrium‐derived data (Goldberg, 1992). The CCM is one of the many
surface complexation models and has been successfully used to describe an array
of chemical reactions on mineral surfaces. These include adsorption, desorption,
and dissolution reactions on soil components. For instance, the CCM has been
used to quantitatively describe P adsorption on 44 noncalcareous soils over a
wide range of pH (4.9–7.6) (Goldberg and Sposito, 1984a). Using the derived
parameters (intrinsic surface protonation dissociation constants, capacitance
density, P packing area parameters), the CCM described the inner‐sphere
adsorption mechanisms (i.e., protonated and nonprotonated P surface species)
quite well using adsorption envelope and isotherm data.
Goldberg and Sposito (1984b) used a similar CCM approach, which consists
of adjustable surface protonation–dissociation constants, surface complexa-
tion constants, and capacitance density parameters, to predict P adsorption on

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 147

aluminum oxide and iron oxide surfaces. Assuming inner‐sphere surface com-
plexation on mineral surfaces, P adsorption envelopes and isotherms on alumi-
num oxide, and iron oxide surfaces were predicted well, resulting in similar
protonation–dissociation constants and surface complexation constants for
both solids.
Nilsson et al. (1992) have also demonstrated the use of CCM to predict P
adsorption behavior on goethite surfaces with the aid of the computer model
FITEQL. The model was able to predict pH‐dependent inner‐sphere adsorp-
tion mechanisms that are represented by three equilibria and intrinsic
constants of protonated and nonprotonated surface species.
Yao and Millero (1996) used the triple layer model to predict P adsorp-
tion on birnessite surfaces. Whereas, a consideration of only inner‐sphere
surface species significantly overestimated the total P adsorption at pH < 7,
a model with outer‐sphere surface species predicted the experimental data in
0.7‐M NaCl solutions fairly well.
Results of equilibrium data modeled using the SCM must be carefully
interpreted because the actual adsorption mechanisms is not proven, high
degrees of freedom in the adjustable parameters allow one to describe
material balance data very well, and the majority of the SCM do not include
surface precipitation and other nonadsorption phenomena as part of the
model description and prediction (Sparks, 1995b).

B. ELECTROPHORETIC MOBILITY MEASUREMENT STUDIES


Copyright © 2007. Elsevier Science & Technology. All rights reserved.

Electrophoretic mobility (EM) measurements are a useful microscopic


approach for not only determining the isoelectric point (IEP) of pure compo-
nents but also for obtaining information that can be used to indirectly
distinguish bulk surface complexes at colloidal–water interfaces. Nonspecific
ion adsorption of indiVerent electrolyte at the outside of the shear plane
(i.e., formation of outer‐sphere complexes via van der Waals forces) generally
does not aVect the IEP but it could cause shifts in the value of EM if the
electrolyte is present at high concentration (Hunter, 1981). The shear plane is
at the outer edge of the inner part of the double layer and near the outer
Helmholtz plane or the Stern layer, depending on the models used to describe
the interface (Hunter, 1981). Inner‐sphere complexes, however, cause shifts in
both EM and IEP due to specific ion adsorption inside the shear plane
(Hunter, 1981). In other words, oxyanion inner‐sphere adsorption does
increase the net negative charge on the surface. In some cases, however,
inner‐sphere adsorption does not cause shifts in EM (Hunter, 1981).
With this knowledge, one can indirectly distinguish the predominant surface
complexes on pure colloidal materials using EM measurements.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
148 Y. ARAI AND D. L. SPARKS

There are some EM data suggesting predominant inner‐sphere P com-


plexes on metal oxides. Microelectrophoretic mobility measurements of P
adsorbed on ferrihydrite, goethite, and boehmite (g‐AlOOH) have shown that
inner‐sphere complexes form due to charge reversal and a lower IEP with
increasing P loading level (Anderson and Malotky, 1979; Arai and Sparks,
2001; Bleam et al., 1991; Hansmann and Anderson, 1985; Tejedor‐Tejedor
and Anderson, 1990).

C. EX SITU SPECTROSCOPIC STUDIES

Ex situ spectroscopic techniques have been extensively utilized to directly


distinguish the adsorption mechanisms of P on clays and clay minerals. Parfitt
et al. (1975) investigated P adsorption complexation on iron oxides (ferri-
hydrite, goethite, lepidocrocite, and hematite) using ex situ infrared (IR)
spectroscopy. The IR spectra of P adsorbed on iron oxides showed the replace-
ment of two singly coordinated surface hydroxyls, suggesting the formation of
bidentate binuclear species. Using the same IR technique, Atkinson et al. (1974)
compared the v3 bands of P adsorption complexes on goethite and other model
systems [e.g., Co(III)‐P solution complexes]. They suggested that the shift in v3
bands of the adsorption complex was due to binuclear bidentate species
(Atkinson et al., 1974). Nanzyo and Watanabe (1982) utilized diVuse reflec-
tance FTIR (DRIFTIR) spectroscopy to investigate P adsorption complexes
on goethite over a wide pH range of 3.3–11.9. Goethite background subtracted
IR spectra showed that: (1) the surface complexes at the same pH did not
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

change with increasing loading levels up to 197 mole g1 and (2) bidentate
bridging complexes were present throughout all pH values based on the IR
bands assigned by Parfitt et al. (1975).
DRIFTIR spectroscopy was also utilized to re‐investigate P adsorption
surface complexation on goethite (Persson et al., 1996). Using the symmetry
rule arguments of P v3 bands, Persson suggested that the monodentate surface
complex predominantly formed between pH 3 and 12.8, but at intermediate
pH, a bidentate complex could not be ruled out.
Martine and Smart (1987) utilized X‐ray photoelectron spectroscopy (XPS)
to investigate phosphate adsorption complexation on goethite at pH 3 and 12,
at high initial P concentrations ([P]i ¼ 5 and 50 mM). They concluded that the
replacement of two A‐type surface hydroxyls via a ligand exchange reaction
suggested bidentate binuclear adsorption.
While ex situ spectroscopic studies have strongly suggested specific
P adsorption (inner‐sphere complexes) on Fe and Al oxides, the results have
been questioned by many researchers because of the creation of artifacts (i.e.,
a structural alteration due to vacuum pressure and/or formation of bulk
precipitates by drying residual adsorbates) under dry and severely evacuated

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 149

sample conditions. Colloidal–water interfaces in natural settings are usually


under near atmospheric pressure. Analyzing samples near environmental
conditions (in situ) is more appropriate to obtain accurate information on
environmental samples (Goldberg and Sposito, 1985).
Accordingly, in the past decade, with the development of in situ molecular
scale techniques such as XAS and ATR‐FTIR, one can directly determine
oxyanion reaction mechanisms on minerals under environmental conditions
that are representative of field settings.

D. IN SITU SPECTROSCOPIC STUDIES


31
P solid‐state magic angle spinning nuclear magnetic resonance (NMR)
spectroscopy coupled with the constant‐capacitance model was utilized to
investigate the hydrolysis of adsorbed P molecules on boehmite (g‐AlOOH)
(Bleam et al., 1991). An inner‐sphere complex was suggested between the pH
range of 4 and 11, and a fully deprotonated P surface complex was also
reported at pH > 9 (Bleam et al., 1991).
Parfitt and Atkinson (1976) combined in situ IR spectroscopy and potentio-
metric titration experiments to examine P adsorption complexation at the
goethite–water interface at pH 3.6 or 5.1. They suggested that binuclear bidentate
complexes were predominantly formed at both pH values. Tejedor‐Tejedor
and Anderson (1990) used in situ cylindrical internal reflection‐FTIR (CIR‐
FTIR) to investigate P adsorption complexation on goethite between pH 4 and
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

8. Comparing the spectra of ferric phosphate solutions with those of adsorption


complexes on goethite, the formation of protonated and non‐protonated biden-
tate binuclear and nonprotonated monodentate mononuclear complexes was
suggested (Tejedor‐Tejedor and Anderson, 1990). Luengo et al. (2006) have used
ATR‐FTIR to elucidate the changes in P surface speciation on goethite surfaces
with time (5–400 min). They reported that non‐protonated and protonated
bidentate species coexist at pH 4.5, and these species form over time rather
than independently. Nonprotonated bidentate species are predominantly formed
at pH 7.5 and 9 with an extra unidentified species at low concentration. The
formation of these species was not dependent on reaction times.
Arai and Sparks (2001) investigated P surface species (reaction time up to
24 h) at the ferrihydrite–liquid (H2O and D2O) interface using ATR‐FTIR
spectroscopy. The IR study was combined with peak deconvolution pro-
cesses to understand the reduction of PO4 Td symmetry that is caused
by surface complexation and/or protonation on surface species. They sug-
gested that inner‐sphere surface complexes were nonprotonated bidentate
binuclear species (Fe2PO4) at pH > 7.5 and the surface complexes might
coexit with diVerent surface species (e.g., monodentate mononuclear) and/or

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
150 Y. ARAI AND D. L. SPARKS

A B

pH 4,
pH 7.5,
1 year
1 year
1025
1035
970
950
1099
1095

978
Normalized absorbance

Normalized absorbance
pH 4, pH 7.5,
1.5 months 1.5 months
1043
1015

1095
941
956 979
1095

pH 7.5, 2 days
pH 4, 2 days
1028

952
1029
1098
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

976
1091

1200 1100 1000 900 1125 1050 975 900


Wavenumber (cm−1) Wavenumber (cm−1)

Figure 2 ATR‐FTIR spectra of P adsorbed on ferrihydrite at pH 4 (A) and pH 7.5 (B) as a


function of residence time. The raw spectra are shown with solid lines, the deconvoluted peaks
with dotted lines, and the fitted curve with open circles.

Fe2PO4Na at pH  7.5. The exact identity of the protonated P inner‐


sphere surface complexes (i.e., protonated monodentate mononuclear com-
plexes and/or protonated bidentate binuclear complexes) forming at pH < 7.5
could not be elucidated due to limitations in mid‐IR range FTIR analysis.
Arai and Sparks carried out the similar experiments up to 1 year to
observe the changes in P surface speciation at the ferriydrite–water interface.
ATR‐FTIR spectra of aged P‐reacted ferrihydrite with Gaussian profile fits
are shown in Fig. 2A and B. In pH 4 and 7.5 samples (Fig. 2A and B,

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 151

respectively), the fitted Gaussian peaks show triply nondegenerate v3 vibra-


tions except for 1.5 months and 1 year samples at pH 7.5, suggesting that the
symmetry of the surface species is C2v and/or lower symmetry. The peak
positions of the v3 vibrations are similar to those observed in the previous
short‐term adsorption study (Arai and Sparks, 2001), indicating that similar
surface complexes might be forming in aged samples. As we previously
reported, deuterium exchange did not cause any significant peak shift
at pH 7.5, but some shifts occurred at pH 4 under similar loading levels
 2.42 mmol m2, suggesting nonprotonated inner‐sphere surface species for
pH 7.5 samples and protonated inner‐sphere surface species for pH 4 sam-
ples (Arai and Sparks, 2001). Determination of exact bonding environments
for pH 4 samples was diYcult based on the FTIR information alone because
several diVerent molecular configurations (e.g., monoprotonated bidentate
binuclear and/or bidentate binuclear with hydrogen bonding to mineral
surfaces) could satisfy C2v and/or lower symmetry with proton associations.
Therefore, we only suggested the formation of protonated inner‐sphere
complexes at pH 4 in aged samples.
While triply nondegenerated v3 vibrations are observed in diVerent aged
samples at pH 4, Gaussian peak fit analyses revealed the presence of an
additional fourth v3 band (979 cm1) in 1.5 months aged pH 7.5 samples
(Fig. 2B). We interpret the fourth peak as the presence of secondary surface
species in addition to the predominant surface species observed in the 2‐day‐
aged sample (a bottom spectrum in Fig. 2B). The peak at 978 cm1 is
probably one of the v3 vibrations arising from the secondary complexes. It is
diYcult to understand whether the secondary surface complexes give two or
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

three v3 vibrations since the strong signals (i.e., three v3 vibrations) from the
predominant surface complexes in the 1.5‐month‐aged samples probably
overlap with the other v3 vibrations from the secondary complexes. At
pH 7.5, ferrihydrite surfaces are predominantly occupied by non‐protonated
bidentate bridging PO4 surface species (Arai and Sparks, 2001), and P‐sorbed/
unreacted ferrihydrite surface sites could possibly be accessible by additional
monoprotonated (HPO2 4 ) aqueous species to form a minor fraction of: (1)
diVused HPO2 4 ions and/or (2) inner‐sphere complexes. In fact, one of the v3
vibrations in the monoprotonated (HPO2 4 ) aqueous species is exhibited
around 989 cm1 (Arai and Sparks, 2001), and diVused HPO2 4 molecules in
ferrihydrite particles could possibly result in a slight peak shift, producing the
forth peak position at 978 and 979 cm1 in 1.5‐month‐aged samples as
observed in Fig. 2B. However, we cannot exclude the possibility of secondary
inner‐sphere complexes forming.
Hesterberg et al. (1999) have used fluorescence yield P K‐edge X‐ray
absorption near edge structure spectroscopy (XANES) to investigate solid‐
state P speciation in North Carolina agricultural soils. P K‐edge XANES
spectra for Fe‐phosphates were characterized by a unique pre‐edge feature

