phosphate sorption in soils and reference oxides under

139
University of Massachusetts Amherst University of Massachusetts Amherst ScholarWorks@UMass Amherst ScholarWorks@UMass Amherst Doctoral Dissertations 1896 - February 2014 1-1-1989 Phosphate sorption in soils and reference oxides under varying Phosphate sorption in soils and reference oxides under varying pH and redox conditions. pH and redox conditions. Nadim Khouri University of Massachusetts Amherst Follow this and additional works at: https://scholarworks.umass.edu/dissertations_1 Recommended Citation Recommended Citation Khouri, Nadim, "Phosphate sorption in soils and reference oxides under varying pH and redox conditions." (1989). Doctoral Dissertations 1896 - February 2014. 6076. https://scholarworks.umass.edu/dissertations_1/6076 This Open Access Dissertation is brought to you for free and open access by ScholarWorks@UMass Amherst. It has been accepted for inclusion in Doctoral Dissertations 1896 - February 2014 by an authorized administrator of ScholarWorks@UMass Amherst. For more information, please contact [email protected].

Upload: others

Post on 24-Feb-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Phosphate sorption in soils and reference oxides under

University of Massachusetts Amherst University of Massachusetts Amherst

ScholarWorks@UMass Amherst ScholarWorks@UMass Amherst

Doctoral Dissertations 1896 - February 2014

1-1-1989

Phosphate sorption in soils and reference oxides under varying Phosphate sorption in soils and reference oxides under varying

pH and redox conditions. pH and redox conditions.

Nadim Khouri University of Massachusetts Amherst

Follow this and additional works at: https://scholarworks.umass.edu/dissertations_1

Recommended Citation Recommended Citation Khouri, Nadim, "Phosphate sorption in soils and reference oxides under varying pH and redox conditions." (1989). Doctoral Dissertations 1896 - February 2014. 6076. https://scholarworks.umass.edu/dissertations_1/6076

This Open Access Dissertation is brought to you for free and open access by ScholarWorks@UMass Amherst. It has been accepted for inclusion in Doctoral Dissertations 1896 - February 2014 by an authorized administrator of ScholarWorks@UMass Amherst. For more information, please contact [email protected].

Page 2: Phosphate sorption in soils and reference oxides under
Page 3: Phosphate sorption in soils and reference oxides under

PHOSPHATE SORPTION IN SOILS AND REFERENCE

OXIDES UNDER VARYING pH AND REDOX CONDITIONS

A Dissertation Presented

by

NADIM KHOURI

Submitted to the Graduate School of the University of Massachusetts in partial fulfillment of the

requirements for the degree of DOCTOR OF PHILOSOPHY

May 1989

Department of Plant and Soil Sciences

Page 4: Phosphate sorption in soils and reference oxides under

c Copyright by Nadim Khouri 1989

All Rights Reserved

Page 5: Phosphate sorption in soils and reference oxides under

PHOSPHATE SORPTION IN SOILS AND REFERENCE

OXIDES UNDER VARYING pH AND REDOX CONDITIONS

A Dissertation Presented

by

NADIM KHOURI

Approved as to style and content by:

j

Peter L. M. Veneman, Chairperson of Committee

tc/ ■fytc--

Michael S. Switzenbaum, Member

Baker, Department Head

lent of Plant and Soil Sciences

Page 6: Phosphate sorption in soils and reference oxides under

Dedication

This dissertation is dedicated

to the memory of my parents,

Mary Debbas Khouri

and Joseph Khouri.

Page 7: Phosphate sorption in soils and reference oxides under

ACKNOWLEDGMENTS

I would like to thank Dr. Peter Veneman for his trust, support,

and friendship throughout the course of my studies at the University

of Massachusetts and for suggesting the research area for this

dissertation.

I am thankful to Drs. John Baker, Allen Barker, Daniel Hillel,

and Michael Switzenbaum for their guidance as members of the

dissertation committee. Collectively, they were extremely patient in

attempting to keep my work focused and my objectives clear.

Individually, I feel very privileged to have shared their friendship

and science.

I feel privileged to have been associated with the Soil

Morphology group. I thank Steve Bodine for the daily discussions we

had and for guiding me through the soils characteristics data used in

Chapters 4 and 5; he has had a major role in generating this basic

data. Judy Bartos artfully performed metal analyses reported in

Chapter 3 and Mohammad Mahinakbarzadeh ran quite a few of the P

sorption isotherms of Chapter 4. Bill Donnelly's help in typing parts

of this dissertation came at a crucial time. Also, to Ron Lavigne, Deb

Kozlowski, Mike Reed and David Lindbo I express my gratitude for the

moments we had in Rm 18.

I am grateful to the following persons who advised and assisted

me during the construction of the experimental setup to control redox

potentials of oxide suspensions: Sami Haddad, Keith Shimeld, and Drs.

David Curran, Les Whitney, and Haim Gunner.

v

Page 8: Phosphate sorption in soils and reference oxides under

I would also like to thank Mrs. Lucy Yin and Mr. Lou Raboin for

conducting the electron microscope-related work of my research. Ed

Ostrowski and Ed Weist ran the "BET" analysis, and the Geology

Department allowed me to use their XRD equipment for the

characterization of metal oxides.

I thank Carol Cumps, Fred Schulten, and Martha Robinson for

repeatedly devoting time, energy and quite a bit of imagination in

addressing my concerns and those of my family. We always felt welcome

at the Office of Foreign Students and Scholars.

In Amherst, I feel privileged to have met some wonderful people.

In particular, I would like to thank Dr. and Mrs. Mack Drake, Val

Veneman, Rev. and Mrs. Philip Ward, and the staff of the Hampshire-

Franklin Day Care Center their help and encouragement made this

dissertation possible.

At the UNDP/World Bank program in Washington I am grateful for

the understanding of my supervisors and colleagues who have allowed me

to pursue a double-life of sorts for the last year. I am particularly

grateful to Carl Bartone for his confidence in me and to Saul

Arlosoroff and Fred Wright for their support.

My daughters, Lina and Maya, will soon realize how essential

their role was in this dissertation. Finally, I would like to express

my gratitude to my wife, Hoda, for her love. I feel that it is from

our companionship that I derived the will to produce this

dissertation, and the strength to accomplish whatever good comes out

of my life.

vi

Page 9: Phosphate sorption in soils and reference oxides under

ABSTRACT

PHOSPHATE SORPTION IN SOILS AND REFERENCE

OXIDES UNDER VARYING pH AND REDOX CONDITIONS

MAY 1989

NADIM KHOURI, B.S., AMERICAN UNIVERSITY OF BEIRUT

M.S., AMERICAN UNIVERSITY OF BEIRUT

Ph.D., UNIVERSITY OF MASSACHUSETTS

Directed by: Professor Peter L. M. Veneman

Despite numerous studies of phosphate (P) retention by soils

there is no unified methodology for the determination of P sorption

capacity (PSC). This dissertation establishes a framework for the use

of chemical models to evaluate PSC obtained through different methods

under varying environmental conditions. Chemical modeling too often is

an exercise in numerical fitting of thermodynamic data to reactions

that do not necessarily apply to field conditions. This study assessed

the PSC of two hydroxides and binary mixtures of these hydroxides

under different redox conditions. Subsequently, the PSC in selected

Massachusetts soils was studied to evaluate the significance of Fe and

A1 in true soils. Finally, the significance of P sorption in the

disposal of wastewater on land was quantified by determination of the

number of acres necessary to treat a certain amount of wastewater and

to ascertain the relative role of PSC determination techniques in this

land limiting constituent analysis procedure.

Two oxyhydroxides, goethite (a-FeOOH) and gibbsite (/-A^O^) were

synthesized and physically and chemically characterized. Phosphorus

vii

Page 10: Phosphate sorption in soils and reference oxides under

sorption capacities for the goethite and gibbsite were approximately 100

and 70 mmol kg~^, respectively, with maximum PSC values at pH 3 to 4.

The sorption additivity hypothesis, stating that each particle in

a mixed system reacts as in a monocomponent system, was validated for

the binary mixture of goethite and gibbsite at an initial P in

solution of 200 mmol m and at different pH values.

The applicability of mathematical and chemical models to land

application system is dependent on the determination of the maximum

capacity of the system under varying natural conditions. Iron is

affected by the oxidation-reduction potential; reducing conditions

solubilize Fe and possibly release sorbed P into solution. However, In

some instances, researchers have reported increases in P sorption

under reducing conditions. The speciation of Al theoretically should

not be affected by changes in soil redox.

To establish the relative contribution of Al and Fe to PSC, the

latter was evaluated under controlled redox conditions in the

laboratory. Special reactor vessels were constructed, filled with

simulated wastewater, and the redox potential microbially maintained

at ranges of +500 to -50 mV, representing oxidizing to reducing

conditions.

A substantial increase in P sorption was observed in reducing

environments containing 1:1 mixtures of goethite and gibbsite. This

increase was due entirely to Fe11 */Fe^ transformations. Compared to

the initial monocomponent systems, reoxidation of previously reduced

minerals increased the PSC but not to the extent of the continuously

reduced suspensions. The principal mechanism of P sorption in the

viii

Page 11: Phosphate sorption in soils and reference oxides under

reduced reactors could not be determined but was attributed to surface

reactions in the solid phase. Gibbsite surfaces may act as catalysts

to facilitate the formation of reduced iron minerals, possibly

vivianite Fe^CPO^).8^0. The research results indicated the

significance of PSC analysis for long term disposal and its

implications for the possible movememt of reduced colloids to the

groundwater.

In the present study, the elements that could be included in

such models are identified by correlating the PSC of representative

Massachusetts soils to selected soil characteristics. Five PSC indices

(Langmuir, Freundlich, Bache-Williams, Interpolation, and Equilibrium)

were derived. PSC was significantly correlated to Fe, Al, organic C,

and pH. Pedotransfer regression equations to predict P sorption in

soils were developed with Al and Fe as regressors.

ix

Page 12: Phosphate sorption in soils and reference oxides under

TABLE OF CONTENTS

Page

ACKNOWLEDGMENTS. v

ABSTRACT. vii

LIST OF TABLES. xii

LIST OF FIGURES.xiii

Chapter

1. INTRODUCTION. 1

2. BACKGROUND. 7

Introduction. 7

The Phosphate Controversy. 7

The Langmuir Equation: An Empirical Model . 8

Empirical vs Mechanistic Models. 12

Phosphorus Retaining Factors in Soils . 13

The Use of Reference Oxide Minerals . 13

Research Needs. 14

3. PHOSPHORUS SORPTION ON REFERENCE OXIDES:

pH AND REDOX EFFECTS. 16

Introduction. 16

Phosphorus Retention Mechanism. 17

Phosphate Sorption by Goethite. 18

Phosphate Sorption by Gibbsite. 19

Phosphate Sorption by Mixtures of Oxides. 20

Redox Effects on P sorption. 22

Materials and Methods . 24

Reference Oxides . 24

Sorption Isotherms. 25

pH Effects. 26

Chemical Characterization. 26

Crystallographic Characterization. 26

Sorption by Mixture of Oxides. 27

Redox Effects. 28

Experimental Apparatus. 28

Experimental Procedures. 32

Phosphate Isotherms. 33

Microscopic Analyses. 33

x

Page 13: Phosphate sorption in soils and reference oxides under

Results and Discussion. 34

Optical and Chemical Characterization

of the Oxides. 34

Phosphate Sorption by Monocomponent Systems. 40

Sorption Isotherms of Binary Mixtures. 43

pH Effects. 43

Redox Effects. 46

Phosphorus Sorption in Oxidized Systems. 48

Phosphorus Sorption in Reduced Systems. 50

Conclusions. 55

4. PHOSPHORUS SORPTION CAPACITY IN ACID SOILS. 58

Introduction. 58

Materials and Methods. 60

Soil Sampling and Characterization. 60

Phosphorus Sorption Capacity. 61

Statistical Analysis. 62

Results and Discussion. 63

Conclusions. 71

5. ROLE OF PHOSPHORUS SORPTION CAPACITY IN

SUITABILITY ASSESSMENT FOR WASTE DISPOSAL. 72

Introduction. 72

Materials and Methods. 74

Wastewater Characteristics. 74

Soils. 77

Phosphorus Sorption Capacity. 77

Assessment of Soil Assimilative Capacity. 81

Results and Discussion. 82

Conclus ions. 89

6. DISCUSSION AND CONCLUSIONS. 90

APPENDIX. 93

REDOX POTENTIAL MEASUREMENT

A TOOL FOR SOIL INVESTIGATIONS. 93

REFERENCES. 108

xi

Page 14: Phosphate sorption in soils and reference oxides under

LIST OF TABLES

Table 3.1. Demonstration of the independence of specific edge

area from the dimension of crystal thickness (D) in

gibbsite, assuming a perfectly hexagonal crystal. 39

Table 3.2. Maximum phosphate sorption capacity of the two

oxides derived by five methods: indirect surface

measurement (BET), geometric calculations (TEM),

batch equilibration (Pexd), Langmuir (PL), and

interpolation (Pj). 42

Table 4.1. Range in properties of 57 surface and subsurface

horizons in 23 representative Massachusetts soils.... 65

Table 4.2. Correlation matrix of selected soil characteristics

and PSC indices from 23 Massachusetts soils. 66

Table 4.3. Selected results of stepwise regression for

different PSC indices and various soil properties.

In each model, all variables are significant

at the 0.15 level. 68

Table 4.4. Summary of statistical parameters for regression

models with Fe^ and Al^, or Fep and

Alp as independent variables. 70

Table 5.1. Representative wastewater characteristics for raw

municipal sewage and secondary effluent. 76

Table 5.2. Range in properties of selected surface and

subsurface horizons of two Massachusetts soil series. 78

Table 5.3. Phosphate sorption capacity of two soils using

different estimation methods. PE y and Pg g refer

to PSC values based on final soil’solution

concentrations of 7 and 8 mg L~ , respectively. 80

Table 5.4. Land limiting constituent analysis for application

of 1,000 nr d_i of raw municipal wastewater

on two representative New England soil series. 83

3 -1 Table 5.5. Acreage required for the disposal of 1,000 m d

of secondary effluent using land limiting constituent

analysis in two representative New England soil

series. 84

Table 5.6. Summary of potential land limiting constituents

as affected by PSC method, soil series, and type

of effluent.

xii

Page 15: Phosphate sorption in soils and reference oxides under

LIST OF FIGURES

Figure 1.1 The chemical modeling process. 3

Figure 3.1 Cross-section of the reactor used for controlling

redox potential in soil suspensions. 29

Figure 3.2 Cross-section of the salt bridge (a), and platinum

electrode (b). 30

Figure 3.3 Diagram of electronic switching mechanism used to

control the Eh in soil suspensions. 31

Figure 3.4 Titrimetric surface acidity characterization of

goethite. 35

Figure 3.5 Titrimetric surface acidity characterization of

gibbsite. 36

Figure 3.6 Summary of major characteristics of the two oxides.. 38

Figure 3.7 Phosphate sorption in goethite, gibbsite, and a 1:1

mixture (1.9 g L” ) of both oxides. Electrolyte:

0.1 M NaCl, pH: 5.5-6.5,

concentrations:

initial solution P

0-500 mmol m” .. 41

Figure 3.8 1:1 Phosphate sorption of goethite, gibbsite and

mixture (1.9 g L” ) at different pH levels;

electrolyte* 0.1 M NaCl, initial P concentration:

100 mmol m~ (53 mmol P Kg” solids). 44

Figure 3.9 Phosphate sorption of goethite, gibbsite and a 1:1

mixture (1.9 g L” ) at different pH levels.

Electrolyte* 0.1 M NaCl, initial P concentration:

200 mmol m” (106 mmol P Kg” solids).. 45

Figure 3.10 Phosphate sorption of goethite, gibbsite and a

1:1 mixture (1.9 g L"^) at different pH levels.

Electrolyte* 0.1 M NaCl, initial P concentration:

300 mmol m” (160 mmol P Kg”^ solids).

Figure 3.11 Phosphate sorption on 1:1 goethite and gibbsite

suspension in simulated wastewater. pH=6.5.

Figure 3.12 Phosphate sorption in goethite and gibbsite (1:1)

suspension in simulated wastewater. Initial

P concentration=300 mmol m .

Figure 3.13 Phosphate sorption in 1.9 g L ^ gibbsite

suspension in simulated wastewater. pH=6.5.

47

49

51

52

xiii

Page 16: Phosphate sorption in soils and reference oxides under

Figure 3.14 Phosphate sorption expressed on volume basis.

pH=6.5. 54

Figure 5.1 Land limiting constituent (LLC) concept in land

treatment design. 75

xiv

Page 17: Phosphate sorption in soils and reference oxides under

CHAPTER 1

INTRODUCTION

Sustainability is increasingly being considered as a prerequisite

for any decision-making in policies affecting all areas of human

activities. In environmental management, the hazards of improper waste

disposal have been recognized by decision-makers and the general

public alike (Time, 1989). According to Turner (1988), the sustainable

use of environmental resources entails what he called "Conservation

Rule One" that states:

"Maintenance of the regenerative capacity of renewable resources,

and avoidance of excessive pollution which could threaten

biospherical waste assimilation capacities and life support

systems."

The evaluation of waste assimilation capacity of a given system,

and at a given time, has to be compared to a reference or "threshold

value". Liverman et al. (1988) emphasized the importance of such

values for the evaluation of sustainability of all natural systems.

In the case of waste disposal on land, the threshold value would be

the point at which the land application system can no more "absorb"

one or more of the constituents of the waste material applied.

Models that are commonly used to describe and predict the

physical, chemical, and biological interactions occurring in the soil-

waste system (Iskandar, 1981) are needed to evaluate the assimilative

capacity for land application systems. This modeling should be aimed

at (i) identifying the major parameters and processes controlling the

soil-waste system, (ii) selecting suitable sites for waste disposal,

(iii) designing waste disposal systems that will promote maximum waste

1

Page 18: Phosphate sorption in soils and reference oxides under

attenuation, and (iv) designing appropriate monitoring systems to

ensure that the land application schemes are performing as planned.

Models that concern the transformation of pollutants in soils are of a

chemical nature and generally involve two levels of abstraction (Fig.

1.1). Firstly, the real-world components relevant to the system are

identified (soil particles, solutes, etc.). Secondly, the chemical

components are abstracted into mathematical systems of equations

generally of a thermodynamic nature. The outcome of the mathematical

manipulations is generally a prediction of the chemical speciation

state, i.e., which valences of components will be dominant. The

results of the chemical models then can be interpreted and validated

against real-world experiments.

One comprehensive experiment rarely can be designed to include

all the steps in Fig.1.1. Nevertheless, the process can be used to

combine the results of individual chemical modeling studies and

communicate with other models such as water transport models.

Mathematical systems that handle chemical speciation issues require

very sophisticated levels of numerical capabilities (e.g. Westall,

1979; 1982). The models generally assume that a system is at

equilibrium and that reactions are governed by thermodynamics. Strict

application of thermodynamics to the soil solution usually is not

justified (Bohn and Bohn, 1986) and a more thorough knowledge of the

chemical reactants, processes, and products under natural conditions

could help refine, adapt, or correct the results of numerical

speciation modeling (Sposito, 1985).

2

Page 19: Phosphate sorption in soils and reference oxides under

Figure 1.1. The chemical modeling process.

3

Page 20: Phosphate sorption in soils and reference oxides under

Processes occurring at the solid/liquid interface tend to defy

direct thermodynamic interpretation (Stumm, 1986; Morel and Gschwend,

1987). Phosphorus is one of the more important pollutants whose

movement is strongly dependent on interactions with solids. It is

considered rather immobile once it is applied to land because of its

strong tendency to adsorb to soil particles. There is evidence,

however, that P is reaching the groundwater (Gschwend and Reynolds,

1987). Field evaluation of P movement in soils is costly and its

applicability often limited. Chemical laboratory systems simulating

the soil-P system can provide an insight into the type of possible

forms of P in natural systems. The overall objective of this research

was to study the fate of P in the soil system, in particular with

respect to the disposal of municipal wastewater on land.

This dissertation is centered on the abstraction phase depicted

in Fig.1.1, as it relates to the retention or sorption of P by soils.

The different chapters and appendix in this dissertation follow a

"scientific article" format and are, therefore, somewhat independent.

In each chapter the specific objectives of the experiments are

identified, accompanied by a review of the appropriate literature and

a description of the methodology addressing each objective.

In Chapter 2, an overview of the major issues pertaining to the

importance of non point sources of P pollution and soil assimilative

capacity is presented as a review of literature.

In Chapter 3, a system of oxide suspensions was used in the

laboratory under various environmental conditions to evaluate

phosphorus sorption capacity (PSC). Iron oxide (goethite; a-FeOOH) and

4

Page 21: Phosphate sorption in soils and reference oxides under

aluminum oxide (gibbsite; -A^O^) minerals were generated in the

laboratory, characterized, and tested for their capacity to retain P

in monodisperse systems under aerated conditions and a variable pH.

This procedure defined the behavior of each parameter in an inorganic,

aerated environment. In a second stage, the oxides were combined in

experiments to determine the extent of particle-particle interactions

and to assess the effect such interactions have on P sorption. The

principal objective was to verify whether or not the oxides behaved

the same in a combined system, as they did separately.

Further simulation of real-world conditions was achieved through

the study of binary mixtures of Fe and Al oxides under various redox

conditions. In enclosed reactors, low redox potentials were obtained

by microbial activity, whereas positive redox potentials were attained

by manipulating the amount of air available in the system. The

objective of this experiment was to verify that Al oxides were

practically unaffected by reducing conditions, and that reduced forms

of Fe caused significant changes in the P sorption capacity of the

system. A review of applicable oxidation-reduction reactions in soils

and the problems inherent to the thermodynamic consideration of redox

reactions in natural systems is presented in the Appendix.

