117
Sorption reactions of 1,4-dichlorobenzene in low organic carbon soils Item Type Thesis-Reproduction (electronic); text Authors Klein, Adam,1959- Publisher The University of Arizona. Rights Copyright © is held by the author. Digital access to this material is made possible by the University Libraries, University of Arizona. Further transmission, reproduction or presentation (such as public display or performance) of protected items is prohibited except with permission of the author. Download date 09/06/2021 17:13:43 Link to Item http://hdl.handle.net/10150/191906

Sorption reactions of 1,4-dichlorobenzene in low organic ......4 rph f dt fr nt bth xprnt 42 5 rph f ,4 DB dt fr l ln xprnt 4 6 rph f hlrd dt fr l ln xprnt 4 7 hlrd Brthrh rv th fttd

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • Sorption reactions of 1,4-dichlorobenzenein low organic carbon soils

    Item Type Thesis-Reproduction (electronic); text

    Authors Klein, Adam,1959-

    Publisher The University of Arizona.

    Rights Copyright © is held by the author. Digital access to this materialis made possible by the University Libraries, University of Arizona.Further transmission, reproduction or presentation (such aspublic display or performance) of protected items is prohibitedexcept with permission of the author.

    Download date 09/06/2021 17:13:43

    Link to Item http://hdl.handle.net/10150/191906

    http://hdl.handle.net/10150/191906

  • SORPTION REACTIONS OF 1,4-DICHLOROBENZENE

    IN LOW ORGANIC CARBON SOILS

    by

    ADAM KLEIN

    A Thesis Submitted to the Faculty of the

    DEPARTMENT OF HYDROLOGY AND WATER RESOURCES

    in Partial Fulfillment of the RequirementsFor the Degree of

    MASTER OF SCIENCEWITH A MAJOR IN HYDROLOGY

    In the Graduate College

    THE UNIVERSITY OF ARIZONA

    1986

  • Roger Bales, Professor ofHydrology and Water Resources

    STATEMENT BY AUTHOR

    This thesis has been submitted in partial fulfill-ment of requirements for an advanced degree at the Univer-sity of Arizona and is deposited in the University Libraryto be made available to borrowers under rules of thelibrary.

    Brief quotations from this thesis are allowablewithout special permission, provided that accurate acknow-ledgment of source is made. Requests for permission forextended quotation from or reproduction of this manuscriptin whole or in part may be granted by the head of the majordepartment or the Dean of the Graduate College when in hisor her judgement the proposed use of the material is in theinterests of scholarship. In all other instances, however,permission must be obtained from the author.

    SIGNED:

    APPROVAL BY THESIS DIRECTOR

    This thesis has been approved on the date shown below.

    R4T/Ag 6Date

  • ACKNOWLEDGMENTS

    I would like to thank my parents, who have always placed a

    high value on my education; without their emotional and

    financial support this would never have been possible. I'd

    like to thank Pat Pascal, who provided the divine guidance

    necessary to get our Varian 7630 working; this research is

    only as good as its detection levels. Many thanks to

    Mama's Pizzeria, for needed nutritional support and a good

    space to think and to Bentley's, for lots of coffee, late

    night crams, and a place to relax and sort this mess out.

    I can't possibly thank Augusta Davis enough; in addition to

    guiding me through the administrative mazes of both the

    University and this department, she has the distinction of

    being the only person to survive the front office for the

    duration of my stay here. Finally, there are Jim, Kris and

    Tom, who suffered through this entire process with me; U of

    A water chemsistry program in its nascent stages is no fun.

    Even though it was often the blind leading the blind, I

    learned alot with you guys, and wish you the best of luck

    in everything. Thanks for an ear to hear my gripes, ques-

    tions, and realizations. This research was performed at

    the University of Arizona under the guidance of Dr. Roger

  • Bales. Partial support for this work was obtained from the

    Motorola Corporation. I would like to express my thanks to

    R. Lee from Motorola inc., and Rebecca Pruitt from Dames

    and Moore for their help on this project.

    iv

  • TABLE OF CONTENTS

    Page

    LIST OF ILLUSTRATIONS vii

    LIST OF TABLES viii

    ABSTRACT ix

    1. INTRODUCTION 1

    2. BACKGROUND SORPTION THEORY 5

    2.1 Hydrophobic Sorption 62.2 Sorption Forces 72.3 Sorption Isotherms 9

    2.4 Sorption Kinetics 102.4.1 Physical Kinetic Processes 112.4.2 Chemical Kinetic Processes 122.4.3 Chemical Composition of Soils . . • 122.4.4 Comparison of Kinetic Models . . . • 13

    2.5 Sorption Partition Coefficients 142.5.1 Effect of Soil Organic Matter

    on Sorption 142.5.2 Kp /K oc Relationships 152.5.3 K oc /Kow Relationships 162.5.4 Effect of Aqueous Solubility on

    Partitioning 192.5.5 Use of Empirical Equations for

    Describing Partitioning 212.6 Soil Clays 23

    2.6.1 Clay Sorption Mechanisms 23

    2.6.2 Clay-Organic Interactions 24

    3. EXPERIMENTAL PROCEDURES AND METHODS 27

    3.1 Experimental Approach 27

    3.2 Soils Description 283.3 Equilibrium Batch Procedures 293.4 Problems with Batch Procedures 323.5 Kinetic Batch Procedures 33

    3.6 Desorption Batch Procedures 333.7 Soil Column Experimental Procedures . . 34

    3.8 Chemical Analysis 37

  • TABLE OF CONTENTS--Continued

    vi

    Page

    4. EXPERIMENTAL RESULTS 38

    4.1 Equilibrium Batch Results 384.2 Kinetic Batch Results 414.3 Soil Column Results 44

    4.3.1 Conductivity Data fromColumn Experiments 44

    4.3.2 1,4-DCB Data 464.4 Anomalies in Column Experimental Results . . 48

    5. DISCUSSION 52

    5.1 Curve-Fitting Analysis of Column Experiments 535.1.1 Models and Input Parameters in CFITIM 535.1.2 Boundary Conditions in CFITIM . . . . 555.1.3 Column Experimental Analysis by CFITIM

    Equilibrium Model 565.1.4 Analysis by non-equilibrium models . 64

    5.2 Use of Empirical Equations to Predict Kp . . 705.3 Comparison of Koc Values 72

    6. CONCLUSIONS 77

    APPENDIX A 83

    APPENDIX B: GRAPHS OF CHLORIDE AND 1,4 DCB EFFLUENT

    APPENDIX C:

    APPENDIX D:

    DATA FROM COLUMN EXPERIMENTS . . . . 84

    DIMENSIONLESS PARAMETERS FROMCFITIM MODELS 90

    GRAPHS OF FITTED VS. OBSERVED EFFLUENTDATA FROM CFITIM EQUILIBRIUM MODELFOR 1,4-DICHLOROBENZENE 91

    APPENDIX E: GRAPHS OF FITTED VS. OBSERVED EFFLUENTDATA FROM CFITIM EQUILIBRIUM MODEL FORCONDUCTIVITY 97

    APPENDIX F: SUMMARY OF PARAMETER EXTIMATES FROMCFITIM MODELS D AND E . . . . . . 103

    REFERENCES 104

  • LIST OF ILLUSTRATIONS

    Figure Page

    1 Schematic representation of columnexperimental setup 35

    2 Sorption isotherm from batch experimentswith soil from borehole 2V2DB 39

    3 Sorption isotherm from batch experimentswith soil from borehole 102SG 40

    4 Graph of data from kinetic batch experiments 42

    5 Graph of 1,4 DCB data fromsoil column experiments 45

    6 Graph of chloride data fromsoil column experiments 45

    7 Chloride Breakthrough Curves with fittedpoints from CFITIM equilibrium model analysis 57

    8 1,4-DCB Breakthrough Curves with fittedpoints from CFITIM equilibrium model analysis 58

    vii

  • LIST OF TABLES

    Table Page

    1 Physical characteristics of 1,4 dichlorobenzene 4

    2 Koc/ Kow empirical relationships 18

    3 Size fractions of soils used in the experiments 30

    4 Parameters from column experiments 49

    5 Parameter estimates from the CFITIMequilibrium model 60

    6 Dispersion coefficients generated by theCFITIM equilibrium model 62

    7 Summary of Alpha Values 69

    8 Summary of Omega Values 69

    9 Koc Values for 1,4-DCB 73

    10 Summary of Kp Values from CFITIMNonequilibrium Models 76

    v i i i

  • ABSTRACT

    The rate and extent of sorption of 1,4-dichloro-

    benzene (1,4-DCB) was studied using column and batch

    experiments. Column experiments with a soil with fraction

    organic carbon (foc) = 0.00086 yielded a soil/water

    partition coefficient (K r ) of 0.41; mass balance on thesorption and desorption limbs of the breakthrough curves

    gave similar Kp's, indicating sorption was readily rever-

    sible. A computer program that fits column effluent data

    to analytical solutions of the advection-dispersion

    equation under different models of sorption behaviour gave

    Kp = 0.46, assuming equilibrium sorption. The breakthrough

    curves for 1,4-DCB showed slight tailing when plotted

    against the fitted data, indicating some slow sorption.

    The time scale for sorption/desorption estimated by this

    program was up to 10-100 times larger than (physical)

    transport times in the column, but was of the same order as

    transport times in the field.

    ix

  • CHAPTER 1

    INTRODUCTION

    Organic solvents are present in groundwater,

    creating a possible hazard to the public's drinking water.

    As the pollutant plumes are discovered and mapped, recla-

    mation strategies will be formulated. Central to this

    reclamation effort is determining how (and if) synthetic

    organic compounds will be transformed, and their movement

    retarded, as they travel in solution through aquifers.

    Retardation of a solute as it travels through an aquifer

    occurs due to sorption reactions. For sorption reactions

    the questions become: 1) how much will the organics sorb to

    aquifer solids, 2) what is the rate at which these com-

    pounds desorb from the solids as the water is progressively

    cleaned, and 3) what are the mechanisms responsible for the

    sorption reactions.