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
152 Y. ARAI AND D. L. SPARKS

near 3 eV (relative energy) that increased in intensity with increasing


mineral crystallinity and was very weak for phosphate adsorbed on goethite
and aluminum phosphate minerals. Due to spectra resemblances of distinct
postedge features, the authors suggested that Ca‐phosphate minerals such as
apatite might be forming in the soils.
Sato et al. (2005) have used a similar approach to understand the
P speciation in silt‐clay fractions of high‐P soils from Southern New York.
On the basis of the spectral similarities between unknown samples and Ca
and Fe reference compounds [e.g., CaHPO4, CaHPO4  2H2O, Ca3(PO4)2,
Ca5(PO4)3OH, and FePO4  2H2O], they suggested that: (1) unamended for-
est soils contain iron‐associated P species, (2) soils with short‐term manure‐
amended soils contain both iron and Ca‐associated P species, and (3) long‐
term manure‐amended soils contain predominantly Ca‐associated P.
Beauchemin et al. (2003) have combined linear combination (LC) XANES
fit analyses with principal component analyses (PCA) to better understand P
speciation in Canadian soils which had slightly acidic and alkaline soil pH
values (5.5–7.6) (Beauchemin et al., 2003). They found sorbed P on Fe and Al
oxides in all the soils. While the presence of hydroxylapatite was suggested in
two slightly alkaline pH soils, Ca‐associated P species were predominant in an
acidic soil. They also reported amorphous iron phosphate minerals in acidic
soils of the A horizon.
While these studies (Beauchemin et al., 2003; Hesterberg et al., 1999; Sato
et al., 2005) showed a useful application of P K‐edge XANES techniques to
understand P speciation in bulk soils and soil silt‐clay fractions, the data
might have been misinterpreted due to self‐absorption eVects on fluorescence
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

yield XANES spectra of reference compounds. To eliminate the self absorp-


tion eVect of concentrated P materials, Toor et al. (2006) have demonstrated
the use of total electron yield XANES data to appropriately understand the
P speciation in biosolids (Toor et al., 2006).
The direct P speciation of soils using LC XANES fit and PCA analyses must
be carefully performed since the results can be misinterpreted due to limitations
using these analyses. Beauchem et al. (2003) pointed out that the LC XANES
analyses of P K‐edge XANES data can be inherently restricted by: (1) the
data quality and (2) correct choice of appropriate standards. Furthermore,
several researchers have indicated that PCA of X‐ray absorption spectra are
not sensitive enough to pick up subtle diVerences in spectral features (e.g.,
diVerences in ligand coordination) (Arai et al., 2006; Beauchemin et al.,
2003). Therefore, the number of significant components suggested by indicator
functions must be only used as a guide rather than drawing conclusions.
In more recent studies, several researchers have used P K‐edge XANES
techniques to understand P adsorption mechanisms at the metal–oxyhydroxide
interface. Khare et al. (2005) investigated the P adsorption mechanisms in
binary adsorbent systems (e.g., mixture of iron oxyhydroxides and aluminum

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 153

oxide) using P K‐edge XANES analyses. Changes in the full width at half‐
maximum height in the white line peak were used to distinguish diVerent
binding mechanisms. While P precipitation was observed in an Al oxide
single‐mineral system as well as a goethite/bohemite binary system, only
adsorption was observed in ferrihydrite and goethite single‐mineral systems.
Arai and Sparks recently conducted fluorescence yield P K‐edge XANES
measurements on diVerent P salts, minerals, and P‐reacted ferrihydrite (sus-
pension density, 1.25 g liter1; I, 0.1‐M NaCl; and initial P concentrations,
1 mM). In Fig. 3A–D, florescence yield XANES spectra of P minerals/salts,
P‐reacted ferrihydrite, and PO4(aq) samples are shown. P K‐edge XANES
spectra of reference iron phosphate minerals, strengite (FeIII(PO4)2H2O),
barbosalite (FeII FeIII2 ðPO4 Þ2 ðOHÞ2 ), and rockbridgeite (ðFe ; Mn ÞFe4
II II III

ðPO4 Þ3 ðOHÞ5 ) are compared in Fig. 3A. Strong white line peaks can be
observed in all spectra as the result of a P 1s electron transition into an
unoccupied valence state of PO4 sp3 hybridized orbitals. Distinctive post-
edge resonance features, at 5–25 eV relative energy, are similar in crystalline
and amorphous strengite but barbosalite and rockbridgeite show diVerent
features. The relative energy regions of pre‐edge doublet features (indicated
by dotted lines between 6 and 3 eV) are similar in strengite, barbosalite,
and rockbridgeite. The pre‐edge region of P K‐edge XANES spectra includes
a sharp white line peak resulting from electronic transitions of the core
electronic states in the conduction band (Behrens, 1992; Fendorf and Sparks,
1996). The electronegativity, number of nearest neighbors, and coordination
environment (i.e., molecular symmetry) of the absorbing atoms could influ-
ence the intensity and the position of the pre‐edge features (Behrens, 1992;
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

Behrens et al., 1991; Wong et al., 1984). On the basis of a P XANES study on
several transition metal phosphates by Okude et al. (1999), the pre‐edge
features in an Fe(III)‐P salt [i.e., Fe4(PO4)3(OH)3] were attributed to interac-
tions between P 3p states [i.e., sp3 hybridization of the phosphate (PO4)
tetrahedral molecule] and 3d5 electronic state of Fe(III). It is also important
to note that the pre‐edge feature is not present when there is no direct
interaction between P tetrahedra and Fe(III) octahedra [i.e., HPO2 4 (aq) spec-
trum in Fig. 3A]. The position and intensity of the pre‐edge features also depend
on the number of d‐electrons in the transition metal associated with phosphate
(Okude et al., 1999). The eVects of d0–d8 electronic state of metal phosphate
salts on the P XANES pre‐edge features were re‐investigated after the study by
Okude et al. (1999), and some of them are reproduced, along with new results
(Fig. 3B). As previously reported by Okude et al. (1999), the pre‐edge intensity
decreases with increasing number of d‐electrons, and the position (indicated
by solid lines in Fig. 3B) shifts to a lower energy, with a gap between d5 and
d6 electronic states (Fig. 3B). Moreover, the peak maxima of the white line shift
to higher energy with decreasing number of d‐electrons. Although vivianite
[i.e., a predominantly d6 Fe(II)‐based P mineral] shows a subtle pre‐edge

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
154 Y. ARAI AND D. L. SPARKS

A B

d0, CaHPO4
Normalized intensity

Normalized intensity
d4, CrPO4
2−
HPO4 (aq)
d5, Strengite
Rockbrigeite (FeIIIPO4·2H2O)
d6, Vivianite
Barbosalite (Fe3II[PO4]2·8H2O)

d7, Co3(PO4)2
Crystalline strengite
d8, Ni3(PO4)2
Amorphous strengite
d10, Zn3(PO4)2

−10 0 10 20 30 2140 2150 2160 2170 2180


Relative energy (eV) Energy (eV)

C D
Normalized intensity

Normalized intensity

11 months 12 months
1.5 days
1.5 days
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

6h 6h

1h 1h

5 min 5 min

−10 0 10 20 30 40 50 −10 0 10 20 30 40 50
Relative energy (eV) Relative energy (eV)

Figure 3 (A) P K‐edge XANES spectra of synthetic amorphous strengite (FeIIIPO4 2H2O),
crystalline strengite (FeIIIPO4 2H2O), barbosalite (FeII FeIII
2 ðOHÞ2 ðPO4 Þ2 ), rockbrigite ((Fe ,
II
II III 2
Mn )Fe 4 (PO4)3(OH)5), and HPO4 (aq). Relative energies are presented with respect to the
absorption edge energy position of ferric phosphate minerals, (B) P K‐edge XANES spectra of
transition metal phosphate salts; dx (x, number of d electrons of the predominant transition
metal). Solid lines and dotted lines are pre‐edge and white line peak maxima, respectively
(modified after Okude et al., 1999), (C) P K‐edge XANES spectra of P‐reacted ferrihydrite at
pH 4.0 as a function of residence time, and (D) P K‐edge XANES spectra of P‐reacted
ferrihydrite at pH 7.5 as a function of residence time.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 155

feature at a similar position (2150 eV) due to a small amount of d5 Fe(III)


impurity, it does not contain a strong pre‐edge feature that is observed in
amorphous strengite [i.e., d5 Fe(III)‐P mineral].
In a similar tetrahedral oxoanion (i.e., sulfate) XANES study, Myneni
(2000) compared the energy position of the pre‐edge feature in aqueous sulfate
and Fe(III)‐SO4 salts [i.e., coquimbite, (Fe,Al)2(SO4)3 9H2O; copiapite,
MgFe4(OH)2(SO4)6 10H2O; jarosite, KFe3(SO4)2(OH)6; ferric ammonium sul-
fate, FeNH4(SO4)2 12H2O] that contain diVerent SO4‐Fe(III) coordination
environments (bidentate binuclear linkages in coquimbite and copiapite, tri-
dentate linkages in jarosite, and hydrogen bonds in ferric ammonium sulfate).
These data indicated that the energy position of the pre‐edge feature in these
Fe(III)‐SO4 salts depends on the number of Fe(III) polyhedra connected to
each sulfate polyhedron surface, and the pre‐edge feature that was absent in the
samples had no direct interaction between Fe(III) polyhedra and sulfate (i.e.,
ferric ammonium sulfate and aqueous sulfate) (Myneni, 2000). In the case of
Fe(III)‐P salts/minerals, the relationships between specific Fe(III)‐O‐P coordi-
nation environments and pre‐edge features have not been well investigated.
Therefore, we further investigated the pre‐edge feature characteristics (i.e.,
shape and energy position) in natural/synthetic Fe(III)‐P minerals (barbosalite,
rockbridgeite, and strengite) containing diVerent P tetrahedral attachment on
iron octahedra (e.g., bidentate binuclear configuration) and aqueous phosphate
(i.e., HPO2
4 ).
Phosphate tetrahedra and Fe(III) octahedra coordination environments in
these minerals were previously studied by Rose et al. (1997). They described
the coordination environments using a notation (NXMY) based on the num-
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

ber of corners of the P tetrahedron in association with the Fe(III) octahedron


and the number of Fe atoms in the second coordination sphere. The notation
NXMY was used to describe the diVerent linkages: N is the number of corners
of the P tetrahedron involved in the linkage, X is the type of linkage between
the polyhedra (i.e., C for corner and E for edge), M stands for the number of
iron atoms in the second coordination sphere of P (for one linkage), and Y is
the type of linkage between iron octahedra (C for corner and F for face), when
more than one Fe(III) octahedra is in the second coordination sphere (Rose
et al., 1997). The linkages are 1C1 (monodentate mononuclear corner attach-
ment on an iron octahedron) for strengite, 1C2F (monodentate binuclear
corner attachment on the center of the face sharing two iron octahedra) for
both barbosalite and rockbridgeite, and 2C2C (bidentate binuclear corner
attachment on two corners sharing iron octahedra) for both barbosalite and
rockbridgeite (Rose et al., 1997).
While the aqueous phosphate spectrum shows no pre‐edge feature [i.e., no
Fe(III)‐O‐P linkages], the strong doublet pre‐edge features are consistently
present at a similar energy range (i.e., 3 to 6 eV) in all spectra regardless of
the diVerent types of Fe(III)‐O‐P linkages (i.e., 1C1, 1C2F, 2C2C) (Fig. 3A).