In Chapter 4, the major soil components that are responsible for

P sorption were identified through a correlation study aimed at

verifying the importance of Fe and Al for P sorption in Massachusetts

soils. The capacity of a soil to retain P was evaluated by a number of

methods that can be used for site evaluation purposes and land

application system design.

5

Page 22: Phosphate sorption in soils and reference oxides under

In Chapter 5, the role of P-soil assimilative capacity was

assessed in the disposal of various types of municipal wastewater. The

particular role of PSC was evaluated and its effect on land

application system design was studied. A comprehensive discussion of

the research results and a summary of its conslusions is provided in

Chapter 6.

The research presented in the following chapters is an attempt to

quantify P sorption capacity of reference oxides and representative

Massachusetts soils by using empirical as well as mechanistic

approaches. The elements required to improve the long-term prediction

of soil assimilative capacity were identified and evaluated in this

study. The experiments presented in the following chapters had the

following overall objectives:

a. to evaluate the effect of varying pH and redox conditions on P

sorption by reference oxides;

b. to develop statistical models of P retention in soil based on

empirically derived PSC parameters;

c. to determine the effect of soil PSC evaluation on the design

of land application systems.

6

Page 23: Phosphate sorption in soils and reference oxides under

CHAPTER 2

BACKGROUND

Introduction

Current efforts in pollution control legislation are aimed at

increasing the role of the land as recipient of waste materials. By

aiming at the "elimination of discharge of pollutants" in water

systems, the Clean Water Act (PL 92-500) has affected land application

by embracing the ecological approach to waste disposal (Noss, 1984).

Land is not only the final depository of waste treatment residues, but

also contributes to the actual treatment of wastewaters (Westman,

1972).

The waste assimilative capacity of a soil is a measure of the

ability of the soil to retain and possibly transform wastewater

constituents (Settergren and Tennyson, 1979). The capacity of soils to

act as "living filters" depends on soil environmental conditions that

govern the physical, microbial, and chemical processes by which

contaminants are transported, sorbed, or degraded in the soil

(Norstadt et al., 1977). Phosphate retention is of particular interest

because wastewaters, in general, are rich in phosphates that can

promote eutrophication in receiving waters (Vollenweider, 1968).

The Phosphate Controversy

Point sources as well as non-point sources contribute P loading

to receiving water bodies. Reduction of this effect can be achieved by

Page 24: Phosphate sorption in soils and reference oxides under

improved point-source control. In a study of the lower Great Lakes,

Switzenbaum and co-workers (1980) estimated that between 11 and 14 %

P loading reduction can be achieved by treating municipal plant

effluents. They also found that 24 to 47 % of the total P loading was

from non-point sources other than direct runoff or atmospheric

precipitation. While it is clear that P seeping through the soil

contributes to this significant non-point fraction, it is difficult to

quantify this contribution.

A survey of current literature shows that there is no consensus

regarding the threat of P percolation through soils. The spectrum of

opinions ranges from a virtually unlimited capacity to retain P

(Canter and Knox, 1985, p.75) to the belief that groundwater seepage

is a significant source of P to surface water bodies (Belanger et al.,

1984). Hook (1983) reviewed both sides of the P controversy and

concluded that field investigations of P movement are difficult to

interpret because of "inherent problems" in sampling and analysis of

soils and soil water, and because of the complexity of P chemistry in

the soils.

While efforts are being made continuously to elucidate P

behavior in soil, several methods for the determination of PSC are

being used currently - although major aspects of P chemistry are still

unknown. Phosphate retention capacity and P release in the soil

solution presently are estimated by one of the models described below.

The Langmuir Equation: An Empirical Model

Any equation that is used for the evaluation P retention capacity

8

Page 25: Phosphate sorption in soils and reference oxides under

of soils should answer the following "real world" questions:

a) If an effluent with a certain concentration of pollutant is

applied to a soil, how much of the contaminant will be

retained by the soil particles?

b) What is the maximum amount of pollutant that can be retained

by the soil particles?

Langmuir originally developed his adsorption model to represent the

retention of gases on solid surfaces. It was only later that this

equation was applied to the adsorption of solutes to soil particles

(Olsen and Watanabe, 1957). The following equation (Langmuir, 1918)

has been used more than any other model to describe and predict the

two parameters identified above:

q = abc/(l+ac) [1]

where, q — amount of sorbed P per unit of soil

a — constant

b — maximum amount of P that can be sorbed per unit of soil

c — concentration of solute, at equilibrium with q.

Experimentally, parameters c and q can be evaluated by

equilibrating soil samples in solution containing different amounts of

P and measuring the amount of P remaining in solution after

equilibrium is attained (Nair et al., 1984). The experimental

procedure has to be performed at constant temperature, which is

probably the reason why the plot of q vs. c is called an "isotherm".

Theoretically, and following equation 1, the isotherm is a smooth

concave curve with a horizontal asymptote equal to the maximum

adsorption value b. This can be demonstrated by considering a modified

form of equation 1:

q - b/[l+(l/ac)] [2]

It is clear that (1/ac) approaches zero, and q approaches b, as c (the

concentration of P in solution) increases. The limit of q when c tends

9

Page 26: Phosphate sorption in soils and reference oxides under

towards infinity is thus b. In other words, b is the maximum sorption

capacity. Veith and Sposito (1977) reported another relationship that

can be derived from equation 1 between q and q/c:

q - -(1/a)(q/c) + b [3]

The term q/c is known as the distribution coefficient. A more

widely used linear transformation of the Langmuir equation is:

c/q — (l/b)c + 1/ab [4]

Equation 4 has been traditionally used to determine maximum sorption

capacity in soils (Olsen and Watanabe, 1957).

The Langmuir equation and its derived forms are based on the

following assumptions (Mansell and Selim, 1981):

1. There is a constant energy of sorption for each contact

surface (no solute-solute reaction, and same intensity of surface

bond).

2. Sorption occurs at localized sites.

3. Maximum sorption corresponds to a complete layer that is one

molecule thick.

These assumptions are similar to the ones advanced by Langmuir

(1918) for the representation of gas-surface interactions. Harter and

Baker (1977) reviewed the major difficulties in transposing these

assumptions to the solute-surface system. A major difference is the

nature of the phenomenon described. The solute system involves the

exchange of a sorbed entity with one or more molecules of solute,

whereas no such exchange is required in the gaseous system. In

addition, in case of high solute concentration, physical bonds could

link solutes to one another in multilayer adsorption (Larsen, 1967)

which contradicts one of the basic assumptions cited. Another

assumption that would not be justified, in general, is the uniqueness

10

Page 27: Phosphate sorption in soils and reference oxides under

of type of bond involved. Soils have more than one type of site for

ion adsorption, leading to a range (not only one) of bonding energies

(Holford et al.f 1974). Finally, it is virtually impossible to

differentiate between adsorption of solutes and surface precipitation,

especially as the concentration of solute increases (Veith and

Sposito, 1977).

Upon testing the Langmuir model with soil samples data many

researchers reported that they did not obtain the linear relationship

of equation 4. Those who did (e.g. Olsen and Watanabe, 1957)

attributed the close fit of their results to the Langmuir equation to

the fact that they had used low concentrations of P. In general, the

results obtained by sorption studies followed a curvilinear

relationship when equation 4 was used. Nevertheless, the Langmuir

equation is generally considered suitable for the empirical

description of P retention in soils and for the interpretation of

analytical data (Harter, 1984).

Numerous variations of the simple Langmuir equation have been

suggested. They can be grouped under the form:

n

q - 2 (at b£ c±)/(l+&± cj_) [5]

i Equation 5 represents the juxtaposition of n simple Langmuir terms.

Values for n of up to n—6 have been reportred to model P sorption by

soils (Goldberg and Sposito, 1985). Holford et al. (1974) obtained a

satisfactory fit for a model where n=2 with soil experimental data.

They interpreted this as indicating the presence of two levels of

bonding energies. However, Sposito (1982) considered the terms a^ and

b^ mere curve-fitting parameters without mechanistic significance.

11

Page 28: Phosphate sorption in soils and reference oxides under

Furthermore, he demonstrated the "Interpolation Theorem":

"Any sorption reaction for which the distribution coefficient is

a finite, decreasing function of the amount sorbed, q, and

extrapolates to zero at a finite value of q, can be represented

mathematically by a two-surface Langmuir equation."

Most soils would have a typical two-layer Langmuir shape with q

linearly approaching zero at increasing values of q/c and q linearly

approaching b (the maximum sorption) at lower values of q/c. The

latter method of maximum sorption capacity determination is most

convenient because it removes the forced linearization of the simple-

Langmuir model and uses the general applicabiity of the two-layer

model.

Empirical vs. Mechanistic Models

In the previous section, the Langmuir equation was presented as

one of the most widely used empirical models of P sorption. Other

empirical models have been reviewed and compared elsewhere (e.g.

Travis and Etnier, 1981; Ratkowsky, 1986). One way of using the

results obtained on a given set of soils to another set is through the

development of statistical relationships between soil components and

the parameters of the empirical models. Another approach is to

develop mechanistic models that have, as parameters, the particular

soil components and characteristics that are effective in P sorption

related within specific chemical reactions representing the sorption

processes that are thought to occur in the soil.

An effective model incorporates the factors that are believed

pertinent to the problem at hand (Hillel, 1977). In the case of P

retention, it is mainly chemical modeling that considers P retaining

factors while mathematical models only describe the retention effect

12

Page 29: Phosphate sorption in soils and reference oxides under

without explaining it (Sposito, 1985). The identification and

quantification of soil properties responsible for P retention will

increase the applicability of empirical models as well as define the

parameters of mechanistic levels.

Phosphorus Retaining Factors in Soils

In any soil system the factors that affect P retention can be

grouped into: (i) the amount of P-retaining components, and (ii)

environmental factors that control the behavior of the P-retaining

components. In non-calcareous soils, Fe and Al are the most effective

such components as was demonstrated in numerous studies (e.g. Syers

and Iskandar, 1981; Singh and Tabatabai, 1977; Laverdiere and Karam,

1984). This has led to the development of land evaluation criteria for

waste-P disposal that are exclusively based on the levels of Al and Fe

in the soil (Breeuwsma et al., 1986).

An appropriate model of P retention in acid soils should thus

include Fe and Al oxides and the environmental conditions that could

affect their behavior. Soil pH and redox conditions have been shown

to significantly affect P retention (e.g. Holford and Patrick, 1979).

Quantifying this effect has been achieved more successfuly for

pH (e.g. Barrow, 1984) than for oxidation-reduction conditions.

The Use of Reference Oxide Minerals

Recent progress in the understanding of P retention can be

attributed largely to the study of aqueous synthetic mineral systems.

Oxides of Fe and Al generally are present in the soil in very low

13

Page 30: Phosphate sorption in soils and reference oxides under

concentrations and therefore can be much more effectively studied in

synthetic suspensions (Dixon et al., 1983). The relevance of such

laboratory studies to actual field conditions naturally is a crucial

consideration.

Goldberg and Sposito (1984a; b) determined that there was

significant correlation between the results obtained by different

researchers working with synthetic oxides to retention data obtained

on a variety of non-calcareous soils. Most of the chemical speciation

models used in the field of environmental planning are based on

thermodynamic data obtained from reactions of pure chemical products

in controlled experiments (Sposito, 1985).

The gap between laboratory studies and in-situ research can be

reduced through the inclusion of a greater number of relevant factors

in laboratory experiments or through the improvement of field

techniques in order to identify the various contributing processes.

The research presented herein is an attempt at the former approach.

Research Needs

In this Chapter, it is argued that the evaluation of PSC models

is hindered by the fact that the chemistry of P sorption is not well

understood. Another obstacle is the difficulty to include appropriate

soil environmental factors.

Oxide-P laboratory studies have improved our knowledge of P

behavior in soil. Such studies, however, should be extended to include

mixtures of well defined adsorbents (Hingston, 1981). In practice,

such studies should result in improved quantification of the retention

14

Page 31: Phosphate sorption in soils and reference oxides under

process in order to be used in transport models (Sposito, 1985) and

site selection for land application of wastes (Veneman, 1982; Kleiss

and Hoover, 1986).

15

Page 32: Phosphate sorption in soils and reference oxides under

CHAPTER 3

PHOSPHORUS SORPTION ON REFERENCE OXIDES:

pH AND REDOX EFFECTS

Introduction

Many areas in environmental research involve the determination of

macroscopic processes using parameters microscopic in nature. For the

evaluation of pollutant retention by soils, two macroscopic approaches

are used commonly. The first is the partition coefficient between

solution and solid components (Lyman, 1982). The second approach is

the use of sorption isotherm data (Tofflemire and Chen, 1976). While

both evaluations are based on specific assumptions concerning single-

particle behavior (e.g. energy of adsorption, maximum monolayer ionic

coverage, particle charge), these microscopic parameters are

quantified by the use of macroscopic, indirect techniques of an

empirical nature (Harter and Baker, 1977).

In practice, sorption coefficients are often used as mere curve -

fitting parameters (in pollution transport models, for example),

precluding any measurement -direct or indirect- of the sorption

phenomenon itself (Jury, 1983). An assessment of sorption processes

based on independently measurable parameters that relate to accepted

mechanistic sorption models (Westall and Hohl, 1980) will improve

significantly the validity of environmental models (Sposito, 1985).

Modeling of P retention in acid soils has been studied

extensively and it was shown that Fe and A1 hydroxides can be used as

16

Page 33: Phosphate sorption in soils and reference oxides under

reference models that mimic the sorption properties of soil (Goldberg

and Sposito, 1984a; b).

Phosphorus Retention Mechanism

Phosphate ions adsorb to and desorb from the oxide surfaces

through ligand exchange (Stumm et al., 1980). This adsorption

mechanism involves the release of ligand hydroxyl ions (0H~) from

metal ions (S) in an oxide mineral, and its replacement with another

ligand, in this case phosphate ions. The reactions are as follows:

S0H(s) + H+ - S0H+2(s) [1]

S0H(s) - S0'(s) + H+(aq) [2]

S0H(s) + H3P04(aq) = SH2P04(s) + »2° [3]

S0H(s) + H3P04(s) - SHp0-4(s) + H2° + H+(aq) [4]

S0“ (s) H- H3P04(aq) = spo"4(s) + H20 + 2H+(aq) [5]

In general (Goldberg and Sposito, 1985):

aS0H(s)+HbP04(aq)b'3+cH+(aq) “ SaHcP04(s)+bH20(s)+<a-b>0H‘(aq) t6)

where b-3 is the degree of protonation of phosphate ions. Each one of

equations [1] to [5] has an equilibrium constant Ka(^s of the form (Morel

and Gschwend, 1987):

^ads ^chem * ^coul

where Kchem — chemical (or intrinsic KSa(intr)) energy term

Kcoul “ coulombic (electrostatic) energy term

[7]

17

Page 34: Phosphate sorption in soils and reference oxides under

In equation [4] for example, the intrinsic constant KSa(intr) can be

written (Goldberg and Sposito, 1984a):

[SHP0-4][H+]

KSa(intr) “ --- exP<-F A1) («] [S0H][H3P04]

where F — Faraday constant (Cmol )

= surface potential (V)

R = molar gas constant (Jmol-^ K"^)

T = absolute temperature (K)

Reactions [1] and [2] are independent of the ion being adsorbed (i.e.

phosphates) and can thus characterize the particular surfaces being

investigated (Hohl et al., 1980).

Phosphate Sorption by Goethite

Goethite was used by Sigg and Stumm (1980) as a model metal

hydrous oxide to illustrate ligand exchange reactions involving

phosphate ions. They concluded that P ions formed "inner sphere

complexes" by binding directly to surface Fe atoms. Atkinson et al.

(1974) studied dry samples of goethite treated with P ions and

estimated that most of P is sorbed as binuclear complexes, to the

surface of goethite. In this type of complex, two atoms of surface Fe

are connected to two oxygen atoms and finally to one P ion. Based on

the total surface area, such an arrangement would lead to an

average coverage of 2.8 +0.7 pmol P m according to Borggaard (1983).

Assuming a monodentate complex between surface 0-H and P ions,

Anderson et al. (1985) estimated the average capacity of goethite for

P coverage at 4.5 /imol m ^. Goldberg (1983) reported values between

18

Page 35: Phosphate sorption in soils and reference oxides under

_ 2 2.5 and 7 /imol P m of goethite. The area occupied by one phosphate

2 2 molecule is 0.235 nm which leads to 7.1 /xmol Pm (Van Riemsdijk and

Lyklema, 1980a) . The capacity of 7 /imol P m~ would thus provide the

complete coverage of goethite with phosphate molecules. Such coverage

would only occur, at adsorption capacity, at the edge surface of

goethite, because it has the 0-H groups that contribute to P-

adsorption (Atkinson et al., 1974).

Phosphate Sorption by Gibbsite

Gibbsite crystals, like goethite, have their reactive 0-H groups

at the edge surface. This surface was found by Van Riemsdijk and

Lyklema (1980a) to represent about 14% of the total BET area and to

-2 -2 adsorb 7.1 /xmol Pm of edge area or 1.0 /imol Pm of BET area.

Based on total BET area, P adsorption capacity of gibbsite was found

to vary between 1.6 and 2.8 /imol P m’^ (Goldberg, 1983).

Van Riemsdijk and Lyklema (1980a; b) studied sorption reactions of

P on gibbsite surfaces at increasing P concentrations and found that

adsorption is the rate controlling P retention mechanism at low P

_ 3 concentrations. At P concentrations greater than 5000 mmol m , P

precipitated on the gibbsite surface (Van Riemsdijk and Lyklema,

1980a; b).

The various sorption models that combine chemical and coulombian

effects have been reviewed extensively (Hingston, 1981; Goldberg and

Sposito, 1985; Sposito, 1985; Morel et al., 1981; Vestall and Hohl,

1980; Dzombak, 1986; Westall, 1987). These models rely on a thorough

characterization of the solid surfaces, including size, reactivity,

and electrostatic parameters (James and Parks, 1982).

19

Page 36: Phosphate sorption in soils and reference oxides under

Empirical and mechanistic methods of sorption evaluation assume a

relationship between experimental, crystallographic, and chemical

representations of the phenomenon. This relationship should, however,

be tested for any set of adsorbent/ion system. Such a verification

would evaluate the sorption capacity of the surfaces and also

characterize these surfaces so that the effects of system

perturbations can be compared to reference systems.

Phosphate Sorption by Mixtures of Oxides

Phosphate-retaining constituents of the soil generally are

assumed to have a simple cumulative effect on the macroscopic sorption

capacity of the soil. This assumption underlies site suitability

assessment for waste disposal and the computation of equilibrium

constants for chemical speciation programs (Goldberg and Sposito,

1984a). No explicit validation of this assumption has been proposed

and recent reviews have emphasized the need to study ion retention by

mixtures of adsorbents (Hingston, 1981). Multicomponent studies can be

of particular interest in systems where changing environmental

conditions affect each one of these components differently.

Common models describing anion sorption, assume a constant

sorption energy for each contact surface (Langmuir equations), or

decreasing bonding energies with increasing site coverage by sorbed

ions (Freundlich equation; Mansell and Selim, 1981). Surface

heterogeneities are not considered by these models and are averaged

out for simplicity (Van Riemsdijk et al., 1986). Extension of such

models to include heterogeneous surfaces is possible by adapting

20

Page 37: Phosphate sorption in soils and reference oxides under

heterogeneous gas-solid models (Sircar, 1984) to the electrolyte-solid

system (Van Riemsdijk et al., 1986). While these models are empirical

in nature, they may lead to an improved description of multicomponent

systems. Parker and Jardine (1986) evaluated the effects of surface

heterogeneities on ion transport models by considering the solid phase

equivalent fractions of exchanging ions and assigning particular

thermodynamic parameters to each of the fractions. Their model

predicted the behavior of surfaces that have heterogeneous ionic

affinities on the same particle. Other models concentrate on

heterogeneity caused by the presence of mixtures of oxides (Honeyman,

1984). Assuming that all particles of each oxide in the system behave

in an average manner, the representation of the whole system of oxides

could be obtained by summing the separate thermodynamic parameters in

the system. The weighting factor of each oxide would be dependent on

its mass fraction in the system. Schwarz et al. (1984) applied this

concept to the evaluation of electrical properties of mixed oxide

suspensions. Using silica-alumina mixtures, they verified that the

zero point of charge of various mixtures was linearly affected by the

weight fraction of the pure components. Kuo and Yen (1988) suggested

replacing the weight fraction with a surface area fraction that would

represent the surface charge density of each oxide.

Whether weight fraction or surface area fraction is used, there

is an underlying assumption of non-interference between oxide

particles. As suggested by Honeyman (1984), such an assumption is not

necessarily justified in all cases. He enumerated the following

interactions that alter the non-interference assumption and,

consequently, the partial weight or surface area concept, (i)

21

Page 38: Phosphate sorption in soils and reference oxides under

coagulation or heterocoagulation that may block surface sites, (ii)

changes in the surface charge, and (iii) oxide dissolution. Honeyman

(1984) used the term "adsorptive additivity" to describe the non¬

interference concept. In the study reported in the following section,

the concept will be referenced as the "sorptive additivity" because of

the difficulty in identifying the phenomenon (adsorption vs.

precipitation) responsible for P retention by oxides.

The sorptive additivity concept should be tested on systems of

mixtures of oxides to aid in the extrapolation from monocomponent

systems to more complicated systems that are closer to the soil

conditions.