    Transport of chemicals (pollutants, salts, tracers,

    etc.) in groundwater can be described by the advection-

    dispersion equation. The one dimensional form of this

    C))_1-(= - v - 240s r61- - b(z 3x

    where C is the aqueous concentration of the sorbate (jig

    equation is:

    (1)

    1

  • sorbate/cm 3 water), v is the average linear (Darcian)

    velocity (cm/sec), and D is the dispersion coefficient

    (cm 2 /sec). The term on the left hand side describes the

    change in concentration in solution of the compound of

    interest with time. The first term on the right hand side

    describes diffusive transport of the compound, the second

    term describes advective transport of the compound, and the

    last term describes losses and gains of the compound due to

    all chemical reactions. For transport of water or a

    conservative tracer such as KC1, this last term is not

    used. Equation 1 assumes steady-state water flow, a

    constant diffusivity, and a constant soil-water content.

    The chemical reactions term, Ekiri, refers to the

    change in concentration of a chemical with time. Changes

    occur because of interactions with other chemicals in the

    water (such as complexation, degradation, or biological

    transformations), or interactions with the solid matrix,

    primarily sorption reactions. When the chemical reaction

    term is limited to sorption reactions, equation 1 becomes:

    _ „ e (2)3A" — X'

    where S is the sorbed concentration (jig sorbate/g soil), p

    is the soil bulk density (g soil/cm 3 soil) and 0 is the

    total porosity of the soil in the column (cm 3 void space/

    cm3 total volume). This current research is concerned with

    determining the rate and extent of sorption reactions in

    relation to groundwater contamination at a site in the

    2

  • southwestern U.S.

    The chemical under study is 1,4-dichlorobenzene

    (1,4-DCB), a volatile, mainly nonpolar, organic compound.

    A list of 1,4-DCB's relevant characteristics is given in

    Table 1. Soils used in the sorption studies have a low

    fraction organic carbon (foc), and were taken from bore-

    holes near the Motorola 52nd St. production plant, in

    Phoenix, Arizona.

    The purpose of this research was to determine

    equilibrium levels of 1,4-DCB sorption onto a variety of

    soils from the field site in order to obtain sorption

    isotherms, and to determine if those isotherms are linear.

    An additional objective was to determine if partitioning

    between the liquid and solid phases takes place within

    seconds, hours, days or weeks, and to obtain an estimate

    for the reaction rate coefficient if partitioning is

    kinetic. Desorption was also examined, to determine if the

    sorption reactions are completely reversible.

    3

  • TABLE I

    Selected Physical Properties of 1,4-Dichlorobenzene

    Molecular Weight 147.01(Weast, 1977)

    4

    Melting Point(Weast, 1977)

    Boiling Point at 760 torr(Weast, 1977)

    Vapor Pressure at 25 °C(Weast, 1977)

    Solid solubility in Water at 25 °C(Verschueren, 1977)

    Supercooled Liquid Solubility(Chiou, 1981)

    Log octanol/water partition coefficient(Leo et. al., 1971)

    Henry's Law Constant

    53.1 °C

    174 °C

    1.18 torr

    79 mg/1

    137 mg/1

    3.39

    0.34 atm/M

  • CHAPTER 2

    BACKGROUND SORPTION THEORY

    In general, sorption refers to a process by which

    ions or molecules accumulate more in one phase than

    another, typically at the boundary between those phases

    (Sposito, 1984). According to Adamson (1982) two pictures

    of liquid/solid sorption exist. In the more simplified

    two-dimensional picture, adsorption is confined to a

    monolayer next to the surface. The more complex picture is

    that of a three-dimensional interfacial layer, multimolecu-

    lar in depth, over which several different sorption

    processes may be occuring simultaneously. Sorption from

    solution in this case is a partitioning between a bulk and

    an interfacial phase of the solution.

    The two driving forces for sorption onto a solid are

    the lyophobic (solvent-disliking) characteristics of the

    solute with respect to the solvent, and a general affinity

    of the solute for the solid (Weber, 1972). For aqueous

    solutions, the first force is termed hydrophobicity (water

    repulsion) or hydrophyllicity (water attraction). Sorption

    which occurs due to the first process is often called

    partitioning, while sorption due to the second process is

    5

  • 6

    often called adsorption. Collectively, these two processes

    are called sorption.

    2.1 Hydrophobic Sorption

    The more hydrophobic a compound is, the more

    likely it is to partition into the solid phase. Since

    water is a polar molecule, and non-polar organics, such as

    1,4-DCB are relatively hydrophobic, compounds such as these

    will move out of water and towards organic compounds in

    soil. This is primarily due to water-water interactions.

    Introduction of an organic molecule into water disrupts the

    configuration of water molecules. In order to achieve

    maximum entropy, water-water interactions will "push" the

    organic out of aqueous solution. As the solute moves from

    the aqueous to the solid phase, a large positive entropy

    develops, due to the dehydration of the solute molecules

    (Schwarzenbach et al., 1981). This type of hydrophobic

    interaction is largely responsible for the sorption of 1,4-

    DCB onto organic material in soil, and is often called

    hydrophobic sorption (Voice and Weber, 1983, Hamaker and

    Thompson, 1972), although more accurately might be termed

    hydrophobic partitioning. Hydrophobic sorption increases

    as compounds become less polar (larger carbon chains or

    fewer substituents), or as water solubilities decrease

    (Hassett and Means, 1980).

  • 2.2 Sorption Forces

    The second force behind sorption reactions is the

    surface-solute attractive force. This force is generally

    divided into three types: physical, chemical and electro-

    static. For hydrophobic solutes (predominantly non-ionic

    neutral compounds), large-scale electrostatic forces do not

    play a part in sorption reactions (Voice and Weber, 1983).

    However, for slightly polar molecules in soils with a high

    clay/organic matter ratio, these forces may be important

    for sorption (Hassett and Means, 1980), and will be discus-

    sed more thoroughly in section 2.6.

    For a completely nonpolar compound, physical

    sorption is due to van der Waals forces, nonspecific, weak,

    electrostatic attractions between molecules. Values for

    van der Waals-type interactions for small molecules are

    generally on the order of 1 to 2 kcal mo1 -1 (Hamaker and

    Thompson, 1972).

    Chemical sorption involves electronic interactions

    between specific sites on the sorbent surface and solute

    molecules. The solute typically will form a monomolecular

    layer over the sorbing surface. This can result in bonds

    that have a large energy of sorption, 15-50 kcal mol -1

    (Hamaker and Thompson, 1972). However, a substantial

    activation energy may be required in order for the reaction

    to occur. Consequently chemisorption is slowly reversible,

    in contrast to physical sorption, which is readily revers-

    7

  • ible.

    The partitioning of organic solutes out of aqueous

    solution onto soil organic components is primarily due to

    both hydrophobic partitioning and van der Waals interac-

    tions, although it is often difficult to distinguish

    between chemical and physical sorption (Voice and Weber,

    1983). It is important to remember that sorption is a

    surface phenomenon, and as such will also be a function of

    the surface properties of the sorbent. Because of this,

    sorption onto clay minerals as well as sorption onto

    organic carbon must be considered. In the present context

    of sorption, one must not forget that we are dealing with

    the three-dimensional, multimolecular depth surfaces

    mentioned before.

    For the purposes of the rest of this paper, sorption

    will be defined as the transfer of a solute (contaminant,

    i.e. 1,4-DCB) from the liquid phase (groundwater) to the

    solid phase (soil particles). Desorption is the corre-

    sponding transfer of a solute from the solid into solution.

    The main effect of sorption is to retard the mean rate of

    the movement of a solute through an aquifer, relative to

    the Darcian velocity of the water, as measured by a

    conservative tracer (one which does not interact with the

    soil particles or the other solutes). In addition,

    sorption may catalyze or inhibit the breakdown of organic

    compounds (Mortland, 1970).

    8

  • 9

    sorption may catalyze or inhibit the breakdown of organic

    compounds (Mortland, 1970).

    2.3 Sorption Isotherms

    One of the major goals of this research project was

    to determine sorption isotherms for 1,4-dichlorobenzene on

    soils from the field site. A sorption isotherm is a ratio

    between the concentration of a solute on the solid phase

    vs. the concentration of a solute remaining in the liquid

    phase, at equilibrium at a given temperature. There are

    several types of equations for describing sorption iso-

    therms; most common are those developed by Langmuir and

    Freundlich, which are nonlinear isotherms. However, for

    modelling contaminant transport in groundwater, the most

    frequently used isotherm is a linear isotherm. While this

    widespread use is generally attributed to the mathematical

    simplicity of the linear isotherm, it should be noted that

    the Langmuir isotherm reduces to linear partitioning

    relationships under conditions of dilute solutions (low

    concentrations of the solutes) (Voice and Weber, 1983).

    Because the present work deals with concentrations in the

    part per million (PPM) and part per billion (PPB) range, a

    linear isotherm is postulated. This isotherm is of the

    form:

    K xC=S

    (3)

    where Kp is the equilibrium partition coefficient (cm 3 /g),

    S is the concentration of solute adsorbed to the solid

  • 10

    phase at equilibrium (ug solute/g soil), and C is the

    equilibrium concentration of the solute remaining in

    solution (pg solute/cm 3 water).

    2.4 Sorption Kinetics

    Historically, when the chemical reaction term of the

    advection-dispersion equation has been considered, the

    kinetics of the reaction(s) have been ignored. Instaneous

    equilibrium was assumed, with the reaction following a

    linear isotherm (equation 3). Within the last twenty

    years, however, more complex models of the sorption

    reaction have been developed, which look at the kinetics of

    the reaction. Central to these models is the mechanism of

    the sorption/desorption reaction and the determination of

    the rate limiting step of the reaction. Kinetic sorption

    necessarily implies competing processes of sorption and

    desorption, which are occuring at the same time. This is

    given as:

    A(aq)-..?.-A(s) (4)

    where A(aq) is desorption into solution, and A(s) is

    sorption onto the sorbate. First order kinetic descrip-

    tions of these reactions are of the form:

    èS ,i,--,----- f x C/rsw - kb x S (5)Oi.

    where kf (sec -1 ) is the forward (sorption) reaction rate

    constant, kb (sec -1 ) is the reverse (desorption) reaction

    rate constant, and rsw is the soil/water ratio (g soil/cm 3

    water).

