PB91-231514
MODELING  MULTIPHASE ORGANIC CHEMICAL TRANSPORT
IN SOILS  AND GROUND WATER
Virginia  Polytechnic Institute and State  University
Blackburg,  VA
Aug 91
                  U.S. DEPARTMENT OF COMMERCE
               National Technical Information Service

-------
                                                    EPA/600/2-91/042
                                                    August 1991
 MODELING MULTIPHASE ORGANIC CHEMICAL TRANSPORT
              IN SOILS AND GROUND WATER
                           by

     J. C. Parker, A. K. Katyal, J. J. Kaluarachchi, R. J. Lenhard,
        T. J. Johnson, K. Jayaraman, K. Unlii and J. L. Zhu

     Center for Environmental and Hazardous Materials Studies
         Virginia Polytechnic Institute and State University
               Blacksburg, Virginia 24061-0404
                     Project CR-814320
                      Project Officers

                         J. S. Cho
            Processes and Systems Research Division
        Robert S. Kerr Environmental Research Laboratory
                   Ada, Oklahoma 74820

                       L. G. Swaby
              Office of Research and Development
             U. S. Environmental Protection Agency
                   Washington, D.C. 20460
ROBERT S. KERR ENVIRONMENTAL RESEARCH LABORATORY
        OFFICE OF RESEARCH AND DEVELOPMENT
       U. S. ENVIRONMENTAL PROTECTION AGENCY
                 ADA, OKLAHOMA 74820

-------
                                       TECHNICAL REPORT DATA
                               (Please read instructions on the reverse before completr
1. REPORT NO.
  EPA/600/2-91/042
                                 2.
                          3.
                                                                        PB91-23151
4. TITLE AND SUBTITLE
MODELING MULTIPHASE ORGANIC CHEMICAL TRANSPORT IN SOILS
AND GROUND WATER
                          5. REPORT DATE
                              August 1991
                                                                  6. PERFORMING ORGANIZATION CODE
7VAUTHOR(S)
J.C.  Parker,  A.K. Katyal,  J.J. Kaluarachchi,  R.J. Lenhard,
T.J.  Johnson,  K. Jayaraman, K. Unlu and J.L.  Zhu
                          8. PERFORMING ORGANIZATION REPORT NO.
9. PERFORMING ORGANIZATION NAME ANO ADDRESS
Center for  Environmental  and Hazardous Materials Studies
Virginia Polytechnic Institute and State University
Blacksburg, Virginia   24061-0404
                          10. PROGRAM ELEMENT NO.

                            CBWD1A
                          11. CONTRACT/GRANT NO.

                            CR-814320
12. SPONSORING AGENCY NAME AND ADDRESS
Robert S. Kerr  Environmental Research Laboratory
U.S.  Environmental Protection Agency
P.O.  Box 1198
Ada,  OK  74820
                          13. TYPE OF REPORT AND PERIOD COVERED
                            Project Report	
                          14. SPONSORING AGENCY CODE
                           EPA/600/15
15. SUPPLEMENTARY NOTES
PROJECT OFFICER:   Jong Soo Cho
FTS:  743-2353
16. ABSTRACT
            Subsurface contamination due to immiscible organic liquids is a widespread problem
     which poses a serious threat to ground-water resources. In order to understand the movement
     of such materials in the subsurface, a mathematical model was developed for multiphase flow
     and multicomponent transport in porous media with water, NAPL and air or any subset of
     these phases. Numerical procedures for solving the system of coupled flow equations, based
     on various formulations of the governing equations, were compared. Accurate representation
     of three-phase permeability-saturation-capillary pressure (£-S-P) relations is crucial to model
     multiphase fluid movement and accurate models for interphase mass partitioning are critical to
     describe species transport. A detailed description of hysteresis in three-phase k-S-P relations
     was reported. Simplified models, which consider effects of nonwetting fluid entrapment, were
     shown to provide a reasonable compromise between accuracy, on the one hand, and efficiency
     and robustness, on the other. Laboratory studies of light and dense NAPLs in a 1 x 1.5 meter
     sand tank, involving measurements of water and NAPL pressures and saturations and
     component concentrations, are described. These studies were used to validate  the
     mathematical model for multiphase flow and transport.
 7.
                                   KEY WORDS AND DOCUMENT ANALYSIS
                    DESCRIPTORS
                                                   b.lDENTIFIERS/OPEN ENDED TERMS
                                            COSATi field,Group
18. DISTRIBUTION STATEMENT

 RELEASE TO  PUBLIC
            19. SECURITY CLASS (This Report I
              UNCLASSIFIED
                                                                                 21. NO. OF "AGES
                                                   20. SECURITY CLASS 
-------
                                 DISCLAIMER

 The results reported in this document have been funded wholly or in part by the United
 States Environmental Protection Agency under Cooperative Agreement No. CR-814320
 to Virginia Polytechnic Institute and  State University.   It has been subjected to the
 Agency's peer and adminstrative review, and it has been approved for publication as an
 EPA document.  Mention of trade names  or commercial products does not constitute
 endorsement or recommendation for use.
                      QUALITY ASSURANCE STATEMENT

All research projects making conclusions or recommendations based on environmentally
related measurements and funded by the Environmental Protection Agency are required
to participate in the Agency Quality Assurance Program.  This project was conducted
under an approved  Quality Assurance Project  Plan.   Information  on  the  plan  and
documentation  of the  quality  assurance activities and results are available  from the
Principal Investigator.
                            ACKNOWLEDGEMENTS

Jong  Soo  Cho  with  the  U.  S.  Environmental  Protection  Agency,  R.  S.  Kerr
Environmental Research  Laboratory in Ada, Oklahoma and  Lou G. Swaby with the
U.S. EPA Office of Research and Development  in Washington, D.C., served as project
officers  whose technical and  administrative assistance throughout  the course of the
study were much appreciated.

During the second half of the project, Bob Lenhard left VPI to work for Battelle Pacific
Northwest Laboratories, where he continued to  contribute to  the project with support
from the Ecological Research  Division, Office of Health and Environmental Research,
U.S. DOE under contract DE-AC06-76RLO as  part of OHER's Subsurface Transport
Program.
                                        11

-------
                                   FOREWORD

EPA is charged by Congress to protect the Nation's land, air and water systems. Under
a mandate of national environmental laws focused on air and water quality, solid waste
management and  the control of toxic substances,  pesticides, noise and radiation, the
Agency strives to  formulate and implement actions which lead to a compatible balance
between human activities and the ability of natural systems to support and nurture life.

The Robert  S. Kerr  Environmental Research  Laboratory  is  the  Agency's center of
expertise for investigation of the  soil and subsurface environment.  Personnel at the
Laboratory are responsible for management of research programs to: (a) determine the
fate, transport and transformation rates of pollutants in the soil, the unsaturated  and
the saturated zones of the subsurface environment; (b) define the processes to be used in
characterizing  the soil and  subsurface environment as a receptor of pollutants; (c)
develop techniques for predicting the effect of pollutants on  ground  water, soil,  and
indigenous  organisms; and (d) define and demonstrate the applicability and limitations
of using natural processes indigenous to the soil and subsurface environment, for the
protection of this resource.

This report describes  the development of a  computer model for multiphase flow  and
transport in  the subsurface environment  and laboratory validation of the model. The
research assesses the  behavior of immiscible phase flow and contaminant transport in
each phase.
                                        Clinton W. Hall
                                        Director
                                        Robert S. Kerr Environmental
                                           Research Laboratory
                                          in

-------
                                    ABSTRACT

 Groundwater contamination due to surface spills or subsurface leakage of hydrocarbon
 fuels,  organic solvents and other immiscible organic  liquids is a  widespread problem
 which poses  a serious threat to groundwater resources. In order to model the movement
 of such materials in the subsurface, it is necessary, in general, to consider flow of water,
 nonaqueous  phase  liquid  (NAPL)  and  air,  and  transport  of  individual chemical
 components,  which  may  move  by  convection and  dispersion  in each  phase.  A
 mathematical model is developed for multiphase flow and multicomponent  transport in
 porous media with water, NAPL  and  air or any subset of these phases. Numerical
 procedures  for  solving the  system  of coupled flow  equations, based  on  various
 formulations  of the  governing equations,  were  compared.  An adaptive solution
 procedure  is  developed  that  automatically  eliminates  or  includes  equations  in
 subdomain regions, to take advantage of the fact that  flow may be near  steady state in
 large parts of a  physical domain. A linear phase-summed component transport equation
 is derived for the case of local equilibrium interphase mass transfer. A semi-decoupled
 solution approach  is  employed with  transport equations  solved serially with  flow
 equations using time-lagged interphase mass transfer terms in the flow  equations. For
 the case of rate-limited interphase mass transfer, a phase-summed transport equation is
 derived in terms of apparent partition coefficients, which are functions of the  current
 concentrations and interphase mass transfer rates, hence requiring  an iterative solution
 procedure.