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
156 Y. ARAI AND D. L. SPARKS

Although the shape of the pre‐edge feature is slightly varied in these Fe(III)‐P
minerals, the presence of direct P‐O‐Fe(III) linkages is indicated in these
minerals. Unlike the Fe(III)‐SO4 reference salts used in Myneni’s S XANES
study (Myneni, 2000), our natural Fe(III)‐P minerals contain: (1) transition
metal impurities [i.e., d6 Fe(II) and d5 Mn(II)] coordinated with PO4 which
weaken the pre‐edge feature seen in Fe(III)‐P compounds and (2) mixed
PO4‐Fe(III) linkages in barbosalite and rockbrigite (i.e., 2C2C and 1C2F).
Therefore, it is diYcult to conclude that the shape, energy position,
and intensity of the pre‐edge features are strictly attributed to the specific
coordination environments between P tetrahedra and Fe(III) octahedra.
Although observed pre‐edge features do not provide clear evidence for
distinguishing monodentate mononuclear from bidentate binuclear linkages,
it is safe to say that the pre‐edge features observed in the Fe(III)‐P solids
arise from a direct interaction between P tetrahedra and Fe(III) octahedra,
and this feature can be used as a signature of covalent bonds between PO4
and iron(III) octahedra. Myneni previously used this feature as an indication
of inner‐sphere SO4 coordination environments on goethite surfaces
(Myneni, 2000). Electrophoretic mobility measurements are often used to
distinguish diVerent complexation mechanisms (i.e., inner‐sphere and outer‐
sphere complexes). However, charge reversal in the presence of oxyanions is
diYcult to interpret since formation of ternary complexes with inert back-
ground electrolyte ions and/or formation of surface precipitates could mask
the experimental results. The pre‐edge features observed in PO4/SO4 K‐edge
XANES measurements can be useful for identifying the inner‐sphere P/S
oxyanion coordination environments on the Fe(III) oxide surfaces.
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

In Fig. 3C and D, XANES spectra of adsorption kinetic samples at pH 4 and


7.5 and H2 PO 2
4 (aq) and HPO4 (aq) samples are shown. The distinctive post-
edge resonance features seen in the Fe(III)‐P reference minerals at 0–30 eV
(Fig. 3A) are absent in all of the sorption kinetic samples (Fig. 3C and D),
indicating no predominant fraction of Fe(III)‐P precipitates. This might be due
to deconstructive interferences of the outgoing multiple scattered photoelec-
tron wave function from predominant two‐dimensional adsorption complexes
which overwhelm weak resonance features of Fe(III)‐P precipitates, if any
form. Several researchers have previously suggested the formation of iron
phosphate precipitates at mineral surfaces (i.e., ferrihydrite, goethite, and
hematite) on P sorption using various ex situ spectroscopic techniques (i.e.,
Auger, DRIFT, XRD, XPS, SIMS, and TEM) (Johansson et al., 1998; Martine
and Smart, 1987; Martine et al., 1988; McCammon and Burn, 1980; Nanzyo,
1986). However, there was no evidence for the formation of ferric phosphate
precipitates under our in situ experimental reaction conditions. Saturation
indices (a maximum 0.932) estimated using the equilibrium constant of
synthetic strengite (Nriagu, 1972) and dissolved total Fe concentrations also
indicate that all systems were undersaturated with respect to synthetic strengite.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 157

Compared to the aqueous PO4(aq) spectra, weak pre‐edge/shoulder fea-


tures (indicated by two dotted lines between 6 and 3 eV) appear after
5 min in the P adsorption samples at both pH values (Fig. 3C and D). It seems
that the pre‐edge features become more pronounced with increasing reaction
times at both pH values (indicated by arrows in Fig. 3C and D). These peak
positions are similar to those in strengite, indicating the formation of inner‐
sphere P tetrahedral linkages on the Fe(III) octahedral structures. This is in
good agreement with previous results of ATR‐FTIR analyses and electropho-
retic mobility measurements (Arai and Sparks, 2001). In Fig. 4, the first
derivative of selected spectra from Fig. 3A–C are summarized. In this figure,
one can clearly see that the pre‐edge features in the adsorption samples become
more pronounced with increasing aging time from 1.5 days to >11 months at
both pH values, and the features in aged adsorption samples (indicated by open
circles) become near doublets which are also observed in amorphous and/or
crystalline strengite, barbosalite, and rockbrigeite (indicated by an arrow). This
suggests that similar inner‐sphere P‐O‐Fe(III) linkages are present in 11‐
months‐aged adsorption samples at pH 4 and 7.5. The pre‐edge features
intensify with aging from 1.5 days to 11 months at both pH values. Since
XANES spectra are normalized with respect to per P atom, the intensified

Crystalline strengite
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

Amorphous strengite
First derivative intensity

pH 4, 11 months

pH 4, 1.5 days

pH 4, 1 h

pH 7.5, 12 months

pH 7.5, 1.5 days

pH 7.5, 1 h

0 10 20 30 40 50
Relative energy (eV)

Figure 4 First derivative of selected P K‐edge XANES spectra from Fig. 3A, C, and D.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
158 Y. ARAI AND D. L. SPARKS

features cannot be explained by an increase in P loading levels. An increase


in the amount of inner‐sphere P‐O‐Fe(III) linkages seems to be the only
explanation to support the pronounced pre‐edge feature. Interestingly, the P
loading levels remained nearly equal (2.73  0.03 mmol m2) after 1.5 days in
the pH 4 system. The lack of a pre‐edge feature in the 1.5‐day sample compared
to the 11‐month sample is probably attributed to the presence of secondary
species (e.g., diVused ions), and these complexes were gradually converted into
more stable inner‐sphere species after 11 months. Similar reaction mechanisms
can be also suggested in the pH 7.5 system. The intensified pre‐edge features
are due to changes in surface speciation [i.e., an increase in inner‐sphere
P‐O‐Fe(III) linkages]. Two possible mechanisms are postulated to explain the
pre‐edge feature in aged samples.
First, an increase in inner‐sphere Fe(III)‐O‐P linkages can be supported by
the formation of PO4‐bridged ferric polymer complexes. The diVused P ions
within ferrihydrite aggregates could possibly react with unreacted ferrihydrite
particles to form PO4‐bridged ferrihydrite polymers. In a similar tetrahedral
oxianion [i.e., arsenate (As(V))] adsorption study on ferrihydrite, Waychunas
et al. (1993) observed large As(V)‐Fe coordination numbers (i.e., >3.02) in
As(V) sorbed ferrihydrite samples based on As and Fe K‐edge EXAFS ana-
lyses. They suggested that ferrihydrite might contain substantially smaller basic
crystalline units that are freely accessible by disordered aggregates as well as
dissolved arsenate anions as long as local geometry permits (Waychunas et al.,
1993). Since AsO4 is often used as an analogue of PO4 due to similar chemical
properties (e.g., solution speciation with respect to protonation constants),
it is possible that PO4 molecules could induce the formation of ferric polymer
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

chains. Since our experimental pH values (4 and 7.5) are below the PZSE of
ferrihydrite (8), the surface charge density of ferrihydrite particles not reacted
with P is positively charged. These positively charged particles could be
electrostatically attracted to P‐sorbed ferrihydrite surfaces which are negatively
charged due to specific P adsorption. Previous electrophoretic mobility
measurements have shown that charge reversal can occur via inner‐sphere P
surface complexation on ferrihydrite (Arai and Sparks, 2001). The particle
interaction could eventually lead to the formation of multi‐Fe(III) polymer
layers/clusters, entrapping/bridging P adsorption complexes within the struc-
ture. Interestingly, Anderson et al. (1985) earlier suggested a similar PO4‐
partitioning mechanism on Fe oxyhydroxide (i.e., goethite) based on XRD/
electron microscopic experimental data (Anderson et al., 1985). An aggregate
order of goethite particles significantly increased after goethite particles were
reacted with phosphate, and isotopically exchangeable P was decreased with
increasing P uptake (Anderson et al., 1985). In aged P‐reacted ferrihydrite
surfaces, similar P‐partitioning mechanisms might have been occurred.
Second, ferric phosphate surface precipitation mechanisms on the ferri-
hydrite surface could also result in an enhancement of pre‐edge features

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 159

[i.e., an increase in Fe(III)‐PO4 linkages]. Ler and Stanforth (2003) reported


that the Zeta (x) potential of P sorbed on a goethite surface at pH 5 with time
reverted to that observed for goethite surfaces without P (Ler and Stanforth,
2003). The long‐term reactions most likely involve the dissolution of ferric
ions from goethite surfaces and the subsequent reaction between ferric (aq)
and surface‐bound P. In our P–ferrihydrite system, it is possible that dissolved
ferric ions are interacting with diVused P ions/surface‐bound P, resulting in
the formation of ferric phosphate surface precipitates on ferrihydrite surfaces
with increasing time.

V. RESIDENCE TIME EFFECTS ON PHOSPHATE


ADSORPTION AND DESORPTION IN SOILS AND
SOIL COMPONENTS

A. RESIDENCE TIME EFFECTS THEORY

In environmental settings, it is important to consider the residence time


(aging) eVect on contaminant bioavailability, transport, and remediation.
Soils and sediments are nearly always at disequilibrium with respect to ion
transformations (Sparks, 1987). The rate of bioavailability can be reduced or
increased in the environment with increasing time (Pignatello and Xing, 1995).
The residence time eVect has also been described by irreversible, hysteretic, and
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

nonsingular reactions, and can be suggested by the observation of two slow


reaction processes: (1) a continuous slow adsorption of the adsorptive onto the
adsorbent with increasing time and (2) a slow desorption of the adsorbate from
the adsorbent with increases in time.
A slow adsorption process usually occurs after an establishment of quasi‐
equilibrium following an initial rapid reaction for a few hours. In thermody-
namics, the definition of slow desorption is that it is diYcult for the deso-
rptive to overcome the total activation energy created during adsorption.
In other words, the activation energy of desorption is less than the sum of
the activation energy for adsorption and the energy formed during adsorp-
tion [Edesorption Eadsorption þ DH, where Edesorption ¼ activation energy for
desorption, Eadsorption ¼ activation energy for adsorption (0), and DH ¼
energy formed during adsorption]. Formation of inner‐sphere complexes via
ligand exchange and/or the transformation of amorphous to crystalline
materials generate a greater DH, increasing the value of the total activation
energy. Slow desorption phenomena are attributed not only to chemical
factors (e.g., ligand exchange and chemisorption) but also to physical factors
(diVusion). Ions trapped in mesopores (interparticle) between aggregates

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
160 Y. ARAI AND D. L. SPARKS

and micropores (intraparticles) within an individual particle fissure are


diYcult to extract and/or desorb with any desorptive.

B. SLOW ADSORPTION

Continuous slow adsorption of P on soils and soil components [i.e.,


amorphous Al(OH)3, natural allophane, ferrihydrite, hematite, goethite,
a‐Al2O3, and kaolinite] over diVerent timescales (hours to months) was
reported by many researchers (Anderson et al., 1976; Barrón and Torrent,
1995; Barrow, 1974, 1985; Beek and van Riemsdijk, 1982; Black, 1942;
Colemann, 1944; Edzwald et al., 1976; Fuller et al., 1993; Hingston et al.,
1974; Hsu and Rennie, 1962; Kafkafi et al., 1967; Madrid and Posner, 1979;
Okajima et al., 1983; Parfitt, 1979, 1989; Ryden and Syers, 1977; van
Riemsdijk et al., 1977; Willett et al., 1988).
While some studies documented continuous slow P adsorption on phyl-
losilicate and aluminum and iron oxyhydroxides over timescales of hours,
others carried out the experiments up to days–months on soils and soil
components. Edzwald et al. (1976) observed rapid P adsorption on kaolinite,
montmorillonite, and illite at pH 7–8 during the initial 4 h, and this was
followed by a slow continuous uptake after 24–72 h (Edzwald et al., 1976).
Madrid and Posner (1979) also documented slow P uptake on goethite after
1–24 h at pH 4.25–10.25. Ryden and Syers (Ryden and Syers, 1977; Ryden
et al., 1973, 1977a) studied long‐term P adsorption on natural and synthetic
goethite, hydrous ferric oxides and four New Zealand soils (Egmontblack
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

loam, Porirua fine sandy loam, Okaihau gravelly clay, and Waikakahi silit
loam). In all adsorbents, an initial fast reaction was completed within 20 h,
followed by slow P adsorption after 192 h. Hsu and Rennie (1962) investi-
gated P sorption on amorphous aluminum hydroxides at pH 3.8, 4.2, and 7.
While short‐term experiments showed only an initial rapid adsorption within
24 h at pH 4.2 and 7, continuous slow adsorption was observed after 528 h
at pH 3.8 in the long‐term experiments.
Black (1942) carried out P adsorption experiments on a Cecil clay for
days. The slow uptake continued from 48 h to 30 days over a wide range of
pH values (2.5 to 8). An increase in P adsorption with time at acidic pH
values was much greater than at alkaline pH values. van Riemsijk et al.
(1977) investigated P adsorption on aluminum oxide and a‐Al2O3 at pH 5
and 6 using inorganic synthetic sewage water. In both solids, an initial fast
adsorption was completed within 1–4 days, and this was followed by a
slow adsorption up to 70 days. Haseman et al. (1950) studied long‐term P
sorption on clay minerals (montmorillonite, illite, and kaolinite) at pH 3–7.
While slow adsorption was observed for all clay minerals up to 200–300 h,
long‐term experiments on gibbsite and goethite showed a slow continuous

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 161

uptake even after 1000 h at pH 3–5. Willet et al. (1988) have documented
long‐term slow P sorption on ferrihydrite aggregates after 90 days at pH 4.
The P migration into the internal adsorption sites of ferrihydrite aggregates
was evident in electron microprobe analyses. Arai and Sparks have recently
investigated the eVect of initial P concentrations (0.6–2.6 mM) on P adsorp-
tion kinetics on ferrihydrite surfaces (suspension density, 0.5 g liter1; I,
0.1‐M NaCl and pH 4 and 7.5). They observed fast adsorption within the
initial 4 h under all reaction conditions (insets of Fig. 5A and B), and the
rapid adsorption was followed by slow continuous adsorption after 650 h.
At steady state (670 h), total P adsorption was consistently greater in the
pH 4 systems than in the pH 7.5 systems at respective initial P concentrations
(Fig. 5A and B).