Redox Effects on P Sorption

The succession of wet-dry cycles in the soil enhances generally

its capacity to retain P. For example, Hill and Sawhney (1981)

reported a regeneration of soil retention sites below septic system

leaching fields receiving intermittent doses of wastewater. Bradley et

al. (1984) found that extractable P decreased in soils that are

waterlogged seasonally. In non-calcareous soils, this increase in P

retention was attributed to the formation of Al-P and Fe-P minerals

(Sah and Mikkelsen, 1986). It is expected that only in cases where

flooding causes an increase in soluble Fe, subsequent drying increases

the P retention capacity of soils (Holford and Patrick, 1981). While

details concerning the exact mechanisms involved in the regeneration

of sorption sites are still unknown, there is little disagreement in

the literature on the effect of wetting and drying cycles. The

effect of flooding on the availability of P, however, is not known.

22

Page 39: Phosphate sorption in soils and reference oxides under

Reducing soil conditions generally are associated with the

release of P into the soil solution (Ponnamperuma, 1972; Savant and

Ellis, 1964). This action is attributed to the reduction of ferric to

ferrous ion and the release of P held by ferric ions (Bradley et al.,

1984). Such conditions are believed to contribute to P contamination

of water sources by land disposal systems in areas with high water

tables (Gilliom and Patmont, 1983). In some cases, reducing conditions

have also been shown to enhance P retention by soils (Holford and

Patrick, 1979). This rather surprising effect is believed to be caused

by the formation, under suitable pH conditions, of a highly reactive

ferrous-ferric iron complex (Khalid et al., 1977).

Previous studies of redox-phosphate interaction have failed to

positively identify P-retaining compounds under reduced conditions

(Willett, 1983), and have considered generally only precipitation-

dissolution equilibrium reactions (Willett, 1985), and not sorption

reactions. Willett and Cunningham (1983) published a report on P-

reference oxide interactions under controlled pH and redox conditions.

They used H2 to decrease redox levels, a method which leads to rapid

redox changes not believed to represent real soil conditions. A

microbially mediated system could improve the simulation of such

conditions and their effect on soil chemical processes.

The objective of this research was to determine the effect of pH

and reduction-oxidation conditions on oxide-P systems in the presence of

chemical constituents of wastewater. This was attained by pursuing the

following specific objectives:

23

Page 40: Phosphate sorption in soils and reference oxides under

(i) characterize the monocomponent goethite and gibbsite

particles, and define their PSC using various methods;

(ii) test the sorption additivity concept in the goethite-

gibbsite system; and

(iii) evaluate the effect of varying pH and redox conditions

on mixtures of goethite and gibbsite oxides and assess

their PSC values.

Materials and Methods

Reference Oxides

Goethite (a-FeOOH) crystals were prepared by reacting a 0.4M

ferric nitrate (Fe(NO^) ^ • 9^0) solution with a NaOH solution at an

OH/Fe ratio of 1.5 for 50 hours (Atkinson et al., 1968). Sodium

hydroxide was then added to adjust the pH to 11.8. The precipitate was

maintained at 60°C for 4 days and dialyzed until no nitrate was

detected in the water bath.

Gibbsite (AI2O3.3H2O) crystals were prepared by adding 2.5M

NaOH solution to a 1.4M aluminum chloride (AICI3.6H2O) solution until

pH 4.6 was reached (Gastuche and Herbillon, 1962). The resulting

suspension was heated to 40° C for 2 hours then dialyzed against

distilled water for 23 days at 60° C with a daily change of distilled

water.

Goethite and gibbsite were identified by X-ray diffraction

(Schwertmann and Taylor, 1977; Hsu, 1977) and characterized by

electron microscopy and BET analysis to facilitate specific surface

24

Page 41: Phosphate sorption in soils and reference oxides under

area computations (McLaughlin et al., 1981). The oxides were stored as

aqueous suspensions throughout the experiment.

Sorption Isotherms

Five-milliliter aliquots of oxide suspensions containing about 75

mg of adsorbent were added to 45-ml polycarbonate centrifuge tubes

containing from 0 to 20 /imol Na2HP0^ before bringing the total volume

to 40 ml with 0.1M NaCL solution. Most tubes were run in duplicate.

Initial pH was adjusted to pH 5. The range of P concentrations

selected (0 to 300 mmol m after equilibration), was relatively low

and associated with adsorption with little possibility for

precipitation to occur especially at the relatively short

equilibration times used (Goldberg, 1983).

The centrifuge tubes were gently shaken for 24 hours on a

reciprocating shaker before centrifugation at 12,500 rpm for 30

minutes. Supernatant solutions were passed through a 0.20 /xm Metricel

filter, and the pH and P content (Murphy and Riley, 1962) were

measured in the filtrate. The amount of P sorbed was calculated by the

difference between the initial and final P in the centrifuge tube

filtrate.

Two parameters of maximum PSC (Pjj= Langmuir maximum, and P^=

Interpolation) were derived from the P sorption isotherm as follows:

(i) Pj^ was calculated from a linear form of the following

equation:

q = a PLc/(l + ac) [1]

where, q = moles of P sorbed per unit weight of soil, a = a constant,

c — moles of P in solution at equilibrium.

25

Page 42: Phosphate sorption in soils and reference oxides under

(ii) Pj was derived graphically by plotting the partition

coefficient (=q/c) against q and determining the intercept with the

q-axis (Sposito, 1982).

pH Effects

The same batch equilibration procedure as above was used to study

pH effects on P sorption. The initial P concentration was identical in

all centrifuge tubes (200 mmol m ) and pH levels were varied between

3 and 11 by the addition of IN HC1 or IN NaOH. The maximum sorption

value obtained in this experiment was considered the maximum

experimental PSC (PeXp)•

Chemical Characterization

The electrochemical properties of the oxides were determined by

acid-base titration and the procedure of Hohl et al. (1980) was

used to determine the intrinsic constants and the point of zero charge

(ZPC) of the oxides.

Crystallographic Characterization

The oxide crystals were observed by transmission electron

microscopy (TEM) to determine (i) the degree of homogeneity of the

crystals and (ii) the average dimensions of goethite and gibbsite

particles. Phosphate sorption capacities were estimated from the

distribution of reactive sites of the goethite (Hingston et al., 1972)

and gibbsite crystals (Van Riemsdijk and Lyklema, 1980a). The

thickness of the oxides was determined by gold shadowing in which gold

26

Page 43: Phosphate sorption in soils and reference oxides under

was sprayed at a 60° angle on oxides mounted on glass slides and the

resulting shadow was measured on electron micrographs. This thickness,

for both oxides, indicates the "edge area" (Van Riemsdijk and Lyklema,

1980a) where most of the 0H(H) groups available for P adsorption are

o located. By assuming an average area of 0.235 nm per mole of P

_ 2 adsorbed, equivalent to 7.1 /imol m (Van Riemsdijk and Lyklema,

1980a), an estimate of the maximum adsorption capacity of the edge

area can be calculated. In general, P adsorption capacity can thus be

expressed as:

q = 7.1 [ A r / (V X)] [9]

where q =

A =

r =

V =

X =

P adsorption capacity (mmol kg~ or /imol g" )

total surface area of oxide (mz)

fraction of total oxide surface area in edge area

total volume of oxides (nr)

density of the oxide (g m-'5)

Surface area was also determined from BET analysis (Everett and

Otterwill, 1969). The BET isotherm for N2 adsorption of each of these

oxides was obtained after drying of duplicate samples , and the BET

value was used to evaluate maximum PSC using equation 9 with the BET

surface value as parameter A (Everett and Otterwill, 1969).

Sorption by Mixture of Oxides

Phosphate sorption in the mixture of oxides was evaluated by

transferring 2.5-ml aliquots from each of the 15-g L stock

suspensions to 40-ml centrifuge tubes and running the same

equilibration, filtration, and P measurement procedures previously

presented.

27

Page 44: Phosphate sorption in soils and reference oxides under

The effect of pH on P sorption by the mixture of oxides was

evaluated at initial P solution concentrations of 100, 200, and 300

-3 mmol m . Adjustment of pH levels was done through the addition of a

few drops of HC1 or NaOH at concentrations of 0.2 or 1.0 M. Adsorbed P

was then determined as described previously.

Redox Effects

Experimental Apparatus. A system to control redox potential

conditions in soil suspensions was adapted from Patrick et al.

(1973). Experimental containers were constructed from clear acrylic

plastic (plexiglas) as shown in Fig.3.1.

Redox potential was measured continuously with a platinum

electrode (Fig.3.2) constructed according to Dirasian (1968) and

tested in a poised ferrous/ferric solution (Light, 1972). The Pt

electrodes were tested periodically during the course of the

experiments. Each electrode was connected to a voltmeter (Accumet

Model 805 MP, Fischer Scientific). The voltmeter had

microprocessor-based features that allowed the use of minimum and

maximum limit controls. If the redox potential in the reaction

container fell below a certain preset value, an electric signal

would activate a solenoid valve (Dynamo single solenoid 4-way valve.

Model 02110k00), which, through an electronic switching mechanism

admitted air into the reactor (Fig.3.3).

28

Page 45: Phosphate sorption in soils and reference oxides under

5

1. Pressure regulator 2. Gas inlet

3. Glass electrode 4. Salt bridge

5. Reference electrode

6. Platinum electrode

7. Brass o-seal connectors 8. Septum stopper

9. Stirring bar

10. Air trap

Figure 3.1. Cross-section of the reactor used for controlling redox

potential in soil suspensions.

29

Page 46: Phosphate sorption in soils and reference oxides under

KC1 agar

Semi-permeable membrane

___ Epoxy cement

t—-Acrylic tubing

_Copper wire

_Silver solder

_Epoxy cement

_Platinum wire

(a) (b)

Figure 3.2. Cross-section of salt bridge (a) and platinum

electrode (b).

30

Page 47: Phosphate sorption in soils and reference oxides under

To pH

meter

pin 14

Figure 3.3. Diagram of electronic switching mechanism used to control

the Eh in soil suspensions.

31

Page 48: Phosphate sorption in soils and reference oxides under

Experimental Procedures. Six reaction vessels were used to

control redox conditions. Each reactor contained the following

mixture:

2.0 L of simulated wastewater

2.0 g of reference oxides

50 mg of dextrose

1.0 g of straw

1.0 g soil.

The composition of simulated wastewater was adapted from Van

Riemsdijk et al. (1977) to represent approximately one tenth of the

strength of municipal wastewater. Per liter of suspension the

following amounts of reagent grade chemicals were added and allowed to

dissolve:

73 mg NaCl

30 mg CaS04.2H20

25 mg CaC^^I^O

20 mg NH4C1

14 mg MgCl2

8 mg NaF

5 mg KC1

4 mg KNO^

5 mg KH2P04

The ingredients in the reactor mixture, minus the oxides, were

allowed to equilibrate. Purpose of this preliminary step was to

promote microbial action and remove the largest amount of P possible

from solution. When P was no longer detected in solution the oxide

suspensions were introduced. Five reactors received 1:1 mixtures of

goethite and gibbsite, and one reactor received gibbsite only. The

equilibrium redox potential in the six containers was adjusted to

reach and maintain one of the following Eh values: +500, +300, 0, or -

50 mV. These values cover the range of redox potentials common in

well-drained to waterlogged soils (Patrick and Mahapatra, 1968). The

reactor vessels were placed on magnetic stirrers allowing continuous

mixing of the suspensions.

32

Page 49: Phosphate sorption in soils and reference oxides under

During the incubation period, the pH was monitored and manually

maintained at pH 6.5 by addition of IN NaOH or IN HC1. Monitoring of

the pH was required several times each day at the beginning of

incubation, but little adjustment was needed by the time redox

potentials stabilized. Holford and Patrick (1981) found the pH

stabilized after 8 days of incubation of suspensions where Eh was

controlled and the pH was not controlled.

Phosphate Isotherms. At the end of each incubation period

(approximately one week after the system stabilized at any given Eh

and pH levels), samples from the stirred suspensions were removed and

transferred to centrifuge bottles for P sorption evaluation. Samples

from the reactors in the reducing treatments were removed

anaerobically with a syringe and added to air-tight centrifuge

bottles containing solutions of P and filled with N gas (Holford and

Patrick, 1981).

Microscopic Analyses

Microprobe analysis was attempted on samples extracted from the

reactors (Lee and Kittrick, 1984; Sawhney, 1973) to determine the

location of P-sorption sites. The degree of resolution of the

instrument was, however, insufficient for the size of the oxides

involved.

To visually inspect the shape of the goethite and gibbsite

minerals, samples of the suspensions were mounted on glass slides and

observed under the transmission electron microscope. Samples from

33

Page 50: Phosphate sorption in soils and reference oxides under

reactors with a reducing environment were dried in a N2-atmosphere.

This was done as a precaution to prevent reoxidation although many

researchers who studied the micromorphology of anaerobic environments

did not provide such anaerobic conditions during sample preparation

(Lee and Kittrick, 1984; Jeanson, 1972).

Results and Discussion

Optical and Chemical Characterization of the Oxides

The goethite crystals exhibited a somewhat poorly defined, lath¬

like structure and tended to aggregate strongly. They appeared similar to

synthetic goethite prepared by Dixon et al. (1983). Gibbsite appeared

as pseudohexagonal, plate-like crystals, the typical form described

by Hsu (1977).

The electrochemical properties of the oxides were determined

based on a theoretical representation of P sorption onto the crystal

surface sites. It is generally accepted that P adsorption occurs

through ligand exchange in accordance to reactions [1] to [5].

According to that theory, the first two protonation-dissociation (or

acid-base) constants are independent of P, depend only on surface

characteristics of the oxides and can be determined by acid-base

titration.

For goethite, the intrinsic acid and base constants were 8.4 and

10.6 respectively, with the ZPC equal to the average of these two

values or 9.5 (Fig.3.4). The constants for gibbsite were equal to 3.2

for acidity, 6.1 for basicity, and 4.6 for ZPC (Fig.3.5).

34

Page 51: Phosphate sorption in soils and reference oxides under

PH

Figure 3.4. Titrimetric surface acidity characterization of goethite.

35

Page 52: Phosphate sorption in soils and reference oxides under

PH

Figure 3.5. Titrimetric surface acidity characterization of gibbsite.

36

Page 53: Phosphate sorption in soils and reference oxides under

These values are not in agreement with others reported in the

literature. Goldberg and Sposito (1984a), for example, reviewed Fe and

A1 oxide data and found that the ZPC of Fe and Al oxides averaged 8.0

and 8.2 respectively. Theoretically, such discrepancies may be

attributed to interactions of the background electrolyte (here,

NaNO^), with the oxide surface instead of just serving as a swamping

electrolyte (Sposito, 1981a). In practice, such interactions have not

been shown to occur between NaNO-j and the oxides of Fe and Al

(Goldberg, 1983).

The major properties of both oxides are summarized in Fig.3.6.

The principal characteristic for P sorption is the specific area, in

particular the specific edge area. The two methods used for the

evaluation of total specific area yielded significantly different

dimensions for goethite, with a 145 % difference between the BET and

the TEM measurements. The tendency of goethite to strongly aggregate

probably caused a decrease in BET surface compared to the specific

area of dispersed crystals. Also, the poor crystallinity of goethite

may have caused overestimation of the specific surface area.

In contrast to goethite, BET and TEM measurements in gibbsite

resulted in practically identical values. The specific area measured

43 m^ gwhich is in the range of values reported in the

literature (Goldberg, 1983).

The edge area was evaluated by gold shadowing and by TEM

calculations. It should be noted that the estimation of specific

surface area for gibbsite is independent of crystal thickness. Both

methods can be used to verify one another. This is illustrated in

Table 3.1 which shows that variable crystal thicknesses (D)

37

Page 54: Phosphate sorption in soils and reference oxides under

(a) Goethite.

72 nm_

5 U^y -3 V Density: 4.3 kg dm

Specific Surface Area: BET - 53 + 1

TEM - 130 + 20 m" g"1

ZPC - 9.5

(b) Gibbsite.

L - 225 + 17 nm; D - 23 + 3 nm

Density: 2.4 kg dm"^

o .1 Specific Surface Area: BET - 43 + 3 m g

TEM - 44 + 4 m2 g'1

Edge Area — Approximately 15 % of total area

ZPC - 4.6

Figure 3.6. Summary of major characteristics of the two oxides.

38

Page 55: Phosphate sorption in soils and reference oxides under

Table 3.1. Demonstration of the independence of specific edge

area from the dimension of crystal thickness (D) in

gibbsite, assuming a perfectly hexagonal crystal.

D L

Total

Specific

Area

Specific

Edge

Area

Edge Area

Percent of

Total Area

2 -1 2 -1 -nm- m g m g %

17 208 57.0 8.0 14

17 225 56.4 7.4 13

17 242 55.9 6.9 12

20 208 49.7 8.0 16

20 225 49.1 7.4 15

20 242 48.5 6.9 14

23 208 44.2 8.0 18

23 225 43.6 7.4 17

23 242 43.1 6.9 16

26 208 40.1 8.0 20

26 225 39.5 7.4 19

26 242 38.9 6.9 18

39

Page 56: Phosphate sorption in soils and reference oxides under

approximately yield the same specific edge area for a given size of

crystal represented by the dimension L (L refers to the diagonal

dimension of a hypothetical hexagon as shown in Fig.3.6b).

Phosphate Sorption by Monocomponent Systems

Of the two oxides, goethite had the highest P sorption capacity

(Fig.3.7). No maximum P sorption plateau was reached, due probably to

P precipitation on the surface of the oxides at the highest P

concentrations used. In order to avoid precipitation, the P adsorption

envelopes (sorption variation with pH) were run at initial

concentrations of 200 mmol m , and the maximum experimental sorption

capacity was determined.

Table 3.2 summarizes the results of the different techniques used

to evaluate P sorption capacity of the two oxides. In goethite the

BET, P_v_, and PT led to comparable results of around 110 mmol P kg"^.

The Pj and especially TEM resulted in higher estimates of PSC. It was

to be expected that would be greater than P^, because the isotherm

did not reach a plateau (see discussion in Chapter 2). As discussed

previously, the poor crystallinity of goethite may cause a much greater

effective surface area and higher PCS values.

Phosphorus sorption capacity values for gibbsite varied from

about 52 to 130 mmol P kg Results from the BET and TEM estimates

where very close because of the similar evaluation of the size of

gibbsite crystals. The difference between Pj^ and P^ is greater in

gibbsite than in goethite probably because of the relatively steeper

slope of the gibbsite isotherm at higher P concentrations (Fig.3.7).

40

Page 57: Phosphate sorption in soils and reference oxides under

PH

OS

PH

AT

E

SO

RB

ED

(MM

OL

/KG

)

Figure 3.7. Phosphate sorption in goethite, gibbsite, and a 1:1

mixture (1.9 g L"^) of both oxides. Electrolyte: 0.LM

NaCl, pH: 5.5- 6.5, initial solution P concentrations:

0-500 mmol m""^.

41

Page 58: Phosphate sorption in soils and reference oxides under

Table 3.2. Maximal phosphate sorption capacity of the two

oxides derived by five methods: indirect surface

measurement (BET), geometric calculations (TEM),

batch equilibration (P^p) , Langmuir (P^), and

interpolation (Pj).

_Maximum_P_Sorption_Capacity_

Oxide BET TEM Pexp PL Px

mmol kg~^

Goethite 113 + 2 277 +43 110 110 165

Gibbsite 52 + 6 53 + 4 75 77 130

42

Page 59: Phosphate sorption in soils and reference oxides under

Sorption Isotherms of Binary Mixtures

The isotherms obtained from the binary mixtures and each of the

monocomponent suspensions are shown on Fig.3.7. The isotherm of the

binary system is located clearly between the goethite and gibbsite

isotherms, with goethite having the highest P sorption capacity.

pH Effects

At the lowest initial P concentration of 100 mmol m , the curves

for goethite and the mixture of oxides were essentially identical.

(Fig.3.8). They showed a decrease in P sorption with increasing pH.

At pH<4.5 gibbsite dissolved but the sites available on the goethite

particles must have been sufficient for the sorption of P in the

binary suspensions. The apparent P-goethite affinity was also the

case at higher pH levels although goethite and gibbsite were present

in equal concentrations. Whether the shape of the curves in Fig.3.8

is an indication of the high affinity of P on goethite could not be

verified. From the fact that at pH>5 the three systems showed

parallel curves, one could conclude that gibbsite contributed also to

the sorption process.

At initial P concentrations of 200 mmol m , the mixture of

oxides had a retention capacity that closely followed the average

between the goethite and the gibbsite monocomponent sorption

capacities (Fig.3.9). The mixture envelope was generally closer to

the goethite evelope than to the gibbsite. However, the concentration

of P in solution was clearly beyond the retention capacity of goethite

particles at all pH levels.

43

Page 60: Phosphate sorption in soils and reference oxides under

SO

HB

EO

PH

OS

PH

AT

E

(MM

OL

/KG

l

130. 0-

120. O-

1 10. O-

100. 0-

90. 0-

80. O-

70. O-1 I

ODSORStiNTs

—■— GOETWITE mix TUBE

~ GI9BSITE

Figure 3.8. Phosphate sorption of goethite, gibbsite and a 1:1 mixture

(1.9 g L ) at different pH levels. Electrolyte: O.LH

NaCl initial P concentration: 100 mmol m"^ (53 mmol P kg" solids).

44

Page 61: Phosphate sorption in soils and reference oxides under

PH

OS

PH

AT

E

SO

RB

ED

(MM

OL

/KG

)

130. 0-

120. O-

1 10. 0-

100. 0-

90. 0-

80. O-

70. O-

60. 0-

SO. 0-

«0. 0-

30. 0-

20. 0-

10. 0-

. 0-

1. o

PCSOR0ENTS —— GOETWITE --«■* MIXTURE

GI93SITE

I II.'. 3. O S. 0 7. 0 9. 0 11.0 13. 0

PH

Figure 3.9. Phosphate sorption of goethite, gibbsite and a 1:1 mixture

(1.9 g L’^) at different pH levels. Electrolyte: 0.1M

NaCl initial P concentration: 200 mmol m (106 mmol • 1

P kg" solids).