 Accurate  representation   of   three-phase  permeability-saturation-pressure  (k-S-P)
 relations is  crucial  to model  multiphase fluid movement  and  accurate  models for
 interphase mass partitioning are  critical to  describe  species transport.  A detailed
 physically-based model for hysteresis  in  three-phase  k-S-P  relations  is described.
 Simplified models, which consider effects of nonwetting fluid entrapment, are shown to
 provide a reasonable compromise between accuracy, on the one hand, and efficiency and
 robustness, on the  other.  Laboratory studies of light  and dense NAPLs in  a 1 x 1.5
 meter  sand tank, involving measurements of water and NAPL pressures and saturations
 and  component  concentrations  are  described  and  were  used  to   validate  the
mathematical model for multiphase flow and transport. Column experiments performed
to study NAPL-water  interphase mass transfer kinetics indicate that existing empirical
models for mass transfer coefficients are reasonably accurate.
                                          IV

-------
                                 CONTENTS

  DISCLAIMER  	  ii

  QUALITY ASSURANCE   	  ii

  ACKNOWLEDGEMENTS	  ii

  FOREWORD	  iii

  ABSTRACT  	  iv

  CONTENTS  	  v

1. Summary of project objectives and results	  1

   1.1 Project overview  	  1
   1.2 Conclusions and recommendations   	  2


2. Numerical analysis of multiphase flow   	  4

   2.1 Introduction  	  4

   2.2 Development of implicit pressure-pressure formulation	  6
      2.2.1 Theory and mathematical formulation  	  6
      2.2.2 Numerical investigations	  17

   2.3 Evaluation  of alternative equation formulations  	30
      2.3.1 Formulation of governing equations  	30
      2.3.2 Solution procedure   	34
      2.3.3 Numerical results   .	39

   2.4 Development of an adaptive solution method  	  47
      2.4.1 Description of ASD algorithm	  49
      2.4.2 Numerical results  	  50

-------
3. Hysteresis in three-phase flow relations	   57

    3.1  Evaluation of general hysteretic formulation  	   57
         3.1.1 Model description  	   57
         3.1.2  Numerical simulation of hysteretic flow  	   69

    3.2  Development and testing of a simplified hysteretic model	   80
         3.2.1  Model description	   81
         3.2.2  Example problems  	   86

    3.3  Experimental verification of hysteretic flow model  	   93
         3.3.1  Static three-phase measurements  	   93
         3.3.2  Two-phase dynamic measurements  	  106
         3.3.3  Three-phase dynamic measurements  	  121
4.  Equilibrium-controlled multiphase transport  	   126

     4.1  Mathematical model for multiphase transport  	   126
          4.1.1 Governing equations  	   126
          4.1.2 Phase-summed equations for local equilibrium transport	   130
          4.1.3 Initial and boundary conditions	   132

     4.2  Numerical model description  	   134
          4.2.1 General solution approach  	   134
          4.2.2 Finite element formulation  	   136
          4.2.3 Interphase mass  transfer and density updating	   138

      4.3 Model applications   	   140
          4.3.1 Example 1  	   140
          4.3.2 Example 2  	   145
          4.3.3 Example 3  	   148
                                          VI

-------
 5.  Transport with nonequilibrium interphase mass transfer	  152

      5.1  Model for nonequilibrium interphase mass transport	  152
           5.1.1 Governing phase transport equations	  152
           5.1.2 Consideration of nonequilibrium interphase mass transfer. . .  .  152
           5.1.3 Solution approach	  156
           5.1.4 Numerical verification	  157

      5.2  Laboratory investigations	  159
           5.2.1 Experimental methods	  159
           5.2.2 Data analysis methods	  163
           5.2.3 Experimental results  	  168
6. Two dimensional laboratory studies of multiphase flow and transport  ....   171

      6.1  Experimental setup	   171
           6.1.1 Cell design	   171
           6.1.2 Gamma attenuation measurements	   171
           6.1.3 Pressure measurements  	   173
           6.1.4 Sample collection and analysis	   174

      6.2  LNAPL experiment	   174
           6.2.1 Experimental methods  	   174
           6.2.2 Model calibration and numerical analysis  	   177
           6.2.3 Results  	  180

      6.3  DNAPL experiment   	   193
           6.3.1 Experimental methods  	   193
           6.3.2 Model calibration and numerical analysis  	   195
           6.3.3 Results  	   197
7.  References	'	   202
Appendix: Publications resulting from this project	   206

                                          vii

-------
             1. SUMMARY OF PROJECT OBJECTIVES AND RESULTS

 1.1 Project Overview

 Spills and subsurface leaks  of immiscible  organic  liquids are  a  frequent  cause of
 groundwater contamination. This project was undertaken to improve our understanding
 of  contaminant  migration  arising  from such  events and  to develop  improved
 quantitative tools for their description. Attention was focussed on three fronts involving:
 (1) the development of efficient and robust numerical methods for solving the difficult
 problem of simulating simultaneous multiphase flow  and transport, and (2) developing
 and numerically implementing  theoretical  and empirical constitutive models governing
 multiphase  flow  and  transport  processes,  and  (3)  performing  laboratory  scale
 experimental studies  to  validate the mathematical  models developed in conjunction
 with the first two objectives.

 Chapter 2 of this report describes the development and testing of numerical methods for
 solving multiphase flow problems.  The  basic governing  equations for three-phase flow
 are presented and various numerical formulations are derived and compared vis  a vis
 their efficiency,  accuracy and robustness.  A  new algorithm is presented and discussed
 which enables substantial reductions in  computational effort by automatic elimination
 and inclusion of elements in the global solution.

 Chapter 3 presents results of numerical and experimental studies undertaken.to develop
 and  test  constitutive models  for  permeability-saturation-capillary  pressure  (k-S-P)
 relations  governing three-phase flow.  A  rigorous,  physically-based hysteretic  k-S-P
 model is described as well as various simplified models.  Numerical comparisons of the
 models are presented and results of static  and dynamic laboratory column experiments
 are compared with the models to evaluate their accuracy.

 Chapter 4  describes  the  theoretical foundation for modeling coupled multiphase flow
 and  multispecies transport with equilibrium interphase mass  transfer,  leading  to  a
 system of flow  equations for each bulk  phase  and phase-summed  species transport
equations for each species. A  numerical formulation for solving the system of equations
is presented based on partial decoupling  of flow and transport equations. Results of
several hypothetical numerical  simulations are  presented to verify  the model and to
demonstrate its applicability to specific problems.

-------
 In Chapter 5,  the  assumption of local equilibrium-controlled  mass transfer is relaxed
 and a formulation is derived which enables the form of the phase-summed equilibrium
 model to be retained by introducing  apparent partition coefficients that depend on the
 mass  transfer  rates  and first-order  mass  transfer coefficients,  as  well as  the  true
 equilibrium partition coefficients. Hypothetical  numerical  simulations are presented to
 verify the  formulation and  to demonstrate effects of nonequilibrium mass  transfer.
 Laboratory experimental studies  are presented which indicate  interphase mass transfer
 can be  described accurately  at  the laboratory scale  by a  first-order  mass  transfer
 relation and that mass transfer coefficients may be estimated from empirical  relations
 previously reported in the literature.

 Chapter 6 presents results of two-dimensional experiments involving simulated spills of
 light and  dense organic liquids.  The design of the  "sand tanks" and procedures for
 measuring fluid saturations and  pressures and component concentrations is described.
 Experimental results are compared with numerical simulations to provide validation of
 the coupled flow and transport model  described in Chapter  4.
1.2  Conclusions and Recommendations

In this report,  we present a mathematical model for multiphase flow and component
transport which is based on fundamental principles. Solution of the resulting set  of
nonlinear coupled differential  equations  is  computationally difficult. The numerical
results have indicated that the accuracy and robustness of multiphase flow problems is
quite sensitive to the method of formulating the numerical model. In particular, it has
been shown that chaining of saturation time derivative terms can lead to mass balance
errors and  other  difficulties that  are greatly reduced  by means  of  an alternative
formulation  that treats  saturation time  derivatives directly as  implicit  functions  of
pressure.  Marked  reductions in  computational effort were achieved by  an  adaptive
solution procedure that  takes advantage of the fact that flow may be near steady state
in large  parts  of the  physical  domain.  Implementation of schemes to  selectively
eliminate individual phase  equations should prove  even more  effective in  reducing
computational effort as well as improving solution robustness.  Other facets  affecting
numerical solution  efficiency,  accuracy   and  robustness,  such as  matrix  solution
methodology, should  be pursued  —  particularly to facilitate practical extensions to
three-dimensional problems.

-------
 Solution  of coupled  multiphase  flow  and  transport  problems introduces  additional
 numerical difficulties. The semi-decoupled phase-summed approach employed  in this
 study is an efficient formulation, although the efficiency undoubtedly comes at the cost
 of accuracy. The most critical aspect of the decoupled  approach is the back-calculation
 of  interphase mass  transfer rates,  since  accumulated small  errors can eventually
 destabilize the solution.  Such problems are greatly diminished by suppressing mass
 transfer calculations  during periods  of  highly transient oil flow.  Since  compositional
 changes are generally small over  short  time  periods, mass balance errors incurred  by
 this approach are very small. Future investigations  to investigate alternative solution
 formulations and to develop algorithms for automatic component mass balance controls
 on the solution should be pursued  (e.g., "re-inject ion" of mass balance errors or iteration
 of transport and flow  with corrected phase transfer terms).

 Accurate representation of k-S-P relations is crucial in order to predict  multiphase fluid
 movement  in the subsurface, and  accurate models for mass interphase partitioning are
 critical to describe species transport. We have developed a detailed model for hysteresis
 in k-S-P  relations which  has proven  to  accurately describe static  laboratory S-P data
 and transient flow response. Simplified models which consider effects of nonwetting fluid
 entrapment provide a reasonable compromise between  accuracy, on the one hand, and
 efficiency  and robustness, on the other. Direct measurements of three-phase relative
 permeabilities for non-monotonic saturation histories  is  a formidable task, but one that
 would be highly desirable to undertake in the future.

 Existing empirical  relations  for estimating column scale mass transfer rate constants
from basic physical  properties of the system yielded  surprising  accuracy.  However.
 under common field conditions, it may be shown  that mass transfer limitations  should
 be of minor importance.  Field scale heterogeneity, however, may lead to  preferred flow
 paths that produce apparent nonequilibrium effects controlled by diffusive mass transfer
 between "fast" and "slow" zones.

Since explicit treatment  of heterogeneity is generally impractical, future  studies  should
address the feasibility of describing field scale behavior using effective  large scale mass
transfer relations.  Likewise,  effective k-S-P relations at the field scale that  implicitly
consider  effects  of heterogeneity  may  well differ  from  laboratory  scale  relations.
Substantial future  efforts will  be needed to more  fully understand  and model the
behavior of large scale heterogeneous systems.