C. SLOW DESORPTION PROCESS AND HYSTERESIS

Many researchers utilizing batch techniques (e.g., replenishment techni-


ques) to observe short‐term (<24 h) P desorption processes from soils and
soil components have reported that the process was often biphasic (a fast
initial desorption followed by slow desorption). Two studies examined P
desorption (>120 h) on synthetic goethite and ferrihydrite using batch
replenishment techniques (diluting the reacted samples at a constant I )
(Madrid and Posner, 1979; Ryden and Syers, 1977). Desorption phenomena
were biphasic, a fast reaction for a few hours was followed by a slow reaction
for over 100 h (Madrid and Posner, 1979; Ryden and Syers, 1977). Biphasic
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

P desorption phenomena were also observed for kaolinite using a similar


batch technique (Bar‐Yosef and Kafkafi, 1978). Overall, the desorption
process was divided into a rapid and a slow first‐order reaction, and the
rate coeYcient for the rapid reaction was about four times higher than the
slower reaction.
The traditional batch technique is not usually suitable to investigate
desorption processes due to: (1) possible readsorption of reaction products
due to their accumulation in the bulk solution, (2) pH fluctuation during
desorption, and (3) establishment of a diVerent equilibrium after each re-
plenishment step. Flow systems and batch techniques (e.g., an ion exchange
resin) that employ a ‘‘sink’’ for the desorbed species prevent the build up of
reaction products in the bulk solution, and are preferable kinetic methods
for studying apparent desorption behavior for long‐time periods (>24 h).
Phosphate release from P‐reacted iron oxyhydroxide (i.e., ferrihydrite,
goethite, and hematite) coated silica was investigated using a flow‐through
method (Freese et al., 1999). After 4400 min of desorption at pH 4, the
relative amount of desorbed P ¼ Pdesorb/Psorb was found to be of the order of
ferrihydrite < goethite < hematite.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
162 Y. ARAI AND D. L. SPARKS

A
120,000
100,000

80,000
100,000
Total P adsorbed (mg kg−1)

60,000

40,000
80,000
20,000
0 08 16 24

60,000
2.6 mM

40,000 1.3 mM
0.6 mM

20,000
0 200 400 600 800 1000 1200
Time (h)

B
90,000
70,000

60,000
80,000
50,000
Total P adsorbed (mg kg−1)

70,000 40,000

30,000

60,000 20,000

10,000
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

50,000 0 8 16 24

40,000

30,000

20,000

10,000
0 200 400 600 800 1000 1200
Time (h)

Figure 5 P adsorption kinetics (suspension density, 0.5 g liter1 and I, 0.1‐M NaCl) at the
ferrihydrite–water interface at (A) pH 4 and (B) pH 7.5 as a function of initial P concentrations
(0.6, 1.3, and 2.6 mM). The adsorption reactions up to 24 h are magnified in insets.

Several kinetic models have been used to successfully describe biphasic


P desorption processes from soils and soil components. These include zero‐
order (Onken and Matheson, 1982), first‐order (GriYn and Jurinak, 1973),
second‐order (Kuo and Lotse, 1973), third‐order (Onken and Matheson, 1982),

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 163

parabolic diVusion (Evans and Jurinak, 1976), and Elovich (Atkinson et al.,
1970; Chen, 1977; Chen and Clayton, 1980). One of the reasons a single‐
kinetic model is highly applicable to describe biphasic reactions is that
experimental conditions (i.e., the length of experimental periods associated
with disequilibrium) are often appropriate for the assumptions of the model
(Sparks, 1989). Aharoni and Suzin (1982) used heterogeneous diVusion
models to describe segments of kinetic processes (initial, intermediate, and
long term). These models were approximated by a sequence of parabolic,
Elovich, and exponential equations.
In some of early P kinetics research, Australian researchers compared
P adsorption and desorption processes on soils and soil components and
investigated the hysteresis eVect, in which an adsorption isotherm curve does
not coincide with a desorption isotherm curve. Using a batch equilibrium
study (<24 h), Barrow (1983b) showed the slight shifts between adsorp-
tion and desorption isotherm curves on an Australian sandy loam. Ryden
and Syers (1977) utilized the same experimental approaches to compare the
P hysteresis eVect on soils and ferrihydrite. Desorption isotherm curves on
the soils and ferrihydrite were greatly shifted from adsorption isotherm
curves, and ferrihydrite showed greater irreversibility than the soils (Ryden
and Syers, 1977). The concentration eVect on hysteresis was also investi-
gated. A wide range of initial P ([P]i ¼ 0.001–1 mM) was reacted with
kaolinite, and then the desorption was investigated using isotopic exchange
(Kafkafi et al., 1967). A hysteresis eVect was observed over all concentration
ranges.
Incubation time highly influences the reversibility of adsorbed P from soils
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

and soil components. Barrow reported that P desorption from Australian


soils was rapid when aging time was short (<24 h), but the desorption process
became much slower when aging time was longer (>24 h) (Barrow, 1979a;
Barrow and Shaw, 1975). Madrid observed similar irreversible P desorption
from goethite (Madrid and Posner, 1979). P desorption decreased with in-
creasing aging time from 1 to 23 h. A similar study was also conducted on a
synthetic goethite and Australian clay loam using longer desorption experi-
ments (>96 h) and varying incubation time. The irreversibility was enhanced
with increasing aging time (Barrow, 1979a; Madrid and Posner, 1979). The
irreversibility could also occur within a short adsorption reaction (<15 s).
In our recent investigation, we have observed similar aging eVects on long‐
term (up to 19 months)‐aged P‐sorbed ferrihydrite at pH 4 and 7.5 (suspen-
sion density, 1 g liter1; I, 0.1‐M NaCl; and initial P concentrations, 1 mM for
the pH 4 system and 0.6 mM for the pH 7.5 system). Two diVerent initial P
concentrations were chosen to achieve nearly 100% adsorption within 2 days,
so that the loading levels for all aged samples were approximately equal at
each pH value (2.7 mmol m2 at pH 4 and 1.62 mmol m2 at pH 7.5) prior
to each desorption experiment. It is important to mention that ferrihydrite

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
164 Y. ARAI AND D. L. SPARKS

A
2000
1800
Total P desorbed (mg kg−1)

1600 2 days
1400 1 month
9 months
1200 19 months
1000
800
600
400
200
0
0 5 10 15 20
Time (h)
B
1600

1400
Total P desorbed (mg kg−1)

1200
2 days
1000 1 month
4 months
800 10 months

600
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

400

200

0
0 5 10 15 20
Time (h)

Figure 6 Residence time eVects on P desorption kinetics from ferrihydrite at pH 4.0 (A) and
pH 7.5 (B). Incubation times of P‐reacted ferrihydrite suspensions prior to desorption experi-
ments are described in legends (suspension density, 1 g liter1; I, 0.1‐M NaCl; and initial
P concentrations, 1 mM for the pH 4 system and 0.6 mM for the pH 7.5 system).

transformation into crystalline phases such as goethite and hematite was not
observed in 1‐year‐aged samples at both pH values via conventional bulk
XRD analyses. Figure 6A and B shows short‐term (24 h) P desorption data at
pH 4 and 7.5, respectively. The stirred flow method was used so that the
adsorbents are exposed to a greater mass of ions than a static batch system,

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 165

and the flowing solution continuously removes reaction products (i.e., desorbed
species) (Sparks, 1989).
Biphasic P desorption reactions were observed at both pH values (Fig. 6A
and B). At pH 4, the initial fast reaction occurred during the first 3–5 h, and
slow reactions followed up to 24 h, and total P desorption decreased from
4.81% to 0.92% with increasing aging time from 2 days to 10 months,
suggesting a residence time eVect. Similarly, biphasic desorption behavior
was observed at pH 7.5 (Fig. 6B). Total desorbable P decreased from 10.07%
to 3.83% with increasing aging time from 2 days to 19 months at pH 7.5. It is
interesting that total desorbable P is much less at pH 4 even though the
initial loading levels at pH 4 are much greater than those at pH 7.5.
The mechanisms responsible for the residence time eVect are not clearly
defined; however, it has been ascribed to: (1) solid‐state diVusion, or intra‐
and interparticle diVusion (Bar‐Yosef and Kafkafi, 1978; Barrow, 1983a;
Bolan et al., 1985; Cabrea et al., 1981; Enfield et al., 1981; Evans and
Jurinak, 1976; Parfitt, 1989; Ryden et al., 1977a; Torrent et al., 1992; van
Riemsdijk et al., 1984; Willett et al., 1988), (2) surface precipitation (e.g., Ler
and Stanforth, 2003), and (3) higher energy binding on surface structures via
chemical reconfiguration (Beek and van Riemsdijk, 1982; Fendorf et al.,
1992; Pignatello and Xing, 1995).

D. SOLID‐STATE, INTER‐, AND INTRAPARTICLE DIFFUSION

In situ soil chemical reactions between ions and solids are ultimately limited
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

by the process of diVusion (McBride, 1994). For example, if the initial oxya-
nion reaction on soil materials is completed rapidly, then the unreacted
oxyanions build up in solution until the rate of diVusion of oxyanions in the
adsorbents equals the overall reaction rate. Fast chemisorption (e.g., the
ligand exchange reaction) takes place at external sites where ions in the bulk
fluid can easily access the sites. The slow adsorption processes occur at sites
within soil particle aggregates, and are often controlled by diVusion processes.
Four diVerent diVusion processes have been proposed: film, mesopore
(2 nm), micropore (<2 nm), and solid‐state diVusion (Pignatello and Xing,
1995). Film diVusion involves the transport of an ion or molecule through a
boundary layer or film (water molecules) that surrounds the particle surface
(Sparks, 1995a). Mesopore (interparticle) diVusion takes place between
aggregates. Micropore (intraparticles) diVusion involves ion penetration in
an individual particle fissure. Solid‐state diVusion involves diVusion in the
solid, and it can be a rate‐limiting step for all diVusion processes. The rate of
these diVusion processes is dependent on the size and chemical properties of
the diVusant and the size, shape, tortuosity, and discontinuity of the particle
pores (Pignatello and Xing, 1995).

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
166 Y. ARAI AND D. L. SPARKS

One way to investigate the diVusion process is to model slow adsorption


processes using models accounting for diVusion reactions. Several modeling
studies have indirectly suggested P solid‐state mesopore and micropore diVu-
sion processes in natural materials (Barrow, 1983a; Enfield et al., 1981; van
Riemsdijk et al., 1984). In an early P adsorption‐modeling study, Barrow
(1983a) developed a mechanistic model to describe slow P adsorption processes
on soils. The model had three components: (1) reaction between phosphate ions
and variable charge mineral surfaces, (2) assumption that soils consisted of
an assemblage of elements, each with diVerent values for the binding constant
and/or initial electrostatic potential, and (3) the assumption that the solid‐state
diVusion processes are driven by the gradients of electrical and chemical
potentials at the adsorbent surface (Manning, 1968; Rickert, 1982). As a result,
the model closely described P adsorption processes as influenced by concentra-
tion, pH, temperature, and reaction time. It was suggested that long‐term (90
days) P adsorption is predominantly attributed to P penetration into the soil
particles. In a later study, Bolan et al. (1985) utilized Barrow’s mechanistic
model to describe slow P adsorption processes on soil components (i.e., amor-
phous iron and aluminum oxides). Models that fit the data well indirectly
suggested that slow uptake of P was caused by redistribution of P into the
interior of amorphous metal oxide particles via solid‐state diVusion. A similar
model successfully described continuous P sorption on a ferric hydroxide gel
(Ryden et al., 1977a). While Barrow’s mechanistic model has been extensively
used to describe oxyanion adsorption phenomena, other models have also been
utilized to investigate slow adsorption processes.
While these modeling approaches indirectly suggest a diVusion‐controlled
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

reaction, they do not provide any information on what soil physical proper-
ties (e.g., crystallinity of adsorbents) are associated with slow adsorption.
Several researchers have pointed out that slow adsorption processes could be
aVected by crystallinity and particle morphology.
Phosphate adsorption on goethite and lepidocrocite was compared using a
batch adsorption technique (Cabrea et al., 1981). Lepidocrocite exhibited
more pronounced continuous slow sorption than goethite, and adsorbed P
was more irreversible in lepidocrocite. Cabrea et al. (1981) suggested that the
enhanced slow reaction and irreversibility were attributed to P micropore
diVusion into the less crystalline nature of lepidocrocite particles. Parfitt
(1989) also indirectly suggested that P particle diVusion was caused by poor
crystallinity and the greater porosity of metal oxides. A comparison of 1‐ and
30‐day P sorption isotherm experiments showed that slow P sorption was
observed in only amorphous goethite and ferrihydrite and not in highly
crystalline goethite (Parfitt, 1989). Later, Torrent et al. (1992) conducted
experiments that supported the above findings. The slow P adsorption on
10 goethite rich soils was described using a modified Frendlich equation
including a rate term (Torrent et al., 1992). The study showed that the extent