45

Page 62: Phosphate sorption in soils and reference oxides under

_ 3 At the higher concentration of 300 mmol P m , the mixture

envelope did not follow the general trend of the monocomponent systems

(Fig.3.10). The monocomponent suspensions had a generally smooth

decrease of P retention with increasing pH levels whereas the mixed

suspension had two features that could not be explained adequately,

(i) around pH 4, the mixture of oxides retained more P than would be

expected by the additivity concept, and (ii) the mixture curve

fluctuated at pH>5 instead of smoothly and gradually decreasing (Fig.

3.10).

There is no specific Fe-Al-P complex reported in the literature

(Lindsay and Moreno, 1960; Stumm and Morgan, 1981) to explain the

apparent interaction between goethite and gibbsite. Samples of the P-

treated mixtures were observed by electron microscopy and did not show

any new deposits on the minerals. There was no difference in the

shape of the oxides, even at low pH levels. We can thus assume that

the only effect of very low pH (<4) is to dissolve gibbsite and

probably increase the potential of the remaining sorption sites to

retain P.

Redox Effects

Three distinct colors developed in response to different redox

conditions in the reactors containing goethite-gibbsite mixtures. The

reactors with oxidizing environments maintained at Eh values of +500

and +300 mV, had a brown color described by 10YR 6/6 (Munsell Color,

1975). The reactors with reducing environments, after about two weeks

46

Page 63: Phosphate sorption in soils and reference oxides under

SO

RB

ED

PH

OS

PH

AT

E

(MM

OL

/KG

l

Figure 3.10. Phosphate sorption of goethite, gibbsite and a 1:1 mixture

(1.9 g L'^) at different pH levels. Electrolyte: 0.1M

NaCl, initial P concentration: 300 mmol m(160 mmol P

kg' solids).

47

Page 64: Phosphate sorption in soils and reference oxides under

at 0 and -50 mV, developed a greenish color of 5Y 5/6, characteristic

of reduced iron in the soil. The reducing environments subsequently

were reoxidized by discontinuing the flow of N2 and allowing air to

enter. The color quickly turned from green to brick red (5YR 5/8). The

reactor containing a monocomponent suspension of gibbsite did not show

any color changes due to alteration in redox conditions.

Observed under the electron microscope, samples from the reducing

reactors showed very few goethite particles remaining,whereas the

shape and abundance of gibbsite particles were unchanged. Gibbsite

remained unchanged in samples from the oxidized reactors, whereas

goethite minerals were rare. Instead, smaller particles that may have

been Fe precipitate formed aggregates somewhat similar to goethite.

Examination of reoxidized samples by X-ray diffraction did not reveal

any Fe in crystalline form. It was surmised that after dissolution and

reoxidation, Fe precipitated as amorphous ferric oxides. To confirm

this interpretation, extracts from three of the reactors were analyzed

for dissolved Fe. The +500 mV treatment had 0.14 mg Fe L”^, upon

reduction to -50 mV, dissolved Fe reached 281 mg L ^. Finally, iron in

the reoxidized reactor dropped to 3 mg L ^. Aluminum varied much less

than Fe between treatments. The levels reached 0.47, 0.53, and 2.28 mg

Al L ^ for the +500 mV, the -50 mV, and the reoxidized reactor

respectively. The increase of Al in the reoxidized reactor can not be

explained.

Posphorus Sorption in Oxidized Systems

Phosphorus sorption (Fig.3.11) reached a maximum of 40 mmol kg

in both oxidized treatments of +300 and +500 mV. This level of PSC

48

Page 65: Phosphate sorption in soils and reference oxides under

140 Sorbed P (mmol/kg)

120 -

100 m

80 e

-50 mV

Figure 3.11. Phosphate sorption on 1:1 goethite and gibbsite

suspension in simulated wastewater. pH—6.5.

49

Page 66: Phosphate sorption in soils and reference oxides under

represents a reduction of more than 50% from the sorption capacity if

the oxides were in a "pure" environment (without biological activity

or wastewater constituents) as reported in Chapter 2. There, the Na

and Cl ions did not compete with phosphate for sorption sites. Of the

major compounds that compete with P for sorption sites, organic acids

produced by the microbial activity may be predominant (Goldberg, 1985;

Evans 1985; Sibanda and Young, 1986).

The effect of pH on P sorption (Fig.3.12) showed a trend of

decreasing sorption at higher pH values although sorption peaks occur

at much lower levels than in the case of suspensions in purely

inorganic systems (Fig.3.9.)

Gibbsite, as a monocomponent suspension, in an oxidizing

environment shows P sorption isotherm similar to the mixture of

goethite and gibbsite (Fig.3.13). This behavior is in agreement with

the findings by Goldberg and Sposito (1984a) that Fe and Al oxides have

a similar P-sorption capacity.

Phosphorus Sorption in Reduced Systems

Binary suspensions of goethite and gibbsite showed a significant

increase in P sorption capacity under reducing conditions (Fig.3.11).

This result was consistent with results reported by Holford and

Patrick (1979) who postulated that at pH 6.5, increased sorption was

due to the precipitaton of fresh Fe(0H)2* Another possibility was the

unintentional precipitation of amorphous Fe(0H)2 because of accidental

oxidation during the experimental procedures (Holford and Patrick,

1979).

50

Page 67: Phosphate sorption in soils and reference oxides under

Sorbed P (mmol/kg)

Figure 3.12. Phosphate sorption in goethite and gibbsite (1:1)

suspension in simulated wastewater.

Initial P concentration—300 anol ■

Page 68: Phosphate sorption in soils and reference oxides under

Sorbed P (mmol/kg)

Figure 3.13. Phosphate sorption in

simulated wastewater.

1.9 g I/1

pH—6.5.

gibbsite suspension in

52

Page 69: Phosphate sorption in soils and reference oxides under

In the system containing gibbsite as a monocomponent sorbent,

there was no apparent effect of redox on P sorption, especially at P

equilibrium concentrations below 200 mmol m (Fig.3.13). This

confirms that, generally, A1 is not significantly affected by

differences in redox potential (Lindsay, 1979). In binary systems the

increase in P sorption with reducing conditions, thus, has to be

explained by Fe and P interactions.

A series of experiments were conducted to study which mechanism

was responsible for the increased sorption. To establish whether the

increased P retention was due only to the oxides or to compounds in

solution, one of the reducing reactors (Eh=-50 mV) was sampled.

Minimizing contamination with O2, the extracted suspension was rapidly

filtered and the filtrate collected in a series of centrifuge tubes

containing increasing levels of P in solution. Because no solids were

_ 3 in the filtrate, the results of P retention were expressed in mmol m

of filtrate. The goethite/gibbsite suspension isotherm (Fig.3.13) was

converted to mmol Pm of suspension allowing comparison

with the filtrate results (Fig.3.14). Although some P appeared

retained by the filtrate, the amount was practically insignificant

compared to the retention capacity of the whole suspension. In

general, the share of sorbed P contributed by the filtrate was around

l/10th of the P sorbed by the whole suspension.

53

Page 70: Phosphate sorption in soils and reference oxides under

Sorbed P (mmol/cu.m.)

Figure 3.14. Phosphate sorption expressed on volume basis. pH

54

Page 71: Phosphate sorption in soils and reference oxides under

Accidental reoxidation of solids in the centrifuge tubes appeared

insignificant. Reoxidation would yield an amorphous Fe oxide of the

type obtained in the two reduced reactors that subsequently were

reoxidized. The isotherm of such products was determined (Fig.3.11).

Although the sorption capacity of the reoxidized mixed suspension is

clearly larger than the capacity of the crystalline oxides (Fig.3.11)

the isotherm of the reoxidized material reached a plateau at around

220 mmol P kg~^ whereas the reduced mixture had a sorption capacity

exceeding 700 mmol P kg~^.

One possible mechanism for the increase in P sorption under

reduced conditions could be the formation of colloidal precipitates of

Fe-P on the surface of gibbsite particles. Morel and Gshwend (1987)

suggested the occurrence of such colloids as an explanation to some of

what they call "strange" behavior of sorbates in natural waters. They

postulated that these colloids have solute-like properties but can

aggregate to form larger particles. The existence of such colloids

would have important applications in the area of non-point source

pollution and the movement of P in soils and ground water (Gschwend

and Reynolds, 1987).

Conclusions

From the study of the goethite and gibbsite monocomponent

systems the following conclusions were drawn:

i. The evaluation of sorption capacity based on particle size gave

comparable results when BET and TEM procedures were used on the

highly crystalline gibbsite minerals, whereas TEM values were

twice as large as BET in the poorly crystalline goethite.

55

Page 72: Phosphate sorption in soils and reference oxides under

ii. Of the empirical, isotherm-derived methods, the P exp and PL

parameters were in agreement for both oxides. These two

parameters, however, are dependent on the particular P

concentrations used during the equilibration phase. On the other

hand, the third empirical parameter, Pj, is found by

extrapolation and thus represents a "potential" that is

independent of the particular concentrations at which it was

evaluated.

iii. The two systems of adsorbents studied here were characterized by

measurable properties that, although not always consistent, can

be used to monitor the changes of surface-ion interactions under

variable environmental conditions.

The study of binary mixtures of the references oxides led to the

following conclusions:

i. For the adsorbents studied, the sorptive additivity concept was

validated for pH values of 6.0 and at levels of solution P that

exceed the sorption capacity of each of the oxides, and

ii. Non-adherence to the additivity concept occurred at initial P

concentrations of 300 mmol m , at pH values of 4.0 to 4.5 and

_ 3

6.5 to 7.0, and in initial P concentrations of 100 mmol m at

pH<4.5. These effects were possibly caused by particle-particle

interation P-precipitation at high solution P levels, and to

excess available sorption sites on goethite at low P levels.

Finally, from a study of mixtures of goethite-gibbsite

suspensions under varying redox potentials it was concluded that:

56

Page 73: Phosphate sorption in soils and reference oxides under

i. An apparatus to study soil chemical reactions under controlled pH

and Redox conditions was successfully adapted to the study of

reference oxides. By controlling redox through microbial activity

the behavior of particular soil components can be studied in

"living" environments that are often "non-thermodynamic" and

cannot be simulated in inorganic media,

ii. At redox potentials exceeding 300 mV, P sorption in 1:1 goethite

and gibbsite suspensions (1.9 g L”^) reached a maximum between 35

to 40 mmol kgA 1.9 g Lsuspension of gibbsite alone showed

a similar maximum P sorption value,

iii. In a reducing environment (Eh = -50 mV), the goethite and

gibbsite mixture exhibited a very sharp increase in P sorption

capacity due to dissolution of goethite and subsequent

precipitation, probably as an amorphous ferrous oxide,

iv. P sorption onto gibbsite appeared independent of the redox

potential.

v. Repeated wetting and drying cycles in soils may result in

degradation of goethite into less crystalline Fe solids that will

increase the P sorption capacity of the soil unless Fe is leached

out.

57

Page 74: Phosphate sorption in soils and reference oxides under

CHAPTER 4

PHOSPHORUS SORPTION CAPACITY IN ACID SOILS

Introduction

Land-use decisions may be facilitated by the application of soil

survey information. Current soil survey interpretations can be

improved by better correlations between available data and selected

soil properties and integrating these "pedotransfer" functions in

appropriate computer modules (Bouma, 1988). Phosphate retention data

are used increasingly to estimate the total phosphorus sorption

capacity (PSC) of a soil in relation to soil fertility (Laverdiere and

Karam, 1984; Ping and Michaelson, 1986) or environmental pollution

(Breeuwsma et al., 1986).

Studies undertaken with acid soils have shown that factors such

as clay content, organic matter, and extractable Fe, Al, and Ca

influenced PSC (White, 1981). No unifying formulations have been

derived, however, of the relationship between PSC and selected soil

properties if research results from different regions are combined

(Berkheiser et al., 1980). Consequently, PSC correlation studies have

been limited to the regional level (Tofflemire and Chen, 1976;

Polyzopoulos et al., 1983; Laverdiere and Karam, 1984). A regional

approach ensures uniformity of analytical methods and soil description

criteria and provides essential information to users of this data.

While correlation studies can lead to predictive regression equations,

58

Page 75: Phosphate sorption in soils and reference oxides under

no cause-effect relationships can be inferred from these type of

analyses.

The role of Fe and Al oxides in enhancing P retention in acid

soils has been well established (Goldberg and Sposito, 1984b). Simple

transformation equations, including extractable Fe and Al, have been

used to obtain PSC as a component in the evaluation of land for the

application of agricultural wastes (Breeuwsma et al., 1986). This

simplified procedure cannot be applied to all acid soils, since the P

retention capacity of some of these soils can be affected by other

parameters such as extractable Ca (Smillie et al., 1987). The

hypothesis that, in acid soils, the main agents of P sorption are Fe

and Al, is worth testing because of the wealth of information

available on these reference compounds in soil survey data bases and

the potential of quantitatively adapting results of Al and Fe oxide

experiments to soil simulation models.

One of the major difficulties in incorporating PSC values in

land suitability assessment procedures is the selection of a PSC index

that is simple, accurate, and field-applicable (Stuanes, 1982).

Sorption indices can be relative or absolute indicators of PSC.

Relative indices compare and rank various soils as to their relative

sorptive capacity, whereas an absolute index, in addition to being

suitable for making comparisons, can be used to calculate actual P

loading rates for wastewater or fertilizer applications.

The equilibrium sorption (Pgl Singh and Tabatabai, 1977;

Loganathan et al., 1987), Langmuir maximum retention (P^; Veith and

Sposito, 1977) and the interpolation maximum (Pjj Sposito, 1982) are

absolute indicators of P sorption in soil. The Freundlich index (PpJ

59

Page 76: Phosphate sorption in soils and reference oxides under

Bache and Morel, 1983) and the Bache-Williams phosphorus index (Pgyl

Williams, 1971) are examples of relative P sorption indicators. While

four of the PSC indices have been used extensively, the more recent Pj

method has received little attention.

The present study is aimed at evaluating the relationship between

PSC and selected soil properties in various acid Massachusetts soils

and to assess the effect of different PSC determination techniques on

the design of waste disposal systems. Specific objectives of this

study are to (i) compare PSC values of selected horizons in

representative Massachusetts soils by different measuring techniques,

(ii) correlate these PSC values with selected soil properties, and

(iii) develop regression equations to predict appropriate PSC values

from these soil parameters, including an evaluation of statistical

models that use Fe and A1 as unique regressors.

Materials and Methods

Soil Sampling and Characterization

Fifty-seven samples were obtained from surface and subsurface

horizons from 23 Massachusetts soil profiles located throughout the

state. Chemical and physical properties were determined on the

fraction of air-dried soil passing a 2-mm sieve. Particle-size

analysis was performed by wet-sieving to determine the sand fraction

and by measurement of the silt and clay fractions either by the

pipette method or the hydrometer procedure (Gee and Bauder, 1986).

When necessary, organic matter was removed by oxidation with H2O2. The

pH was measured in 1:1 (w:v) suspensions of soil in 0.01M CaC^.

% 60

Page 77: Phosphate sorption in soils and reference oxides under

Cations extracted by a 1M ammonium acetate solution at pH 7.0, were

determined by atomic absorption spectrometry (Soil Survey Staff,

1972). Cation exchange capacity (CEC) was obtained by distillation of

ammonium retained on the soil from pH 7.0 buffered ammonium acetate

(CECpHy) or by the sum of extractable cations (CECp^g ^ Survey

Staff, 1972). Extractable acidity was determined by a barium chloride-

triethanolamine method (Peech et al., 1962). Iron and A1 concen¬

trations were measured by atomic absorption analyses of extracts of

dithionite-citrate at pH 7 (Fe^, Al^) and pyrophosphate at pH 10

(Fep, Alpj Soil Survey Staff, 1972). Base saturation was calculated

from the measured values for extractable bases and CEC, and organic C

was determined by acid-dichromate digestion (Nelson and Sommers,

1986).

Phosphorus Sorption Capacity

The P-sorption procedure was adapted from Nair et al. (1984).

Duplicate samples were shaken for 24 hours in 0.01 M CaC^ solutions

containing K^PO^ at initial concentrations of 0 to 516 mmol m (0-16

mg L”^). Phosphorus in solution was determined after equilibration at

25 ± 1°C (Murphy and Riley, 1962). Final P at equilibrium subtracted

from the initial P value yielded the amount sorbed. The sorption

isotherms relating P sorbed to P in solution were used to calculate

the following PSC parameters:

(i) P£ was obtained from the sorption isotherm by determining the

level of P sorbed corresponding to the estimated P levels of the soil

solution. An equilibrium level of 216 mmol Pm (7 mg L ) was

61

Page 78: Phosphate sorption in soils and reference oxides under

assumed, which is the approximate P content of raw wastewater (U.S.

EPA, 1981).

(ii) P^ was calculated from a linear form of the following

equation:

q = a PLc/(l + ac) (1)

where, q = moles of P sorbed per unit weight of soil, a = a constant,

c = moles of P in solution at equilibrium.

(iii) Pj was derived graphically by plotting the partition

coefficient (=q/c) against q and determining the intercept with the

q-axis. (Sposito, 1982).

(iv) The Freundlich index (Pp) was determined by regression of a

linear, logarithmic transformation of equation 1:

log q = log (Pp) + r log (c) (2)

where Pp and r are emperical constants with 0<r<l.

(v) The value for the Bache-Williams index (Pgy) was derived from

the following equation:

PBW = / loS c' (3)

where qr = sorbed P (in mg/lOOg soil) after addition of 50 mmol P kg 3

soil, and c' = soluble P (in mmol m ) at equilibrium after addition of

50 mmol P kg^ soil.

Statistical Analysis

The data were evaluated statistically by deriving a correlation

matrix between principal soil characteristics and various PSC values

(SAS Institute, Inc., 1985). From this matrix, variables with the

highest correlation were used in a stepwise, multiple regression

62

Page 79: Phosphate sorption in soils and reference oxides under

by stepwise elimination or by inclusion of variables (Draper and

Smith, 1981).

Results and Discussion

Ranges in values for the soil properties and the various PSC

procedures are summarized in Table 4.1. The PSC methods were

correlated significantly (PC0.001) with one another (Table 4.2). In

this respect, PE exhibited the best overall correlation (r>0.75) with

the other PSC parameters. Holford (1979) discussed that a high degree

of correlation between P indices did not imply that they were

equivalent. In fact, the effectiveness of any PSC index depended on

its particular application such in as plant uptake or in P transport

(Holford, 1982).

A significant (P<0.001), positive correlation was found between

each of the PSC indices and Fe^, Al^, Fep, Alp, CEC, extractable

acidity, and in most cases organic C. Phosphate retention was

correlated negatively with pH and base saturation (P<0.01). No

significant correlation was found between various PSC values and

different particle size fractions. Extractable Ca was not correlated

significantly with Pj and P^, and only correlated weakly (P<0.10) with

Pp, Pgy. and Fg- Except for the PL value, all PSC indices had higher

correlation coefficients with extractable Al than with extractable Fe.

This Al-PSC correlation was reported also in previous studies

(Vijayachandran and Harter, 1975; Diggle and Bell, 1984; Laverdiere

and Karam, 1984; Ping and Michaelson, 1986).

For each PSC index, the correlation coefficients obtained with

the different Fe and Al extractants were comparable (Table 4.2).

Sodium dithionite removes amorphous and crystalline Fe compounds,

63

Page 80: Phosphate sorption in soils and reference oxides under

whereas pyrophosphate extracts the amorphous and organic forms of Fe

(Bascomb, 1968). These extractants are less discriminating in the case

64

Page 81: Phosphate sorption in soils and reference oxides under

Table 4.1. Range in properties of 57 surface and subsurface horizons

in 23 representative Massachusetts soils.