-------
               2.  NUMERICAL ANALYSIS OF MULTIPHASE FLOW

 2.1  Introduction

 Groundwater  contamination due to surface spills or subsurface leakage of hydrocarbon
 fuels, organic solvents  and other  immiscible organic liquids is a widespread problem
 which poses  a serious  threat  to groundwater  resources.  In order  to  describe  the
 movement of  such materials in the subsurface, it is necessary in general to account for
 separate phase flow of  nonaqueous phase liquid (NAPL)  as well as water and air.  A
 substantial obstacle to modeling multiphase flow has been the difficulty in determining
 constitutive  relationships  for three-phase  flow.    Recent  work  pertaining  to  the
 parametric description  and calibration of three-phase permeability-saturation-pressure
 relations including hysteretic effects  has diminished these difficulties to a significant
 extent (e.g., Parker et al., 1987; Lenhard and Parker, 1988; Parker and Lenhard, 1987).

 Numerical models for NAPL migration in the vadose zone and groundwater have been
 presented recently by various researchers based on  finite difference methods (Abriola
 and  Pinder,  1985b;  Faust,  1985; Falta  and Javandel, 1987) and on finite element
 methods (Osborne and Sykes, 1986,  Kuppusamy  et al., 1987).  Most of these efforts
 have focused on basic aspects of numerical model formulation with little attention to
 the effects of various techniques on computational efficiency and solution  accuracy.  A
 variety of formulations  of the governing equations for NAPL and water flow have been
 taken as  a starting  point for  numerical model  development.   Abriola  and Pinder
 (1985a,b), Osborne and Sykes (1986), Kuppusamy et al.,  (1987) and Kaluarachchi and
 Parker (1989) used flow equations in terms of fluid  pressures with mass storage terms
expressed as the product  of time derivative of fluid pressure and saturation-pressure
 derivatives.  The  formulations reported by Faust (1985)  and Forsyth (1988) have oil
pressure and water saturation as the dependent variables.  Forsyth and Sammon (1986)
presented  a  three-phase  flow  model  with  non-zero  gas  phase  pressure  gradients
employing oil pressure, oil saturation and water saturation as the primary unknowns
 and  Thomas and  Thurnau (1983) reported  a similar three-phase reservoir simulation
model in terms of oil pressure, water saturation and gas saturation.

With highly nonlinear single phase flow and transport problems, it has  been shown that
upstream  weighting  techniques can improve nonlinear convergence behavior especially

-------
for problems  with hyperbolic tendencies (Huyakorn  and Nilkuha, 1979;  Kaluarachchi
and Parker, 1989). Newton-Raphson iteration methods also offer greater robustness for
highly  nonlinear  problems, but  at a  cost  of  considerable  additional  computation
(Huyakorn and Finder,  1978; Huyakorn and Finder, 1983).   The influence coefficient
method of evaluating element matrices has been employed in single phase finite element
simulations to improve computational efficiency  (Huyakorn  et al., 1984;  Kaluarachchi
and Parker, 1987), but has not been applied to multiphase flow models.

Any numerical solution  to flow and transport  equations should have the capability of
conserving mass in the system under the applied  boundary conditions. Most numerical
studies  of multiphase flow simulations  have not  addressed this issue.  Abriola (1984)
investigated mass balance accuracy  for a finite difference  simulator during continuous
NAPL  infiltration only.  Long  term mass conservation  behavior for  pulse  injection
scenarios, which  are expected  to  be more troublesome,  has not  been  investigated.
Difficulties associated with mass conservation require careful attention to the handling
of saturation-pressure derivatives in the problem formulation.

In most practical problems, large changes in fluid  pressures  and saturations do not occur
throughout the spatial domain at a given time step.  Computational effort is thereby
inefficiently spent solving equations in  areas where  little  activity occurs rather  than
concentrating effort in the more active  zones. However,  since the  locations  of active
zones change with time, schemes designed to take advantage of these variations must be
capable of automatically adapting to the current  conditions.  One method for doing so
involves adaptive  grid refinement in which the numerical mesh is automatically refined
in active  zones and expanded  in  inactive zones.  Another general approach which
employs a fixed grid  varies the solution  approach within the domain to gain efficiency.
Adaptive  implicit methods have been  described by  Thomas and Thurnau (1983),
Forsyth  and  Sammon  (1986)   and Forsyth  (1988)  that  involve solving  implicit
formulations of the governing equations in parts  of the domain where changes in the
primary variables are large, while in the rest of the  domain a  computationally less
intensive but  less robust implicit pressure-explicit saturation (IMPES) formulation is
employed.

In this chapter,  we will  discuss various formulations and  techniques  for  modeling
multiphase flow problems to evaluate their efficiency and accuracy.

-------
 2.2  Development of Implicit Pressure- Pressure Formulation

 2.2.1  Theory and mathematical formulation

 Governing equations  for multiphase flow.  In the present  analysis, we assume pressure
 gradients in the gas phase  to  be negligible so that gas pressure  remains effectively
 constant at atmospheric  pressure.  Assuming also, an incompressible porous medium
 and constant fluid properties, the flow equations for water (w) and organic liquid phase
 (o) may be written in summation convention for a two-dimensional Cartesian domain as

                            h       \\       dhw      dh                    ,
                                              dh       dh
where i, are the Cartesian spatial coordinates  (t=l, 2), Kwij and Koij are conductivity
tensors for water  and oil  (i.e., NAPL), respectively;  hw and h0  are  water height-
equivalent pressure heads of water and oil (Parker et. al., 1987) respectively; pro is the
ratio of oil to water  density; Uj =  dz/dij is a unit  gravitational  vector  where z is
elevation, and t is the time.  The terms Cow> C00, etc. are fluid capacities defined by

                                  dSD
                                        P>V=°>W                             (2-2)
where  is the porosity of the medium, 5p is the saturation of phase p, and hg is the q
phase head.

Darcy velocities in water and oil phases may be written in the form

                          qwi = - Kwij       + «,                            (2.3a)
and
                                      ,flk
                                                                             (2.3b)
where  qwi and  qoi are velocity components along Cartesian coordinates i for water and
oil phases.  Phase conductivites are assumed to be described by

-------
                                                                             (2.4a)

                                           o                                 (2.4b)

where krw and kro are relative permeabilites of water and oil, respectively, rjro is the
viscosity ratio between oil and water and Ktw^ is the saturated conductivity tensor for
water.  We  shall  assume  here that  the  coordinate  system is  oriented  with  the
conductivity tensor, or otherwise that off-diagonal components may be disregarded, so
that Ktwij = 0 for t ? j.

Initial and boundary conditions for  each phase p may be written as

                     hp(xi, 0)  = hpl(xt)     on R for t = 0                     (2.5a)

                     hp(xi} t) =  fcp2(z,.,  t)   on S1 for t > 0                     (2.5b)

                     9P, »,- =  QPi(*i> *)     °n 52 for t> 0                     (2.5c)

where n, is the outward normal unit vector at the boundaries.  Here (2.5a) denotes the
initial conditions described by hpl  over the entire region R and  (2.5b) denotes type-1
boundary conditions along the  boundary segment 5j.  Equation (2.5c) denotes flux-type
boundary conditions with Qpi(zi,t)  representing normal boundary fluxes along boundary
segment 52 for phase p.

Upstream weighted finite element formulation.  We will apply the upstream weighting
technique described by Huyakorn  and Nilkuha (1979)  in conjunction with Galerkin's
weighted  residual method where terms involving spatial derivatives of the governing
equations are weighted using unsymmetric upstream  weighting functions.  Denote the
symmetric  basis  functions and unsymmetric upstream weighting functions as Nj and
Wj,  respectively, which are given  by Kaluarachchi and Parker (1989) for the case of
linear  quadrilateral elements.  The exact solutions for hw and  h0  are assumed to  be
approximated by
                                    #/ (e, T,) hwl (t)                           (2.6a)
                                 =1
and

-------
                   .  h. (zt, t)=rNj (e, 77) hol (t)                           (2.6b)
where  hwj and hoj are the values of  hw and h0 at node 7,  and e and 77 refer to local

coordinates within the element.



Using  the weighted  residual  method based on  Galerkin's  principle,  eq.  (1)  can be

written as
 f  W, £ fay ( ^ + u, )) dR - J AiC^ ^ tf -}

 fl      ' ^         j       '      R                R
                                                                            (2.7a)
 f  w  9 ( K   (dhw          \\ ,    r   „  oh0 ,   r    r   dhw            .     ,
 J    ' ~dx (  oij (~dz~    pro  j))     J    '  °° ~dt     ~\    'C°w ~dt  dR = 0     v2-7b)
 R      ' ^         }           '      R               R



After further simplification using Green's theorem, (2.7 may be written in matrix form

as

                                         w \  I  r cx^i i    o \    fc1^            /oo\
                                      "dTJ  + ™(-5TJ=^F-)            (2-8a)


                                                        = {F0}              (2.8b)
where


                                      BWjdN^
                           v1-  f  TS   dWrdN,  JD
                        = e£  J   ^ -g^- -g^  ^                          (2-8c)
                                                                            (2.8d)
                                                                            (2.8e)
                        = E  f CW^JV, iIA                                (2.8f)

                          C = 1
                        = E  f C^/^J ^                               (2.8g)

                          e = 1

-------
                     E°u = E  f  CowNjNj dR                               (2.8h)
                           e = 1 X
                 n  I t      AW          r          \
             = EN Kwij ^ «,. d* - J  W/fcl dS\                        (2.8i)
               * *•" ^ \  ia                   n         /
=  £(• J  ** '" TE1"' dR' I  tr"°' *)
  e — 1 \  D            *          r«         /
                                                                            (2.8J)
where n is the total number of elements.  Here Se refers to the segment of the surface of
element e where flux-type boundary conditions prevail and Re is the element volume.