  • 1 1

    Kinetic sorption models are divided into two basic

    categories; in one, physical processes are assumed respon-

    sible for the kinetic rate of the reaction, and in the

    second category chemical processes are assumed to be the

    rate controlling step.

    2.4.1 Physical Kinetic Processes

    The physical process models partition soil water

    into mobile and immobile regions (Rao et al., 1979).

    Convective-dispersive transport of solutes through the soil

    takes place only in the mobile region, although diffusion

    through the stagnant region is also taking place. The

    sorption reactions are assumed to be instantaneous. It is

    the rate at which the sorbate approaches active sites on

    the soil surface in the stagnant region that controls the

    rate at which the solute will sorb onto the soil. The

    solute must first diffuse through stagnant water films to

    reach soil surface sites before the instantaneous sorption

    reaction can occur. According to the model of Rao et al.

    (1983), solute transfer between the mobile and stagnant

    regions is described by Fick's second law for diffusion.

    van Genuchten (1981) divides sorption sites into two

    fractions, one which is in close contact with the mobile

    liquid, and one which is in contact only with immobile

    water. It is hypothesized that the larger pores contain

    the sites in contact with the mobile liquid, and therefore

    these sites will experience instantaneous sorption, while

  • 12

    the smaller intraaggregate pores are in contact with the

    immobile liquid fraction, and therefore these sites will

    experience diffusion-controlled sorption. Rao et al.

    (1983) point out that most laboratory column experiments

    are conducted with sieved soil, which will have no aggre-

    gates and small particle sizes. However, no matter how

    much processing the soil undergoes before it is used in

    experiments, it appears that it will always contain some

    microaggregates. Thus it seems that the physical process

    model cannot be eliminated from the discussion describing

    the results from these experiments.

    2.4.2 Chemical Kinetic Processes

    The chemical process models (Cameron and Klute,

    1977) divide sorption sites on soils into two types: type 1

    sites, where chemicals sorb rapidly, producing an instanta-

    neous equilibrium at that site, and type 2 sites, where

    chemicals sorb more slowly, resulting in a kinetic reac-

    tion. A subset of this model is one where all sites are

    assumed to undergo kinetic-type reactions. Conceptual

    justification for this type of model is based on the

    heterogeneous nature of soil.

    2.4.3 Chemical Composition of Soils

    The soil solid phase consists of crystalline

    primary and secondary minerals (mostly layer silicates and

    metal hydroxides), mineral colloids (mostly oxides and

  • 13

    hydroxides of silicon, iron, manganese and aluminum) and

    organic particles (Ahlrichs, 1972). Soil organic matter

    consists of carbohydrates, proteins, fats, waxes, resins,

    pigments, and low molecular weight compounds physically

    associated with humic acids (Kenaga and Goring, 1983);

    together these components form amorphous humic colloids.

    Typically these are characterized by high molecular weight,

    aromatic structures, and acidic hydroxyl and carbonyl

    functional groups (Kenaga and Goring, 1983). See Ahlrichs

    (1972) for a functional group analysis of organic matter

    from two soils. These components are present in different

    combinations and ratios in different soils. Thus, a

    chemical moving through the soil/water environment may

    react instantaneously with organic matter and slowly with

    mineral clays, or rapidly with one type of humic substance

    and slowly with a different type. The fraction assumed

    responsible for kinetic sorption reactions has not been

    isolated.

    2.4.4 Comparison of Kinetic Models

    Rao et al. (1979) evaluated one model from each

    category, to determine how well they predict experimental

    results. Both models evaluated assume a two-site sorp-

    tion/desorption system, where sorption onto the type one

    sites is instantaneous, and sorption onto the type two

    sites is nonlinear, and kinetically significant. These

    authors concluded that both models can adequately describe

  • 14

    the assymetrical breakthrough curves(BTCs) obtained

    experimentally for their solute, but that different sets of

    parameters were required to predict the BTCs at two

    different input concentrations. van Genuchten (1981)

    contains a mathematical description of both types of

    models; he notes that mathematically, they both reduce to

    the same partial differential equation.

    2.5 Sorption Partition Coefficients

    Researchers have long sought to demonstrate rela-

    tionships between fundamental characteristics of organic

    compounds and soil components. Lambert et al. (1965) and

    Lambert (1967,1968) were some of the first to emphasize the

    importance of the soil organic matter fraction (OM) to

    sorption reactions.

    2.5.1 Effect of Soil Organic Matter on Sorption

    Lambert presumed that all of soil sorption was a

    partitioning of the solute onto soil organic matter and,

    with that assumption, calculated a Kp based on soil OM

    rather than total soil mass:

    Kp = Cp m /C w (6)

    where Cm is the concentration of solute on soil organic

    matter (jig solute/g soil OM), and C w is the concentration

    of solute in water (pg solute/g water)(Lambert, 1965). His

    experimental results showed that this partition coefficient

    is relatively constant for a given solute across a range of

  • 15

    different soils, and thus is a characteristic of that

    solute. Deviations from this model will occur if 1) all of

    the soil OM (as determined by standard TOC analysis) does

    not participate in the sorption reactions, or 2) if the

    chemical exhibits some anomalous behaviour, such as a pH

    dependance or ion exchange reactions. This type of

    behaviour is primarily a function of a chemical's struc-

    ture, and its substituent groups. Kenaga and Goring (1979)

    expanded on this by stating that deviations from this model

    will occur because a) there are inherent differences

    between soils in the sorption characteristics of its

    organic matter, and h) there may be an impact on sorption

    due to other soil properties (i.e. swelling clay content.

    By restricting the model to uncharged, organic molecules,

    deviations due to ion exchange reactions are eliminated.

    2.5.2 Kp /K oc Relationships

    Work by later researchers (Schwarzenbach et al,

    1981, Karikchoff, 1979) has verified Lambert's research;

    the more common form of the relationship is:

    Koc = Kp/foc (7)

    where K oc is a partition coefficient normalized for organic

    carbon, or alternatively is the partition coefficient onto

    a hypothetical 100% organic sorbent, and fo c is the

    fraction organic carbon of the actual sorbent. Karikchoff

    (1984) states that "(based on available documentation) it

    can be safely generalized that for uncharged organic

  • 16

    compounds of limited water solubility (

  • 17

    log K e = a(log Ke l ) + a (9)where Ke and K e l are the distribution coefficients between

    a solute and two different, immiscible liquid phases. a

    and a are constants.

    Several researchers have derived empirical K oc /K ow

    relationships for different compounds. These are given in

    Table 2; note that they are all of the same form as

    equation 8. Schwarzenbach et al .(1981) state that their

    relationship is valid only for dilute solutions, and for

    soil with f oc > 0.001, and that the constant a in equation

    9, giving the slope of the linear regression for equation 8

    is a function of the free energy of transfer of the solute

    from the aqueous to the nonaqueous phase.

    Karickhoff (1981) demonstrates the thermodynamic

    basis for equation 14, showing that for sufficiently dilute

    systems, the proportionality constant for K oc /K ow is the

    ratio of the fugacity coefficients for the solute dissolved

    in octanol, and the solute bound in natural organic matter.

    For this linear relationship to be valid, the fugacity

    ratio must be independent of the solute; thus compounds not

    showing this independence may not be amenable to this type

    of treatment. The earlier work by Karickhoff (equation 10)

    as well as the work of Kenaga and Goring (equation 11) were

    not justified by the authors on the basis of any physico-

    chemical theory. In fact, the work by Kenaga and Goring

    was only a log-linear fit of Kp vs. Kow values obtained in

  • 18

    TABLE II

    EMPIRICAL KOW/KOC RELATIONSHIPS

    Equation

    No. Ref.

    log K oc = 1.00 x log Kow + 0.21 (10) Karikchoff(10 aromatic and chlorinated (1979)

    hydrocarbons)

    log Ko c = 0.544 x log Kow + 1.377 (11) Kenaga and(45 assorted organic compounds) Goring (1979)

    (12) Hassett and(22 polynuclear aromatics) Means (1980)

    10 9 Koc = 0.72 x log K ow + 0.49(13 halogenated alkenes and

    benzenes)

    log Koc = 0.989 x log Kow - 0.346(5 polynuclear aromatics)

    log K oc = 1.029 x log Kow - 0.18(13 pesticides)

    (13) Schwarzenbach(1981)

    (14) Karikchoff(1981)

    (15) Rao (1982)

    10 9 Koc = 1.00 x log K ow - 0.317

  • 19

    the literature for 45 different compounds; this included

    the data from Karickhoff (1979).

    2.5.4 Effect of Aqueous Solubilityon Partitioning

    Chiou et al. (1982,1983) developed a linear regres-

    sion for log K ow and log S, where S refers to aqueous

    solubility. Theoretical considerations in this derivation

    centered on innaccuracies in reported K ow and S values.

    Determination of K ow from octanol/water mixtures will

    create an aqueous phase saturated with octanol. This will

    result in higher than actual aqueous solubilities due to

    the presence of the organic in the aqueous phase. This

    will lead to a decrease in the actual Kow. This effect

    will not happen in soil/water mixtures, because very little

    of the soil organic matter is soluble in water. Under

    ideal conditions (i.e. solubility in water and octanol-

    saturated water are the same) the linear regression should

    be:

    log K ow = -1.0 x log S + 0.92 (16)

    Nonideal conditions will produce a downward deviation from

    the ideal line; the effect is most pronounced for compounds

    with very low water solubilities.

    Further deviations will occur for solutes which are

    solids at room temperature (1,4-DCB is in this category).