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 167

of the slow continuous reaction was correlated to the ratio between micropore
surface area and total surface area, as well as impurities due to amorphous
Fe oxides (Feox). Willet et al. (1988) found: (1) long‐term (>200 days) adsorp-
tion and (2) rapid (<24 h) migration of phosphate (mesopore diVusion) into
the aggregated particles of ferrihydrite using electron microprobe analyses.
Long‐term P adsorption kinetics at pH 4 exhibited an initial fast reaction
followed by slow sorption up to 260 days, and a maximum sorption capacity
of 1.2 mM g–1. The distribution of P moved toward the core of ferrihydrite
aggregate particles as time increased from 1 h to 1 day.
Electron microscopic studies have also shown that the crystalline
morphologies of goethite are related to P irreversibility (Torrent et al.,
1990). After sequential acid (5‐M HCl and 0.5‐M H2SO4)/base (0.1‐M
KOH) extractions of adsorbed P from 31 diVerent goethites, the amount
of nonextractable P (P was adsorbed at pH 6 for 24 h) was correlated to the
observation of solid morphologies [thin, multidomainic laths (V‐shaped
interdomainic grooves), and slit‐shaped macropores of multidomainic laths].
DiVusion‐controlled reactions can also be suggested by a low energy of
activation (<42 kJ mol1), Ea (Sparks, 1989). The results from a temperature‐
dependent sorption/desorption study can be applied to the Arrhenius equation
(k ¼ AeEa =RT ) to obtain Ea (Sparks, 1995a). The Ea of the slow reaction
between phosphate and diVerent morphologies of goethite was investigated
(Torrent, 1991). The slow sorption was described by a modified Frendlich
equation including time and activation energy (Ea) terms. Activation energy
values ranged from 38 to 80 kJ mol1, suggesting that the P sorption reaction
might involve a penetration of P into the diVerent crystalline matrices.
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

E. SURFACE PRECIPITATION

Whereas an adsorption complex has a two‐dimensional surface structure,


surface precipitates involve a three‐dimensional growth of the adsorbate on
the adsorptive surface. The magnitude of the precipitation growth is a func-
tion of the saturation of the adsorptive in the bulk fluid. When the adsorptive
is supersaturated with respect to the precipitate in the bulk fluid, a bulk
precipitate forms rapidly. However, when the system is even undersaturated
with respect to the bulk precipitate, precipitates could still form. Such a
precipitate is called a surface precipitate.
Surface precipitates are categorized into one of three groups based on
formation mechanisms. These are: (1) polymerization of the adsorptive at
the surface, (2) coprecipitation which forms between dissolved coions and
the adsorbate in the bulk solid, and (3) precipitates composed of ions from
the bulk fluid (e.g., hydrolysis products) (Chisholm‐Brause et al., 1990;

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
168 Y. ARAI AND D. L. SPARKS

Farley et al., 1985). Polymerization occurs as the result of continued chemi-


sorption. Newly created sites, by monolayer coverage of the adsorbate,
facilitates the formation of the precipitates (multilayer surface coverage).
Coprecipitation processes can also be facilitated by an increase in the surface
coverage of the adsorbent, and then the precipitation of adsorbate ions form
with dissolved coions of the solids (Sparks, 1995b). Several hypotheses have
been suggested to describe the formation of surface precipitates. The first
theory states that the dielectric constant of the solution near the surface is
less than that of the bulk fluid because the mineral surfaces are polar due to
the dipole moment of adsorbed water molecules (O’Day et al., 1994). Since
the activity of the ions in the bulk fluid is inversely proportional to the
dielectric constant of water, the lowered dielectric constant can facilitate the
formation of ion activity products.
The second theory states that the activity of the solid phase is less than unity
(Sposito, 1986). The activity of the pure solid (i.e., FePO4) can be assumed to
be 1; however, if a mixed solid (FexAl1xPO4) is formed, the activity is less than
1. FexAl1xPO4 will precipitate prior to FePO4. The following empirical equa-
tion describes formation of the surface precipitate, which is favored over the
pure solubility products:
ðIAPÞi ¼ gi Xi Kiso ð5Þ
where gi is the activity coeYcient of the solids i, Xi is the mole fraction of the
solid, and Kiso is the bulk precipitate of the pure minerals i (van Riemsdijk
and Lyklema, 1980).
The third theory states that sterically similar sites promote further nucle-
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

ation (McBride, 1991). Sterically similar sites indicate that the crystal lattice
has energy barriers which are reduced to facilitate nucleation processes.
While mechanisms and theories have been postulated for the formation of
surface precipitates, many researchers have reported indirect (macroscopic
data) and direct (spectroscopic data) evidence for aluminum‐ and iron‐P sur-
face precipitates on metal oxides and clays. Chen et al. (1973) indirectly sug-
gested a nucleation of a new P‐containing phase (i.e., P surface precipitate)
at the kaolinite–water interface. They observed a disequilibrium phenomenon
for the P adsorption envelope on kaolinite where the adsorption maximum
at pH 4 increased with increasing time from 1 day to 23 days when P in the
bulk solution was undersaturated with respect to aluminum phosphate. van
Riemsdijk and Lyklema (1980) also suggested that P adsorption on gibbsite
resulted in the formation of a potassium aluminum phosphate precipitate. This
was due to: (1) the increase in the rate of excess sorption (i.e., beyond monolayer
coverage) with increasing supersaturation, (2) enhanced P adsorption with
increasing potassium in the bulk solution, and (3) increased surface area after
P adsorption (van Riemsdijk and Lyklema, 1980). Ler and Stanforth (2003)
reported on changes in the Zeta (x)‐potential of P sorbed on goethite at pH 5

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 169

with time. The x values initially dropped due to chemisorption of P, but


gradually returned with time to those of goethite without sorbed P. They
suggested that the long‐term reactions most likely involved the dissolution of
ferric ions from goethite surfaces and the subsequent reaction between ferric
(aq) and surface‐bound P, resulting in the formation of ferric phosphate surface
precipitates.
Direct evidence for P surface precipitates has been suggested from micro-
scopic and spectroscopic investigations. Electron microscopy, electron diVrac-
tion, and XRD were utilized to investigate long‐term (40 days) P sorption on
amorphous Al(OH)3 and a‐Al2O3 at pH 5. Researchers observed the formation
of a new solid aluminum‐phosphate mineral (Sterrettite‐like) at aluminum
oxide surfaces (van Riemsdijk et al., 1977). Nanzyo (1984, 1986) investigated
P adsorption on amorphous aluminum and iron oxides over a wide pH range
(4–9) using diVuse reflectance infrared spectroscopy. He suggested that the
formation of amorphous aluminum phosphate‐like and amorphous iron
phosphate‐like products increased with decreasing pH based on the observa-
tion of similar IR spectra between the adsorption complex and synthetic
minerals (i.e., aluminum phosphate and iron phosphate). XPS, STEM, and
XRD investigations have shown that a mixture of goethite and tinticite
(Fe4(PO4)4(OH)6 7H2O) are formed after 18 days of P adsorption (25  C and
[P]i ¼ 0.01 mM liter1) at pH 3 (Jonasson et al., 1988). Multispectroscopic
techniques (Auger, XPS, scanning SIMS, and TEM) have been applied to
investigate P adsorption on natural goethite containing SiO2, Al2O3, and
Mn2O3 impurities ([P]i ¼ 1 mM, 90 days, and 60  C) (Martine et al., 1988).
The mineral griphite (Fe3Mn2(PO4)2.5(OH)2) was found in isolated crystallites.
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

Martine et al. suggested that P fixation in soils may be controlled by precipi-


tation reactions on Fe oxides. Similarly, Wang and Tzou (1995) observed
an Fe‐P precipitate‐like compound on a P‐reacted hematite surface. Ex situ
Mössbauer analysis showed quadrupole splitting for P‐adsorbed hematite
([P]i ¼ 155 mg g1, pH 3, 24 h, and 298 K). Such splitting resembled the
spectra observed for vivianite (Fe(II)3(PO4)2) and strengite (Fe(III)(PO4))
(McCammon and Burn, 1980; Wang, 1987). Since XRD analysis did not detect
the mineral phase, the mineral phase was presumed to be amorphous.
Slow P adsorption on allophane with continuous silicon release has been
indirectly linked to a precipitation reaction by many researchers. A reaction
with P caused a disruption in the allophane structure by displacing structural
silicon. Exposed reactive sites might facilitate P precipitation as aluminum
phosphate (Nanzyo, 1987; Parfitt, 1989; van Riemsdijk and Lyklema, 1980).
Induced silicon release during phosphate adsorption has also been reported
on soils and natural clay minerals (i.e., kaolinite, natural allophane, ferri-
hydrite, goethite, and allophanic clay) (Low and Black, 1950; Mattson and
Hester, 1935; Parfitt, 1989; Reifenberg and Buckwold, 1954). Rajan (1975)
observed silicon release during P adsorption on allophanic clay which was

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
170 Y. ARAI AND D. L. SPARKS

pretreated with selenite to remove labile silicate. P isomorphic substitution


of structural silicon in allophanic clay has been also suggested. Veith and
Sposito (1977) also reported the formation of aluminum phosphate surface
precipitates on allophatic materials after P adsorption. The formation of
X‐ray amorphous analogues of variscite and Na‐montebrasite were demon-
strated by reacting Al2O3 nH2O, synthetic allophanes and allophanic soils
with P at varying acidity. The formation of amorphous Al‐phosphates was
favored by both the large values of the hydration number in Al2O3 nH2O
and the increase in the acidity of the added P solution. The formation of
amorphous Al‐phosphate by secondary precipitates has been indicated by:
(1) an observed slow reaction rate between P and Al coatings, (2) the
immediate and significant increase of silicate in solution when aluminosili-
cate minerals were reacted, and (3) a significant amount (above monolayer
coverage) of P adsorption.
While numerous macroscopic, microscopic, and spectroscopic studies have
speculated on the formation of P surface precipitates at mineral surfaces, there
are some questions about these findings. These doubts have been raised
because: (1) macroscopic data do not directly suggest a mechanism for surface
precipitate formation, (2) high P concentration (>1 mM) might cause a super-
saturated condition with respect to bulk precipitates, and (3) ex situ conditions
in XRD, XPS, SEM, and TEM analyses might create artifacts with residual P
in the samples. Macroscopic and microscopic data must be carefully combined
with in situ spectroscopic techniques to draw better conclusions about the
formation of surface precipitates.
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

F. HIGHER ENERGY BINDING THROUGH CHEMICAL RECONFIGURATION

Real soil environments are not in equilibrium and always undergo slow
chemical changes to reach equilibrium (Koskinen and Harper, 1990; Steinfield
et al., 1989). As the materials become more stable (i.e., lowering solubility or
resulting multidentate complexes) with increasing aging time through transfor-
mations, the release of the adsorbate becomes slower, as evidenced by slow
desorption phenomena. Lowering the Gibbs free energy with increasing entropy
converts the products to more stable compounds. As a result, the activation
energy of the desorption phenomena can be expressed as Ed ¼ Ea þ Q.
This theory can be categorized into two groups: (1) a chemical transformation
from two‐dimensional surface complexes to three‐dimensional precipitates and
(2) a reconfiguration within two‐dimensional surface complexes.
Stumm and Morgan (1995a) suggested a chemical transformation theory,
the Oswald‐step rule. This theory proposes that the nucleation process
involves the formation of the least stable solid phase (highly soluble amor-
phous materials) first followed by the formation of a more stable solid phase
(insoluble crystalline materials).

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 171

The stability of the surface complexes can be altered from monodentate to


multidentate configuration with increasing time. The more the bonds that are
associated with the surface functional group of the solid, the greater the
stability. The chemical transformation from mononuclear to binuclear bridg-
ing complexation was suggested to describe the hysteresis phenomena for
P adsorption on soils and soil components (Kafkafi et al., 1967; Mengel,
1985); however, there is no molecular scale evidence to support this chemical
reconfiguration theory.