Evaluation  of  element matrices.  An  important consideration in the  finite  element
formulation is the handling of nonlinear  terms in element integrals such that element
matrices may be computed in an efficient manner. In this work, we use the linear shape
functions  described  to evaluate soil  properties within  the elements.  For example,
capacity terms  are represented by
                                  4
                         C/_  4\ _ \~^  AT (~q                                   (f) Q-\
                      pg \xi> l) ~ 2—i   B P9                                  \&-s&)
where


in which C|, denote Gauss point capacities and h*, and hs0 refer to Gauss point pressure
heads.  In a typical  approximation of nonlinear parameters by  (2.9), local nodal heads
would  be used.   Our  experience indicates  that  numerical instability  and  poor
convergence  could occur  with  nodal  heads especially with problems involving sharp
fronts or material nonhomogeneity. Present scheme is more appropriate  as Gauss point
are somewhat optimal  sampling  points in the finite element method than the nodal
values  (Zienkiewicz,  1986).   Once this procedure  is  adopted with  all  other  soil
properties, it is possible to evaluate the influence coefficient matrices for each integral in
(2.8c) to (2.8j). In  the present analysis,  we  consider the  formulation  for  linear
rectangular elements with sides parallel to the principle axes.  Equation (2.8c) may be
modified as

-------
Equations (2.8d) - (2.8j) may be modified in a similar fashion.  Element matrices in
(2.8a) and (2.8b) can now be written as


                               A», + 4 £  (0")S ^                      (2.Ha)
                                        9=1
                                                                            (2.nd)
                             4  .=1
In (2.11g)  and (2.11h), flux-type boundary conditions have been  omitted and will be
treated separately. Also  superscript g in each  parameter  corresponds  to  the values
evaluated at a given Gauss  point  g.  Matrices [0*]£, [IP]', [V\eg and  [G]eg in (2.11)
represent influence coefficient  matrices which can be further simplified as

                                                                            (2.i2a)

                                          «                                  (2.l2b)
                                      10

-------
                                                                             (2.12c)
where i and y represent coordinate directions,  r denotes matrices involving  regular
linear shape  functions  and u  denotes those with  contributions from the upstream
weights  QJ and  /?,-.  Note  that matrices [V\ are  independent of  upstream weights.
Influence coefficient matrices in (2.12a) to (2.12c) are described by Kaluarachchi and
Parker (1989).

Treatment  of nonlinearities by Pi card method.   The Picard  method is  one of the
commonly used schemes to solve a set of nonlinear ordinary differential equations by a
suitable finite difference approximation. With moderately nonlinear flow and transport
problems, the Picard scheme provides rapid convergence.  A distinct advantage of the
Picard scheme is its simplicity  and lower  computational effort  per iteration than  more
sophisticated schemes.

The  Picard iterative  scheme  involves  approximating  the resulting finite  element
formulation by a finite difference approximation.  First equations (2.8a) and (2.8b) can
be written as
where
                             (A-'}  0
                              0  [A°]
                     [!?} =
                                                                              (2.13)
(2.14a)
(2.14b)
                                                                             (2.14c)
Equation (2.13) can be written in finite difference form as
                                                                             (2.14d)
                                                                              (2.15)
                                      11

-------
where matrices at time (k + 9} are evaluated using the heads hk + e with

                                                                          (2.16)
in which k and k+1 denote the previous and current time levels, Aifc + i is the current
time step and 6 is a time weighting factor.  For a fully stable numerical solution, it is
recommended to use  6=1 corresponding to the  fully-implicit backward  difference
scheme.  Use of (2.15) combined with (2.16) to obtain the current time step head values
is straightforward.

Treatment of nonlinearities  by Newton-Raphson method.  The Newton-Raphson method
is often recommended for highly nonlinear problems for which the Picard scheme may
fail or  provide slow convergence (Huyakorn and Finder, 1983). We may rewrite (2.8a)
and (2.8b) at matrix element level as

                    rwi = A0J   ~ hoJ)  ~ ~rwl                      (2.18a)
                               -oJ
            wJ               unoJ
                               s  (ft i . Vo]] = -r'ol                      (2.18b)
                             un
where  r and r + 1 are the iteration  numbers at k + 1  time step and which may be
written as
                               ££^1*31. M.                    (2.19a)
                                  &tk + i    ohwj  dhwj                    v     i
                                    12

-------
                          l - hkoj)
                            (2.19b)
                      dh..,j
                            +
                               XV
                        wJ
+
    E°
      IJ
                                            dhwj
                                                                             (2.19c)
               _
               — tloL
                     -gr -
                     dh
                        oj
                                              •
                                                                             (2.19d)
In (2.19a) to (2.19d), terms with superscript k refer to previous time step values where
the quantities are known.  The subscript L is a dummy index denoting the appropriate
influence coefficient matrix to be  selected for  a particular set of / and J.   Once the
recently updated values of heads are used to evaluate quantities in (2.19), the final set
of simultaneous equations to be solved by an iterative method can be given as
                                                                              (2.20)
where
                              drwl  ,r
                                —
                                                                             (2.21a)
                                                                             (2.21b)
                                       13

-------
Mass lumping.  For most nonlinear flow  problems and especially for multiphase  flow
problems, the use of a consistent mass matrix with integrals given by (2.8e) to (2.8h)
will cause  instability  in  the solution  that  would  produce  poor convergence.   To
overcome these  difficulties and to obtain a more stable solution, a common approach is
to  diagonalize the  mass matrix  -- a procedure referred to as mass lumping.   The
procedure is given for (2.8e) as
                              NgNjdRC°WW  for  I=J
                                                                  (2.22a)
                     g =
                          Bfi = 0  for 7 /  J
                                                                  (2.22b)
Similar procedures are followed for the remaining equations given by (2.8f) to (2.8h).

Flux-type  boundary  conditions.   For  flux-boundary conditions  given  by  (2.5c),  the
original finite element formulation given by (2.11g)  and (2.11h) can be  modified with
very little  computational effort for both Picard and  Newton-Raphson schemes.  These
modifications in  Equations (2.8i) and (2.8j) for both schemes  can be written for any
given phase as
-I
2
-------
          Cp'g = (1 - w)  Cpkg  +  wCpkg+l     P,q=o,w                         (2.24)

where C*g  is the  weighted capacity,  superscripts  k and  k -f 1  denote  previous  and
current time levels, and w is a time- weighting coefficient varying between 0 and 1.

Mean head  analytic scheme.  This  scheme was initially suggested  and used by Osborne
and Sykes (1986) and is given as

                Cp'q = Cpq (/C V)   P,q=o,w                             (2.25)
where
with superscripts k and A -f 1 denoting previous time step and current iteration values of
pressure heads, respectively.

Modified Chord-Slope Scheme.  According to the standard chord-slope scheme (Abriola,
1984), the capacity Cpq (p, q= o, w) is defined as
Due to very  poor convergence  under highly nonlinear conditions  from  preliminary
investigations with a standard chord-slope method, this method was abandoned in favor
of the modified scheme which introduces time-weighting in a manner given by (2.24)
with w = 0.5.  The expression for this scheme is given by
                -f  Sp(hkp, &5 + >) - $„(/£+', fcj) - Sf(hli, hkg]                      (2.27)

                                      15

-------
 Updating nodal pressure beads.  During each iteration, nodal pressure heads need to be
 updated for  the next iteration. With the Picard method, a  common procedure is to
 simply  employ (2.16). With highly nonlinear flow problems,  schemes of this type may
 not  be fully  effective  as  the heads  are  updated without  due consideration  to  the
 maximum convergence error for the entire mesh.  The updating method introduced by
 Cooley  (1971) which introduces an optimal relaxation scheme, which accounts for  the
 maximum convergence error for the entire mesh, was used in conjunction with both the
 Picard and Newton-Raphson schemes. This method is briefly described below. Let
where N is the number of nodes in the mesh
                                          hrp
Step 1
                               =1        Jfc=0
Step 2
                     *,   = (3 + Sp)/(3 +  |5P|)      5P >  • 1
Step 3
                                         ep+1
where e™ax is the maximum change in the hp allowed during any iteration.

The convergence criterion used in this study for a given phase p (= o, to) is as follows
where  r+1 and r refer to current and previous iterations, v is the allowable relative
convergence error and cp is the allowable absolute convergence error.
                                      16

-------
 Automatic adjustment of time step is undertaken depending on the number of iterations
 required for nonlinear convergence.  If the number of iterations is less than  /mtn, the
 time step  is increased  by a  factor 1 + F and  if greater  than /mox,  the time step is
 decreased by a factor 1/(1 + F).  In the  simulations in this paper, satisfactory results
 were obtained using 7min=4, Imax=l2 and  F < 0.04.