    The solubilities used in the log-linear regressions in

    Table 2 must be the solubility of the supercooled liquid

  • 20

    (Karickhoff, 1981, Chiou, 1982, 1983) and not the solubil-

    ity of the solid, due to the nonideal solution behaviour of

    the supercooled liquid. This is often termed the melting

    point effect. Downward deviations from ideal behaviour

    will occur because supercooled liquids have reduced aqueous

    and solvent solubility. Thus, while the aqueous solubility

    decreases, the octanol/water partition coefficient remains

    the same. For 16 aromatic compounds, Chiou (1982) found

    the linear regression under the assumption of ideal

    behaviour to be:

    log K ow = -1.004 x log S + 0.34 (17)

    Taking into account the non-idealities the equation

    becomes:

    log Kow = -0.862 x log S + 0.710

    (18)

    Karickhoff (1981) sought to compensate for the

    melting point effect by addition of a "crystal energy

    term":

    log Koc = -a 10 9 X501 + b + "crystal energy term" (19)

    where X so l is the mole fraction solubility of the solute in

    water. This term is a function of the melting point and

    the entropy of fusion for the solute; for the compounds in

    Karickhoff's study this term was empirically determine to

    be -0.00953(mp-25), where rap is the melting point of the

    solute in °C. Note that for compounds which are liquids at

    room temperature, this term vanishes. The regressions for

  • 21

    5 condensed ring aromatics without and with the crystal

    energy term are:

    log K oc = -0.594 x log Xsol - 0.197 (20)

    log K oc = -0.921xlog X s0 1-0.00953 (mp-25)-1.405 (21)

    Not only does this term improve the fit of Karickhoff's

    data to a linear regression, it also moves the slope

    (-0.921) much closer to unity, the coefficient under ideal

    conditions. After applying equation 21 to literature data,

    it was found (Karickhoff, 1984) that the equation worked

    well for both aliphatic and aromatic chlorinated hydro-

    carbons of low molecular weight, but for highly chlorinat-

    ed, high molecular weight compounds (such as DDT), the

    equation tended to overestimate sorption.

    2.5.5 Use of Empirical Equations forDescribing Partitioning

    Variations in the equations in Table 2 are due to

    several factors. The logK oc /logK ow linear dependence is

    predicated on strict hydrophobic sorption; polar group or

    ionic interactions are assumed negligble. Soil organic

    carbon is presumed to be the dominating component governing

    sorption; the impact of swelling clays in the soil on

    sorption is also assumed to be negligble. Hamaker et

    al.,(1972) suggest that variations in Koc values for

    different soils may be due to variations in the composition

    of the organic matter complex. High organic matter soils

  • 22

    may see a reduction in the sorption surface per unit weight

    of organic matter, due to a "piling up" of the organic

    matter. Conversely, soils with a very low organic matter

    content may show a significant amount of sorption by the

    mineral phase. Thus, high organic matter soils may show

    low K oc values, while very low organic matter soils may

    show a high K oc value. Finally, computations of Kow for

    large molecules based on addition of K ow 's for the substi-

    tuents may overestimate the actual K ow for these larger

    molecules (Karickhoff, 1984). Despite this variability, we

    have seen that the coupling of linear regressions with

    thermodynamic or linear free-energy theory can improve the

    fit of experimental data, and allow for the extrapolation

    of the regression to different classes of compounds.

    The importance of the equations is that an a priori

    estimate of the sorption partition coefficient for a given

    compound can be developed, based on information that is

    available in the literature. This can then be used as a

    very good first approximation to determine if the mobility

    of a given organic chemical poses a threat to water

    sources. In doing this, care must be taken in choosing the

    appropriate relationships from listed equations; this

    choice must be based on a careful consideration of both the

    physical properties of the compound of interest, and the

    properties of the sorbing medium.

  • 23

    2.6 Soil Clays

    Because of the high surface areas of clays and col-

    loids, these parts of the soil may also play a role in

    sorption of organics. There are extensive data available

    that show that sorption of pesticides onto pure clays is

    significant; however it must be remembered that both

    organic matter present in soils and exchangeable cations

    present in the soil solution may profoundly modify the

    sorption properties of the clays with respect to organics

    (Hamaker, 1972).

    2.6.1 Clay Sorption Mechanisms

    Sorption of organics onto clays happens by a variety

    of mechanisms (Mortland, 1970). Ion exchange reactions may

    occur with the exchangeable cations on the clay mineral

    surface; this mechanism is pH dependent, and is significant

    primarily for organics with amine or carbonyl groups.

    Sorption of polar, nonionic organics can occur through ion-

    dipole interactions. Again, cations in the exchange

    complex serve as sorption sites for this mechanism. The

    greater the cation's affinity for electrons, the greater

    the energy of interaction with polar groups on organic

    molecues; polyvalent cations are more electrophilic than

    monovalent cations, and thus transition metals will form

    the strongest bonds by this mechanism. For cations with

    high solvation energies, sorption occurs through a "water

    bridge"; functional groups on the organic will hydrogen

  • 24

    bond with water that is part of the solvation sphere of the

    cation.

    These mechanisms are much more important for sorbing

    organics than the classical view of clay-organic sorption,

    that held that sorption is due to hydrogen bonding between

    oxygen atoms and hydroxyl groups of the silicates. These

    weaker hydrogen bonds become important only for large

    organic molecules, which will form many hydrogen bonds, and

    when the cation exchange complex of the clay is saturated

    with cations of low solvation energy. In this latter case,

    organic molecules are in competion with water for sorption

    sites. Compounds that are not polar enough to hydrogen

    bond with clays in the presence of water may complex on the

    clay surface when water is removed, i.e. by drying in the

    air, or by alternate wetting and drying conditions in the

    field. Under these conditions, water content is a very

    important factor in determining whether or not clay-organic

    interactions take place.

    2.6.2 Clay-Organic Interactions

    Perhaps the most significant alteration in the

    sorbing properties of clays occurs when either organic

    compounds in the soil organic matter itself, or in solution

    sorb to the mineral surface. The mechanisms for this

    reaction have already been discussed. Ahlrich (1972) notes

    that clay crystals are often coated or mixed with decompos-

    ing organic matter. McBride et al. (1977) observed that

  • 25

    smectite clay, which first underwent exchange reactions

    with different alkylammonium ions, sorbed 12 times as much

    phenol as the untreated clay; benzene sorbed to an even

    greater extent than phenol. The mechanism of interaction

    was with the interlamellar oxygens in the silicate layers.

    1,2-dichlorobenzene and 1,2,4-trichlorobenzene were not

    sorbed by this system, presumably because of their large

    size. Larger alkylammonium ions may provide enough silicate

    layer separation to allow this mechanism to work on

    chlorobenzenes.

    While it has been demonstrated in the laboratory

    that mechanisms exist for sorption of organics to mineral

    surfaces, in natural environments, organics are more likely

    to partition onto organic components than onto clay

    surfaces, due to hydrophobic sorption, and the hydrophilic

    nature of the clay surface. Thus, the most important

    factor in the field determining sorption onto clays is the

    clay/organic matter ratio. Alternatively, we have seen

    that clay-organic complexes may sorb compounds that didn't

    sorb to pure clays. (Hasset and Means, 1980) studied

    sorption of 1-napthol, a 2 ring aromatic with an OH group,

    onto many different soils. Their findings indicated that

    sorption onto one group of soils produced a reasonable fit

    to a log Koc = alog Kow + b equation, while sorption onto

    another set deviated from this relationship. This second

    set had a much lower organic carbon/montmorillonite ratio

  • than the first set, indicating that for soils with low

    organic carbon contents, sorption of organics onto clays

    may be important.

    26

  • CHAPTER 3

    EXPERIMENTAL PROCEDURES AND MATERIALS

    The techniques employed for studying sorption

    reactions in this research were batch and column experi-

    ments.

    3.1 Experimental Approach

    In the batch method, a sorbent (soil) is suspended

    by agitation in a sorbate solution (1,4-DCB dissolved in

    water), until equilibrium is reached. The suspension is

    separated by centrifugation, and the solution analyzed for

    changes in concentration, presumably due to sorption of the

    sorbate (1,4-DCB) onto the sorbent. Initial concentrations

    are determined by blanks, tubes containing the sorbate

    solution without the sorbent, that undergo the same

    treatment as the soil/solution suspension. This method is

    used to determine sorption isotherms. By analyzing the

    aqueous solution prior to equilibration, this method can

    also be used to determine the rate of sorption.

    The column method involves a continuous flow of

    sorbate solution through a column packed with the sorbent.

    Effluent at the bottom of the column is collected and

    analyzed for sorbate concentration. When the sorbate

    27

  • concentration in the effluent reaches the initial concen-

    tration of the influent, the system is at equilibrium. At

    this point, either the influent solution is changed to

    contaminant-free water (to determine desorption characte-

    ristics of the system) or the experiment is terminated.

    Results of column experiments are commonly expressed in

    dimensionless variables C/Co vs. pore volumes. A pore

    volume is equal to the total pore space in the column and

    C/Co is equal to the exit concentration of the sorbate

    divided by the influent concentration.

    Because of the volatile, hydrophobic nature of 1,4-

    dichlorobenzene, care was taken at all points in the

    experiments to avoid headspace in any solution containing

    1,4-DCB, and thereby eliminate volatilization losses. For

    this work, it was found that use of screw-cap bottles with

    teflon-faced silica septa were most suitable for minimizing

    headspace.

    3.2 Soils Description

    Soils near the Motorola 52nd street site are derived

    from colluvium-alluvium and lie on Precambrian and Tertiary

    bedrock. The colluvium-alluvium ranges in thickness from a

    few centimeters to more than nine meters. In general, the

    upper materials are clayey sand with gravel, which are

    either very dense or cemented. Soil porosities range from

    23 percent to 32 percent. Results from pumping tests

    indicate that the average hydraulic conductivity of the

    28

  • colluvial-alluvial aquifer is 42 feet per day (1.5 x 10 -2

    cm per day), and average linear velocities in the aquifer

    were determined to be between 0.25 and 2.5 feet per day.

    For this research, all of the solid material used

    for experimentation will be referred to as soils. Soils

    were taken from drilling cores in both the saturated and

    unsaturated zones; the depths of the cores used ranged

    between 15-39 feet below the ground's surface. Soil

    DM104SG was two miles from the field site, and had very

    little clay, while soil DM102SG was from the courtyard of

    the Motorola plant, and had a larger amount of fine sand

    and clay. The soils were air dried for several days, and

    then oven dried at 105 °C for 24-48 hours. The soils were

    deaggregated and decemented by repeated hitting and

    grinding with a mortar and pestle. Finally, the soils were

    sieved using Soil Conservation Service soil sieves, ranging

    between a size 250 sieve (0.061 mm mesh size) and a size 5

    sieve (3.96 mm mesh size). A breakdown of the size

    fractions of the soils is given in Table 3. Fraction

    organic carbon was determined only for soil DM104SG, and

    was 0.00086.