Soil

Property

Unit Range Mean Standard

Deviation

Sand % 6-99 53 23

Silt % 1-75 38 17

Clay % 0-50 10 10

Fen % 0.01-2.49 0.37 0.64

Fed % 0.13-4.73 1.18 0.97

A1P % 0.01-1.01 0.21 0.20

A1d % 0.02-1.15 0.26 0.22

Organic C % 0.1-11.4 1.4 2.2

Base sat. % 1-99 15 26.2

pH 1 3.0-6.6 4.6 0.8

extract. Ca cmol(+) kg 0.01-12.5 1.2 2.5

exchang. acid. cmol(+) kg' 1.1-50.3 13.7 14.4

CECpH 8.2

CECpH 7.0

cmol(+) kg” 1.1-50.4 15.0 14.2

cmol(+) kg”F 0.9-34.7 10.0 8.5

PSC parameters:

fe mmol kg”F 1.9-66.5 16.4 15.2

PL mmol kg” 4.5-69.6 16.7 13.6

Pt mmol kg 8.1-145.3 32.4 33.4

pp 8-1552 295 292

PBW 7-99 25 20

65

Page 82: Phosphate sorption in soils and reference oxides under

Table 4.2 Correlation matrix of selected soil characteristics

and PSC indices from 23 Massachusetts soils.

pH FeP Alp F°d Ald 0RG. C

0.62424

(0.0001)

-0.09768

(0.5232)

-0.19977

(0.1936)

0.14576

(0.2793)

-0.09868

(0.4693)

-0.03353

(0.8044)

-0.38851

(0.0092)

-0.37032

(0.0145)

-0.13488

(0.3262)

-0.27843

(0.0415)

-0.49601

(0.0001)

0.84900

(0.0001)

0.82908

(0.0001)

0.77666

(0.0001)

0.50059

(0.0005)

0.67503

(0.0001)

0.94367

(0.0001)

0.45965

(0.0017)

0.75145

(0.0001)

0.35099

(0.0074)

0.41155

(0.0016)

Pf Pi P| Pf PBW

-0.21017

(0.1348)

-0.12873

(0.3489)

-0.14539

(0.3772)

-0.25935

(0.0514)

-0.21389

(0.1101)

EXTC. Ca

-0.38997

(0.0051)

-0.44615

(0.0008)

-0.40379

(0.0132)

-0.35475

(0.0079)

-0.35439

(0.0079)

pH

0.77587

(0.0001)

0.64302

(0.0001)

0.47457

(0.0070

0.71277

(0.0001)

0.72209

(0.0001)

FeP

0.86167

(0.0001)

0.63825

(0.0001)

0.63424

(0.0002) 0.90377

(0.0001)

0.89926

(0.0001)

Alp

0.72994

(0.0001)

0.67847

(0.0001)

0.45995

(0.0032)

0.67255

(0.0001)

0.67733

(0.0001)

0.88158

(0.0001)

0.69705

(0.0001)

0.64502

(0.0001)

0.33750

(0.0001)

0.87752

(0.0001)

Ald

0.51805

(0.0001)

0.69779

(0.0001)

0.65562

(0.0001)

0.35700

(0.0064)

0.41361

(0.0014)

ORGANIC C

0.78741

(0.0001)

0.75523

(0.0001)

0.91824

(0.0001)

0.89493

(0.0001) P£

0.89778

(0.0001)

0.66499

(0.0001)

0.68638

(0.0001) pl

0.57564

(0.0001)

0.57959

(0.0001) P|

0.90504

(0.0001) pf

Page 83: Phosphate sorption in soils and reference oxides under

of A1 (McKeague et al., 1971). The amorphous forms of Fe and Al oxides

are associated clearly with P sorption (Johnson et al. , 1986). Stuanes

(1984) cautioned against using extractable Fe and Al to indirectly

derive PSC, chiefly because of the great variation in published

results. For land application of wastes, the usefulness of soil Fe and

Al correlations with PSC is complicated further by the possibility

that the Fe and Al that retain P are present already in the waste

applied (Chang et al., 1983).

Stepwise regression, including exchangeable Ca, organic C, pH,

Al^, Fe^, Al^, and Fep as independent variables, resulted in

prediction equations (Table 4.3). Validation of these models required

additional analytical and field work, which was beyond the scope of

the present study. All models were significant at the 0.1% level. The

equation for Pj had the least predictive ability. In addition to

possible interactions between soil parameters not considered in these

regression equations, the low correlation may result from inherent

inaccuracies in determining Pj. The estimation of this parameter was

facilitated by the use of concentrated P solutions to obtain the P-

isotherms, but at high concentrations other phenomena such as

precipitation may affect the results. Nair et al. (1984), therefore,

recommended that P-equilibrium studies be performed only at low P

levels. On the other hand, regression equations are empirical, and the

gain in additional data achieved by the use of higher P levels in the

equilibration solutions may outweigh the need to know which P-

retention mechanism predominates.

67

Page 84: Phosphate sorption in soils and reference oxides under

Table 4.3. Selected results of stepwise regression for different PSC

indices and various soil properties. In each model, all

variables are significant at the 0.15 level.

Model R2

PE ~~ 18.3 + 60.1Ald - 3.6pH 0.800***

PL = -18.4 + 8.9Fed + 5.3C + 4.8pH - 2.4Ca 0.847***

Pi = -8.6 + 147.4Ald - 27.5Fe 0.620***

Pv = 16.4 + 1183A1H - 22Ca P 0.848***

PBW = 3.27 + 76.02Ald + 2.54Fed - 1.27Ca 0.902***

***, significant at the 0.001 probability level.

Fed and Ald, dithionite-extractable Fe and Al.

Fep and Alp, pyrophosphate-extractable Fe and Al.

68

Page 85: Phosphate sorption in soils and reference oxides under

Dithionite-extractable A1 was a component in four of the five

regression equations (Table 4.3). Autocorrelated variables that

individually correlated highly with the PSC indices, often were not

retained in the stepwise regression procedure. Conversely, calcium,

which correlated very poorly with any of the indices, was retained in

three of the equations, evidence that Ca contributed to that part of

the variation not covered by Al-related parameters. In our study, the

contribution of Ca was, however, always correlated negatively with PSC

indicating a pH-type of effect rather than the positive correlation

reported by Smillie et al. (1987).

If extractable Fe and Al were used as independent variables to

o predict the PSC indices, a distinct difference between the R values

of the models of the various PSC indices was observed (Table 4.4).

Most of the two-variable models, including either Fe^ and Al^ or Fep

and Alp, were highly significant, but Fe and Al together covered only

38 to 50% of the variability in the P^ and Pj values. On the other

hand, Fe and Al covered 74 to 88% of the total variation in each of

the other three PSC methods. Relative PSC indices, Pp and Pgy, can be

predicted with a fairly high degree of confidence from dithionite

extractable Fe and Al. This result confirms earlier work by Burnham

and Lopez-Hernandez (1982) who found a good relationship between Pgy

and soil classification at the series level.

The correlation between extractable sesquioxides (Fe and Al) and

predictors of absolute PSC (Pg, Pg» an<^ Pj) was not as well defined as

that of the relative PSC predictors. The regression results (Tables

4.3 and 4.4) indicated that inclusion of a pH-related factor

significantly improved the predictive potential of the models.

69

Page 86: Phosphate sorption in soils and reference oxides under

Table 4.4. Summary of statistical parameters for regression models

with Fe^ and Al^, or Fep and Alp as independent

variables.

- Model considered -

PSC

Parameter

No. of

samples Fed j A1d

RZ

Fe„ + Al Pr2 P

39 0.766*** 0.743***

PL 41 0.501*** 0.490***

Pi 29 0.378** 0.424***

PF 43 0.874*** 0.818***

PBW 43 0.880*** 0.809***

-t— aa A A. AT » > ° ignificant at the 0.01 and 0.001 probability levels.

respectively.

and Al^, dithionite-extractable Fe and Al.

p and Alp, pyrophosphate-extractable Fe and Al.

70

Page 87: Phosphate sorption in soils and reference oxides under

Computer modeling to assess differences in assimilative capacity

between soils can be facilitated by inclusion of transfer functions

based on relative PSC values. Modeling of actual P sorption in soils

is complicated and should consider additional variables in the

transfer functions. These parameters need to be based on more

mechanistic models of P sorption.

Conclusions

Phosphorus sorption capacity was correlated highly with

dithionite and pyrophosphate extractable Fe and Al, pH, exchangeable

Ca, and organic C. Models including Fe and Al, exclusively, were

effective in explaining P sorption as measured by PE, Pp, and PBW

methods, and were less significant in the case of PL and Pj. The

latter was attributed, in part, to inaccuracies inherent to the

determination of these PSC parameters. All PSC parameters were

correlated significantly with one another. The results emphasize the

need to refine analytical techniques determining P^ and Pj indices,

because these parameters lead to estimates for absolute retention

capacities, compared to Pp and PBW> which give only a relative index

of soil P retention. More accurate determination of Pj is of special

interest because it leads to higher estimates of PSC than the other

absolute PSC indices, which underestimate the absolute PSC value.

71

Page 88: Phosphate sorption in soils and reference oxides under

CHAPTER 5

ROLE OF PHOSPHORUS SORPTION CAPACITY IN

SUITABILITY ASSESSMENT FOR WASTE DISPOSAL

Introduction

Land application of waste materials is a practice encouraged by

legislation to control environmental pollution (Massey, 1983).

Common waste application methods include overland flow, sludge

application, injection wells, evaporation ponds, sanitary landfills,

waste compost application, and on-site sewage disposal.

Except for contained systems, such as lined landfills, waste

application techniques rely on physical, chemical, and microbiological

processes in the soil to treat waste. Pollutant treatment processes

include volatilization, ion exchange, adsorption, precipitation,

oxidation, reduction, biodegradation, or uptake by living organisms.

Ideally, all these components should be considered in the assessment

of the assimilative capacity of a particular soil for waste disposal

(King, 1985).

The role of soil scientists in land application of wastes is to

assess the suitability of a particular site, especially the

determination of the soil assimilative capacity for various waste

constituents (Loehr, 1985). In practice, assimilative capacity is

evaluated by considering the factors that affect treatment efficiency.

Loehr and Overcash (1985) considered the following site specific

factors of importance: climate, topography, hydrogeology, vegetation,

and edaphic conditions. The latter include percolation rate, pH,

72

Page 89: Phosphate sorption in soils and reference oxides under

organic matter content, cation exchange capacity (CEC), and phosphorus

sorption capacity (PSC). These characteristics may be used to evaluate

the potential effect of general site conditions on the efficacy of

land treatment (Veneman, 1982; Steele et al., 1986). With respect to

soil chemical properties that contribute to pollutant attenuation,

organic C is used routinely to estimate soil assimilative capacity for

organic compounds (Lyman, 1982), whereas cation exchange capacity

often is indicative of heavy metal assimilative capacity (Houck et

al., 1987). To determine potential phosphorus sorption, however, a

myriad of methods are used to estimate the total assimilative capacity

(Sibbesen, 1981; Travis and Etnier, 1981).

Determination of the PSC generally has been concerned with

defining either a "maximum PSC" (Olsen and Watanabe, 1957; Stuanes,

1984; Van Riemsdijk and van der Linden, 1984) or an index parameter

that can be used to evaluate the relative PSC of soils (Bache and

Williams, 1971). Comparison of the relative PSC of two sites is

technically less challenging than the determination of an absolute PSC

value for each soil. Ultimately, however, absolute values of PSC are

needed in the integrated design of waste disposal systems.

Land limiting constituent (LLC) analysis is a convenient way of

integrating the two major technical considerations in land disposal,

waste characteristics and site assimilative capacity (Loehr and

Overcash, 1985). The concept of LLC analysis is outlined in Fig. 5.1.

The loading rate of each waste component is estimated and divided by

the projected waste assimilative capacity of the soil. The resulting

area-value is a measure of the total soil mass required to retain,

73

Page 90: Phosphate sorption in soils and reference oxides under

detoxify, or otherwise absorb waste components in an environmentally

sound fashion in accordance with the nature of each waste component.

The LLC procedure determines which waste consitituent has the highest

land area requirements for disposal. After an initial LLC analysis is

completed, two actions are possible depending on the nature of the

LLC: (i) further treatment of the effluent at the source in order to

reduce the value of LLC, or (ii) acquisition of additional land based

on LLC land area requirements.

In this study, a LLC analysis was used to evaluate whether or not

PSC was the limiting factor in land disposal of municipal wastewater

on acid soils. Specific objectives of the study were to: (i) compare

the results of different methods to determine PSC in two

representative Massachusetts soils, (ii) determine the effect of the

PSC determination method on potential waste loading rate, and (iii)

determine the sensitivity of system design for land disposal of

municipal wastewater to soil phosphorus loading capacity.

Materials and Methods

Wastewater Characteristics

Characteristics of raw municipal sewage and effluent from a

secondary treatment plant were used in this study. A daily flow of

1000 m^ was assumed, corresponding to the expected sewage flow for a

population of 3,000 people (U.S. EPA, 1981). Average wastewater

characteristics were obtained from the literature (Table 5.1). The

concentration of inorganic P reported (8mg L ^) is in the higher limit

of average P levels in raw wastewater. Annual loads of relevant

wastewater components were calculated by multiplying the elemental

74

Page 91: Phosphate sorption in soils and reference oxides under

WASTE CHARACTERISTICS

volume of flow

chemical composition

PH oxygen demand

solids content

I

I

I

SITE ASSIMILATIVE CAPACITY

kg ha"^ yr~^

I I

-| II V

AREA REQUIREMENT

ha

(LLC is waste consitutent

with largest area requirement)

I

I

I

WASTE LOADING

kg yr

I

I

SITE CHARACTERISTICS

climatic conditions

topography

geology, groundwater

soil properties

vegetation

LLC

REDUCTION

AT SOURCE

II II II V

ENGINEERING

DESIGN

LAND

ACQUISITON

Figure 5.1. Land limiting constituent (LLC) concept in land treatment

design (after Loehr and Overcash, 1985).

75

Page 92: Phosphate sorption in soils and reference oxides under

Table 5.1. Representative wastewater characteristics for raw municipal

sewage and secondary effluent (Source: Thomas and Law, 1977;

U.S. EPA, 1981).

PARAMETER RAW SEWAGE SECONDARY EFFLUENT E m-3 _ gm

Solids 450 320

COD 500 100

BOD 200 40

TOTAL N 40 20

INORG. P 8 7

B 1 0.8

Cd 0.02 0.004

Cr 0.20 0.12

Cu 0.10 0.06

Fe 1.0 0.5

Hg 0.01 0.005

Mn 0.14 0.13

Ni 0.08 0.05

Pb 0.50 0.25

Zn 0.40 0.12

76

Page 93: Phosphate sorption in soils and reference oxides under

composition by the total volume of the wastewater. Waste treatment was

assumed to occur through overland flow.

Soils

Two soils were selected for this study. The first was a Quonset

sandy loam (Typic Udipsamment, mixed, mesic) and the second a

Lanesboro silt loam (Typic Dystrochrept, coarse-loamy, mixed, mesic).

A summary of the major characteristics of these soils is in Table 5.2.

Phosphorus Sorption Capacity

The PSC of the major horizons of each soil was determined by

performing batch equilibrium sorption studies at 22° C to obtain P-

isotherms (Nair et al., 1984). Duplicate soil samples passing a 2-mm

sieve were equilibrated for 24 hours in Kl^PO^ solutions with initial

-1 - 3 concentrations ranging from 0 to 16 mg P L (0 - 516 mmol m ) .

The amount of P remaining in solution after equilibration was

determined (Murphy and Riley, 1962) and the amount of P sorbed was

calculated by subtracting the amount of final P in solution from the

initial P. The resulting P-sorption isotherms were used to evaluate

the following three parameters of "maximum" PSC:

(i) Equilibrium Sorption (Pg):

The P£ was obtained from the sorption isotherm by matching the

level of sorbed P with an assumed P-level expected in the soil

solution after application of the waste (Loehr et al., 1979). In this

simulation, we considered final concentrations of 7 and 8 mg P L

77

Page 94: Phosphate sorption in soils and reference oxides under

Table 5.2. Range in properties of selected surface and subsurface

horizons of two Massachusetts soil series.

Quonset Lanesboro

Soils:

Sand, % 51-86 37-69

Silt, % 12-39 25-53

Clay, % 1.5-8 5-12

PH , 4.3-4.9 3.1-4.7

CEC, cmol(+).kg” 3.4-18.0 3.5-48.8

Fed, % 0.52-0.83 0.67-4.73

A^, % 0.11-0.47 0.11-0.88

Organic C, % 0.2-3.1 0.3-11.4

PSC (mmol kg~l):

7.7-26.6 8.5-54.2

PL 6.7-39.0 10.1-69.6

pi 15.3-145.3 27.1-122.7

78

Page 95: Phosphate sorption in soils and reference oxides under

(ii) Langmuir-Maximum Retention (P^):

This index is the theoretical maximum adsorption capacity as

described in the equation:

q — aPLc/(l+ac) [1]

where, q = moles of P sorbed per unit weight of soil, a = constant,

c = moles of P in solution, at equilibrium. Values for P^ were

calculated from a linear transformation of Equation 1 (Olsen and

Watanabe, 1957).

(iii) Interpolation Maximum (Pj):

This method refers to the interpolation theorem proposed by

Sposito (1982). The maximum PSC was found by defining the partition

coefficient:

Kd = q / c [2]

and plotting Kd against q. Pj was then determined by extrapolating

to intercept the q-axis.

A weighted average of PSC (Tofflemire and Chen, 1976) for each

soil was computed by considering a one-meter reactive depth of soil

and applying the following equation:

n

t = (Sdj^ x qi) /100 [3]

i

where, t — average maximum PSC (g kg ^), i = horizon down to 1 m

depth, n = total number of horizons down to 1 m, d = thickness of

horizon i (cm), q = maximum PSC of horizon i (g kg ^). To transform

the results from g P kg ^ soil to kg P ha ^ an average soil bulk

density of 1.3 Mg m'3 was assumed. These calculations were repeated

for each of the three methods of estimating PSC (Table 5.3).

79

Page 96: Phosphate sorption in soils and reference oxides under

Table 5.3. Phosphate sorption capacity of two soils using different

estimation methods. P£ y and P£ g refer to PSC values

based on final soil solution concentrations of 7 and 8 mg

L~ , respectively.

PHOPHATE SORPTION PARAMETERS

SOIL HORIZON PE y PE g PL P-j-

- mg p kg'1 -

Quonset A 825 860 1207 4500

Bwl 700 760 963 3300

C 237 250 209 475

profile average 379 403 444 1386

Lanesboro A 1275 1375 2156 3750

Bwl 1680 1760 2087 3800

Cr2 263 294 313 840

profile average 615 660 717 1608

80

Page 97: Phosphate sorption in soils and reference oxides under

Assessment of Soil Assimilative Capacity

The following wastewater constituents were considered potentially

limiting and were included in the LLC analysis: total N, Cd, Cr, Cu,

Mn, Ni, Pb, and Zn, and inorganic-P (Table 5.1). The soil assimilative

capacity of the waste constituents included in the LLC analysis, was

evaluated using methods discussed hereafter.

To simulate the N-assimilative capacity, the maximum wastewater

loading rate (W) was calculated (Loehr et al., 1979):

W= [C+A(p-ET)-c'p]/(Y-A) [4]

where, C = plant N-uptake (about 150 kg yr"'*'), A = allowable N

concentration in percolating water (10 g m ), p = precipitation (1 m

yr-^), ET = evapotranspiration (0.6 m yr~^), Y = concentration of N in

-3 applied wastewater effluent (g m ), and c' = N concentration in

_ 3

precipitation p (0.5 g m ). Values for p and ET that are

representative of central Masschusetts were obtained from Yuretich et

al., (1986). The resulting W (in m yr-^) was used to calculate the

maximum N loading rate (kg ha ^ yr ^):

M = W x Y [5]

where, M = maximum allowable N loading. It was assumed that 35% of

applied N would be lost through denitrification (U.S. EPA, 1981). The

area of land required was determined by dividing the value for M by

the waste-N generation rate G (kg yr ^).

The soil assimilative capacity for heavy metals, in general, can

be obtained from laboratory column studies (Fuller and Warrick, 1985),

or by single (Tirsch and Baker, 1976) or repetitive (Amacher et al.,

1986) batch equilibrium experiments from which sorption isotherms can

be derived. In this simulation, we used maximum application levels of

81

Page 98: Phosphate sorption in soils and reference oxides under

heavy metals that in the field produced no toxicity in plants. These

levels have been termed the "USDA Guidelines" (Loehr et al., 1979).

The magnitude of the values generally are affected by the CEC or the

texture of the soil. Recommended maximum allowable levels (Overcash

and Pal, 1979) for soils with CEC values < 5 cmol (+) kg"1 (Quonset)

and 5 to 15 (Lanesboro), were used in the simulation (Tables 5.4 and

5.5).

Results and Discussion

Not included in the LLC analysis, but nevertheless of practical

importance were site-specific factors such as topography and land use,

soil hydraulic characteristics and soil pH (Fig. 5.1). Sites with a

high chemical retention capacity may have a low permeability that

could limit the volume of liquid wastes applied. A neutral, well-

buffered soil is often considered ideal for land application (Overcash

and Pal, 1979). The relatively low pH of most southern New England

soils could be a limiting factor for the disposal of wastes except

that the pH can be raised by co-disposal with alkaline waste

materials or by liming. In this study, it was assumed that the site-

specific factors were equally non-limiting in both sites or could be

corrected by practical management methods.

The LLC analysis indicated the importance of N in sewage as a

potential pollutant, because N was limiting in both soils (Table 5.4).

Secondary treatment significantly reduced the N content (Table 5.5),

thereby, lessening the area-limiting role of N. Overall, this

reduction decreased land requirements by more than 20% for the Quonset

soil and 55% for the Lanesboro soil. The assumption that N-

82

Page 99: Phosphate sorption in soils and reference oxides under

Table 5.4. Land limiting constituent analysis for application of

1,000 nr d~ of raw municipal wastewater on two

representative New England soil series.

Constituent

Loading

rate

kg yr_1

- Quonset

Site

assimilative

capacity

kg ha~^ yr"^

Area

required

ha

- Lanesboro -

Site Area

assimilative required

capacity

kg ha'l yr~^ ha

Total N 15,000 385 39 385 39

Inorganic P

7

29,000

105 28 172 17

PL 115 25 206 14

PI 360 8 418 7

Cd 7 0.2 35 0.4 18

Cr 70 20 3.5 40 2

Cu 40 5 8.0 10 4

Mn 50 15 3.3 20 2.5

Ni 30 2 15 4 7.5

Pb 180 20 9 40 4.5

Zn 150 10 15 20 7.5

83

Page 100: Phosphate sorption in soils and reference oxides under

Q 1 Table 5.5. Acreage required for the disposal of 1,000 m d of

secondary effluent using land limiting constituent

analysis in two representative New England soil series.

Quonset - Lanesboro

Constituent

Loading

rate

kg yr"1

Site

assimilative

capacity

kg ha”^ yr"^

Area

required

ha

Site

assimilative

capacity

kg ha-^ yr"^

Area

require*

ha

Total N 7,500 577 13 577 13

Inorganic P

7

2,600

99 26 160 16

PL 115 23 202 13

PI 360 7 418 6

Cd 5 0.2 25 0.4 12.5

Cr 40 20 2 40 1

Cu 20 5 4 10 2

Mn 50 15 3 20 2.5

Ni 20 2 10 4 5

Pb 90 20 4.5 40 2.5

Zn 35 10 3.5 20 2

84

Page 101: Phosphate sorption in soils and reference oxides under

assimilative capacity was independent of soil type, resulted in

identical land area requirements for both soils. As a result, in the

case of land disposal of raw sewage, where N was the LLC (Table 5.4),

no difference was found between soils in the amount of land required

(39 ha).