 2.2.2 Numerical investigations

 Several example problems will be analyzed  to  investigate various aspects of the
 numerical  model.  The first  example will focus attention  on  the  effects of various
 capacity computation techniques on mass balance  accuracy.  The second example will
 evaluate the efficiency of  Picard and Newton-Raphson iteration schemes combined with
 upstream weighting in analyzing a highly  nonhomogeneous soil profile. The last example
 will  address the  effects  of different fluid properties on the  flow of  NAPL in a two-
 dimensional flow  domain.
TABLE 1.  Soil and Fluid Properties Used in Simulations
Parameter
n
a
sm
K.w
P.O
(iow
pn
nn

Example 1 Case
A B
3.25 3.25
0.05 0.05
0.0 0.0
50.0 50.0
1.8 3.0
2.25 2.5
0.8 0.8
2.0 2.0
0.40 0.40
Example 2 Layer
1 2 3
3.0 1.8 4.0
0.07 0.01 0.1
0.05 0.2 0.0
40.0 0.5 60.0
1.8 1.8 1.8
2.25 2.25 2.25
0.8 0.8 0.8
2.0 2.0 2.0
0.40 0.40 0.40
Example 3 Case
A B C D
2.1 2.1 2.1 2.1
0.007 0.007 0.007 0.007
0.02 0.02 0.02 0.02
2.1 2.1 2.1 2.1
1.8 1.8 1.8 1.8
2.25 2.25 2.25 2.25
1.2 1.2 0.8 0.8
2.0 0.5 2.0 0.5
0.43 0.43 0.43 0.43
All units are given in centimeters and hours. Isotropic conductivity is assumed with Km
                                       17

-------
Example  ii   One-dimensional  Infiltration  and  redistribution  in uniform  soil  The
problem analyzed corresponds to a vertical soil column 100 cm long with an oil-free
initial condition in equilibrium with a water table located 75 cm below the top surface.
The simulation was performed in two stages.  In stage 1,  oil was allowed to infiltrate
into the system under a water equivalent oil head of 3 cm until a total of 5 cm3 cm ~2 of
oil had entered.  Boundary conditions for the water  phase were zero-flux at the top  and
a constant head of 25 cm at the bottom during infiltration.  A zero flux oil condition
was  employed at the lower boundary.  During stage 2, oil was permitted to redistribute
up to t=100 hours with zero-flux conditions for oil at both boundaries and conditions for
water  the same as during infiltration.  Soil and fluid properties  used in the simulation
are given  in Table 2.1 corresponding to nonhysteretic three-phase relations described in
Chapter 3.
                 E
                 U
                 Q.
                 01
                D
t  -   0. 09  h
Chord—si ope
Mean  Head
                     100
                         0. 0   O. 2    0. 4    0. 6   O. 8    1.0
                                  Saturat i on
Fig.  2.1   Initial water saturation distribution and total liquid saturation at the end of
          infiltration for Example 1A.
                                      18

-------
The initial condition of zero oil saturation was assigned by fixing the initial oil head at
each node to the critical oil head, hc0T, using the following expression which derives from
the form of the scaled constitutive model

                                                                              (2.28)
where 0ao and 0OW are fluid-dependent scaling factors defined in Chapter 3. As pointed
out by  Parker and Lenhard (1990), a jump  condition in the water saturation vs.  air-
water capillary head function, Sw(haw),  will occur during  the  transition  from a two-
phase air-water system to a three-phase air-oil-water system in circumstances for which
NAPL contamination diminishes the surface tension of water  relative to the pristine
state. The latter condition corresponds to the criterion
                     I/flu, +
                                                         (2.29)
                E
                U
                o_
                01
               D
                      20  -
40 -
                     60 -
                     80 -
                    100
              MQon  Hood
              Chord—s1opQ
              Equ ili br i um
                        0. O    0. 2   0. 4    0. 5    0. 8
                                  Saturot. i on
                                         1. 0
Fig.  2.2  Water and total liquid saturations at the end of redistribution  period for
          Example 1A and theoretical equilibrium distribution.
                                       19

-------
To avoid numerical problems associated  with this jump condition, a phase  updating
scheme is adopted  at the end of each time step to index whether the node  is a two-
phase  or three-phase  system (i.e., oil is absent  or present).   Once  a three-phase
condition occurs, reversion to a two-phase condition is not allowed.  The criterion for a
node to change from a two- to three-phase system is that h0  >  hcj. It is important to
note that in the part  of the domain remaining  as an air-water system, capacity and
conductivity terms  related to the oil phase will become zero and the oil flow equation
solution  reduces  to the identity 0  = 0.   To  avoid this problem, we assign minimum
cutoff  values to  capacities and oil relative permeabilities.  Minimum  values  used for
capacities Cwo, C00 and  Cow are 10~6 cm"1 and the minimum kro is 10'6, which were
determined by trial to provide accurate results.
        Table 2.2
Summary of Results for Example 1
Capacity
Computation Scheme
Time average
analytic, u> = 0
Time average
analytic, a/ = 0.5
Time average
analytic to = 1.0
Mean head analytic

Modified chord slope

Period T
I
R
I
R
I
R
I
R
I
R
CPU Time, ' s
Case A Case B
PMB*
PMB*
105
444
92
449
123
434
104
483
81
411
80
421
80
433
89
428
93
460
Oil Mass
Error
Case A
PMB*
PMBt
1.6
4.2
-0.8
-1.8
4.2
5.2
0.1
3.4
Balance
, %
Case B
-30.8
-30.6
11.6
-11.6
-20.8
-22.0
-9.2
9.8
-5.4
-6.4
        *An IBM 3094 system was used for simulations.
        tl, infiltration of 5 cm of oil; R, redistribution up to 100 hours.
        tPMB, extremely poor mass balance (> 70%).
                                      20

-------
To evaluate effects of the jump condition associated with transition from two to three-
phase systems, we consider two cases in this example from £!//? = 1 for Case  A and
531//3 = 0.73 for Case B.  Simulations were performed using the time-averaged analytic
capacity scheme with time-weighting of w = 0.0, 0.5 and 1.0, and with the mean head
analytic scheme and the modified chord-slope method.  Upstream weights used in all
simulations were 0.2.  The finite element  mesh used in the simulation consists  of 100
elements with a uniform spacing of 1 cm and the time step varied between 0.001 to 0.5
h with a time increment factor of 4%.

The duration of the infiltration stage was approximately 0.09 h for Case A and 0.1 h for
Case B. The initial water content distribution and liquid saturation distribution at the
end of the infiltration stage  using the  mean head analytic and  modified  chord-slope
schemes are illustrated in Figure 2.1 for Case A.  It may be noted that the initial water
saturation  at the upper surface was  very  low (Sw K 0.05).   The water  saturation
distribution at  the end of infiltration  was almost identical to the  initial conditions.
Computed total liquid saturation  distributions  at  the end of infiltration  were almost
identical for the two schemes shown, as well as for the time-averaged analytic schemes
for w = 0.5 and 1.0,  which are not shown.   Although the infiltration fronts are very
sharp, no  numerical  difficulties were  encountered owing  to  the effectiveness  of the
upstream weighting procedure.

Saturation distributions at the end of 100 h of redistribution are illustrated in Figure 2.2
along with equilibrium distributions computed from hydrostatics for an oil volume of 5
cm subject to the imposed boundary conditions. Here, equilibrium condition is defined
as the fluid distributions at which the total head gradient of both phases with respect to
elevation approaches zero.  Results for  mean head and chord slope schemes are quite
comparable as  were  time-averaged  analytic scheme results  for w =  0.5 and  1.0.
Convergence toward  the  equilibrium  results provides  a  verification check for the
numerical  model. It  is noted,  however,  that whereas  the equilibrium distribution
indicates no oil above a depth of 43 cm the numerical  model predicts an average oil
saturation  remaining above this depth  of 7-8%  even though oil velocities at 100 h are
practically zero.  This "residual oil" content predicted by the numerical model reflects
the equivalent of an oil "field capacity" due to rapidly diminishing drainage rates under
gravitational gradients alone.  It is important to note that this oil is not immobile in
any strict sense and is predicted without invoking hysteresis, fluid entrapment or fluid-
dependent  residual saturations.  Due to ambiguity in the term residual oil, which  is

                                      21

-------
often used  to  imply. occluded or discontinuous  oil,  a more  descriptive  term for  this
phenomenon may be nondrainable oil.

A  comparison of  execution  times  and  mass balance  errors  for the  simulations of
Example 1  are summarized in Table 2.2.  The mass balance error was  always higher
during redistribution than  infiltration.  For Case A, all of the capacity computation
schemes provided satisfactory mass balance results, except for the analytic scheme with
w = 0.0 which produced very poor  results.  Optimum time-weighting for the analytic
scheme appears to be between 0.5 and 1.0.  Mass balance results for Case  B with £!//?
 ^  1 were consistently poorer than those for Case A.  The modified chord-slope scheme
yielded clearly superior mass balance results with  an increase in computational effort of
about 10% over the analytic schemes. The higher accuracy of the modified chord-slope
scheme when  £!//? ^  1  is attributable to the  fact that it is the  only method  that
properly evaluates saturation  derivatives during  the jump condition associated  with
transition from a two-  to three-phase system.

Example 2l One-dimensional infiltration and redistribution jn nonhomogeneous soil.  The
main objective of this example is to evaluate the efficiency of the  Picard and Newton-
Raphson schemes in combination with different  upstream weights under conditions of
strong nonlinearity  and heterogeneity in material  properties.   The geometry of  the
problem is  the same as that of Example 1  but the  soil profile is assumed  to be
comprised of three layers.  Layer properties are given in Table 2.1.  The middle layer
represents a finer grained material than  the other materials with a saturated hydraulic
conductivity contrast of two orders of magnitude.  Fluid properties used are the same as
for  Example 1A with £!//? = 1. Initial and boundary  conditions are also identical to
those of the previous  example except that redistribution time was continued to  150
hours.  Simulations were carried  out using both Picard and Newton-Raphson methods
each repeated  for four sets  of upstream weighting factors (0,  0.2, 0.5, 0.8).  The mesh
and time steps used in this example is identical to  that of Example 1.

Simulation results are illustrated  in Figures 2.3 and 2.4 for Picard and Newton-Raphson
analyses with an  upstream weighting factor of 0.5. The  extreme variability with depth
in initial water contents (Figure 2.3) produces a numerically difficult  problem on which
to test the capabilities of various numerical schemes.  Simulated total liquid saturation
distributions at  the end of infiltration  exhibit  very sharp  fronts  and are virtually
indistinguishable  for the two iteration methods (Figure 2.3) as is the case at the end of

                                     22

-------
the redistribution period  (Figure 2.4a).  At 150 h oil has penetrated to the water table
but liquid velocities have become so small that redistribution has virtually ceased.  The
equilibrium liquid distribution shown in Figure 2.4b predicts no oil above a depth of 55
cm which is 5 cm below the clay layer.  The simulations show fully half of the oil in the
system remains  above this  depth at  150 hours.  This  "nondrainable oil saturation"
averages about 12% for the upper two layers which is considerably higher than found for
the homogeneous soil in  Example 1. This is attributable to the higher oil retention
within the fine layer itself and to its impeding effect on drainage from the overlying
layer.