    3.3 Equilibrium Batch Procedures

    Equilibrium batch experiments were conducted with

    soil and 1,4-dichlorobenzene to determine sorption iso-

    therms. The size fraction of the soil that was less than

    29

  • TABLE III

    Soil Size Fractions from various boreholes and depths

    SIZE FRACTIONS(mm)

    Percent in stated size range

    DM102SG 2V2DB 2V2DB DM104SG

    15-16 16.5-18.3 18.3-20.5 39feet feet feet feet

    X > 3.96 50.0 75.8 72.1 29.2

    3.96 > X >1.98 15.8 7.9 12.7 23.5

    1.98 > X >0.991 10.8 6.1 7.5 16.3

    0.991 > X >0.495 7.8 4.1 3.7 10.6

    0.495 > X >0.124 12.2 4.5 2.5 15.1

    X < 0.124 3.3 1.5 1.5 5.2

    TOTAL 99.9 99.9 100.0 99.9

    30

  • 0.124 mm (Table 3) from boreholes 2V2DB 16.5-18.3 feet,

    2V2DB 18.3-20.5 feet, and DM102SG 15-16 feet were used for

    the equilibrium batch experiments. A concentrated stock

    solution was prepared by dissolving a weighed amount of

    1,4-DCB in a 40 ml screw-capped bottle filled with a known

    weight of methanol, typically giving concentrations between

    500 and 2000 PPM. Use of methanol facilitates the dissolu-

    tion of 1,4-DCB in water. A known volume of the concen-

    trated methanol stock (20-300 pl) is then added to to 250

    ml of water in a screw-capped bottle. Both the volume of

    methanol and volume of water were multiplied by their

    respective densities (0.794 and 1.0 g/cm 3 ); this procedure

    gave concentrations between 0.1 and 2 ppm. This solution

    is added to a known mass of soil in 35 ml screw-capped

    centrifuge tubes and reweighed; the soil-to-water ratio

    (g soil/g water) was between 0.4 and 0.6. The soil and

    water are mixed for time periods ranging between 5 and 7

    days. The tubes are centrifuged for 20 minutes at 4000

    RPM. The liquid is then poured into 2 dram (7.4 ml) vials,

    filling approximately half the vial, and weighed. Pentane

    is added to fill the remaining volume, and the vial was

    reweighed, giving a water/pentane ratio of 1.65-2.5:1; the

    pentane is used to extract the organic compounds from the

    water. The 2 dram vials are shaken for 1 minute (150

    shakes), and then the pentane is analyzed in a gas chromat-

    ograph for the presence of 1,4-DCB.

    31

  • 3.4 Problems with Batch Procedures

    Volatilization losses occured in pouring the 1,4-

    DCB-water stock into the centrifuge tubes. By sampling the

    stock solution before and after the decanting, it was

    determined that losses were less than five percent. Major

    difficulties were encountered in trying to eliminate

    headspace within the centrifuge tubes. In order to ensure

    complete saturation of the soil, the tubes were filled

    approximately 90 percent full of water, and were shaken

    until all the soil in the tubes went into suspension. The

    caps were removed, and the tubes were then filled to the

    top with water. The maximum soil/water ratio compatible to

    good mixing was used. Since the caps of the sample tubes

    were removed after the 1,4-DCB was added (in order to fill

    the tubes with water and thereby eliminate any headspace),

    it is likely that some losses of 1,4-DCB occured in the

    sample tubes, both through displacement of the air in the

    sample tube and due to partitioning from the liquid to the

    air. However, no other method was developed that accom-

    plished the tasks of complete mixing of the soil with the

    solution, and elimination of the headspace in the centri-

    fuge tubes. Henry's law can be used to determine the

    partial pressure of 1,4-DCB from the headspace in the

    vials, and combining this with the ideal gas law, the

    amount of 1,4-DCB lost from the headspace can be determin-

    ed; this turned out to be 1.5 percent, using the highest

    32

  • 33

    concentration (2 PPM) used in the experiments. The

    calculations for this are given in appendix A.

    Because 1,4-DCB sorption was measured by looking at

    the disappearance of the compound from solution, losses

    during an experiment would result in an overestimation of

    total sorption. However, these calculations show that the

    total loss of 1.4-DCB during the mixing should not have an

    impact on sorption calculations. Additional complications

    may arise due to the presence of suspended soil particles

    in the soil water after centrifugation, possibly leading to

    an understatement of the extent of the sorption.

    3.5 Kinetic Batch Procedures

    The kinetic batch experiments were prepared in the

    same manner as the equilibrium batch experiments. The

    soil/water suspension was mixed for time periods ranging

    between 1 hour and 1 week, with centrifugation and sampling

    typically occurring after 1 hour, 1 day, and 1 week. 1,4-

    DCB/water solutions used ranged from 0.1 to 2.0 ppm. In

    addition to the soils mentioned above, soil from borehole

    DM104SG was used for the kinetic batch experiments.

    3.6 Desorption Batch Procedures

    An attempt was also made to run desorption batch

    experiments. The procedure was the same as for the

    equilibrium batch experiments. After one week the tubes

    were centrifuged, and the liquid solution was decanted.

  • The tubes were then filled with distilled, deionized water,

    remixed (through the use of a sonicator), and allowed to

    equilibrate for another week. The decanted solution was

    analyzed for the presence of 1,4-DCB, in order to determine

    the equilibrium level of sorption, as well as to aid in a

    mass balance accounting of the 1,4-DCB. After one week,

    the tubes were analyzed in the same manner as the equili-

    brium sorption experiments. Mass balance revealed that in

    almost all cases, the 1,4-DCB in solution at the end of

    desorption equilibrium exceeded the amount of 1,4-DCB which

    was calculated to be present in the entire tube at the

    beginning of the desorption experiment. These problems

    could not be resolved, and further work along these lines

    was discontinued.

    3.7 Soil Column Experimental Procedures

    Column experiments (see schematic, figure 1) were

    conducted at various concentrations of 1,4-dichlorobenzene,

    and at different flow rates. The concentrations ranged

    from 0.15 ppm to 1.5 ppm., and the flow rates were typi-

    cally 1 ml/min although they did dip as low as 0.5 ml/min.

    A concentrated solution of 1,4-DCB and methanol was

    mixed in a 40m1 vial; this solution was then added to a 20

    liter bottle, nearly full with distilled, deionized water.

    In addition, the 20 liter bottle had 30 grams of KC1 in it

    (a 2 x 10 -2 M solution), which acted as a conservative

    tracer. This bottle was then topped off with distilled,

    34

  • li fll

    STOGY\

    Sc L v.-rzoN

    SAMPLTAG-Loota

    C...OL.utflN

    sprrItumCrLOOP

    1 ,i,

    35

    C.OND M.-C1r_vi.T1DZTEC:roK

    Figure 1. Schematic representation of the soil columnexperimental setup

  • deionized water, and sealed with a rubber stopper that had

    been wrapped with teflon. The stopper had two 1/8" ID

    stainless steel tubes through it; one reaching near the

    bottom of the stock bottle, and the other only partially

    through the rubber stopper, acting as an air vent to admit

    air as water is pumped out. The stock solution was mixed

    for at least 24 hours with a stirrer and magnetic bar

    before running any of the experiments.

    The solution was pumped into the column using a

    peristaltic pump and 1/16" ID tygon tubing. A "t" fitting

    was at the head of the column, which allowed flow both into

    the soil column, and into a 2.3 ml sampling loop before the

    soil column. This allowed monitoring of the feed (in-

    fluent) solution before it entered the column. At the

    outlet of the column was a conductivity detector, which

    measured the breakthrough of the KC1. After the detector

    was a 2.3 ml sampling loop made of 1/8" ID stainless steel

    tubing. This sampling loop was used to collect samples for

    GC analysis. The soil column was a 30.5 cm long by 2.7 cm

    ID glass column, for a total volume of 174 cm 3 ; the cross-

    sectional flow area was 5.7 cm 2 . For flow rates between

    0.5 and 1.0 ml/min, this will give a Darcian flow velocity

    through the column of between 0.09 and 0.17 cm/min, or 4.1-

    8.2 feet per day. By contrast, the average linear velocity

    of groundwater in the shallow colluvium-alluvium underlying

    the field site was estimated to be between 0.25 - 2.5 feet

    36

  • 37

    per day. The column was packed with the fraction of soil

    from borehole DM104SG 39' that passed through a 0.991 mm

    soil sieve, but which was retained on a 0.246 mm soil

    sieve. The weight of the soil used in the packing was 167

    grams, for a bulk density of 0.96. 98 cm 3 of water was

    required to fill all the pore space in the column; thus,

    for these experiments, one pore volume was equal to 98 cm 3 .

    Total porosity was 0.56.

    3.8 Chemical Analysis

    All chemical analyses were conducted on a Varian

    3760 gas chromatograph, using an Ni63 electron capture

    detector. A Hewlett Packard integrator integrated the

    electronic detector signal. The column in the GC is a 6

    ft., 1/4 in. ID Supelco glass column, with a 3% 0V-17

    packing. The carrier gas was nitrogen, at a flowrate of 40

    ml/min. The column was run at 75 °C, the detector at 300

    °C and the injection port at 130 °C. A Hamilton 1710 10 ul

    syringe was used; all injections were 3 pl. Settings on the

    integrator were typically: chart speed = 0.5 cm/min, peak

    width = 0.04, threshold = 0 and attenuation = 2, although

    for lower concentrations of 1,4-DCB, a lower attenuation

    and threshold were used. Under these conditions, 1,4-DCB

    peaks occurred between 4.15 and 4.3 minutes, depending upon

    how long the packing had been in the column.