VI. FUTURE RESEARCH NEEDS

This chapter has described past and recent research eVorts to understand
P retention and release mechanisms in soils and soil minerals. Although
numerous P adsorption and desorption studies and several benchmark spec-
troscopic studies have provided vital information about P reactivity in soils
and sediments, uncertainties still persist in predicting the long‐term fate and
transport of P in soil–water environments. An historical understanding
of P reactivity on model adsorbents should lead to new investigations of:
(1) P reaction dynamics in more complex model systems (e.g., mixed adsor-
bents and adsorbates) and (2) P solid‐state speciation in high‐P‐containing
soils and sediments. Although several researchers have begun to investigate
the eVects of competitive ligands on P reactivity (Borggaard et al., 2005;
Geelhoed et al., 1997, 1998; Hawke et al., 1989; Hongshao and Stanforth,
2001; Johnson and Loeppert, 2006; Liu et al., 1999), our understanding of
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

P reactivity with respect to P speciation in natural materials is still limited to


operationally defined extractable fractions via chemical extractions. There are
not many studies characterizing P solid‐state speciation in natural materials.
Using modern microscopic and spectroscopic techniques (e.g., STEM,
synchrotron‐based X‐ray fluorescence spectroscopy and synchrotron‐based
XRD), P solid‐state speciation in soils and sediments can be better character-
ized, and these research findings will lead to a better understanding of the
P retention and release in heterogeneous soils and mixed‐model adsorbent
and ligand systems. Such comprehensive research results could be helpful
in designing more eVective in situ remediation technologies and nutrient
management programs to enhance environmental quality.

REFERENCES

Aharoni, C., and Suzin, Y. (1982). Application of the Elovich equation to the kinetics of
chemisorption. Part 3. Heterogeneous microporosity. J. Chem. Soc., Faraday Trans. 1 78,
2329–2336.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
172 Y. ARAI AND D. L. SPARKS

Anderson, M. A., and Malotky, D. T. (1979). The adsorption of protolyzable anions on hydrous
oxides at the isoelectric pH. J. Colloid Interface Sci. 72, 413–427.
Anderson, M. A., Ferguson, J. F., and Gavis, J. (1976). Arsenate adsorption on amorphous
aluminum hydroxide. J. Colloid Interface Sci. 54, 391–399.
Anderson, M. A., Tejedor‐Tejedor, M. I., and Stanforth, R. R. (1985). Influence of aggregation
on the uptake kinetics of phosphate by goethite. Environ. Sci. Technol. 19, 632–637.
Arai, Y., and Sparks, D. L. (2001). ATR‐FTIR spectroscopic investigation on phosphate adsorp-
tion mechanisms at the ferrihydrite‐water interface. J. Colloid Interface Sci. 241, 317–326.
Arai, Y., Livi, K. J. T., and Sparks, D. L. (2005). Phosphate reactivity in long‐term poultry
litter‐amended Southern Delaware sandy soils. Soil Sci. Soc. Am. J. 69, 616–629.
Arai, Y., McBeath, M., Bargar, J. R., Joye, J., and Davis, J. A. (2006). Uranyl adsorption and
surface speciation at the imogolite‐water interface: Self‐consistent spectroscopic and sur-
face complexation models. Geochim. Cosmochim. Acta 70, 2492–2509.
Atkinson, R. J., Hingston, F. J., Posner, A. M., and Quirk, J. P. (1970). Elovich equation fro the
kinetics of isotopic exchange reactions at isotopic exchange reactions at solid‐liquid inter-
faces. Nature 226, 148–149.
Atkinson, R. J., Parfitt, R. L., and Smart, R. S. C. (1974). Infrared study of phosphate
adsorption on goethite. J. Chem. Soc., Faraday Trans. 1 70, 1472–1479.
Barrón, E. A. V., and Torrent, J. (1995). Organic matter delays but does not prevent phosphate
sorption by cerrado soils from Brazil. Soil Sci. 159, 207–211.
Barrow, N. J. (1974). The slow reactions between soil and anions. 1. EVects of time, temperature,
and water content of a soil on the decrease in eVectiveness of phosphate for plant growth. Soil
Sci. 118, 380–386.
Barrow, N. J. (1979a). The description of desorption of phosphate from soil. J. Soil Sci. 30,
259–270.
Barrow, N. J. (1979b). Three eVects of temperature on the reactions between inorganic phos-
phate and soil. J. Soil Sci. 30, 271–279.
Barrow, N. J. (1980). DiVerences among some North American soils in the rate of reaction with
phosphate. J. Environ. Qual. 9, 644–648.
Barrow, N. J. (1983a). A mechanistic model for describing the sorption and desorption of
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

phosphate by soil. J. Soil Sci. 34, 733–750.


Barrow, N. J. (1983b). On the reversibility of phosphate sorption by soils. J. Soil Sci. 34, 751–758.
Barrow, N. J. (1985). Reactions of anions and cations with variable‐charge soils. In ‘‘Advances
in Agronomy’’ (N. C. Brady, Ed.), Vol. 38, pp. 183–230. Academic Press, Inc., New York.
Barrow, N. J. (1987). ‘‘Reactions with Variable‐Charge Soils.’’ Martnus NijhoV Publishers,
Dordrecht/Boston/Lancaster.
Barrow, N. J., and Shaw, T. C. (1975). The slow reactions between soils and anions: 2. EVect of
time and temperature on the decrease in phosphate concentration in the soil solution. Soil
Sci. 119, 167–177.
Barrow, N. J., and Shaw, T. C. (1977). The slow reactions between soil and anions: 6. EVect of
time and temperature of contact on fluoride. J. Soil Sci. 124, 265–278.
Bar‐Yosef, B., and Kafkafi, U. (1978). Phosphate desorption from kaolinite suspensions. Soil
Sci. Soc. Am. J. 42, 570–574.
Beauchemin, S., Hesterberg, D., Chou, J., Beauchemin, M., Smard, R. R., and Sayers, D. E.
(2003). Speciation of phosphorus in phosphorus‐enriched agricultural soils using X‐ray
absorption near‐edge structure spectroscopy and chemical fractionation. J. Environ. Qual.
32, 1809–1819.
Beek, J., and van Riemsdijk, W. H. (1982). Interactions of orthophosphate ion with soil. In ‘‘Soil
Chemistry B, Physico‐Chemical Models’’ (G. H. Bolt, Ed.), pp. 259–284. Elsevier Science
Publication, Co., Amsterdam.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 173

Behrens, P. (1992). X‐ray absorption spectroscopy in chemistry. 2. X‐ray absorption near edge
structure. Trac.‐Trends Anal. Chem. 11, 237–244.
Behrens, P., Felsche, J., Vetter, S., SchulzekloV, G., Jaeger, N. I., and Miemann, W. (1991).
A XANES and EXAFS investigation of titanium silicalite. J. Chem. Soc., Chem. Commun.
10, 678–680.
Black, C. A. (1942). Phosphate fixation by kaolinite and other clays as aVected by pH,
phosphate concentration, and time of contact. Soil Sci. Soc. Am. Proc. 6, 123–133.
Bleam, W. F., PfeVer, P. E., Goldberg, S., Taylor, R. W., and Dudley, R. (1991). A 31P solid‐
state nuclear magnetic resonance study of phosphate adsorption at the boehmite/aqueous
solution. Langmuir 7, 1702–1712.
Bloom, P. R. (1981). Phosphorus adsorption by an aluminum‐peat complex. Soil Sci. Soc.
Am. J. 45, 267–272.
Bolan, N. S., Barrow, N. J., and Posner, A. M. (1985). Describing the eVect of time on sorption
of phosphate by iron and aluminium hydroxides. J. Soil Sci. 36, 187–197.
Borggaard, O. K., Raben‐Lange, B., Gimsing, A. L., and Strobel, B. W. (2005). Influence of
humic substances on phosphate adsorption by aluminium and iron oxides. Geoderma 127,
270–279.
Bowden, J. W., Nagarajah, S., Barrow, N. J., Posner, A. M., and Quirk, J. P. (1980). Describing
the adsorption of phosphate, citrate and selenite on the variable charge mineral surface.
Aust. J. Soil Res. 18, 49–60.
Burkholder, J. M., and Glasgow, G. B., Jr. (1997). Pfiesteria piscicida and other Pfiesteria‐like
dinoflagellates: Behavior, impacts, and environmental controls. Limnol. Oceanogr. 42,
1052–1075.
Cabrea, F., Arambarri, P. D., Madrid, L., and Toca, C. G. (1981). Desorption of phosphate
from iron oxides in relation to equilibrium pH and porosity. Geoderma 26, 203–216.
Cathcart, J. B. (1980). World phosphate reserve and resources. In ‘‘The Role of Phosphorus in
Agriculture’’ (F. E. Khasawneh, E. C. Sample, and E. J. Kamprath, Eds.), pp. 1–18. Soil
Science Society of America, American Society of Agronomy, Crop Science Society of
America, Madison, Wisconsin.
Chen, S. H. (1977). Dissolution rates of phosphate rocks. Soil Sci. Soc. Am. Proc. 41, 656–657.
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

Chen, S. H., and Clayton, W. R. (1980). Application of Elovich equation to the kinetics of
phosphate release and sorption by soils. Soil Sci. Soc. Am. J. 44, 265–268.
Chen, Y. R., Butler, J. N., and Stumm, W. (1973). Adsorption of phosphate on alumina and
kaolinite from dilute aqueous solutions. J. Colloid Interface Sci. 43, 421–436.
Chien, S. H., Savant, N. K., and Mokwunye, U. (1982). EVect of temperature on phosphate
sorption and desorption in two acid soils. Soil Sci. 133, 160–166.
Chisholm‐Brause, C. J., Hayes, K. F., Roe, A. L., Brown, G. E., Jr., Parks, G. A., and Leckie,
J. O. (1990). Spectroscopic investigation of Pb(II) complexes at the g‐Al2O3/water interface.
Geochim. Cosmochim. Acta 54, 1897–1909.
Chitrakar, R., Tezuka, S., Sonoda, A., Sakane, K., Ooi, K., and Hirotsu, T. (2006).
Phosphate adsorption on synthetic goethite and akaganeite. J. Colloid Interface Sci.
298, 602–608.
Colemann, R. (1944). The mechanism of phosphate fixation by montmorillonitic and kaolinitic
clays. Soil Sci. Soc. Am. Proc. 9, 72–78.
Dalal, R. C. (1977). Soil organic phosphorus. In ‘‘Advances in Agronomy’’ (N. C. Brady, Ed.),
Vol. 29, pp. 83–117. Academic Press, New York.
Diaz, O. A., Anderson, D. L., and Hanlon, E. A. (1993). P mineralization from histosols of the
everglades agricultural area. Soil Sci. 156, 178–185.
Edzwald, J. K., Toensing, D. C., and Leung, M. C. (1976). Phosphate adsorption reaction with
clay minerals. Environ. Sci. Technol. 10, 485–490.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
174 Y. ARAI AND D. L. SPARKS

Enfield, C. G., Phan, T., Wolters, D. M., and Ellis, J. R. (1981). Kinetic model for phosphate
transport and transformation in calcareous soils I and II. Soil Sci. Soc. Am. J. 45,
1059–1064.
Evans, R. L., and Jurinak, J. J. (1976). Kinetics of phosphate release from a desert soil. Soil Sci.
121, 205–211.
Farley, K. J., Dzombak, D. A., and Morel, F. M. M. (1985). A surface precipitation model for
the sorption of cations on metal oxides. J. Colloid Interface Sci. 106, 226–242.
Fendorf, S. E., and Sparks, D. L. (1996). X‐ray absorption fine structure spectroscopy.
In ‘‘Methods of Soil Analysis’’ (D. L. Sparks, Ed.), Vol. 5, pp. 377–416. Soil Science
Society of America, Madison, WI.
Fendorf, E. E., Fendorf, M., Sparks, D. L., and Gronsky, R. (1992). Inhibitory mechanisms of
Cr(III) oxidation by d‐MnO2. J. Colloid Interface Sci. 153, 37–53.
Freese, D., Weidler, P. G., Grolimund, D., and Sticher, H. (1999). A flow‐through reactor with
an infinite sink for monitoring desorption processes. J. Environ. Qual. 28, 537–543.
Fried, C. R., and Shapiro, G. (1956). Phosphate supply pattern of various soils. Soil Sci. Soc.
Am. Proc. 20, 471–475.
Fuller, C. C., Davis, J. A., and Waychunas, G. A. (1993). Surface chemistry of ferrihydrite: Part
2. Kinetics of arsenate adsorption and coprecipitation. Geochim. Cosmochim. Acta 57,
2271–2282.
Gale, P. M., Reddy, K. R., and Graetz, D. A. (1992). Mineralization of sediment organic matter
under anoxic conditions. J. Environ. Qual. 21, 394–400.
Geelhoed, J. S., Hiemstra, T., and Rimsdijk, W. H. V. (1997). Phosphate and sulfate adsorption
on goethite: Single anion and competitive adsorption. Geochim. Cosmochim. Acta 61,
2389–2396.
Geelhoed, J. S., Hiemstra, T., and van Riemsdijk, W. H. (1998). Competitive interaction
between phosphate and citrate on goethite. Environ. Sci. Technol. 32, 2119–2123.
Goldberg, S. (1992). Use of surface complexation models in soil chemical system. In ‘‘Advances
in Agronomy’’ (D. L. Sparks, Ed.), Vol. 47, pp. 233–329. Academic Press, San Diego.
Goldberg, S., and Sposito, G. (1984a). A chemical model of phosphate adsorption by soils: 2.
Nonclacareous soils. Soil Sci. Soc. Am. J. 48, 779–783.
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