A summary  of mass balance errors  and CPU times  for Picard and Newton-Raphson
simulations with all upstream weighting factors  are given in Table 2.3.  The different
numerical schemes provide similar accuracy  with  respect to mass balance with larger
errors occurring during redistribution periods.  For the case of zero weights, the Picard
scheme failed to converge after 12 time steps  during the redistribution  period.  The
Newton-Raphson  scheme  converged,  but required greater  computational effort than
when upstream weighting was employed. The Picard scheme was able  to converge only
after upstream weighting  was increased to 0.2 and thereafter remained insensitive to the
weighting.  Mass  balance was  affected  little by  upstream weights  confirming  that
upstream weighting serves mainly to improve convergence.
E
0
I
£
0.
01
a
                           20 -
                           40 -
                           60 -
                           BO -
                          100
 t  - 0. 122 h
	   Picard
 °   Nowton-Raphson
                            0.0   0. Z   0.4   0.6  0. B   1.0
                                   Saturation
Fig. 2.3   Initial water saturation distribution and total liquid saturation at the end of
          infiltration for Example 2.
                                      23

-------
                    E
                    U
                    I
                   r
                   *j
                    CL
                    01
                   D
                        20  -
40 -
BO -
                        80  -
                       100
                               D   Newton—Rophson
                          0. O   0. 2    0. 4   0. 6   0. 8   1.0
                                -  -  -  FEM
                                          Equi1i brium
                          0. O   O. 2   Q. 4   O. 6   0. 8   1. O
                                   Saturat i on
Fig.  2.4  Water and total liquid saturations at the end of redistribution  period for
         Example   2.  (Top)  Comparison  between  Picard  and  Newton-Raphson
         schemes. (Bottom) Comparison between numerical (150 hours) and predicted
         equilibrium distributions.
                                     24

-------
 One important .aspect of these simulations was to investigate whether increasing the
 weights would deteriorate the accuracy.  The results in Table 2.3 do not indicate such
 deterioration and show that a highly nonlinear problem of this type can be effectively
 handled with a low weight of 0.2.  It is evident  however that computational cost can be
 almost 1/3 more with the Newton-Raphson scheme while  obtaining the same degree  of
 accuracy as the  Picard method. The results indicate that the Newton-Raphson method
 did not improve convergence substantially when upstream  wieghting was used and took
nearly the same number of iterations per time  step as  the Picard scheme.  The reason
for increased computational cost  of the  Newton-Raphson method is mainly  due to
number of  terms  needed per  node at each  iteration.  For  the  Picard scheme,  6
coefficients must be evaluated, namely  krw, kro> Cww,  C00, Cwo and Cow, whereas the
Newton-Raphson scheme  requires  12  more  terms  corresponding   to  the  partial
derivatives of the former terms with respect to both hw and  h0.  Computational costs
decreased slightly with increased upstream weights for the Newton-Raphson method but
increased somewhat for the Picard scheme beyond a weight of 0.2.
              Table 2.3  Summary of Results for Example 2
Oil Mass
Balance Error, *
Upstream
Weighting
Factors
0.0

0.2

0.5

0.8

% CPU Time, s
Period t
1
R
1
R
1
R
1
R
P
-3.7
NC*
-2.8
-6.0
-1.6
-4.0
-1.6
-5.2
NR
-3.6
-6.5
-2.8
-4.4
-2.8
-6.4
1.6
-5.0
P
78
NC*
70
282
71
315
71
312
NR
99
485
88
413
89
414
89
405
               *P, Picard scheme; NR, Newton-Raphson scheme.
               tl, infiltration of 5 cm of oil; and R, redistribution up to 150 hours.
               *NC, No convergence after 12 time steps during redistribution.
               An IBM 3094 system was used in the simulation.
                                    25

-------
                       A    5m     B
      E
      in
                         Water   table
                                                                       E
                                                                       n
                                      23 m
                       Fig. 2.5 Geometry of flow domain for Example 3.
Example 3: Two-dimensional infiltration and redistribution for varying fluid properties.
The objectives of this example are to evaluate the model for a field scale problem and to
investigate the effects of fluid density and viscosity on the movement of NAPL.  The
geometry of the flow domain which has an initial water table  sloping from left to right
with a gradient of 2/23 is shown in Figure 2.5.  Soil and fluid properties used in the
simulations are given in Table 2.1.  Note that Cases A-D involve variations in the oil
density ratio  (pTO =• 0.8 and 1.2) and viscosity ratio (r)ro  = 0.5 and 2.0).   Oil  was
assumed to infiltrate along segment AB in Figure 2.5 under a small head of 1.0 cm to
simulate an oil spill.  The boundaries at the ends and below the water table for the
water phase correspond to type-1 conditions with water heads stipulated, while other
boundaries were zero-flux type for water.  For the oil phase, all boundaries except AB
were zero-flux.  For all cases a total of 7.5  m3 m"1  of oil was permitted to infiltrate,
following which redistribution was allowed up to a total  simulation time of 200 days.
The finite element mesh used in the simulation consisted of 345 nodes and 308 elements.
Time step sizes vary between 0.01 to 2.5 days with a time  increment  factor of 4%.
Both Pi card and Newton-Raphson schemes were used  with an upstream weight of 0.5 to
solve the set of nonlinear equations.  Mass balance errors  at the end of the simulations
for both  schemes  remained almost  identical  and varied  between  -2  to  -6%.
Computational efficiencies however were much different between each scheme where the
                                     26

-------
Newton-Raphson scheme required approximately 25 to 35% more computer time. The
Newton-Raphson scheme was slightly faster in convergence than  the Picard scheme in
comparison to the average number  of iterations taken per time  step but as discussed
earlier, the former scheme requires many more computations per iteration due to larger
number of nonlinear parameters associated per node.

Cumulative oil infiltration with time during the infiltration stage is shown for the four
cases in Figure 2.6 where the infiltration times were 40.1, 10.1, 57.7 and 13.4 hours for
cases A-D, respectively.  The effects  of fluid density and viscosity on infiltration rate are
clearly evident. Analysis of the results  indicates  that the duration  of  the  infiltration
stage is  almost  exactly proportional to the relative  dynamic  viscosity  T)ro/pro,  as
expected on  theoretical  grounds, since the oil plume  does not  encounter any material
inhomogeneities or the water table during the infiltration period.  These results can also
be used as a verification exercise of  the numerical model.   Although  the  infiltration
times varied  for the different fluids,  at the end of the infiltration stage the distributions
of oil for all  cases were  almost identical with a maximum penetration depth about  50
cm above the water table.
                              10   2O    3D   40   so    60
Fig. 2.6   Cumulative infiltration versus time for Example 3A-D.
                                      27

-------
                                             ro
                                                  =  2-0
                                            t  =   200  d
Fig. 2.7  Oil saturation contours after 200 days for Example 3A.
                                              t  =   200  d
                                                  ro
                                                  ro
1.  2
O.  5
Fig. 2.8  Oil saturation contours after 200 days for Example 3B.
                                   28

-------
                                                   r-o
           =   0. 8
           =   0. 5
            t  =  200  d
Fig. 2.9   Oil saturation contours after 200 days for Example 3C.
                                              'r-o
                                                       0.  8
                                                       2.  0
                           0.  05
t   =   200  d
Fig. 2.10   Oil saturation contours after 200 days for Example 3D.

                                     29

-------
Figures 2.7 -  2.10 show oil saturation contours at the end of the simulations (t = 200
days).  Cases A  and  B  which  involve  dense NAPLs of  high and  low  viscosity,
respectively, exhibit substantially different behavior  (Figures 2.7 - 2.8).  Whereas the
low  viscosity  NAPL has fully penetrated  the aquifer and spread laterally about 14 m
along the bottom boundary, the more viscous NAPL does not reach the bottom of the
aquifer and remains confined to a soil volume 2/3 as large as that of the dense NAPL
with a center of gravity well above the original water table level. While the plume for
the  low viscosity dense NAPL is skewed in the direction of water flow,  the viscous
plume remains nearly symmetric and is evidently unaffected by the groundwater flow.

For  the light  NAPLs,  limited penetration  of the aquifer is observed as expected due to
the bouyancy effect (Figures 2.9-2.10).  Effects of viscosity ratio are similar  to those for
the  dense NAPLs. The low  viscosity  plume has spread  laterally to a much greater
extent and exhibits downgradient migration over the  water table.   By contrast, the
viscous plume remains much more compact, maintains a much shallower center of mass
and  is virtually unaffected by the groundwater flow.

For  all simulations, it was observed that NAPL velocities at the end  of 200 days were
extremely small, indicating that in  real field situations, very slow movement may be
observed over long periods of time.  The reason for this slower movement is that as the
oil plume spreads over a larger area, oil saturations and hence oil permeabilities become
smaller.
2.3  Evaluation of Alternative Equation Formulations

In this section, we consider various alternative methods of formulating the governing
equations for flow to assess their relative merits in  terms of efficiency,  accuracy and
robustness.  Five  different formulation methods will  be evaluated, which are described
in the following section.

2.S.I  Formulation of governing equations

Method  1;  Pressure  formulation with  saturation  time  derivatives.  Assuming  an
incompressible porous medium and incompressible and compositionally uniform liquids,
the governing equations for flow in a three fluid  phase porous medium may be written

                                      30

-------
as
                                                                              <2-30a>
where ^ is porosity; Sp is the p-phase fluid saturation for p = o (oil)  or w (water); t is
time; x, is the t-direction coordinate; Kpij is the p-phase conductivity tensor (i,j = 1,2,3)
for phase p; hp =  Pp/pwg is the water height-equivalent pressure head of the p-phase
where Pp is the p-phase pressure, pw is the density of water, and  g is gravitational
acceleration; prp is the specific gravity of the p-phase (prw =  1); and  Uj = dz/dXj is the
unit  gravitational  vector where z is elevation.   For  brevity,  we  will refer  to the
nonaqueous phase  liquid simply as "oil."  Phase conductivities are described by (2.4),
initial and boundary  conditions are specified by (2.5), and  Darcy fluxes are given by
(2.3).