  • CHAPTER 4

    EXPERIMENTAL RESULTS

    4.1 Equilibrium Batch Results

    The results of the equilibrium batch experiments are

    shown in figures 2 and 3. The formulas used to calculate

    points along the sorption isotherm were as follows:

    C = Co x E (22)

    S = Co x (1-E)/r sw (23)

    where C = Equilibrium concentration of 1,4-DCB insolution (pg 1,4-DCB/cm 3 water)

    Co= Initial concentration of 1,4-DCB in solution(big 1,4-DCB/cm 3 water)

    E = Fraction of total 1,4-DCB remaining in solutionat equilibrium

    S = Equilibrium concentration of 1,4-DCB on thesoil (pg 1,4-DCB/g soil)r sw = Soil/water ratio in the centrifuge vials

    (g soil/g water)

    The partition coefficient for 1,4-DCB calculated from the

    sorption isotherm for the soils from borehole 2V20B is

    1.40. At the soil/water ratios used (0.4-0.6 g/cm 3 ) this

    was approximately 40 percent of the total 1,4-DCB added to

    the sample tubes. The partition coefficient for the soil

    from borehole 102SG is 3.33.

    It is apparent from Figures 2 and 3 that equilibrium

    partitioning for these experiments is linear, as was

    38

  • SORPTION ISOTHERM FOR 1,4—DCBSOIL 2V2DB

    Cb.0 0.5 1.0 1.5 2.0ug 1,4—DCB/cm3 water

    Figure 2. Sorption isotherm for 1,4-dichlorobenzene insoil 2V203

  • SORPTION ISOTHERM FOR 1,4—DCBSOIL 102SG

    ociho 0.5 1.0 1.5

    ug 1 ,4—DCB/cm3 water

    C71

    Figure 3. Sorption isotherm fOr 1,4-dichlorobenzene in

    soil 102SG

  • 41

    originally postulated. A Freundlich isotherm is given by:

    S = KC 1 / n (24)

    where 1/n and K are the parameters. When S and C are

    plotted on a log-log scale, the slope is equal to 1/n.

    This procedure was carried out for the data from Figure 2,

    and gave a value for 1/n of 1.05, further confirming that

    sorption of 1,4-DCB at these low concentrations is linear.

    f oc was not measured for the isotherm soil samples. The

    foc for the soil used in the column experiments (DM104SG)

    was 0.00086. Using this value, soil 2V2DB had K oc for 1,4-

    DCB of 1400 cm 3 water/g soil and soil DM102SG had K oc of

    3300 cm 3 water/g soil.

    4.2 Kinetic Batch Results

    Results from the kinetic batch experiments are shown

    in Figure 4. Sorption takes place fairly rapidly, with

    most of the sorption occuring within the first day.

    However, it does appear that there is a small amount of

    time-dependent (kinetic) sorption.

    Sorption was 20-50 percent complete on the samples

    after 1 hour, and almost 100 percent complete after 24

    hours. Thus, in some cases, over half of the total

    sorption occured within the first hour. However, a

    significant amount of sorption occured after the first

    hour, sorbing at a different rate. This indicates a

    kinetic sorption process is operating here.

  • 47

    SORPTION KINETICS OF1,4—DCB1.0

    0.8

    0 . 0 o 50 100 150

    D---SOIL FROM 102SG TIME (hrs.)

    A---SOIL FROM 2V2DB4,---60IL FROM 2V2DB0---SOIL FROM DM104, CONC. = 1.0 PPM----CONCENTRATION = 0.2 PPM

    CONCENTRATION = 2.0 PPM

    Figure 4. Results of kinetic batch experiments with 1,4-dichlorobenzene

  • Five of the seven kinetic batch experiments show an

    increase in 1,4-DCB in solution between times of 24 hours

    and 120 hours (the last two data points on the graphs).

    This increase ranges between 4 and 12 percent of Co for

    these five experiments. This is too high to be accounted

    for by experimental and analytical uncertainty, although

    this uncertainty could have contributed to the rise. The

    rise also cannot be attributed to volatilization losses of

    1,4-DCB during the soil saturation and experimental mixing

    part of the experiment, since volatilization losses would

    result in an overestimation of sorption (see chapter 3),

    not a decrease in sorption. An exception to this would be

    if the 24 hour samples experienced more volatilization

    losses than the 96 hour samples; in this case, the 24 hour

    samples would have overestimated sorption, and the later

    (96 hour) samples would indicate the actual levels of

    sorption. However, the rise in 1,4-DCB in solution for the

    96 hour samples is too systematic to attribute it to

    something as random as volatilization losses; apparently

    some other mechanism is operating here.

    Despite these problems with the kinetic data, Figure

    4 does indicate a small amount of time-dependent (kinetic)

    sorption. Except for the experiments with soil 102SG

    (which has a higher partition coefficient), the equilibrium

    levels of sorption in these experiments were 60-75 percent

    of Co. This is in good agreement with the equilibrium

    43

  • batch experiments, which gave equilibrium sorption levels

    of 60-67 percent of Co. This includes soil DM104, the soil

    used for the column experiments. This indicates that there

    may be a similarity between soil DM104 and soil 2V2DB, the

    soil used for the equilibrium batch experiments. Visual

    inspection of Figure 4 indicates the kinetic part of the

    sorption process beginning somewhere between 70-90 percent

    of Co.

    4.3 Soil Column Results

    The 1,4-DCB data from all five column experiments

    are shown in Figure 5, and the conductivity data are shown

    in Figure 6. At all the concentrations, the rising limbs

    of the breakthrough curves for 1,4-DCB were nearly identi-

    cal, indicating that sorption of 1,4-DCB's at these

    concentrations is not concentration dependent. Appendix B

    contains the breakthrough curves for both 1,4-DCB and

    conductivity from each experiment.

    4.3.1 Conductivity Data from Column Experiments

    Since KC1 is a conservative (nonreactive) tracer,

    the sorption term in the advection-dispersion equation

    drops out. Under these conditions, the exact analytical

    solution to equation 2 shows that the breakthrough of a

    conservative solute at the end of a column (i.e. the

    appearance of the solute in the effluent solution) will

    follow a normal distribution. Thus the time for the

    44

  • 45BREAKTHROUGH CURVES FOR

    —DCB1.2

    0 --EXP. 10 --EXP. 3A—EXP. 41

    0—EXP. 113 — EXP. 3A—EXP. 4Q—EXP. 54-EXP. 6

    0.0

    0.8 0—EXP. 541k—EXP. 6

    04

    40.4 21

    0.0

    Figure 5. Breakthrough curves for 1,4-DCB from columnexperiments 1,3,4,5,6

    BREAKTHROUGH CURVES FORCONDUCTIVITY1.2

    Figure 6. Breakthrough curves for KC1 from columnexperiments 1,3,4,5,6

  • effluent KC1 concentration to reach 50 percent of Co is the

    time when one pore volume (98 cm 3 ) has eluted from the...

    column. Compared to this value the rising limbs of the

    conductivity curves give C/Co of 0.5 at times ranging

    between 0.94 and 1.15 pore volumes, with an average of 1.05

    for the five column experiments. The same calculations

    were carried out on the descending limbs (desorption part)

    of the breakthrough curves; the number of pore volumes

    eluted ranged between 0.96 and 1.21, with an average of

    1.07 for the five experiments. Thus both the sorption and

    desorption parts of the conductivity curves give estimates

    of a pore volume close to the measured value of 98 cm 3 .

    4.3.2 1,4-DCB Data

    Soil-water partition coefficients for a sorbate can

    be obtained from column experiments by simple mass balance

    on breakthrough curves from the experiments. For a sorbate

    (1,4-DCB in this case), the area above its breakthrough

    curve when C/Co = 1.0 is equal to the total mass of sorbate

    contained in the column. For this analysis, pore volumes

    must be expressed in cm 3 and Co in ug sorbate/cm 3 water.

    When C/Co is equal to 1.0, the sorbed phase in the column

    is at equilibrium, and the solution phase concentration is

    equal to Co. Since the total void space is equal to one

    pore volume (98 cm 3 for these experiments), the total

    sorbate in aqueous solution is equal to one pore volume

    times Co, expressed as mass of sorbate. As noted in

    46

  • section 4.3.1, the conductivity of the effluent should be

    50 percent of Co at the 1 pore volume point; if this

    conductivity curve is symmetric, then the area above the

    conductivity curve is equal to one pore volume times Co,

    i.e. an area corresponding to the amount of sorbate in

    aqueous solution in the column at equilibrium. Thus the

    area between the breakthrough curves for conductivity and

    sorbate is equal to the total sorbate in the column minus

    the sorbate in solution, which is the total mass of sorbate

    retained on the soil phase (sorbed). Dividing this sorbed

    mass by the mass of soil in the column gives S, the sorbed

    concentration at equilibrium. Dividing this by Co, the the

    aqueous solution concentration at equilibrium (C) gives K p ,

    the equilibrium soil-water partition coefficient. The same

    analysis was done on the sorption and desorption fronts of

    the breakthrough curves.

    Areas above the breakthrough curves were determined

    by a Keuffel and Esser compensating polar planimeter, for

    both the sorption and desorption portions of the break-

    through curves. Soil-water partition coefficients cal-

    culated by this technique were between 0.38 and 0.48 for

    all five column experiments for the sorption part, and

    between 0.34 and 0.51 for the desorption part. The

    similarities between the sorption and desorption partition

    coefficients indicates that these sorption reactions are

    47

  • 48

    almost completely reversible. Table 4 gives a listing of

    some important parameters from the column experiments.

    4.4 Anomalies in Column Experimental Results

    One problem encountered in running the column

    experiments was that the concentration of the stock

    solution varied over time (see Appendix B). It is unlikely

    that the stock concentration varied directly in the feed

    bottle; since stock concentration sampling was carried out

    after the stock passed through the peristaltic pump, it is

    reasonble to assume that the change in concentration over

    time is due to either sorption onto the tygon tubing in

    the pump, or diffusion of 1,4-DCB through the tubing. The

    Co value used for calculating the dimensionless concentra-

    tion was the highest concentration of 1,4-DCB recorded by

    sampling the feed solution at the head of the column.

    Graphs using a Co value that was the average of the

    concentrations of 1,4-DCB recorded by sampling the feed

    solution at the head of the column did not differ signifi-

    cantly from the graphs in Appendix B.