When components other than N were considered (Table 5.5), the land

requirement for the Lanesboro soil was 38% less than for the Quonset

series. The latter, because of lower CEC and PSC values, exhibited

less assimilative capacity than the Lanesboro. The relative role of

Cd in LLC analyses appeared dependent on the PSC evaluation technique

(Table 5.6). Cation exchange capacity based guidelines for heavy metal

application vary greatly among regions and states. More restrictive

guidelines for sludge application than the ones determined in this

study, result in lower estimated assimilative capacities and in larger

land areas.

Different methods of PSC evaluation led to varying area

requirements for P sorption. Theoretically, the equilibration PSC

method should be the only index to be used because, until now, it is

the only parameter that derives from the examination of the mechanism

of sorption. If it is accepted that a state of equilibrium is

established between sorbed and solution ions at any given

concentration of ions, then it should be impossible for the solid

particles to reach complete coverage with sorbed ions. Although this

might happen at very high concentrations of P. The use of P^ and Pj

as PSC indices assumes that even at relatively low P levels (7 and 8 g

m ^) total coverage of the particle may occur. Although this is known

85

Page 102: Phosphate sorption in soils and reference oxides under

Table 5.6. Summary of potential land limiting constituents as

affected by PSC method, soil series, and type of

effluent.

PSC

parameter

- Quonset - - Lanesboro -

Raw municipal Secondary Raw municipal Secondary

sewage effluent sewage effluent

PE

PL

PI

N>Cd P>Cd N P>Cd

N>Cd Cd>P N P>Cd>N

N>Cd Cd N Cd>N

86

Page 103: Phosphate sorption in soils and reference oxides under

to be theoretically false, there is an interest in using indices that

reflect field conditions, even if these indices are only empirical.

In the case of secondary effluent application to the Quonset

soil, the estimated maximum PSC values (Table 5.3) resulted in area

requirements of 26, 23 and 7 ha, representing a decrease of about 70%

in area requirement because of differences in PSC determination.

Table 5.6 shows that P was not limiting for raw waste applications.

For secondary effluent, however, P exhibited a land limiting role.

The Pg method resulted in P becoming the LLC, indicating the need for

land acquisition to provide adequate land area for treatment. If the

P^ maximum was considered, P remained the principal LLC in the

Lanesboro soil, whereas Cd was of primary concern in the Quonset soil,

because of its lower heavy metal retention. Phosphorus was not of any

LLC importance if the Pj method was used.

The lack of standard methodology for the evaluation of PSC in

soils has resulted in wastewater treatment designs where P is not

considered in the LLC analysis (e.g., Houck et al., 1987). Another

common approach is to consider that P content for the percolating soil

water is negligible and that only runoff P is of environmental

importance. Modeling of P loading through the use of the universal

soil loss equation has been successful in the evaluation of land

spreading of animal manure (Moore and Madison, 1985). The significance

of P loading through percolation may become more evident if large

amounts of waste are applied on land as is quite common in western

Europe. The extent of this seepage-P loading under various soil and

climatic conditions seems to be the major data lacking for the use of

PSC values in the design of land application systems.

87

Page 104: Phosphate sorption in soils and reference oxides under

The present study indicates that the PSC evaluation methods used

can have significant effect on the design of land application systems.

As to the applicability of these methods, Schweich and Sardin (1981)

found a correlation between batch equilibration experiments and column

studies. Interpretation of field experiments has generally been

difficult because of technical problems in the sampling of seepage

soil water (Uebler, 1984) and the uncertainties in extrapolating

laboratory data to the field.

Phosphorus generally is considered an immobile constituent (Loehr

and Overcash, 1985) that will accumulate in soil until the maximum PSC

is reached. Persistently high waste P removal rates have been detected

in the field as well in the laboratory. Nagpal (1985) in a column P-

loading study of almost 3 years, obtained 60 to 90% P removal at the

end of the experiment. Almost from the beginning, however, the P-

concentration in the aqueous extracts from the columns was 1 to 2 mg

L~^. Sorption processes in the soil environment apparently remained

significant but did not result in total P removal.

The contribution of seepage P on surface water quality is

unknown. Recent studies show that river water quality generally is

improving in the USA, whereas lakes and reservoirs are deteriorating

due to non-point source pollution (Humenik et al., 1987). Changes in

lake quality often are related to minute increases in P levels because

this element frequently is the growth limiting nutrient. In evaluating

the potential polluting effect of P, this element should not be

considered an immobile species, but rather a "somewhat mobile"

species. Perhaps a certain allowable seepage P load should be

88

Page 105: Phosphate sorption in soils and reference oxides under

considered as in the case of nitrate. This consideration will become

possible only when more information on the relative importance of

runoff and seepage P is gathered. If only runoff is significant, P

does not affect the design of land application systems through LLC

analysis. If dissolved P is significant, models of P movement through

soils could be adapted to field conditions in order to define a more

meaningful maximum PSC than the one generated by Langmuir-type

isotherms (Stuanes and Enfield, 1984).

Conclusions

i. In addition to facilitating engineering design of land

application systems, LLC analysis offers a suitable framework for

comparing methods to evaluate assimilative capacity of particular

soils.

ii. Phosphorus overloading can be significant when wastewaters are

disposed of through overland flow. The chemical characteristics

ot two representative southern New England soils indicate a

potential for P pollution due to seepage, in addition to runoff,

especially for secondary wastewater application,

iii. No routine PSC method is accepted commonly. Current methods

result in an underestimation of the soil retentive capacity.

Engineering design of waste application systems is affected

strongly by the particular PSC method used.

89

Page 106: Phosphate sorption in soils and reference oxides under

CHAPTER 6

DISCUSSION AND CONCLUSIONS

The detrimental effect of P in the environment is manifested

through its effect on surface water eutrophication mainly. While all

the mechanisms controlling eutrophication are not elucidated yet, it

is known that P is in almost all cases the limiting nutrient for

eutrophication to take place.

There are two accepted mechanisms of P entry in surface water

bodies: through runoff of P enriched soil particles, and through

internal enrichment by the release of sediment-P. A third route,

seepage of P through soil, is generally less accepted. In all three

routes the reaction of P at solid surfaces controls the release of P

in quantities that can be used by aquatic organisms that initiate

eutrophication.

The present research focused on elements that allow the

quantification and prediction of P interactions with solid surfaces

under varying pH and redox conditions. In the first experiment,

goethite and gibbsite were used as reference oxides under increasingly

complex conditions to simulate wastewater-P-solids effects. In

reducing environments a sharp increase in P retention capacity of

the mixture of oxides was observed. This increase was not observed in

monocomponent gibbsite systems. It is believed that this increase was

due to the dissolution of Fe (III) and subsequent precipitation of a

90

Page 107: Phosphate sorption in soils and reference oxides under

ferrous-P compound (possibly vivianite). Gibbsite might have played a

catalytic role to precipitate the Fe(II)-P compound on its surface.

In the second experiment, empirical parameters of PSC of selected

non-calcareous Massachusetts soils were determined and pedotransfer

functions were statistically derived. Extractable Fe and A1 were

essential regressors in the predictive equations for the five PSC

indexes that were studied. Based on the conclusions of the first

experiment, it would seem advisable to alter the relative importance

of Fe and A1 parameters in the empirical models for the long term

prediction of soil PSC.

Under conditions of systematic wastewater application Fe could be

mobilized and transfered to lower levels of the profile, or into the

groundwater. The PSC of the soil would thus be expected to decrease

with time. This decrease would not only be caused by saturation of

retention sites of the soil particles, but also due to Fe leaching

from the upper soil horizons. Furthermore, the Fe that is leached out

of the soil might have a PSC much higher than that of the soil. Such a

high PSC would mean that reduced Fe-P complexes could present a means

of P loading in surface waters through groundwater. Mobile forms of P

would influence the design of land application systems.

Traditionally, PSC of a soil has rarely been considered a

limiting factor in the design of municipal wastewater land application

systems. Where P was considered, it was generally by assuming that P

was immobile in a soil horizon until the PSC of this horizon became

saturated with P. This approach was used to compare PSC with other

factors that should enter in the design of the size of land

application systems. Depending on the particular PSC evaluation

91

Page 108: Phosphate sorption in soils and reference oxides under

method, P could become a limiting factor for the application of

effluents of secondary treatment plants. Other limiting factors were N

and Cd.

A conceptual change might be required to further reduce the

potential hazards from P in the aquatic environment. This change would

require labelling P as "somewhat mobile", assigning a limit

concentration of P in seepage water and then calculating the

permissible rate of application of wastewater after all the other

potential pollutants have been considered as well. Nitrate application

rates are currently being evaluated using this concept.

In conclusion, the work presented here established the potential

impact of waste-P application on soils and the effect of such an

impact on the design of a land application system. Such an approach

should increasingly be introduced in the design of management systems

to limit the impact of wastewater application on the environment.

92

Page 109: Phosphate sorption in soils and reference oxides under

APPENDIX

REDOX POTENTIAL MEASUREMENT

A TOOL FOR SOIL INVESTIGATIONS

Introduction

Oxidation-reduction reactions are characterized by electron

transfer. Such reactions involve inorganic, organic, and biochemical

reactants and take place in all types of natural systems (Bohn, 1971).

In the soil system, oxidation-reduction (redox) reactions affect soil

formation and classification (Collins and Buol, 1970), soil fertility

(Schwab and Lindsay, 1983a,b), and chemical transport (Holford and

Patrick, 1979). This appendix is a review of the principal methods of

redox potential measurement, their application in the various fields

of soil science, and their use in laboratory experiments involving the

control of redox and pH levels.

Rationale for Redox Potential Measurement

Redox Reactions

At equilibrium, any reaction involving the transfer of electrons

(e) can be considered the summation of two half reactions in the

following manner (Sposito, 1981b):

mAQX + nH+ + e = pA^ed + 9H2°(1) (reduction) [1]

sBred - tBQX + e (oxidation) [2]

mAox + sBred + nH+ - pA^ + tBQX + qH20 (redox) [3]

93

Page 110: Phosphate sorption in soils and reference oxides under

Where AQX and Are(^ are the oxidized and reduced states of a chemical

species A (same for B) and e is the electron in solution that is

transfered from B QX (electron donor) to AQX (electron acceptor).

Two parameters -electron activity, and electrode potential- have

been suggested for the measurement of the relative tendency of an

environment (generally a solution) to attract or release electrons.

These two parameters are briefly discussed in the following two

sections.

Electron Activity

Considering that the electron is a reagent in oxidation and

reduction equations, a parameter pe can be defined as follows:

pe = -log{e) [4]

where (e) is the electron "activity" at equilibrium, in either

equation [1] or [2]. The term activity is not used here in its strict

chemical sense because the electrons are not free in solution.

Nevertheless, such a representation is useful because it expresses the

"electron pressure" in terms that are conceptually similar to the

treatment of pH in acid-base equilibria (Stumm and Morgan, 1981).

Consequently, a high pe would indicate a low electron activity and

high oxidizing conditions (i.e. regularly higher tendency to take

electrons). Low pe values indicate that reducing conditions prevail.

Electrode Potential

Similarly to pH, pe is actually determined by the electrochemical

measurement of an Electromotive Force (Emf) in volts. Eh is the

94

Page 111: Phosphate sorption in soils and reference oxides under

electrode redox potential and represents the voltage of the redox

reaction when measured against the standard hydrogen electrode.

The electrode potential Eh represents in fact the free energy

change ( G), (products - reactants) of each half reaction in calories,

G of equation [1], for example, would be:

G = G° + RT In[{A}/{B}] [5]

where G° is the free energy change at unit activity and {A} and {B}

the activities of the oxidizing chemical and its reduced form

respectively. By converting equation [5] to electrical energy the

Nernst equation is obtained:

Eh = Eq + RT/nF ln[{A}/{B}] [6]

where EQ = potential when the redox couple is at unit activity, and F is

the constant of Faraday in heating units.

Standard Electrode Potential (EQ) values for naturally occurring

redox reactions can be found in the literature (Stumm and Morgan,

1981).

As mentioned in the beginning of this section, pe values can be

derived from Eh measurements:

pe - (F/2.3 RT) Eh [7]

at 25°C, equation [7] can be written:

pe = Eh/0.0591 [8]

Lindsay (1979) found that, in a given soil suspension, pe

generally decreases when pH is raised; the value of pe + pH remaining

the same. He suggested that pe + pH be used as a parameter to describe

the redox status of suspensions.

95

Page 112: Phosphate sorption in soils and reference oxides under

Equilibrium vs. Non-Equilibrium Systems

In the previous sections the redox potential of systems that are

in equilibrium has been discussed. According to Bohn (1968),

equilibrium is reached when "the net donation and acceptance of

electrons is zero". But soils and other surface natural systems are

rarely in equilibrium for the following reasons: (i) the heterogeneity

of soil components, (ii) the exchanges with the surroundings, (iii)

the temperature and pressure variations and (iv) the slow reaction

rates. These factors would preclude the application of equilibrium

thermodynamics to redox reactions in the soil.

In ground water chemistry, such a conclusion was reached by

Lindberg and Runnells (1984) who found no internal equilibrium between

the various redox couples in 600 ground water samples. The difference

between measured and calculated Eh reached 1000 mV in many cases.

Measured pH was determined through the use of a platinum electrode and

calculated Eh was derived from the use of sample ionic concentrations

in published equilibrium equations.

Equations combining all significant redox couples in a non-

equilibrated system have been presented by Bohn (1971). It has been

demonstrated, however, that in natural systems the "final" redox

potential is not determined by all the redox couples present, but only

by the ones of significant concentration. Some redox couples do not

affect solution Eh, even at high concentrations. Runnells et al.

(1987), for example, showed that the Se(VI)/Se(IV) dynamics in

solution were not affected by Eh.

The redox couples that are considered to affect the soil redox

potential mostly are chemical entities that derive from the following

96

Page 113: Phosphate sorption in soils and reference oxides under

elements: iron, manganese, sulfur, nitrogen, oxygen and hydrogen

(Bohn, 1971). Sposito (1981a) added carbon to this list and, in the

case of contaminated soils, mentioned arsenic, chromium, copper and

lead. These "redox-significant" elements can even be used as indirect

indicators of approximate Eh values (Stumm and Morgan, 1981).

Even if the quantification of redox potential is a difficult and

rather "inaccurate" exercise, thermodynamics can define the trends and

possibilities of redox couples in the soil system. Furthermore, Eh

measurement is the only suitable means to quantify redox conditions in

reducing environments (Bohn, 1971). This is discussed in the following

section.

Waterlogged vs. Well-Drained Soils

Of the main contributors to soil Eh mentioned in the previous

section, oxygen is the most significant element in systems that are

exposed to the atmosphere. Oxygen is one of the most powerful

oxidizing agents in natural systems. When a soil is waterlogged, the

submerged portion -except for a thin layer in contact with air- is

generally depleted of oxygen and goes from oxidizing conditions to

reducing conditions (Patrick and Mahapatra, 1968). Waterlogging would

thus greatly widen the range of expected Eh values. While aerated

soils usually have Eh values of 400 to 700 millivolts, waterlogged

soils may have an Eh of as low as -300 volts. The wide range of Eh

values, the high concentration of reduced components and the

difficulties in the measurement of oxygen content all contribute in

making Eh the preferred tool for measuring the oxidation reduction

97

Page 114: Phosphate sorption in soils and reference oxides under

status of waterlogged soils (Patrick and Mahapatra, 1968). Reddy and

Patrick (1983) noted that it is not enough to study the effect of

water logging on the reactivity and mobility of soil constituents but

that it was necessary to relate such results to the O2 content of

water.

In general. Eh measurements in well-drained soils have been

unsatisfactory for a number of experimental reasons as well as the

narrowness of redox potential range (Ponnamperuma, 1972). A more

fundamental reason to favor the credibility of Eh measurements in

flooded soils is related to the relative stability of microbial

dynamics in such closed systems. For well oxidizing environments,

other methods such as O2 uptake from solution can be used (McBride,

1987).

The Biological Factor

This general discussion on redox potential would be incomplete

without considering the role of microbial activity in redox reactions.

Most redox reactions that occur in natural systems need microbial

organisms as mediators (Stumm and Morgan, 1981). These organisms do

not actually oxidize or reduce a substance, they make redox reactions

-that are thermodynamically possible- actually occur.

A thermodynamically possible reaction is one where conditions

favor the formation of products that are more stable than reactants.

Microbes cannot affect the thermodynamics of redox reactions, only

their kinetics. This catalytic role involves the acceleration of the

redox reactions and the enhancement of coupling between reduction and

oxidation half reactions (Lovley, 1987). In this respect, a

98

Page 115: Phosphate sorption in soils and reference oxides under

thermodynamic description of redox reactions should be more applicable

in a closed biological system (such as a flooded soil) than in an

open, aerated system (Sposito, 1981b).

In general, any effect on redox due to waterlogging or other

changes in the soil environment is in fact due to a shift in microbial

population and activity (Patrick and Mahapatra, 1968).

Redox potential studies and applications

In this section, the principal methods that have been used for Eh

measurement will be reviewed. First, the in-situ redox studies and

their limited applications will be discussed; then, the applications of

Eh values in laboratory measurements will be covered in greater

detail, and the different methodologies employed for Eh and pH control

will be presented and discussed.

In-Situ Eh Measurements

In a review article, Pormamperuma (1972) acknowledged the

usefulness of Eh measurements only as "semi-quantitative" indicators

of oxidation-reduction status, with limited diagnostic value for

agronomic interpretation. He mentioned that, in the past, in-situ

redox potential measurements have mainly dealt with the following

areas of investigation: (i) study of drainage problems, (ii) Mn, P and

Si release during waterlogging, (iii) solubilization and deposition of

minerals.

More recent applications of in-situ Eh measurements include: (i)

diagnosis of an aquic moisture regime in profiles without low chroma

99

Page 116: Phosphate sorption in soils and reference oxides under

color (Vepraskas and Wilding, 1983), (ii) correlation of Eh values

with natural vegetation productivity along a catena (De Laune et.

al., 1983), and nutrient distribution in submerged soils (Pedrazzini,

1983).

In-situ redox potential is usually measured with a voltmeter

using a palatinum electrode and a reference electrode (Stumm and

Morgan, 1981). The installation and retrieval of the Pt electrode is

often problematic in flooded soils and permanently placed electrodes

have been found to decrease in accuracy with time (Bailey and

Beauchamp, 1971; Fox and Rolfe, 1982). Use of a KCl-salt bridge

facilitates collection of year-round Eh measurements in seasonally dry

soils. (Pickering and Veneman, 1984; Veneman and Pickering 1983).

In-situ quantification of redox potential is used cautiously by

soil investigators, mainly because of the difficulties in measurements

and the limited use of Eh values obtained. Fox and Rolfe (1982)

devised a method to transport soil samples under N2 atmospheric

conditions, to the laboratory where Eh measurements can be performed

more accurately.

Laboratory Redox Potential Studies

Laboratory redox investigations usually are aimed at confirming

or refuting the theoretical behavior of chemical entities in the soil

as depicted by thermodynamically-derived stability diagrams (Lindsay

et al., 1981). The first attempts to test "calculated" stability

diagrams were done in aqueous solutions with the exclusion of soil

particles. Under such conditions Collins and Buol (19/0) found that Mn

and Fe followed the predicted results satisfactorily when each

100

Page 117: Phosphate sorption in soils and reference oxides under

chemical was used alone, but not when a mixture of Fe and Mn was used,

indicating the complexity of chemical interactions that can be

expected in mixed systems.

Bohn (1968) reported that many of the earlier soil redox studies

were unsatisfactory because of measurement defects caused by the use

of inadequate electrodes. He tested the response of platinum, gold,

and graphite electrodes to different oxidation states and found that

platinum was the most suitable. Platinum electrodes are still the most

widely used electrodes for Eh measurement (Mueller et al., 1985).

In the following sections some of the most recent methods used

for laboratory Eh-pH investigations will be discussed, and their

principal findings will be presented.

Soil Suspension Studies

In these studies the soil particles are kept in suspension to get

uniform conditions in all the systems studied and to expose a maximum

particle surface area (Letey et al., 1981).

Patrick et al. (1973) devised an apparatus that allowed the

control of pH and Eh conditions in a soil suspension. The main

features of the apparatus included: platinum electrodes for Eh

monitoring, a pH electrode, and two meter relays for Eh and pH

control. Redox potential was controlled by air (to increase Eh) and

nitrogen (to obtain a low potential). The pH was adjusted by the

controlled addition of either 0.5 N HCL or 0.5 N NaOH. Solenoid valves

permitted the automatic control of Eh to pre-set values. Attainable Eh

values with this system were reported to vary between +600 and -250 mV

(Patrick et al., 1973).

101

Page 118: Phosphate sorption in soils and reference oxides under

The set-up described above has been used to study particle

interactions with mineral ions, organic compounds, and gaseous

compounds. Of the main redox-relevant chemicals cited previously, soil

Fe and Mn have been studied most. Gotoh and Patrick (1974) studied the

effect of Eh and pH on Fe species in soil suspensions. Redox potential

was increased by the supply of one of several mixtures of O2 and N2,

and decreased by the supply of glucose to increase microbial activity

(leading to more reducing conditions). Iron fractionation and analysis

methods under reducing conditions were presented and the results

suggested that Fe activity in a flooded soil was mainly controlled

by the solubility of ferric oxyhydroxide.