The unknowns in Method 1  are taken to  be water and oil head, hw and ft0, which will be
treated here using  a finite element model with linear basis functions.  Time derivatives
will be  handled  using  a  finite  difference  approximation. Water  and oil  relative
permeabilities,  krw and  fcro,  and saturations,  Sw  and 50,  are  treated  as nonlinear
functions of phase pressures which are  updated iteratively  for the current time step.
Details of the numerical methods will be  described in a later section.

Method 2l  Pressure formulation with split time derivatives.  Method 2 also involves oil
and water pressure heads as unknowns, but in this case saturation time derivatives are
split into two terms each written as pressure  time derivatives.  Noting that Sp = f(hw,
h0) and expanding  the time dervatives of (2.30) yields

                    -  dh* _L  n....  ^o _  A.  I v    \ dhw  f  tt A           (2.31a)

where pg-phase fluid capacities are defined by

                                      31

-------
                    .   Cpg=t ^                 p, q =o, w                    (2.31c)
                               9
Initial and boundary  conditions are treated exactly as in Method 1 as given by (2.31).
In Method 2, water and oil relative permeabilities, krw and kro, and fluid capacities, CM,
are treated as nonlinear functions of phase pressures  which are updated iteratively.

Method  2l Capillary  pressure  formulation.  In  Method 3,  fluid pressure  heads are
replaced by capillary heads defined for fluid pair pq by  hpg = hp - hg. Substituting air-
oil  and oil-water capillary heads  into  (2.30) and noting that  in  the present  analysis
dhjdii = 0 by assumption yields
                                  d
d
                             0          tf   \   .,       -oo 1 1                  /0 ookl
                           ~dt  =   H;   K0ij 1 Wi ~  -fir J                   (2.32b)
In this method, air-oil and oil-water capillary heads are taken as the primary variables
and phase saturations and relative permeabilities are treated  as nonlinear functions of
the capillary heads. Initial and boundary conditions for the model are defined by

          hao(*i> °) = fc-oifo)                     on R at i =  0                 (2.33a)
          >W(*,, 0) = howl(Xi)                    on R at t =  0                 (2.33b)
          hao(z{, t) =  ha^(xi}t)                    on Si for t> 0               (2.33c)
          **.(«•» *) = hortfaJ)                   on S2 for <> 0               (2.33d)
           0               (2.33e)
             s         n
          q0. n, =  q0n                           on S4 for t> 0               (2.33f)

where  (2.33a) and  (2.33b)  denote  initial  conditions,   (2.33c) and  (2.33d)  indicate
prescribed capillary head boundary conditions, and (2.33e) and (2.33f)  specify flux type
boundaries.

Method 4 Semi-diffusive oil pressure-water saturation formulation.  Diffusive forms of
mass flow equations  may be  written by expanding pressure derivative  terms into
saturation gradients and pressure-saturation derivatives yielding homogeneous equations
in terms  of  fluid  saturation.   Here, we consider  a  "semi-diffusive" two-phase model
which is formulated in terms of water saturation and oil pressure.  The basic form of the

                                       32

-------
governing equations in this formulation is the same as used in Method 1.  However, the
unknowns  are now taken to be water saturation,  Sw, and oil pressure head,  h0.  Other
quantities  in  the equations are treated as nonlinear functions of the primary variables.
Water  relative permeability is  expressed as a function  of  water saturation, and oil
relative permeability is expressed as a function of water saturation and oil head via the
constitutive relations described  in Chapter  3.   Oil  saturation  is computed from the
equality: S0=St-Sw where total  liquid saturation,  5,,  is computed as a function of oil
pressure head. Initial and boundary conditions for Method 4 are prescribed by

          Sw(xit 0) = Swl(xi)                      on R at t = 0                 (2.34a)
          h0(xi, 0) = hol(zi)                       on R at t = 0                 (2.34b)
          SJz,, t) = 5w2(x,-,t)                     on Si for t > 0               (2.34c)
          h0(Xi, t) = hM(xi,t)                      on S2 for t> 0               (2.34d)
          q«,. n,- = qu,                            on S3 for t> 0               (2.34e)
            i        n
          q0.  n, =  q0                            on S4 for t > 0               (2.34f)
            i         n

where (2.34a)  and (2.34b) denote initial conditions, (2.34c) and  (2.34d) indicate type-1
boundary conditions, and (2.34e) and (2.34f) specify flux-type boundaries.

Method &   Semi-diffusive water pressure-liquid saturation formulation. Method 5 is a
semi-diffusive scheme in terms of total liquid saturation, St=S0 + Sw,  and water pressure
head, hw.  The governing equations for this method are identical to  those in Method 1.
The transformation  of the governing equations to obtain St and hw as  the primary
variables  is achieved  by  treating all  other  variables in  the equations  as nonlinear
functions of the primary variables.  Oil pressure head, hol is expressed as a function of
St via the constitutive  model described in Chapter 3; and Sw is expressed as a function
of hw and St.   Water and the oil relative permeabilities,  krw and  kro, are expressed in
terms of hw and 5,  also via the  constitutive relations.  Initial and boundary conditions
for Method 5  are specified as

          Stfa, 0) = 5tl(z.)                       on R at t = 0                 (2.35a)
          MX,-, 0) = hwl(Xi)                      on R at t = 0                 (2.35b)
          St(xi} t) = St2(Xi,t)                      on Sj for t> 0               (2.35c)
          hjii, t) = hrffat)                     on S2 for t> 0               (2.35d)
          q«,. n, = q^                          on S3 for t > 0               (2.35e)
          q0.  n,- =  q0n                           on S4 for t > 0               (2.35f)

                                       33

-------
where (2.35a) and (2.35b)  denote initial conditions, (2.35c) and (2.35d) indicate type-1
boundary conditions, and (2.35e) and (2.35f) specify flux-type boundaries.

2.3.2  Solution procedure

The governing equations for each of the above formulations have been  solved using a
finite element method.   Spatial derivative terms in the flow  equations are weighted by
unsymmetric upstream- weighting functions developed by Huyakorn and Nilkuha (1979),
while the remaining terms are handled using linear basis functions.   In  this paper, we
shall  restrict  ourselves to linear quadrilateral  elements.  In  the interest of brevity,
details of the  numerical methods  will be given for  Method 1 only.  Computational
procedures employed for  the other equation formulations were similar.

Finite element formulation. The exact solutions for the liquid  pressure heads in (2.30)
over an element are approximated by trial functions of the form
                                      4
                                                                            (2-36a)
                                     7=1

                                     4
                          *>(***)=  2>Xf,*KX*)                         (2-36b)
                                    7=1
where hwj and h0j are pressure heads for water and oil, respectively, at node 7; Nj is the
linear basis function for node 7; and £ and r; are the local coordinates.  Applying Green's
theorem and the  standard Galerkin principle to (2.30)  with upstream-weighting for
spatial terms yields
      R                           R

                         dR    +   lqnwNjdS=0       /, 7=1, 2, ...N    (2.37a)
         R          '             S
                                     34

-------
         +   f NINr7T dR  +  Uno NrdS=Q      I, J=l, 2, ...  N     (2.37b)
             J          at          J
             R                     S

where  Wj is an upstream-weighted shape function  described  by Kaluarachchi and
Parker (1989) and N is the number of nodes in the domain.  Since only diagonal  terms
in the  conductivity tensor are considered in the present analysis, the double subscript
notation  is henceforth dropped.

As in Section 2.2, we expand the matrices  in the first two integrals of (2.37) into four
submatrices by  permitting the  hydraulic  conductivity  within an  element to  vary
through the upstream-weighting function such that
                                  4
                      KP.(*i, ') = E wfal) KP                              (2-38)
                                 7=1

where  Kp1 denotes the hydraulic conductivity attributed to node /.  Previously, we
approximated Kp  as values corresponding  to 2x2 Gauss  points nearest each node.  In
the present  analysis,  we take Kp  values  corresponding exactly  to  node points.
Substituting (2.38) into (2.37) and integrating gives
                                   JO
                  4./J h*J + BIJ  -dr  ~ F»I  = °                         [2-39a)
                                         ' Fol  = 0                       (2-39b)
where

                                                                           (2.40a)
      = E f f E [^]C/ <  +  4 E IP?/ K1  }
         e=l \    /=!                  g=l           I
       = E f f E I^]e/ ^x +  4 E [Uf, Ai  )                       (2.40b)
         e=l \   7=1                 7=1         y /
                                     35

-------
  %=E  ₯                                                           (2-40c)
        e=l   4

            n   , 4
  FU>I=  ~ 5Z  ^ 52 \G\e]KL   +  boundary flux term                       (2.40d)
           Si  27=i

          n   ,    4
 FoI=-J2  f ProU [£]7-Ko -I-  boundary flux term                        (2.40e)
         e=l  ^    7=1        tf
where m and d are the width  and length of the rectangular element, respectively.
Influence coefficient matrices, [U*]f, [Uv]f, [G\je and coefficient vectors for boundary
flux terms are given by Kaluarachchi and Parker (1989).  In the subsequent analysis,
the boundary flux terms in (2.40) have been omitted for simplicity.  A mass lumping
procedure  is employed to diagonalize the mass  storage matrix in the flow equations
which is described earlier.