    Because of the change in stock concentration over

    time, it was difficult to determine when the breakthrough

    curves for 1,4-DCB had reached equilibrium. This is the

    reason the declining limb of the breakthrough curves (the

    desorption portion) for the different experiments are not

    identical. Each experiment ran for a different amount of

    time once it reached "equilibrium", in an effort to

  • TABLE IV

    Summary of Parameters from Column Experiments

    Column Conc.Exp. (ppm)

    SORPTION

    FlowRate(ml/min)

    KD(cmsig)

    PORE VOLUMESwhen C/Co when influent

    = 0.50 pulse stops

    1 0.25 0.85 0.44 1.02 3.123 0.50 1.3 0.38 1.15 3.244 1.00 1.1 0.38 1.02 4.715 1.50 1.1 0.39 1.10 5.876 0.50 1.0 0.48 0.94 5.57

    AVERAGE 0.41 1.05

    DESORPTION1 0.25 0.70 0.34 1.213 0.50 0.81 0.44 1.144 1.00 0.79 0.51 0.965 1.50 0.85 0.51 0.966 0.50 0.55 0.34 1.06AVERAGE 0.43 1.07

    49

  • determine the actual point in time when equilibrium was

    reached. The actual desorption lines have the same slope

    and degree of tailing; they are simply offset from each

    other by the amount of time each experiment ran once it

    reached equilibrium. This is also true for the break-

    through curves for the conservative tracer, KC1. These

    curves are nearly identical, and the offset of the desorp-

    tion parts are again due to the length of time the system

    was allowed to remain at equilibrium. However, it does

    appear that experiments 1 and 3 may not have run for a long

    enough time.

    The jaggedness of the breakthrough curves near

    equilibrium creates another problem in analyzing the data.

    Since the mass balance approach to calculating partition

    coefficients requires integration of the area above the

    breakthrough curves, it was important to determine when

    these curves were at equilibrium, in order to know where to

    integrate. The graphs from experiments 3, 4 and 5 (see

    appendix B) have the conductivity and 1,4-DCB curves

    crossing each other near C/Co = 1.00, and the area between

    these curves was used for the integration. The Kp's gene-

    rated from these experiments were similar. The break-

    through curves for experiments 1 and 6 for 1,4-DCB never

    reached C/Co = 1.00, resulting in some uncertainty in the

    area integrated; this may have caused the variability in

    the partition coefficients calculated for the column

    50

  • experiments using the mass balance approach. The jagged-

    ness of the breakthrough curves near equilibrium also makes

    it difficult to visually determine if there are any kinetic

    effects involved in the sorption. While the conductivity

    and 1,4-DCB curves have nearly parallel rising and falling

    limbs (see appendix B), some tailing does occur at the top

    of the rising 1,4-DCB limb and the bottom of the descending

    1,4-DCB limb. The nearly parallel rising and falling limbs

    indicate that sorption is mostly an equilibirium phenome-

    non, and the slight tailing indicates there may be some

    kinetic effect involved in sorption of 1,4-DCB onto the

    column soil.

    51

  • CHAPTER 5

    DISCUSSION

    When equation 2 is coupled with equation 3, we get:

    DY-c-c (25)-cY -7( 2 - -5 19 3the one-dimensional advection-dispersion equation for

    equilibrium sorption. As explained in chapter I, the first

    term on the right hand side is a dispersion term, the

    second term is an advection term, and the third term is a

    sorption term. As before, C refers to the aqueous concen-

    tration of the sorbate, p is the soil bulk density and 0-is

    the total porosity of the soil in the column. The two

    parameters in this equation, D, the dispersion coefficient

    (cm 2 /sec) and Kp, the partition coefficient ( cm 3 /g), must

    be either measured or estimated before the equation can be

    used to simulate solute transport in groundwater. One of

    the reasons for generating sorption isotherms in the

    laboratory is to determine the partition coefficient, which

    can then be used with equation 25. Equation 25 is often

    written as:

    P, 3c_ D y)—(2, - v-b7c (26)where R, the retardation factor given by:

    R = 1 +(pK)/Q (27)

    52

  • 53

    For a linear sorption isotherm, a retardation factor is the

    ratio between the time it takes 50 percent of Co of the

    solute to pass through the exit point of the column', and

    the time it takes 50 percent of Co of the conservative

    tracer (1 pore volume) to pass through the exit point. The

    inverse of the retardation factor is often called the

    relative velocity, relating the velocity of the sorbate to

    the velocity of the conservative tracer.

    5.1 Curve-Fitting Analysis of Column Experiments

    van Genuchten and coworkers (1980, 1981, 1984) have

    developed a nonlinear least-squares curve-fitting procedure

    that can be used to estimate the different parameters in

    the analytical solution to the one-dimensional advection-

    dispersion equation directly from observed effluent data

    from column experiments. In addition to the equilibrium

    sorption case (van Genuchten, 1980), newer versions of

    their computer programs handle kinetic sorption (van

    Genuchten, 1981) and sorption experiments conducted in the

    field (Parker and van Genuchten, 1984). Analysis of the

    present research was done with the 1981 program, called

    CFITIM.

    5.1.1 Models and Input Parameters in CF1TIM

    CFITIM works with analytical solutions to equation

    26, for five different conceptual models of sorption

    behaviour. A different set of dimensionless variables for

  • 54

    each model is introduced into equation 26, and the trans-

    formed equation is solved for the exit concentration of the

    pollutant as a function of time. These fitted values are

    compared with experimentally derived data, and parameters

    in the analytical solution are adjusted for a best fit of

    the observed data. A list of the dimensionless variables

    is given in Appendix C.

    The 5 models are labeled A-E. Model A is an

    equilibrium model, model B is a physical nonequilibrium

    diffusion model, model C is a physical nonequilibrium with

    anion exclusion model, model D is a nonequilibrium two-site

    chemical kinetics model, and model E is a nonequilibirum

    one-site chemical kinetic model. Models B and D were based

    on the two models already discussed in sections 2.4.1 and

    2.4.2. Once the dimensionless variables are introduced

    into equation 26 the analytical solutions to models B,C,D

    and E mathematically reduce to the same equation. This

    equation has five dimensionless parameters; R is the

    retardation factor, P is the Peclet number, Beta (a) is a

    parameter related to the ratio of the different types of

    sorption sites (kinetic vs. equilibrium, mobile vs.

    immobile), Omega is a parameter related to the rate

    constant for the kinetic portion of sorption and Pulse is

    the length of time (in pore volumes) the influent solution

    flowed into the column. Even though the analytical

    solution for these four models is the same, the dimension-

  • 55

    less variables a and Omega take on different meanings in

    the different models. The equilibrium model (A) solution

    requires just three parameters as inputs; the retardation

    factor, the Peclet number, and the pulse. The one-site

    kinetic model only requires four parameters as inputs, as a

    is always set equal to the inverse of the retardation

    factor (see Appendix C). For any one of the models, only

    one of the parameters must be known; the others are

    estimated using the least-squares curve-fitting technique.

    5.1.2 Boundary Conditions in CFITIM

    Two different column exit boundary conditions are

    available in CFITIM. The first was used, which is that the

    change in concentration (concentration gradient) of the

    pollutant at any time at an infinite distance from the

    inlet is zero; this assumes the presence of a semi-infinite

    soil column. The second one (not used) is that the

    concentration gradient at the lower end of the column is

    zero. While the second condition may intuitively seem to

    be the correct approach, this leads to a condition of a

    continuous concentration distribution at x = L, the lower

    boundary. The upper boundary condition leads to a situation

    where there is a discontinuous concentration distribution

    at the inlet position, which would contradict the physical

    model presented by the second lower boundary condition.

  • 56

    5.1.3 Column Experimental Analysis by CFITIMEquilibrium Model

    As discussed in section 5.1.1, CFITIM requires only

    one of five (one of three in the equilibrium model)

    parameters as a known input. In addition to the known

    variable, CFITIM also requires estimates of the remaining

    parameters as inputs. The only other data CFITIM requires

    is the observed effluent concentration data, given as

    dimensionless values in the form of C/Co vs. pore volumes.

    5.1.3.1 CFITIM Equilibrium Model Graphs. Figure 7

    contains conductivity breakthrough curves generated with

    the equilibrium model of CFITIM with pulse input as the

    known variable; both observed data and fitted points are

    plotted on each graph. Looking at Figure 7, we see again

    that the experiments are internally consistent. Except for

    the dip in conductivity concentration that occurs in each

    experiment at the same time the system starts to reach

    conductivity equilibrium, the model provides a good fit to

    the observed conductivity data.

    The 1,4-dichlorobenzene breakthrough curves gene-

    rated under the same conditions are given in Figure 8; they

    show the observed data tailing off from the fitted data.

    The tailing is indicative of a kinetic effect; this

    observation is in agreement with the qualitative observa-

    tions of the kinetic batch experiments (Figure 5, section

    4.2). Visual inspection of the graphs indicates that with

  • 1 and fitted conduc-xperiments 1,3,4,5,6model. Only the

    Figure 7. Comparison of experimentativity data from column eusing CFITIM equilibriumobserved data is plotted.

    1.0

    0.8

    0.6o

    0.4

    0.2 -

    0.0 1-r- r LT 111- 1-1 I 1- 1-r1 1 -1 • -14 8PORE VOLUMES

    12

    0.2

    -1 r r-r- r- r 1 Tr r" rrl0.0 {Minh r 1 -1 r-r r-r r .:111111f if

    0 4 8 12PORE VOLUMES

    1.0 -

    0.8 -

    0.6 -o0 -...

    0.4 -

    0.2 -

    0.0

    0.0 ri r-r-r -r-r-r r-r- r-r tn-1-1

    4 8 12PORE VOLUMES

    1.0

    0.8

    0.50o

    0.4

    0.2

    0.012

    40 —EXP. 10—EXP. 3A—EXP. 4O—EXP . 5*—EXP. 6

    57

  • 0.8

    58

    0.60 7 00 0

    0.4

    0.2

    0.0 -r ri j I T-1' j 7 1-1-1-14 8 12PORE VOLUMES

    Lc:

    r—m-st

    1.0 -1

    0

    a- r

    0.8

    0.6 -_(-4 -

    C.4 -

    (?