Goldberg, S., and Sposito, G. (1984b). A chemical model of phosphate adsorption by soils:
1. Reference oxide minerals. Soil Sci. Soc. Am. J. 48, 772–778.
Goldberg, S., and Sposito, G. (1985). On the mechanism of specific phosphate adsorption by
hydroxylated mineral surfaces: A review. Commun. Soil Sci. Plant Anal. 16, 801–821.
GriYn, E. A., and Jurinak, J. (1973). Test of a new model for the kinetics of adsorption‐
desorption processes. Soil Sci. Soc. Am. J. 37, 869–872.
Hansmann, D. D., and Anderson, M. A. (1985). Using electrophoresis in modeling sulfate,
selenite, and phosphate adsorption onto goethite. Environ. Sci. Technol. 19, 544–551.
Harris, W. G., Wang, H. D., and Reddy, K. R. (1994). Dairy manure influence on soil and
sediment composition: Implications for phosphorus retention. J. Environ. Qual. 23,
1071–1081.
Haseman, J. F., Brown, E. H., and Whitt, C. D. (1950). Some reactions of phosphate with clays
and hyfrous oxides of iron and aluminum. Soil Sci. 70, 257–271.
Hawke, D., Carpenter, P. D., and Hunter, K. A. (1989). Competitive adsorption of phosphate
on goethite in marine electrolytes. Environ. Sci. Technol. 23, 187–191.
Hayes, K. F., Papelis, C., and Leckie, J. O. (1988). Modeling ionic strength eVects of anion
adsorption at hydrous oxide/solution interface. J. Colloid Interface Sci. 125, 717–726.
Helyar, K. R., Munns, D. N., and Burau, R. G. (1976a). Adsorption of phosphate by gibbsite.
I. EVects of neutral chloride salts of calcium, magnesium, sodium and potassium. J. Soil
Sci. 27, 307–314.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 175

Helyar, K. R., Munns, D. N., and Burau, R. G. (1976b). Adsorption of phosphate by gibbsite.
II. Formation of a surface complex involving divalent cations. J. Soil Sci. 27, 315–323.
Hesterberg, D., Zhou, W., Hutchison, K. J., Beauchemin, S., and Sayers, D. E. (1999). XAFS
study of adsorbed and mineral forms of phosphate. J. Synchrotron Rad. 6, 636–638.
Hingston, F. J., Atkinson, R. J., Posner, A. M., and Quirk, J. P. (1967). Specific adsorption of
anions. Nature 215, 1459–1461.
Hingston, F. J., Posner, A. M., and Quirk, J. P. (1974). Anion adsorption by goethite and
gibbsite. II. Desorption of anions from hydrous oxides surfaces. J. Soil Sci. 25, 16–26.
Hongshao, Z., and Stanforth, R. (2001). Competitive adsorption of phosphate and arsenate on
goethite. Environ. Sci. Technol. 35, 4753–4757.
Hsu, P. H., and Rennie, D. A. (1962). Reactions of phosphate in aluminum systems. I.
Adsorption of phosphate by x‐ray amorphous aluminum hydroxide. Can. J. Soil Sci. 42,
197–209.
Hunter, R. J. (1981). Zeta potential in colloid science, principles and applications. In ‘‘Colloid
Science: A Series of Monographs,’’ pp. 219–257. Academic Press, San Diego.
Johansson, U., Holmgren, A., Forsling, W., and Frost, R. (1998). Isotopic exchange of kaolinite
hydroxyl protons: A diVuse reflectance infrared Fourier transform spectroscopy study.
Analyst 123, 641–645.
Johnson, S. E., and Loeppert, R. H. (2006). Role of organic acids in phosphate mobilization
from iron oxide. Soil Sci. Soc. Am. J. 70, 222–234.
Jonasson, R. G., Martin, R. R., Giuliacci, M. E., and Tazaki, K. (1988). Surface reactions of
goethite with phosphate. J. Chem. Soc., Faraday Trans. 1 84, 2311–2315.
Kafkafi, U., Posner, A. M., and Quirk, J. P. (1967). Desorption of phosphate from kaolinite.
Soil Sci. Soc. Am. Proc. 31, 348–353.
Khare, N., Hesterberg, D., and Martin, J. D. (2005). XANES investigation of phosphate
sorption in single and binary systems of iron and aluminum oxide minerals. Environ. Sci.
Technol. 39, 2152–2160.
Koskinen, W. C., and Harper, S. S. (1990). The retention process: Mechanisms. In ‘‘Pesticides in
the Soil Environment: Processes, Impacts, and Modeling’’ (H. H. Cheng, Ed.), pp. 51–77.
Soil Science Society of America, Inc., Madison, WI.
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

Kotak, B. G., Kenefick, S. L., Fritz, D. L., Rousseau, C. G., Prepas, E. E., and Hrudely, S. E.
(1993). Occurrence and toxicological evaluation of cyanobacteral toxins in Alberta lakes
and farm dugouts. Water Res. 27, 495–506.
Kuo, S., and Lotse, E. G. (1972). Kinetics of phosphate adsorption by calcium carbonate and
Ca‐kaolinite. Soil Sci. Soc. Am. Proc. 36, 725–729.
Kuo, S., and Lotse, E. G. (1973). Kinetics of phosphate adsorption and desorption from
hematite and gibbsite. Soil Sci. 116, 400–405.
Lawton, L. A., and Codd, G. A. (1991). Cyanobacteria (Blue‐green algae) toxins and their
significance in UK and European waters. J. Inst. Water Environ. Manag. 5, 460–465.
Ler, A., and Stanforth, R. (2003). Evidence for surface precipitation of phosphate on goethite.
Environ. Sci. Technol. 37, 2694–2700.
Levin, E. D., Schmechel, D. E., Burkholder, J. M., Glasgow, H. B., Jr., Deamer‐Melia, N. J.,
Moser, V. C., and Harry, G. J. (1997). Persisting learning deflects in rats after exposure to
Pfiesteria piscida. Environ. Health Perspect. 105, 1320–1325.
Lindsay, W. L. (1979). Phosphate. In ‘‘Chemical Equilibria in Soils,’’ pp. 163–209. John Wiley,
New York.
Liu, F., He, J., Colombo, C., and Violante, A. (1999). Competitive adsorption of sulfate and
oxalate on goethite in the absence or presence of phosphate. Soil Sci. 164, 180–189.
Lookman, L. D. F., Merckx, R., Vlassak, K., and van Riemsdijk, W. M. (1995). Long‐term
kinetics of phosphate release from soil. Environ. Sci. Technol. 29, 1569–1575.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
176 Y. ARAI AND D. L. SPARKS

López‐Hernandez, D., and Burnhan, C. P. (1974). The covariance of phosphate sorption with
other soil properties in some British and tropical soils. J. Soil Sci. 25, 207–216.
López‐Pineiro, A., and Navarro, A. G. (1997). Phosphate sorption in Vertisols of Southwestern
Spain. Soil Sci. 162, 69–77.
Low, P. F., and Black, C. A. (1950). Reactions of phosphate with kaolinite. Soil Sci. 70,
273–290.
Luengo, C., Brigante, M., Antelo, J., and Avena, M. (2006). Kinetics of phosphate adsorption
on goethite: Comparing batch adsorption and ATR‐IR measurements. J. Colloid Interface
Sci. 300, 511–518.
Madrid, L., and Posner, A. M. (1979). Desorption of phosphate from goethite. J. Soil Sci. 30,
697–707.
Maguire, R. O., Sims, J. T., and Coale, F. J. (2000). Phosphorus fractionation in biosolids‐
amended soils: Relationship to soluble and desorbable phosphorus. Soil Sci. Soc. Am. J. 64.
Manning, J. R. (1968). ‘‘DiVusion Kinetics for Atoms in Crystals.’’ D. Van Nostrand and
Company Inc, Princeton, New Jersey.
Martine, R. R., and Smart, R. S. C. (1987). X‐ray photoelectron studies of anion adsorption on
goethite. Soil Sci. Soc. Am. J. 51, 54–56. Springer‐Verlag, New York.
Martine, R. R., Smart, R. S., and Tazaki, K. (1988). Direct observation of phosphate precipita-
tion in the goethite/phosphate system. Soil Sci. Soc. Am. J. 52, 1492–1500.
Mattson, S., and Hester, J. B. (1935). The laws of soil colloidal behavior: XV. The degradation
and the regeneration of the soil complex. Soil Sci. 39, 75–84.
McBride, M. B. (1989). Reactions controlling heavy metal solubility in soils. In ‘‘Advances in
Soil Science’’ (B. A. Stewart, Ed.), Vol. 10, pp. 1–56. Springer‐Verlag, New York.
McBride, M. B. (1994). Review of chemical principles. In ‘‘Environmental Chemistry of Soils,’’
pp. 3–30. Oxford University Press, New York.
McBride, M. B. (1997). A critique of diVuse double layer models applied to colloid and surface
chemistry. Clays Clay Miner. 45, 598–608.
McBride, M. M. (1991). Processes of heavy metal and transition metal sorption by soil minerals.
In ‘‘Interactions at the Soil Colloid‐Soil Solution Interface’’ (G. H. Bolt, M. F. D. Boodt,
M. H. B. Hayes, and M. M. Mcbride, Eds.), Vol. 190, pp. 149–176. Kluwer Academic
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

Publishers, Dordrecht.
McCammon, C. A., and Burn, R. G. (1980). The oxidation mechanism of vivianite as studied by
Mössbauer spectroscopy. Am. Miner. 65, 361–366.
Mengel, K. (1985). Dynamics and availability of major nutrients in soils. In ‘‘Advances in Soil
Science’’ (B. A. Stewart, Ed.), pp. 96–105. Springer‐Verlag, New York, Berlin, Heidelberg,
Tokyo.
Muljadi, D., Posner, A. M., and Quirk, J. P. (1966). The mechanism of phosphate adsorption by
Kaolinite, Gibbsite, and Pseudoboehmite. J. Soil Sci. 17, 212–229.
Munns, D. N., and Fox, R. L. (1976). The slow reaction which continues after phosphate
adsorption: Kinetics and equilibrium in some tropical soils. Soil Sci. Soc. Am. J. 40, 46–51.
Myneni, S. C. B. (2000). X‐ray and vibrational spectroscopy of sulfate in earth materials.
In ‘‘Sulfate Minerals Crystallography, Geochemistry, and Environmental Significance’’
(C. N. Alpers, J. L. Jambor, and D. K. Nordstrom, Eds.), Vol. 40, pp. 113–172. Mineral-
ogical Society of America, Washington, DC.
Nanzyo, M. (1984). DiVused reflectance infrared spectra of phosphate sorbed on alumina gel.
J. Soil Sci. 35, 63–69.
Nanzyo, M. (1986). Infrared spectra of phosphate sorbed on iron hydroxide gel and the sorption
products. Soil Sci. Plant Nutr. 32, 51–58.
Nanzyo, M. (1987). Formation of noncrystalline aluminum phosphate through phosphate
sorption on allophanic soils. Commun. Soil Sci. Plant Anal. 18, 735–742.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 177