Patrick and Henderson (1980) compared Fe and Mn behavior in

suspension and soil free systems. They described a method for sample

collection and preservation in a reducing N2 atmosphere and found that

Mn solubility was dependent almost completely on pH with little Eh

effect, while Fe transformations were more Eh-dependent. They also

found that Fe reduction in soil suspension occured at higher Eh and pH

values than in soil-free systems, probably due to microbial activity

(Patrick and Henderson, 1980).

An apparatus with the same features as described previously was

used by Schwab and Lindsay (1983a). They added CO2 to the oxidizing

mixture in order to keep a constant CO2 content and control the pH of

the calcareous soil suspension used. They found that in calcareous

soils Fe4* * activity was controlled by FeCO^ (siderite) when pe + pH

is less than 8 and by Fe3(0H)g (ferric hydroxide) when pe + pH is

greater than 8 (Schwab and Lindsay, 1983a). As for Mn, they found that

102

Page 119: Phosphate sorption in soils and reference oxides under

Mn solubility in a calcareous soil was controlled by MnCO-^

(rhodocrosite) when pe + pH 16 (Schwab and Lindsay, 1983b). They

also used a system developed by Reddy et al. (1976) to study the

effect of Eh and pH on uptake of Fe and Mn by plants. Manganese uptake

was found to decrease at pH values below 7, probably because of

competition from Fe++ (Schwab and Lindsay, 1983a; b).

While research is still in progress to elucidate many of the

remaining unanswered questions concerning Fe (Savant and McClellan,

1987) and Mn behavior in soil, a number of soil chemical substances

have been studied under varying Eh-pH conditions.

Sims and Patrick (1978) studied the effect of redox on the

distribution of certain micronutrients. They found that Fe, Mn, Zn,

and Cu that get solubilized under reduced conditions, became complexed

by organic matter occurring in the controlled reduced environment. In

sulfur rich soils, heavy metals solubility was found to be enhanced

because of increased acidity resulting from reducing conditions.

Sorption and mobility of phosphate in an acid soil suspension

were investigated by Holford and Patrick (1979). At pH values between

5 and 8, phosphate release was found to increase with more reducing

conditions. Also, at any given redox potential value (from -150 to

+500) increasing pH caused decreasing P levels in the soil solution.

It was concluded that Eh and pH were affecting P availability by

influencing Fe phosphates solubility (Holford and Patrick Jr., 1979).

The exact effect of Eh-pH on soil P, however, is still not elucidated

and needs more investigation, especially in view of the existing need

to assign soil P sorption indexes for the rational use of soils

(Burnham and Lopez-Hernandez, 1982).

103

Page 120: Phosphate sorption in soils and reference oxides under

As to organic compounds, Gambrell et al. (1984) demonstrated that

reduced conditions in soil suspensions can enhance the degradation of

DDT and Permethrin.

The study of equilibrium between the different species of a soil

component often involves a gaseous phase. Hunt et al. (1980) modified

the system of Patrick et al. (1973) to collect and analyze gaseous

emanations from incubated soil suspensions. The system consisted of

two independent units, one aerobic and the other anaerobic. This study

demonstrated the interaction between aerobic and anaerobic processes

for the production of ethylene, an important step in the C-cycle (Hunt

et al., 1982).

Gaseous products involved in the soil N-cycle were studied by

Letey et al. (1981), who modified Eh and pH levels in incubated jars of

soil slurries. They found that ^0 was reduced more rapidly than N0^.

In addition. Smith et al. (1983) found that, under certain redox

conditions, the soil could act as a sink for ^0.

In general, redox studies of stirred suspensions are useful and

have incorporated the soil factor in the study of chemical equilibrium

in soil solutions. In many cases, however, there is a need to study

redox potential of soil samples that are more closely related to field

conditions where soil particles are not in suspension (Patrick and

Henderson,1981).

Soil Column Studies

Some of the recent studies of soil Eh effects have dealt with

monitoring the changes in redox status in soil columns subjected to

varying physical and biochemical conditions.

104

Page 121: Phosphate sorption in soils and reference oxides under

In a column experiment, Baligar et al. (1980) found that redox

potential increased with increasing moisture stress and decreasing

bulk densities. Their results were related to soil aeration status.

However, no definite relationship was found between soil texture and

Eh. Moisture-Eh relationships were studied by Callebaut et al.

(1982). Their column studies showed that oxygen diffusion rate (ODR)

above a water table increased when the water table was lowered.

Oxygen-rich soil layers usually gave inaccurate Eh readings and no

definite relationship between water table depth and Eh was

established.

The gradual establishment of reducing conditions after saturation

of soil columns was studied by Farooqi and De Mooy (1983). They

observed that the peak in microbial activity coincided with the lowest

pe + pH values and occured after 48 hours of saturation. Microbial

activity and Eh were found related to the availability of organic

matter, especially C in the soil (Farooqi and De Mooy, 1983). The role

of organic matter was also investigated by Ghosh et al. (1982)

using sterile mixtures of soil, metals and humus. Potentiometric

titrations showed that humic substances can change the valence status

of Fe, Cu and Mn thus affecting the general redox conditions in the

soil even without the action of microorganisms (Ghosh et al., 1982).

The soil column studies cited above, monitored Eh as affected

by changing soil environments. To study the effect of redox

potentials on soil constituents it is necessary to be able to control

Eh levels. Patrick and Henderson (1981) devised a system for the

control of redox potential in compacted soil cores by the use of

105

Page 122: Phosphate sorption in soils and reference oxides under

saturation water and based on the ORD-Eh relationship mentioned

previously (Callebaut et al., 1982). The apparatus was used to study

the behavior of Fe and Mn at variable reducing conditions. The results

of the core study were in agreement with the soil suspension studies;

Mn and Fe were reduced at Eh levels of +400 and +200 mV respectively.

Soil texture was found to be a major concern for the success of the

procedure suggested; samples with a fine texture had to be mixed with

sand to permit successful Eh adjustment (Patrick and Henderson, 1981).

This of course introduces a new factor that may significantly alter

the natural soil conditions.

Conclusions

Reactions involving electron exchange occur in all natural

environnmental systems. This appendix covered the measurement and

control of redox potential and reviewed some of the applications to

soil investigations. The main findings and conclusions can be

summarized as follows:

i. The theory behind Eh quantification is sound (and as valid as pH

measurements) but field measured Eh values are still considered

semi-quantitative, principally because the theoretical state of

thermodynamic equilibrium is practically never present in the

natural environment.

ii. Eh measurement is most adapted to estimate "electron pressure" in

reducing environments and under controlled laboratory conditions.

iii. In-situ studies have yielded significant results, mainly

concerning the behavior of Fe and Mn in soils. Field

measurements, however, are limited because of practical

106

Page 123: Phosphate sorption in soils and reference oxides under

difficulties and do not allow for study where the redox potential

is controlled.

iv. Laboratory studies of the effect of Eh-pH levels on various soil

components concluded that some of the major effects are

modification of surface-water interfaces, and the dissolution of

solids (mainly oxides).

v. The results of soil column studies are questionable because they

require changing the nature of the soil under study.

107

Page 124: Phosphate sorption in soils and reference oxides under

REFERENCES

Amacher, M.C. , J. Kotuby-Amacher, H.M. Selim, and I.K. Iskandar. 1986.

Retention and release of metals by soils: evaluation of several

models. Geoderma 38:131-154.

Anderson, M.A., H.I. Tejedor-Tejedor, and R. R. Stanforth. 1985.

Influence of aggregation on the uptake of phosphate by goethite.

Environ. Sci. Technol. 19:632-637.

Atkinson, R.J., R.L. Parfitt, and R.C. Smart. 1974. Infra-red study of

phosphate adsorption on goethite. J. Chem. Soc. Faraday Trans. I.

1472-1479.

Atkinson, R.J., A.M. Posner, and J.P. Quirk. 1968. Crystal nucleation

in Fe(III) solutions and hydroxide gels. J. Inorg. Nucl. Chem.

30:2371-2376.

Bache, B.W. , and E.G. Williams. 1971. A phosphate sorption index for

soils. J. Soil Sci. 22:289-301.

Bailey, L.D., and E.G. Beauchamp. 1971. Nitrate reduction and redox

potentials measured with permanently and temporarily placed

platinum electrodes in saturated soils. Can. J. Soil Sci. 51: 51-

58.

Baligar, V.C., F.D. Whisler, and V.E. Nash. 1980. Soybean seedling

root growth as influenced by soil texture, matric suction and

bulk density. Commun. Soil Sci. Plant Anal. 11:903-915.

Barrow, N.J. 1984. Modeling the effects of pH on phosphate sorption by

soils. J. Soil Sci. 35:283-297.

Bascomb, C.I. 1968. Distribution of pyrophosphate-extractable iron and

organic carbon in soils of various groups. J. Soil Sci. 19:251-

268.

Belanger, T.V., D.F. Mikutel, and P.A. Churchill. 1984. Groundwater

seepage nutrient loading in a Florida lake. Water Res. 19:773-

781.

Berkheiser, V.E., J.J. Street, P.S.C. Rao, and T.L. Yuan. 1980.

Partitioning of inorganic orthophosphate in soil water systems.

CRC Critical Rev. Environ. Control 10:179-224.

Bohn, H.L. 1968. Electromotive force of inert electrodes in soil

suspensions. Soil Sci. Soc. Am. Proc. 32:211-215.

Bohn, H.L. 1971. Redox potentials. Soil Sci. 122:39-45.

108

Page 125: Phosphate sorption in soils and reference oxides under

Bohn, H.L., and R.K. Bohn 1986. Solid activity coefficients of soil

components. Geoderma 38:3-18.

Borggaard, O.K. 1983. Effect of surface area and mineralogy of iron

oxides on their surface charge and anion-adsorption properties.

Clays Clay Minerals 31:230-232.

Bouma, J. 1988. When the mapping is over, then what? p. 3. Soil

Resources: Their inventory, analysis and interpretation for use

in the 1990's. Intern. Interactive Workshop, 22-24 March, 1988.

Univ. Minnesota, Minneapolis, MN.

Bradley, J., I. Vimpany, and P.J. Nicholls. 1984. Effects of

waterlogging and subsequent drainage of a pasture soil on

phosphate and oxalate-extractable iron. Aust. J. Soil Res.

22:455-461.

Breeuwsma, A., J.H.M. Wosten, J.J. Vleeshouver, A.M. Van Slobbe, and

J. Bouma. 1986. Derivation of land qualities to assess

environmental problems from soil surveys. Soil Sci. Soc. Am. J.

50:186-190.

Burnham, C.P., and D. Lopez-Hernandez. 1982. Phosphate retention in

different soil taxonomic classes. Soil Sci. 134:376-380.

Callebaut, F., D. Gabriels, W. Minjauw, and M. De Boodt. 1982. Redox

potential, O2 diffusion rate and soil gas composition in

relation to water table in two soils. Soil Sci. 134:149-156.

Canter, L.W., and R.C. Knox. 1985. Septic tank system effects and

groundwater quality. Lewis Publishers, Inc. Chelsea, MI, 336 p.

Chang, A.C., A.L. Page, R.H. Sutherland, and E. Grgurevic. 1983.

Fractionation of phosphorus in sludge-affected soils. J. Environ

Qual. 12:286-290.

Collins, J.F., and S.W. Buol. 1970. Effects of fluctuations in the

Eh-pH environment and/or manganese equilibria. Soil Sci. 110:111

118.

De Laune, R.D., C.J. Smith, and W.H. Patrick, Jr. 1983. Relationship

of marsh elevation, redox potential, and sulfide to Spartina

alterniflora productivity. Soil Sci. Soc. Am. J. 47: 930-935.

Diggle, A.J., and L.C. Bell. 1984. Movement of applied phosphorus

following the mining and revegetation of mineral sands on

Australia's east coast. Aust. J. Soil Res. 22:135-148.

Dirasian, H. 1968. Electrode potentials - Significance in biological

systems. Part 1: Fundamentals of measurement. Water Sewage Works

40:420-425.

109

Page 126: Phosphate sorption in soils and reference oxides under

Dixon, J.B., D.C. Golden, F.G. Calhoun, and P.R. Buseck. 1983.

Synthetic aluminous goethite investigated by high resolution

transmission electron microscopy. Proc. Meetings Electron Microscopy 41:192-193.

Draper, N.R., and H. Smith. 1981. Applied regression analysis. 2nd

ed., John Wiley, New York. 709 p.

Dzombak, D. A. 1986. Toward a uniform model for the sorption of

inorganic ions on hydrous oxides. Ph.D. diss. Massachusetts

Inst. Technology, Cambridge, MA.

Evans, A. Jr. 1985. The adsorption of inorganic phosphate by a sandy

soil as influenced by dissolved organic compounds. Soil Sci.

140:251-255.

Everett, D.H., and R.H. Otterwill (eds.). 1969. Surface area

determination. Butterworths, London. 405 p.

Farooqi, M.A.R., and C.J. De Mooy. 1983. Effect of soil saturation

on redox potential and microbial activity. Commun. Soil Sci.

Plant Anal. 14:185-197.

Fox, R., and G. Rolfe. 1982. A method for collecting soil solutions

from flooded soils. Soil Sci. Soc. Am. J. 46:1332-1333.

Fuller, W.H., and A.W. Warrick. 1985. Soils in waste treatment and

utilization. Vol. I and II. CRC Press, Inc., Boca Raton, FL.

Gambrell, R.P., C.N. Reddy, V. Collard, G. Green, and W.H. Patrick,

Jr. 1984. The recovery of DDT, kepone, and permethrin added to

soil and sediment suspensions incubated under controlled redox

potential and pH conditions. J. Water Pollut. Control Fed.

56:174-182.

Gastuche, M.C., and A. Herbillon. 1962. Etude des gels d'alumine:

crystallisation en milieu desionise. Bull. Soc. Chim. Fr. 1404-

1412.

Gee, G.W., and J.W. Bauder. 1986. Particle-size analysis. In: A. Klute

(ed.) Methods of soil analysis. Part 1. 2nd ed. Agronomy

9:383-412. Am. Soc. Agron., Madison, WI.

Ghosh, K. , A. Chattopadhyay, and C. Varadachari. 1982. Electron

exchange behavior of humic substances with Fe, Cu, and Mn. Soil

Sci. 135:193-196.

Gilliom, R.J., and C.R. Patmont. 1983. Lake phosphorus loading from

septic systems by seasonally perched groundwater. J. Water

Pollut. Control Fed. 55:1297-1305.

110

Page 127: Phosphate sorption in soils and reference oxides under

Goldberg, S. 1983. A chemical model of phosphate adsorption on

oxide minerals and soils. Ph.D. diss. University of California, Riverside, CA.

Goldberg, S. 1985. Chemical modeling of anion competition on goethite

using the constant capacitance model. Soil Sci. Soc. Am. J. 49:851-856.

Goldberg, S., and G. Sposito. 1984a. A chemical model of phosphate

adsorption by soils: I. Reference oxide minerals. Soil Sci. Soc.

Am. J. 48:772-778.

Goldberg, S., and G. Sposito. 1984b. A chemical model of phosphate

adsorption by soils: II. Non calcareous soils. Soil Sci. Soc. Am. J. 48:779-783.

Goldberg, S., and G. Sposito. 1985. On the mechanism of specific

phosphate adsorption by hydroxylated mineral surfaces: A review.

Commun. Soil Sci. Plant Anal. 16:801-821.

Gotoh, S., and W.H. Patrick, Jr. 1974. Transformation of iron in a wa¬

terlogged soil as influenced by redox potential and pH. Soil

Sci. Soc. Am. Proc. 38:66-71.

Gschwend, P.M., and M.D. Reynolds. 1987. Monodisperse ferrous

phosphate colloids in an anoxic groundwater plume. J. Contam.

Hydro1. 1:309-327.

Harter, R.D. 1984. Curve-fit errors in Langmuir adsorption maxima.

Soil Sci. Soc.Am. J. 48:749-752.

Harter, R.D., and D.E. Baker. 1977. Applications and misapplications

of the Langmuir equation. Soil Sci. Soc. Am. J. 48:779-783.

Hill, D.E., and B.L. Sawhney. 1981. Removal of phosphorus from

wastewater by soil under aerobic and anaerobic conditions. J.

Environ. Qual. 10:401-405.

Hillel, D. 1977. Computer simulation of soil water dynamics: a

compendium of recent work. Int. Dev. Res. Center, Ottawa, 214 p.

Hingston, F.J. 1981. A review of anion adsorption, p. 51-90. In: M.A.

Anderson, and A.J. Rubin(eds.). Adsorption of inorganics at solid-

liquid interfaces. Ann Arbor Science, Ann Arbor, MI.

Hingston, F.J., A.M. Posner, and J.P.Quirk. 1972. Anion adsorption by

goethite and gibbsite: I. The role of the proton in determining

adsorption envelopes. J. Soil Sci. 23:177-192.

Hohl, H., L. Sigg, and W. Stumm. 1980. Characterization of surface

chemical properties of oxides in natural waters, p. 1-31. In:

M.C. Kavanaugh, and J.O. Leckie (eds.). Particulates in water.

Adv. in Chem. Series, No. 189. Washington, DC.

Ill

Page 128: Phosphate sorption in soils and reference oxides under

Holford, I.C.R. 1979. Evaluation of soil phosphate buffering indices.

Aust. J. Soil Res. 17:495-504.

Holford, I.C.R. 1982. The comparative significance and utility of the

Freundlich and Langmuir parameters for characterizing sorption

and plant availability of phosphate in soils. Aust. J. Soil Res. 20:233-242.

Holford, I.C.R., and W.H. Patrick, Jr. 1979. Effects of reduction and

pH changes on phosphate sorption and mobility in an acid soil.

Soil Sci. Soc. Am. J. 43:292-297.

Holford, I.C.R., and W.H. Patrick Jr. 1981. Effects of duration of

anaerobiosis and reoxidation on phosphate sorption

characteristics of an acid soil. Aust. J. Soil Res. 19:69-78.

Holford, I.C.R., R.W.M. Wedderbum, and C.E.G. Mattingly. 1974. A

Langmuir two-surface equation as a model for phosphate adsorption

by soils. J. Soil Sci. 25:243-255.

Honeyman, B.D. 1984. Cation and anion adsorption at the oxide/solution

interface in systems containing binary mixtures of adsorbents: an

investigation of the concept of adsorptive additivity. Ph. D.

diss. Stanford University, Stanford, CA. UMI No. 8420552.

Hook, J.E. 1983. Movement of phosphorus and nitrogen in soil following

application of municipal wastewater, p 241-255. In: Chemical

mobility and reactivity in soil systems. Soil Sci. Soc. Am.

Special Publication number 11. Madison, WI.

Houck, C.P., S.P. Putnam, and R.C. Loehr. 1987. Initial operations

report on sludge farm. J. Environ. Engin. 113:970-978.

Humenik, F.J., M.D. Smolen, and S.A. Dressing. 1987. Pollution from

nonpoint sources. Where we are and where we should go. Environ.

Sci. Technol. 21:737-742.

Hunt, P.G., T.A. Matheny, R.B. Campbell, and J.E. Parsons. 1982.

Ethylene accumulation in southeastern coastal plain soils: Soil

characteristics and oxidative-reductive involvement. Commun. Soil

Sci. and Plant Anal. 13:267-278.

Hunt, P.G., T.A. Matheny, J.E. Parsons, and R.B. Campbell. 1980. An

aerobic-anaerobic interactive system for soil suspensions. Soil

Sci. Soc. Am. J. 44:1328-1329.

Hsu, P.H. 1977. Aluminum hydroxides and oxyhydroxides p. 99-138. In:

J.B. Dixon and S. Weed (eds.) Minerals in soil environments. Soil

Sci. Soc. Am,. Madison, WI.

Iskandar, I.K. 1981. Modeling wastewater renovation. John Wiley, New

York. 410 p.

112

Page 129: Phosphate sorption in soils and reference oxides under

James, R.O., and G.A. Parks. 1982. Characterization of aqueous

colloids by their electrical double-layer and intrinsic surface

chemical properties, p. 119-216. In: E. Matijevic Surface Colloid

Sci., Vol. 12, Plenum Press, New York.

Jeanson, C. 1972. Etude microscopique de depots de fer, de manganese

et de calcium dans un sol experimental, leur association avec des

microorganismes. Rev. Ecol. Biol. Sol. 9:479-489.

Johnson, D.W. , D.W. Cole, H. Van Miegroet, and F.W. Homg. 1986.

Factors affecting anion movement and retention in four forest

soils. Soil Sci. Soc. Am. J. 50:776-783.

Jury, W.A. 1983. Chemical transport modeling: current approaches and

unresolved problems. In: D.W. Nelson (ed.) Chemical mobility and

reactivity in soil systems. Soil Sci. Soc. Am. Special Publ. No.

11., Madison, WI.

Khalid, R.A., W.H. Patrick Jr., and R.D. Delaune. 1977. Phosphorus

sorption characteristics of flooded soils. Soil Sci. Soc. Am. J.

41:305-310.

King,L.D. 1985. Design and management considerations in applying

minimally treated wastewater to land. p. 15-26. In: Utilization,

treatment, and disposal of waste on land. Proc. Workshop.

Chicago, IL. 6-7 December, 1985. Soil Sci. Soc. Am., Madison, WI.

Kleiss, H.J., and M.T. Hoover. 1986. Soil and site criteria for on¬

site systems, p. 111-128. In: K.W. Brown (ed.) Utilization,

treatment, and disposal of waste on land. Soil Sci. Soc. Am.

Madison, WI.

Kuo, J-H., and T.F. Yen. 1988. Some aspects in predicting the point of

zero charge of a composite oxide system. J. Colloid Interface

Sci. 121:220-225.

Langmuir, I. 1918. The adsorption of gases on plane surfaces of glass,

mica, and platinum. J. Am. Chem. Soc. 40:1361-1403.