    -

    '0.2 7

    0 4 8 12PORE VOLUMES

    40

    1.0 -1

    C.2 - ,-1-. (

    **L. ] *

    Co*

    r

    I_

    0.8 - 12 **1

    O• 0 r- .$1--rl-r-r-l-r, f T -r-r-, 7 7-1 IT 111g n f4 8 12 --.1a 1***

    0.6 ikt *I

    PORE VOLUMES-1

    "....o

    1.0 7 4(....)

    3(----Y_I--, (...)! p

    0.8 14'-1

    3 "i

    0.6 3o

    C.-3 -.'--- -(..3

    O4- A

    0.4 -11

    :40.2

    -n\

    0.0 471** 'i *

    0 4 8 1::FORE: VOL UMES

    0—EXP. 1D— EXP. 3L.—EXP. 4.0—EXP. 54.-EXP. 6

    0.2-

    0.00.0 4fil41.1, I 1- 1 Ij rI12 4 8

    POP.L \'01.L1M LS

    Figure 8. Comparison of experimental and fitted 1,4-dichlorobenzene data from column experiments1,3,4,5,6 using the CFITIM equilibrium model.Only the observed data is plotted.

  • 59

    the exception of experiment 6, the kinetic effect seems to

    occur after sorption is 80-90 percent complete. This is

    the point where the rising limb of the breakthrough curves

    for the observed 1,4-DCB concentrations begins to tail off

    from the breakthrough curves for the fitted 1,4-DCB

    concentrations, and is roughly the same point at which the

    kinetic batch experiments began to show kinetic sorption,

    as discussed in section 4.2. Tailing in the graph from

    experiment 6 occurs much earlier; inspection of the graph

    shows that this is probably due to the 1,4-DCB concentrat-

    ion never reaching the Co level, perhaps because of

    sorption of the influent solution by the tygon tubing.

    Despite this, the same kinetic effect is in evidence,

    before the concentration of 1,4-DCB in the effluent reaches

    an equilibrium level.

    5.1.3.2 CFIT1M Equilibrium Model Parameter Estimates.

    Estimated parameters for the equilibrium model runs are

    shown in Table 5. As noted in section 5.1.1, this model

    requires only estimates of the retardation factor and the

    Peclet number as inputs, if the influent pulse is known.

    The equilibrium model was run under several different

    conditions. First, both 1,4-DCB and conductivity data were

    run with only the pulse input as a known variable (case

    numbers 1 and 3, respectively, in Table 5). The conduc-

    tivity data were also run with the retardation factor (R)

  • TABLE V

    DATA FROM EQUILIBRIUM MODEL

    CASE

    COLUMN EXPERIMENT NUMBER 1

    NAME VALUE S.E.COEFF.

    1 Conductivity data,pulse known

    PECLETR

    21.5321.132

    9.0920.040

    2 Conductivity data R=1.0 PECLET 8.784 2.7233 DCB, pulse known PECLET 14.685 2.678

    1.796 0.0404 DCB, with Peclet from 1 R 1.817 0.0365 DCB, with Peclet from 2 R 1.757 0.048

    COLUMN EXPERIMENT NUMBER 31 Conductivity,

    pulse knownPECLET

    R191.921

    1.14928.3820.005

    2 Conductivity R=1.0 PECLET 25.193 14.1583 DCB, pulse known PECLET 157.203 2.300

    R 1.801 0.0254 DCB, with Peclet from 1 R 1.793 0.0045 DCB, with Peclet from 2 R 1.616 0.000

    COLUMN EXPERIMENT NUMBER 41 Conductivity data,

    pulse knownPECLET

    R149.025

    1.05116.4900.004

    2 Conductivity R=1.0 PECLET 122.944 27.9783 DCB, pulse known PECLET 157.207 2.713

    1.799 0.0304 DCB, with Peclet from 1 R 1.548 0.0035 DCB, with Peclet from 2 R 1.502 0.001

    COLUMN EXPERIMENT NUMBER 51 Conductivity data PECLET 69.629 17.504

    pulse known R 1.030 0.0152 Conductivity R=1.0 PECLET 71.743 19.1143 DCB, pulse known PECLET 58.802 0.657

    R 1.769 0.0194 DCB, with Peclet from 1 R 1.506 0.0025 DCB, with Peclet from 2 R 1.500 0.001

    COLUMN EXPERIMENT NUMBER 61 Conductivity data,

    pulse knownPECLET

    R96.0421.102

    26.1750.014

    2 Conductivity R=1.0 PECLET 42.899 15.7763 DCB, pulse known PECLET 59.713 1.047

    R 1.746 0.0304 DCB, with Peclet from 1 R 1.503 0.0025 DCB, with Peclet from 2 R 1.500 0.000

    60

  • 61

    input as a known variable (case 2). R was assigned a value

    of 1.0, the theoretical value for KC1 in the absence of ion

    exclusion effects. The 1,4-DCB data were then run twice,

    with the Peclet number input as a known variable. The

    first run (case 4) used the value for the Peclet number

    generated from the conductivity data with only the pulse

    input as a known variable (case 1). The second run (case 5)

    used the value for the Peclet number generated from the

    conductivity data when that data had R input with a value

    of 1.0 (case 3). The table gives the estimates of the

    different parameters generated by the equilibrium model,

    and the standard error coefficient involved in the estima-

    te.

    With pulse as the only known input variable, 1,4-DCB

    data from all five experiments gave a retardation factor

    between 1.77-1.80. Converting the average value to Kp

    using equation 26 with p=0.96 and 0=0.56 yields a partition

    coefficient of 0.46. This agrees well with the sorption

    partition coefficients generated by mass balance on the

    breakthrough curves (see Table 4); these coefficients were

    between 0.38 and 0.48.

    The Peclet number is one of the dimensionless

    variables introduced into equation 26; its conversion into

    the dispersion coefficient is given by:

    D = vL/P (28)

  • 62

    TABLE VIo

    Dispersion coefficients generated by CFITIM equilibriummodel, ordered by velocity

    COLUMNEXPERIMENT

    D for KC1data

    (cm 2 /sec)

    D for 1,4-DCBdata(cm 2 /sec)

    Velocity(cm/sec)

    1 3.39x10-3 4.98x10-3 2.43x10-3

    6 0.92x10-3 1.47x10-3 2.86x10-3

    4 6.48x10-4 6.15x10-4 3.14x10-3

    5 1.37x10-3 1.63x10-3 3.14x10-3

    3 5.95x10-4 7.20x10-4 3.72x10-3

    AVERAGE 1.38x10-3 1.88x10-3 3.06x10-3

  • 63

    where D is the dispersion coefficient (cm 2 isec), v is the

    Darcian velocity (cm/sec), L is the column length (cm) and

    P is the Peclet number. Darcian velocities for the column

    experiments can be obtained by dividing the flow rate by

    5.7 cm 2 , the cross-sectional , area of the column. Veloci-

    ties and dispersion coefficients for each column experiment

    are given in Table 6. For each column experiment, disper-

    sion coefficients generated with the 1,4-DCB data and those

    generated with the conductivity data are similar. The

    dispersion coefficients for the column experiments vary

    over an order of magnitude, between 6.0 x 10 -4 and 5.0 x

    10 -3 cm 2 /sec.

    The dispersion coefficient is usually represented by

    D =vxa (29)

    where a is the dispersivity (cm). Comparing equations 28

    and 29 we see the dispersivity is equal to the column

    length divided by the Peclet number. The dispersion

    coefficient is linearly related to velocity, while the

    dispersivity is a property more intrinsic to a flow system,

    conceptually thought of as a characterisitic mixing length.

    Thus dispersivity is more often reported in the literature;

    Anderson (1979) gives typical values of the dispersivity in

    column experiments of 0.1-1 cm; all the values for the

    CFITIM runs fall within this range.

    Inputting different parameters as known variables

    during the equilibrium model runs had only a small effect

  • 64

    on the model results. Inputting R = 1.0 for the conductiv-

    ity data (case 2) resulted in lower estimates for the

    dispersion coefficients, and a higher ratio of standard

    error to parameter value. With the exception of experiment

    3, the change in dispersion coefficient was not more than a

    factor of 3. Inputting the Peclet number as a known

    variable for the 1,4-DCB data also had only a small effect.

    With the exception of experiment 1, the major effect was to

    significantly reduce the standard error of estimation of

    the retardation factor. However, for each column experi-

    ment, the R calculated for all three cases involving 1,4-

    DCB data (2,4 and 5) varied at most 17 percent. The

    largest change in Kp under these conditions was for column

    experiment 4, cases 3 and 5, which had Kp of 0.46 and 0.29,

    respectively.

    5.1.4 Analysis by Non-equilibrium Models

    The data were also analyzed using the nonequilibrium

    models in CFITIM. Model E, the one-site chemical kinetic

    model, assumes all sorption sites are equal, with the rate

    of sorption onto the sites controlled by a first-order

    chemical process. The first-order kinetic rate equation

    for this model is:

    a(kC - S) (30)

    where C and S have already been defined. Comparison of

    equation 30 with equation 5 (section 2.4) reveals that k in

    equation 30 is equal to the sorption (forward) rate

  • 65

    constant divided by the desorption (reverse) rate constant

    times the inverse of the soil/water ratio (rs w ):

    k = (kf/kb)/rsw (30)

    k has dimensions of cm 3 water/g soil, and by definition,

    is equal to the soil-water partition coefficient, K. a

    then, is equal to the desorption rate constant, kb (sec -1 )

    in equation 5. rs w is equal to the soil bulk density

    divided by the porosity; for the column experiments this

    was equal to 1.71.

    Model E was used with the Peclet number from the

    first conductivity data analysis (case 1 from Table 5)

    input as a known variable. The partition coefficients

    generated by this model varied between 0.4 and 0.55, with

    an average of 0.46. This is the same average obtained from

    the equilibrium model, indicating that the kinetic effect

    on sorption is small.

    Pulse values obtained from this analysis deviated

    slightly from the experiment