Nanzyo, M., and Watanabe, Y. (1982). DiVuse reflectance infrared spectra and ion‐adsorption
properties of the phosphate surface complex on goethite. Soil Sci. Plant Nutr. 28, 359–368.
Nilsson, N., Lövgren, L., and Sjöberg, S. (1992). Phosphate complexation at the surface of
goethite. Chem. Spec. Bioavail. 4, 121–130.
Nriagu, J. O. (1972). Solubility equilibrium constant of strengite. Am. J. Sci. 272, 476–484.
O’Day, P., Parks, G. A., and Brown, G. E., Jr. (1994). X‐ray absorption spectroscopy of cobalt
(II) multinuclear surface complexes and surface precipitates on kaolinite. J. Colloid Inter-
face Sci. 165, 269–289.
Okajima, H., Kubota, H., and Sakuma, T. (1983). Hysteresis in the phosphorus sorption and
desorption processes of soils. Soil Sci. Plant Nutr. 29, 271–283.
Okude, N., Nagoshi, M., Noro, H., Baba, Y., Yamamoto, H., and Sasaki, T. A. (1999). P and
S K‐edge XANES of transition‐metal phosphates and sulfates. J. Electron. Spectrosc.
101–103, 607–610.
Olsen, S. R., and Watanabe, F. S. (1957). A method to determine a phosphorus adsorption
maximum of soils as measured by the Langmuir isotherm. Soil Sci. Soc. Am. Proc. 21,
144–149.
Onken, A. B., and Matheson, R. L. (1982). Dissolution rate of EDTA‐extractable phosphate
from soils. Soil Sci. Soc. Am. J. 46, 276–279.
Ozanne, P. G. (1980). Phosphate nutrition of plants‐a general treatise. In ‘‘The Role of Phos-
phorus in Agriculture’’ (F. E. Khasawneh, E. C. Sample, and E. J. Kamprath, Eds.),
pp. 559–589. Soil Science Society of America, American Society of Agronomy, Crop
Science Society of America, Madison, WI.
Parfitt, R. L. (1979). The nature of the phosphate goethite (a‐FeOOH) complex formed with
Ca(H2PO4)2 at diVerent surface coverage. Soil Sci. Soc. Am. J. 43, 623–625.
Parfitt, R. L. (1989). Phosphate reactions with natural allophane, ferrihydrite and goethite.
J. Soil Sci. 40, 359–369.
Parfitt, R. L., and Atkinson, R. J. (1976). Phosphate adsorption on goethite (a‐FeOOOH).
Nature 264, 740–741.
Parfitt, R. L., and Smart, R. S. C. (1978). The mechanism of sulfate adsorption on iron oxides.
Soil Sci. Soc. Am. J. 42, 48–50.
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

Parfitt, R. L., Atkinson, R. J., and Smart, R. S. C. (1975). The mechanism of phosphate fixation
by iron oxides. Soil Sci. Soc. Am. Proc. 39, 837–841.
Parry, R. (1998). Agricultural phosphorus and water quality: A U.S. Environmental Protection
Agency perspective. J. Environ. Qual. 27, 258–261.
Persson, P., Nielsson, N., and Sjöberg, S. (1996). Structure and bonding of orthophosphate ions
at the iron oxide‐aqueous interface. J. Colloid Interface Sci. 177, 263–275.
Pierzynski, G. M., Logan, T. J., and Traina, S. J. (1990a). Phosphorus chemistry and mineralo-
gy in excessively fertilized soils: Solubility equilibria. Soil Sci. Soc. Am. J. 54, 1589–1595.
Pierzynski, G. M., Logan, T. J., Traina, S. J., and Bigham, J. M. (1990b). Phosphorus chemistry
and mineralogy in excessively fertilized soils: Quantitative analysis of phosphorus‐rich
particles. Soil Sci. Soc. Am. J. 54, 1576–1583.
Pignatello, J. J., and Xing, B. (1995). Mechanisms of slow sorption of organic chemicals to
natural particles. Environ. Sci. Technol. 30, 1–11.
Rajan, S. S. S. (1975). Phosphate adsorption and the displacement of structural silicon in
allophane clay. J. Soil Sci. 26, 250–256.
Reifenberg, A., and Buckwold, S. J. (1954). The release of silica from soils by the orthophos-
phate anion. J. Soil Sci. 5, 106–115.
Rickert, H. (1982). ‘‘Electrochemistry of Solids—An Introduction.’’ Springer, Berlin.
Rose, J., Flank, A., Masion, A., Bottero, J., and Elmerich, P. (1997). Nucleation and growth
mechanisms of Fe oxyhydroxide in the presence of PO4 ions. 2. P K‐edge EXAFS study.
Langmuir 13, 1827–1834.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
178 Y. ARAI AND D. L. SPARKS

Ryden, J. C., and Syers, J. K. (1977). Desorption and isotopic exchange relationship of
phosphate sorbed by soils and hydrous ferric oxide. J. Soil Sci. 28, 596–609.
Ryden, J. C., Syers, J. K., and Harris, R. F. (1973). Phosphorus in runoV and streams. In ‘‘Advances
in Agronomy’’ (N. C. Brady, Ed.), Vol. 25, pp. 1–45. Academic Press, New York.
Ryden, J. C., Mclaughlin, J. R., and Syers, J. K. (1977a). Time‐dependent sorption of phosphate
by soils and hydrous ferric oxides. J. Soil Sci. 28, 585–595.
Ryden, J. C., Syers, J. K., and Mclaughlin, J. R. (1977b). EVects of ionic strength on chemi-
sorption and potential‐determining sorption of phosphate. J. Soil Sci. 28, 62–71.
Sanchez, P. A., and Uehara, G. (1980). Management considerations for acid soils with high
phosphorus fixation capacity. In ‘‘The role of phosphorus in agriculture’’ (F. E. Khasawneh,
E. C. Sample, and E. J. Kamprath, Eds.), American Society of Agronomy, Madison,
Wisconsin.
Sato, S., Solomon, D., Hyland, C., Ketterings, Q. M., and Lehmann, J. (2005). Phosphorus
speciation in manure and manure‐amended soils using XANES spectroscopy. Environ. Sci.
Technol. 39, 7485–7491.
Sei, J., Jumas, J. C., Olivier‐Fourcade, J., Quiquampoix, H., and Staunton, S. (2002). Role of
iron oxides in the phosphate adsorption properties of kaolinites from the Ivory Coast.
Clays Clay Miner. 50, 217–222.
Shang, C., Stewart, J. W. B., and Huang, P. M. (1992). pH eVect on kinetics of adsorption of
organic inorganic phosphates by short‐range ordered aluminum and iron precipitates.
Geoderma 53, 1–14.
Shapiro, R. E. (1957). The eVect of flooding on the availability of P and nitrogen. Soil Sci. 85,
190–197.
Sharpley, A. N., and Rekolainen, S. (1997). Phosphorus in agriculture and its environmental
implications. In ‘‘Phosphorus Loss from Soil to Water’’ (H. Tunney, O. T. Carton,
P. C. Brookes, and A. E. Johnston, Eds.), pp. 1–54. Cab International, New York.
Sharpley, A. N., Foy, B., and Whiters, P. (2000). Practical and innovative measures for the
control of agricultural phosphorus losses to water: An overview. J. Environ. Qual. 29, 1–9.
Sheppard, S. C., and Racz, G. J. (1984). EVects of soil temperature on phosphorus extractability.
1. Extractions and plant uptake of soil and fertilizer phosphorus. Can. J. Soil Sci. 64, 241–254.
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

Snoeyink, V. L., and Jenkins, D. (1980). Precipitation and dissolution. In ‘‘Water Chemistry,’’
pp. 243–315. John Wiley & Sons, Inc., New York.
Sparks, D. L. (1987). Dynamics of soil potassium. In ‘‘Advances in Soil Science’’ (B. A. Stewart,
Ed.), Vol. 6, pp. 1–63. Springer Verlag, New York.
Sparks, D. L. (1989). ‘‘Kinetics of Soil Chemical Processes.’’ Academic Press, Inc., San Diego.
Sparks, D. L. (1995a). ‘‘Environmental Soil Chemistry.’’ Academic Press, Inc., San Diego.
Sparks, D. L. (1995b). Sorption phenomena on soils. In ‘‘Environmental Soil Chemistry,’’
pp. 99–139. Academic Press, San Diego, CA.
Sposito, G. (1986). Distinguishing adsorption from surface precipitation. In ‘‘Geochemical
Processes at Mineral Surfaces’’ (J. A. Davis and K. F. Hayes, Eds.), pp. 217–228. American
Chemical Society, Washington, DC.
Sposito, G. (1989). Soil adsorption phenomena. In ‘‘The Chemistry of Soils.’’ Oxford University
Press, New York, Oxford.
Steinfield, J. I., Francisco, J. S., and Hase, W. L. (1989). ‘‘Chemical Kinetics and Dynamics.’’
Englewood CliVs, Prentice Hall.
Stumm, W., and Morgan, J. J. (1995a). ‘‘Aquatic Chemistry, Chemical Equilibria and Rates in
Natural Waters.’’ John Wiley & Sons, Inc., New York.
Stumm, W., and Morgan, J. J. (1995b). Theory of elementary processes. In ‘‘Aquatic Chemistry,
Chemical Equilibria and Rates in Natural Water,’’ pp. 69–76. John Wiley & Sons, Inc.,
New York.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
PHOSPHATE REACTION DYNAMICS 179

Tejedor‐Tejedor, M. I., and Anderson, M. A. (1990). Protonation of phosphate on the surface of


goethite as studied by CIR‐FTIR and electrophoretic mobility. Langmuir 6, 602–611.
Toor, G. S., Hunger, S., Peak, J. D., Sims, J. T., and Sparks, D. L. (2006). Advances in the
characterization of phosphorus in organic wastes: Environmental and agronomic applications.
In ‘‘Advances in Agronomy’’ (D. L. Sparks, Ed.), Vol. 89, pp. 1–72. Elsevier, San Diego.
Torrent, J. (1991). Activation energy of the slow reaction between phosphate and goethites of
diVerent morphology. Aust. J. Soil Res. 29, 69–74.
Torrent, J., Barron, V., and Schwertmann, U. (1990). Phosphate adsorption and desorption by
goethite diVering in crystal morphology. Soil Sci. Soc. Am. J. 54, 1007–1012.
Torrent, J., Schwertmann, U., and Barron, V. (1992). Fast and slow phosphate sorption by
goethite‐rich natural materials. Clays Clay Miner. 40, 14–21.
Vaithiyanathan, P., and Correll, D. L. (1992). The Rode river watershed: Phosphorus distribu-
tion and export in forest and agricultural soils. J. Environ. Qual. 21, 280–288.
van der Zee, S. E. A. T. M., and van Riemsdijk, W. H. (1988). Model for long‐term phosphate
reaction kinetics in soil. J. Environ. Qual. 17, 35–41.
van Riemsdijk, W. H., and de Haan, F. A. M. (1981). Reaction of orthophosphate with a sandy
soil at constant supersaturation. Soil Sci. Soc. Am. J. 45, 261–266.
van Riemsdijk, W. H., and Lyklema, J. (1980). Reaction of phosphate with gibbsite (Al(OH)3)
beyond the adsorption maximum. J. Colloid Interface Sci. 76, 55–66.
van Riemsdijk, W. H., Weststrate, F. A., and Beek, J. (1977). Phosphate in soils treated with
sewage water: III. Kinetics studies on the reaction of phosphate with aluminum com-
pounds. J. Environ. Qual. 6, 26–29.
van Riemsdijk, W. H., Boumans, L. J. M., and de Haan, F. A. M. (1984). Phosphate sorption by
soils: I. A model for phosphate reaction with metal‐oxides in soil. Soil Sci. Soc. Am. J. 48,
537–541.
Veith, J. A., and Sposito, G. (1977). Reactions of aluminosilicates, aluminum hydrous oxides,
and aluminum oxide with o‐phosphate: The formation of x‐ray amorphous of variscite and
montebrasite. Soil Sci. Soc. Am. J. 41, 870–876.
Wang, M. K. (1987). Synthetic and characterization of iron phosphate and aluminum‐
substituted iron phosphate compounds. J. Chin. Agric. Chem. Soc. 25, 398–411.
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

Wang, M. K., and Tzou, Y. M. (1995). Phosphate sorption by calcite, and iron‐rich calcareous
soils. Geoderma 65, 249–261.
Wang, H. D., Harris, W. G., and Yuan, T. L. (1991). Noncrystalline phosphates in Florida
phosphatic soils. Soil Sci. Soc. Am. J. 55, 665–669.
Waychunas, G. A., Rea, B. A., Fuller, C. C., and Davis, J. A. (1993). Surface chemistry of
ferrihydrite: Part 1. EXAFS studies of the geometry of coprecipitated and adsorbed arsenate.
Geochim. Cosmochim. Acta 57, 2251–2264.
Willett, I. R., Chartres, C. J., and Nguyen, T. T. (1988). Migration of phosphate into aggregated
particles of ferrihydrite. J. Soil Sci. 39, 275–282.
Withers, P. J. A. (1996). Phosphorus cycling in UK agriculture and implications for water
quality. Soil Use Manage. 12, 221–228.
Wong, J., Lytle, F. W., Messmer, R. P., and Maylotte, D. H. (1984). K‐edge absorption spectra
of selected vanadium compounds. Phys. Rev. B 30, 5596–5610.
Yao, W., and Millero, F. J. (1996). Adsorption of phosphate on manganese dioxide in sea water.
Environ. Sci. Technol. 30, 536–541.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.
This page intentionally left blank
Copyright © 2007. Elsevier Science & Technology. All rights reserved.

Advances in agronomy. (2007). Retrieved from http://ebookcentral.proquest.com


Created from qut on 2019-07-26 00:03:33.

S-ar putea să vă placă și