Larsen, S. 1967. Soil Phosphorus. Adv. Agron. 19:151-206.

Laverdiere, M.R., and A. Karam. 1984. Sorption of phosphorus by some

surface soils from Quebec in relation to their properties.

Commun. Soil Sci. Plant Anal. 15:1215-1230.

Lee, F.Y., and J.A. Kittrick. 1984. Electron microprobe analysis of

elements associated with zinc and copper in an oxidizing and an

anaerobic soil environment. Soil Sci. Soc. Am. J. 48:548-554.

Letey, J., N. Valoras, D.D. Focht, and J.C. Ryden. 1981. Nitrous oxide

production and reduction during denitrification as affected by

redox potential. Soil Sci. Soc. Am. J. 45:727-730.

113

Page 130: Phosphate sorption in soils and reference oxides under

Light, T.S. 1972. Standard solution for redox potential measurements. Anal. Chem. 44:1038-1039.

Lindberg, R.D., and D.D. Runnells. 1984. Ground water redox

reactions: An analysis of equilibrium state applied to Eh

measurements and geochemical modeling. Science 255:925-927.

Lindsay, W.L. 1979. Chemical equilibria in soils. J. Wiley Inc., New

York. 449p.

Lindsay, W.L., and E.C. Moreno. 1960. Phosphate phase equilibria in

soils. Soil Sci. Soc. Am. Proc. 25:177-182.

Lindsay, W.L., M. Sadiq, and L.K. Porter. 1981. Thermodynamics of

inorganic N transformation. Soil Sci. Soc. Am. J.45:61-66.

Liverman, D.M., M.E. Hanson, B.J. Brown, R.W. Meredith. 1988. Global

sustainability: toward measurement. Environ. Management 12:133-

143.

Loehr, R.C. 1985. The role of soil science in the utilization,

treatment, and disposal of waste, p. 305-318. In: Utilization,

treatment, and disposal of waste on land. Proc. Workshop.

Chicago, IL. 6-7 December, 1985. Soil Sci. Soc. Am., Madison, WI.

Loehr, R.C., W.J. Jewell, J.D. Novak, W.W. Clarkson, and G.S.

Friedman. 1979. Land application of waste. Vol. I and II. Van

Nostrand Reinhold Co., New York.

Loehr, R.C., and M.R. Overcash. 1985. Land treatment of wastes:

concepts and general design. J. Environ. Engin. 111:141-160.

Loganathan, P., N.O. Isirimah, and D.A. Nwachuku. 1987. Phosphorus

sorption by Ultisols and Inceptisols of the Niger delta in

southern Nigeria. Soil Sci. 139:330-338.

Lovley, D.R. 1987. Organic matter mineralization with the reduction

of ferric iron: A review. Geomicrobiol. J. 5:375-399.

Lyman, W.J. 1982. Adsorption coefficient for soils and sediments, p.

4.1-4.33. In: W.J. Lyman, W.F. Reehl, and D.H. Rosenblatt, (eds.)

Handbook of chemical property estimation methods. Mc.Graw Hill,

New York.

Mansell, R.S., and H.S. Selim. 1981. Mathematical models for

predicting reactions and transport of phosphorus applied to

soils. In: I.K. Iskandar (ed.) Modeling waste water renovation.

J. Wiley, New York.

Massey, D.T. 1983. Features of federal law that encourage adoption of

land application of wastewater. J. Water Res. Planning Management

109:121-133.

114

Page 131: Phosphate sorption in soils and reference oxides under

McBride, M.B. 1987. Adsorption and oxidation of phenolic compounds

by iron and manganese oxides. Soil Sci. Soc. Am. J. 51:1466- 1472.

McKeague, J.A., J.E. Brydon, and N.M. Miles. 1971. Differentiation

of forms of extractable iron and aluminum in soils. Soil Sci. Soc. Am. Proc. 35:33-38.

McLaughlin, J.R., J.C. Ryden, and J.K. Syers. 1981. Sorption of

inorganic phosphate by iron and aluminum containing components. J. Soil Sci. 32:31-36.

Moore, I.C., and F.W. Madison. 1985. Description and application of an

animal waste phosphorus loading model. J. Environ. Qual. 14:364-

369.

Morel, F.M.M. 1983. Principles of aquatic chemistry. John Wiley, New

York. 319 p.

Morel, F.M.M., and P.M. Gschwend. 1987. The role of colloids in the

partitioning of solutes in natural waters, pp 405-422. In: W.

Stumm (ed.) Aquatic Surface Chemistry. J. Wiley, New York.

Morel, F.M.M., G.J. Yeasted, and J.C. Westall. 1981. Adsorption

models: a mathematical analysis in the framework of general

equilibrium calculations, p. 263-294. In: M.A. Anderson, and A.J.

Rubin (eds.). Adsorption of inorganics at solid-liquid interfaces.

Ann Arbor Science, Ann Arbor, MI.

Mueller, S.C., L.H. Stolzy, and G.W. Fick. 1985. Constructing and

screening platinum microelectrodes for measuring soil redox

potential. Soil Sci. 139:558-560.

Munsell Color. 1975. Munsell soil color charts. Munsell Color Comp.

Baltimore, MD.

Murphy, J., and J. P. Riley. 1962. A modified single solution method

for the determination of phosphate in natural waters. Anal. Chim.

Acta 27:31-36.

Nagpal, N.K. 1985. Long-term phosphorus sorption in a Brunisol in

response to dosed-effluent loading. J. Environ. Qual. 14:280-285.

Nair, P.S., T.J. Logan, A.N. Sharpley, L. E. Sommers, M.A. Tabatabai,

and T.L. Yuan. 1984. Interlaboratory comparison of a standardized

phosphorus adsorption procedure. J. Environ. Qual. 13:591-595.

Nelson, D.W., and L.E. Sommers. 1986. Total carbon, organic carbon,

and organic matter. In: A.L. Page et al. (ed.) Methods of soil

analysis. Part 2. 2nd ed. Agronomy 9:539-579. Am. Soc. Agron.,

Madison, WI.

115

Page 132: Phosphate sorption in soils and reference oxides under

Norstadt, F.A., N.P. Swanson, and B.R. Sabey. 1977. Site design and

management for utilization and disposal of organic wastes, p.

348-376. In: L.F. Elliot, and F.J. Stevenson (eds.) Soil for

management of organic wastes and wastewaters. Soil Sci. Soc. Am., Madison, WI.

Noss, R. 1984. The US approach to water pollution control. Environ. Int. 10:217-223.

Olsen, S.R., and F.S. Watanabe. 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.

Overcash, M.R., and D. Pal. 1979. Design of land treatment systems for

industrial wastes; theory and practice. Ann Arbor Sci. Publishers

Inc., Ann Arbor, MI.

Parfitt, R.L., A.R. Fraser, J.D. Russell, and V.C. Farmer. 1977.

Adsorption on hydrous oxides. II. Oxalate, benzoate and phosphate

on gibbsite. J. Soil Sci. 28:40-47.

Parker, J.C., and P.M. Jardine. 1986. Effects of heterogeneous

adsorption behavior on ion transport. Water Res. Research

22:1334-1340.

Patrick, W.H. Jr., and R.E. Henderson. 1980. Reduction and reoxidation

cycles of manganese and iron in flooded soil and in water

solution. Soil Sci. Soc. Am. J. 45:855-859.

Patrick, W.H. Jr., and R.E. Henderson. 1981. A method for controlling

redox potential in packed soil cores. Soil Sci. Soc. Am. J.

45:35-38.

Patrick, W.H. Jr., and I.C. Mahapatra. 1968. Transformation and

availability to rice of nitrogen and phosphorus in water-logged

soils. Adv. Agron. 20:323-359. Am. Soc. Agron., Madison, WI.

Patrick, W.H. Jr., B.G. Willingood, and J.T. Moraghan. 1973. A

simple system for controlling redox potential and pH in soil

suspensions. Soil Sci. Soc. Am. J. 37:331-332.

Pedrazzini, F.R. 1983. Mineral nitrogen distribution, redox potential,

and pH of rice paddy soils profiles in northern Italy. Commun.

Soil Sci. Plant Anal. 14:309-320.

Peech, M., R.L. Cowan, and J.H. Baker. 1962. A critical study of the

BaC^-triethanolamine and ammonium acetate methods for

determining the exchangeable hydrogen content of soils. Soil Sci.

Soc. Am. Proc. 26:37-40.

Pickering, E.W., and P.L.M. Veneman. 1984. Moisture regimes and

morphological characteristics in a hydrosequence in central

Massachusetts. Soil Sci. Soc. Am. J. 48:113-118.

116

Page 133: Phosphate sorption in soils and reference oxides under

Pi-ng» C.L., and G.J. Michaelson. 1986. Phosphorus sorption by major

agricultural soils of Alaska. Commun. Soil Sci. Plant Anal.

17:299-320.

Ponnamperuma, F.N. 1972. The chemistry of submerged soils. Adv.

Agron. 24:29-88. Am. Soc. Agron., Madison, WI.

Polyzopoulos, N.A., V.Z. Keramidas, and H. Kiosse. 1983. Phosphate

sorption by some Alfisols of Greece as described by commonly used

isotherms. Soil Sci. Soc. Am. J. 49:81-84.

Ratkowsky, D.A. 1986. A statistical study of seven curves for

describing the sorption of phosphate by soil. J. Soil Sci.

37:183-189.

Runnells, D.D., R.D. Linberg, and J.H. Kempton. 1987. Irreversibility

of Se(VI)/Se(IV) redox couple in synthetic basaltic ground water

at 25°C and 75°C. Mat. Res. Soc. Symp. Proc. 84:723-733.

Reddy, C.N., A. Jugsujinda, and W.H. Patrick, Jr. 1976. System for

growing plants under controlled redox potential-pH conditions.

Agron. J. 68:987-988.

Reddy, K.R., and W.H. Patrick, Jr. 1983. Effects of aeration on

reactivity and mobility of soil constituents. In: Chemical

mobility and reactivity in soil systems. SSSA Special

Publication No. 11. Soil Sci. Soc. Am., Madison, WI.

Sah, R.N., and D.S. Mikkelsen. 1986. Transformations of inorganic

phosphorus during the flooding and draining cycles of soil. Soil

Sci. Soc. Am. J. 50:62-67.

SAS Institute, Inc. 1985. SAS/STAT guide for personal computers,

version 6. SAS Institute, Inc, Cary, NC. 378 p.

Savant, N.K., and R. Ellis Jr. 1964. Changes in redox potential and

phosphorus availability in submerged soil. Soil Sci. 98:388-394.

Savant, N.K., and G.H. McClellan. 1987. Do iron oxides influence soil

properties and nitrogen transformations in soils under wetland

rice-based cropping systems? Commun. Soil Sci. Plant Anal.

18:83-113.

Sawhney, B.L. 1973. Electron microprobe analysis of phosphate in soils

and sediments. Soil Sci. Soc. Am. Proc. 37:658-660.

Schwab, A.P., and W.L. Lindsay. 1983a. Effect of redox on the

solubility and availability of iron. Soil Sci. Soc. Am. J.

47:201-205.

Schwab, A.P., and W.L. Lindsay. 1983b. The effect of redox on the

solubility of manganese in a calcareous soil. Soil Sci. Soc.

Am. J. 47:217-220.

117

Page 134: Phosphate sorption in soils and reference oxides under

Schwarz, J.A. , C.T. Driscoll, and A.K. Bhanot. 1984. The zero point of

charge of silica-alumina oxide suspensions. J. Colloid Interface

Sci. 97:55-61.

Schweich, D., and M. Sardin. 1981. Adsorption, partition, ion exchange

and chemical reaction in batch reactors or in columns- A review.

J. Hydro1. 50:1-33.

Schwertmann, D., and R.M. Taylor. 1977. Iron oxides, p.145-180. In:

J.B. Dixon, and S.B. Weed (eds.) Minerals in soil environments.

Soil Sci. Soc. Am., Madison, WI.

Settergren, D., and L.C. Tennyson. 1979. The importance of soil in

evaluating potential wastewater disposal sites in the Missouri

Ozarks. Water Res. Bulletin. 15:1353-1364.

Sibanda, H.M., and S.D. Young. 1986. Competitive adsorption of humus

acids and phosphate on goethite, gibbsite and two tropical soils.

J. Soil Sci. 37:197-204.

Sibbesen, E. 1981. Some new equations to describe phosphate sorption

by soils. J. Soil Sci. 32:67-74.

Sigg, L. and W. Stumm. 1980. The interaction of anions and weak acids

with the hydrous goethite surface. Colloids Surfaces 2:101-

117.

Sims, J.L. , and W.H. Patrick, Jr. 1978. The distribution of

micronutrient cations in soil under conditions of varying redox

potential and pH. Soil Sci. Soc. Am. J. 42:258-262.

Singh, B.B., and M.A. Tabatabai. 1977. Effects of soil properties on

phosphate sorption. Commun. Soil Sci. Plant Anal. 8:97-107.

Sircar, S. 1984. Adsorption of gases on heterogeneous adsorbents. J.

Chem. Soc. Faraday Trans. 1. 80:1101-1111.

Smillie, G.W. , D. Curtin, and J.K. Syers. 1987. Influence of

exchangeable calcium on phosphate retention by weakly acid soils.

Soil Sci. Soc. Am. J. 51:1169-1172.

Smith, C.J., M.F. Wright, and W.H. Patrick, Jr. 1983. The effect of

soil redox potential and pH on the reduction of nitrous oxide. J.

Environ. Qual. 12:186-188.

Soil Survey Staff. 1972. Soil survey laboratory methods and procedures

for collecting soil samples. Soil Survey Investigations Report

No.1. U.S.D.A. Soil Cons. Serv., Govern. Printing Office,

Washington, DC.

Sposito, G. 1981a. The operational definition of the zero point of

charge in soils. Soil Sci. Soc. Am. J. 45:292-297.

118

Page 135: Phosphate sorption in soils and reference oxides under

Sposito, G. 1981b. The thermodynamics of soil solutions. Oxford

Clarendon Press. Oxford, England.

Sposito, G. 1982. On the use of the Langmuir equation in the

interpretation of "adsorption" phenomena: II. The "two-surface"

Langmuir equation. Soil Sci. Soc. Am. J. 46:1147-1152.

Sposito, G. 1985. Chemical models of inorganic pollutants in soils.

CRC Crit. Rev. Environ. Control 15:1-24.

Steele, D.P., G.W. Petersen, and D.D. Fritton. 1986. Soil potential

ratings for on-site wastewater disposal. J. Environ. Qual. 15:37-

44.

Stuanes, A.O. 1982. Phosphorus sorption by soil; a review, p. 145-152.

In: A.S. Eikum and R.W. Seabloom (eds.) Alternative wastewater

treatment. D. Reidel Publishing Co., Dordrecht, The Netherlands.

Stuanes, A.O. 1984. Phosphorus sorption of soils to be used in

wastewater renovation. J. Environ. Qual. 13:220-224.

Stuanes, A.O., and C.G. Enfield. 1984. Prediction of phosphate

movement through some selected soils. J. Environ. Qual. 13:317-

320.

Stumm, W. 1986. Coordinative interactions between soil solids and

water - An aquatic chemist's point of view. Geoderma 38:19-30.

Stumm, W., R. Kummert, and L. Sigg. 1980. A ligand exchange model for

the adsorption of inorganic and organic ligands at hydrous oxides

interfaces. Croatica Chemica Acta. 53:291-312.

Stumm, W.J., and J. Morgan. 1981. Aquatic chemistry. 2nd ed. Wiley

Interscience Publication, John Wiley, New York.

Switzenbaum, M.S., J.V. De Pinto, T.C. Young, and J.K. Edzwald. 1980.

A survey of phosphorus removal in lower Great Lakes municipal

treatment plants. J. Water Pollut. Control. Fed. 25:2628-2633.

Syers, J.K., and I.K. Iskandar. 1981. Soil-phosphorus chemistry, p.

571-599. In: I.K. Iskandar (ed.) Modeling wastewater renovation,

land treatment. John Wiley, New York.

Time. 1989. Endangered Earth. Planet of the Year. Time Magazine

133:25-72.

Thomas, R., and J.P. Law. 1977. Properties of waste waters, p. 46-72.

In: L.F. Elliott and F.J. Stevenson (eds.) Soils for management

of organic waste and waste waters. Am. Soc. Agron., Madison, WI.

119

Page 136: Phosphate sorption in soils and reference oxides under

Tirsch, F.S., and J.H. Baker. 1976. Heavy metal reactions in soils, p.

II.1~II.26. In: D.D. Adrian (ed.) Suitability of New England soil

in accepting waste effluents, phase I. Water Res. Research

Center. Univ. Massachusetts, Amherst, MA.

Tofflemire, T.J., and M. Chen. 1976. Evaluation of phosphate

adsorption capacity and cation exchange capacity and related data

for 35 common soil series in New York state. Technical Report No.

45, NY State Dept, of Environ. Conservation, Albany, NY.

Travis, C.C., and E.L. Etnier. 1981. A survey of sorption

relationships for reactive solutes in soil. J. Environ. Qual.

10:8-17.

Turner, K. (ed.) 1988. Sustainable Environmental Management,

Principles and Practices. Belhaven Press, London/Westview Press,

Boulder, CO. 292 pages.

Uebler, R.L. 1984. Effect of loading rate and soil amendments on

organic nitrogen and phosphorus leached from wastewater soil

absorption system. J. Environ. Qual. 13:475-479.

U.S. Environmental Protection Agency. 1981. Land treatment of

municipal wastewater. EPA 625/1-81-013. U.S.E.P.A., Center for

Environmental Research Information, Cincinnati, OH.

Van Riemsdijk, H., G.H. Bolt, L.K. Koopal, and J. Blaakmeer. 1986.

Electrolyte adsorption on heterogeneous surfaces: adsorption

models. J. Colloid Interface Sci. 109:219-228.

Van Riemsdijk, W.H., and J. Lyklema. 1980a. Reaction of phosphate with

gibbsite (A1(0H)^) beyond the adsorption maximum. J. Colloid

Interface Sci. 76:55-66.

Van Riemsdijk, W. H., and J. Lyklema. 1980b. The reaction of phosphate

with aluminum hydroxide in relation with phosphate bonding in

soils. Colloids and Surfaces 1:33-44.

Van Riemsdijk, W.H., and A.M.A. van der Linden. 1984. Phosphate

sorption by soils: II. Sorption measurement technique. Soil Sci.

Soc. Am. J. 48:541-544.

Van Riemsdijk, W.H., F.A. Westrate, and J.Beek. 1977. Phosphates in

soils treated with sewage water: III. Kinetic studies on the

reaction of phosphates with Al compounds. J. Environ. Qual. 6:26-

29.

Veith, J.A. , and G. Sposito. 1977. On the use of the Langmuir equation

in the interpretation of "adsorption" phenomena. Soil. Sci. Soc.

Am. J. 41:697-702.

120

Page 137: Phosphate sorption in soils and reference oxides under

Veneman, P.L.M. 1982. The septic tank as a wastewater treatment and

disposal facility. Water Resour. Res. Center Publ. No. 128.

University of Massachusetts, Amherst, MA. 45 p.

Veneman, P.L.M., and E.W. Pickering. 1983. Salt bridge for redox

potential measurements. Commun. Soil Sci. Plant Anal.

14:669-677.

Vepraskas, M.J., and L.P. Wilding. 1983. Aquic moisture regimes in

soils with and without low chroma colors Soil Sci. Soc. Am. J.

47:280-285.

Vijayachandran, P.K., and R.D. Harter. 1975. Evaluation of phosphorus

adsorption by a cross section of soil types. Soil Sci. 119:119-

126.

Vollenweider, R.A. 1968. Scientific fundamentals of the eutrophication

of lakes and flowing waters. Organization for Economic

Cooperation and Development. Technical Report DA5/SCI/68.27,

Paris, France.

Westall, J. 1979. MICROQL: I. A chemical equilibrium program in basic.

Swiss Federal Inst, of Technology, Duebendorf, Switzerland.

Westall, J.C. 1987. Adsorption mechanisms in aquatic surface

chemistry. 3-32. In: W. Stumm (ed.) Aquatic Surface Chemistry.

Wiley, NY.

Westall, J.C. 1982. FITEQL: A computer program for determination of

chemical equilibrium constants from experimental data. Report 82-

01, Dept, of Chemistry, Oregon State Univ., Corvallis, OR.

Westall, J., and H. Hohl. 1980. A comparison of electrostatic models

for the oxide/solution interface. Adv. in Colloid and Interface

Sci. 12:265-294.

Westman, W.E. 1972. Some basic issues in water pollution control

legislation. Am. Scientist 60:767-773.

White, R.E. 1981. Retention and release of phosphate by soil and soil

constituents. CRC Critical Rev. Applied Chem. 2:71-114.

Willett, I.R. 1983. Oxidation-reduction reactions pp.417-416 In:

Soils, An Australian Viewpoint. Division of Soils, CSIRO, CSIRO:

Melbourne/Academic Press: London.

Willett, I.R. 1985. The reductive dissolution of phosphated

ferrihydrite and strengite. Aust. J. Soil Res. 23:237-244.

Willett, I.R., and R.B. Cunningham. 1983. Influence of sorbed

phosphate on the stability of ferric hydrous oxide under

controlled pH and Eh conditions. Aust. J. Soil Res. 21:301-308.

121

Page 138: Phosphate sorption in soils and reference oxides under

Yuretich, R.F., P.J. Stekl, and S.J. Pohanka. 1986. Hydrogeology and

mineral weathering reactions in watersheds of central

Massachusetts. Publication No. 152. Water Res. Research Center.

Univ. Massachusetts, Amherst, MA.

122

Page 139: Phosphate sorption in soils and reference oxides under