Treatment of nonlinearitv.  Owing to the treatment of time derivative terms in Method
1, Picard iteration cannot be used, and a Newton-Raphson iteration procedure is used
for comparison for all five methods. In residual form,  (2.40) may be written as



                     roj=  AoIJhoJ+  BIJ -$  -  FOI                   (2.41b)

Applying the Newton-Raphson procedure gives
                                    ^o/+1- hof) = - r./                  (2.42a)
                                        +*- h0f) = - r.f                  (2.42b)

where  r and r-fl  refer  to  the previous  and  current  iterations, respectively.  The
derivatives in (2.42) can be expressed as
                                     36

-------
                 ,   9AwIL          dSu.j        dFwI
             =  WL ~      + [   n       ]/   ~                             (2'43b)
                 ,   9AoIL
              =
                               IJ
                                            Q O

                                         n

where Ai = 
-------
      Table 2.4  Soil and fluid properties used in simulations.
Parameter
n
a, cm"1
5m
KSWI cm ^l"1
Pao
Pow
Pro
r)ro
*
1
3.00
0.05
0.00
20.00
3.20
1.455
O.SO
O.SO
0.40
Example
2
3.00
0.05
0.00
20.00
3.20
1.455
1.20
O.SO
0.40
3
3.00
0.05
0.00
50.00
3.20
1.455
O.SO
2.00
0.40
          Table 2.5  Total time steps during infiltration.

Example 1
1 104
2 95
3 100

2
107
9S
111
Method
3
104
95
100

4
NC
NC
101

5
103
94
99
NC =  no convergence
                             38

-------
2.3.3  Numerical results

Three example problems involving infiltration  and redistribution of organic liquids in
partially  water saturated soils  will  be investigated to  compare  the computational
efficiency, mass balance  error and robustness  of  the different equation formulations.
Assumed  fluid and  soil parameters in  the  constitutive  model  for  permeability-
saturations-capillary pressure relations are given in  Table 2.4 for the three problems.

Example 1^.  This problem involves vertical infiltration and redistribution of light oil in
a 200  cm long column initially free of organic  liquid and in  equilibrium with a  water
table at a depth  of 100  cm. For Methods 1, 2 and 4,  oil infiltration was  permitted
under  zero oil  pressure, while in Method 3, air-oil  capillary pressure was fixed at zero.
In  Method  5,  the top  boundary was simulated  by maintaining  the total  liquid
saturation,  St  at unity.   In all  simulations, infiltration was  continued until a total
volume  of 3.95  cm3  cm"2 was  absorbed.  Following  the  infiltration  period,  fluid
redistribution was allowed with zero oil or water flux at the upper boundary.

At  the lower boundary, zero oil flux and a  fixed water head of  -1-100 cm was assumed.
This was achieved directly for Methods  1, 2 and 5  by specifying zero oil flux  and  fixing
hw  at  100 cm.   For Methods 3 and 4, there is no way to directly specify the desired
boundary  conditions  so  surrogate  approaches  were employed.  For Method 3,  the
capillary  pressures were fixed at  the  lower boundary in  such a way  as to impose the
correct water head  and to force zero oil  saturation.  The air-oil  capillary head was
assigned a value hao — ha -  hcj where ha = 0 (atmospheric gas pressure) and hcj  is the
oil head  at which  S0 -» 0, which for the soil and  fluid properties in this problem is equal
to 31.25 cm. Thus, hao = -31.25 cm at the lower  boundary and the oil-water capillary
head,  how = h0 - hw, is 100 -  31.25  = 68.75 cm.  For Method  4, with dependent
variables 5^ and A0, the lower boundary conditions were specified by setting Sw =  1 and
h0 = h".  This  effectively stipulates zero oil saturation at the bottom, but does not fully
specify the water head as will be discussed later.

Spatial discretization  in  all  cases was  achieved  using  42 elements with  86  nodes.
Simulations were  started with  an  initial time step of 0.0002 h at  the beginning of
infiltration  and and 0.0001 h  at the beginning  of redistribution using a time-step
acceleration factor, F,  of 1.03 and  a maximum time-step of 0.1 h.
                                      39

-------
                 200.CO
                 150.00 -
 E
 o

 -^ 100.00 -


 '•£       3

     50.00 ^
                                               iquid
                           Water
                   O.CO ' !.	
                      0.00   0.25    0.50    0.75     1.07-

                               Saturation

Fig. 2.11  Fluid saturation distribution at the end of infiltration for Example 1.
                200.00
   150.00 -

 E
o


tf 100.00 H
              CD
             IE
                 50.00 -
                  0.00
                             N    Joto!  liquid
                              \
                                  /
                                  Water
                     0.00    0.25    0.50    0.75    1.03

                               Saturation

Fig. 2.12  Fluid saturation distribution at the end of redistribution for Example 1.
                                  40

-------
          Table 2.6  Toted iterations during infiltration.

Example
1
2
3

1
450
418
476

2
2S9
269
324
Method
3
456
423
476

4
NC
NC
298

5
299
282
324
NC =  no convergence
   Table 2.7  CPU seconds using IBM 3090 during infliltration.

Example
1
2
3

1
30
28
32

2
23
22
27
Method
3
30
29
33

4
NC
NC
25

5
26
25
29
NC = no convergence
                            41

-------
           Table 2.8  Total time steps during redistribution.
Example
1
2
3
Method
12345
2,201 2,186 2,202 NC . MB
2,202 2,186 2,208 NC MB
2,188 NC 2,187 2,187 2,187
NC = no convergence achieved
MB  =  mass balance error exceeded 10% and the solution was halted
           Table 2.9  Total iterations during redistribution.
Method
12345
1
2
3
3,218 2,922 3,483 NC MB
3,273 3,192 4,398 NC MB
4199 NC 4191 3619 4514
NC = no convergence achieved
MB = mass balance error exceeded 10% and the solution was halted
                               42

-------
 Table 2.10  CPU seconds using IBM 3090 for redistribution period.
Example
1
2
3
Method
1 2 3
211 222 222
215 249 279
270 MB 274
4 5
NO 210
NC MB
261 331
NC =  no convergence achieved during infiltration problem
MB =  mass balance error exceeded 10% and solution halted

     Table 2.11  Maximum percent mass balance error for oil.
Example
1
2
3
1
0.50
0.34
0.75
2
1.87
3.70
(d)
Method
3
0.47
(b)
0.71
4
NC
NC
0.10
5
(a)
(c)
0.07
  NC =  no convergence during infiltration
  (a)   Mass balance error of 0.6% prior to oil front reaching water table;
        thereafter mass balance error became greater than 10%
  (c)   Mass balance error of 0.1% prior to oil front reaching water table;
        thereafter mass balance error became greater than 10%
  (b)  Mass balance error of 0.34% prior to oil front reaching lower boundary;
        thereafter oil mass lost by flow through boundary
  (d) • Mass balance error gradually increasing to greater than 10%
                             43

-------
 Oil saturation distributions at the end of oil infiltration (0.095 h) and 200 h after  the
 start  of redistribution are shown in Figures 2.11 and 2.12, respectively, for Method 1.
 Results for other  methods except Method 4 were visually identical.  A comparison of
 the number of time-steps and iterations and computational time during the infiltration
 period are given in Tables 2.5 - 2.7 for all methods.  Similar information, along with  the
 maximum mass balance error during the redistribution period, are compared in Tables
 2.8-2.11.

 Methods  1 and 3  exhibited similar convergence behavior, with an average of about  4.3
 iterations/time-step during infiltration and 1.5 iterations/time-step during redistribution
 (Tables 2.5 - 2.9),  and yielded maximum mass balance errors  of  about  0.5% (Table
 2.11).   Method   5   exhibited  more  rapid  convergence   during  infiltration   (2.7
 iterations/time-step) but similar behavior during redistribution to Methods 1 and  3.  In
 contrast,  mass balance error was markedly higher at almost  2%.  Total CPU time for
 infiltration plus redistribution periods (Tables 2.7 and 2.10) within a few percent for all
 three methods.

 Method 5, the semi-diffusive scheme  in terms of St and hw, performed well until the oil
front  reached the water table, but produced large mass balance errors thereafter.  To
permit comparison with other methods prior to this  occurrence, Tables 2.8-2.11 indicate
results  for the entire redistribution  period of 200  h.   The  difficulty encountered  by-
Method 2.5 may be  due to nonlinearities caused by the boundary  conditions.  Fixing
water head (^=100 cm) and total liquid saturation  (St=l) does not  uniquely specify oil
pressure which may take on values h0 > 0.  We conjecture that for all nodes with 5f=l
this  lack  of constraint on h0,  which  is not one of the dependent  variables in  the
formulation, leads to the observed numerical errors after the oil front reaches the  water
table.

Method 4,  the semi-diffusive scheme in  terms of Sw and /i0,  experienced serious
convergence difficulty during the infiltration period.  The boundary  conditions imposed
at the bottom nodes of the column for Sw and h0 does not guarantee that hw=WQ  cm
will be  imposed on  these nodes during the current  iteration.  When hw > 0 and 5^=1,
the constitutive model gives hw > h0.  We conjecture that this lack of  constraint on  hw,
which is effectively treated as a nonlinear  coefficient of the  primary  variables, causes
instability at nodes below the water table.
                                      44

-------
Example 2..  This  case is  the same  as the previous example except that  the organic
liquid is denser than water  (Table 2.4).  Initial and boundary conditions imposed on this
problem are taken  identical to those of Example 1.  Because of the high fluid density,
oil will penetrate through the capillary fringe and eventually reach the lower boundary
of the system.  For Methods 1,  2 and 5, the zero oil-flux condition will  force  oil to
accumulate at the boundary while water outflow is permitted under constant  water
head. However, the lower boundary conditions employed for Methods 3 and 4 have the
effect of forcing zero oil saturation at the lower boundary which will result in oil loss
from  the lower boundary.  This problem cannot be avoided as there  is no way to
directly impose the desired boundary conditions for the latter methods.  We will take
these problems into consideration in the following discussion.
                   200.00
                                         \
                                          Total   liquid
                   150.00 -
                            i
                            j
                   100.00 -
                CD
                    50.00 -
                     o.oo
                                         Water
                          0.00     0.25     0.50     0.75     1.00
                                      Saturation
Fig. 2.13 Fluid saturation distribution at the end of infiltration for Example 2.
                                   45

-------
                      200.00  TTT
                      150.00 -
                   CJ
                      100.00 H
                   0
                      50.00 -^
                       0.00 -r
                                       lotol  liquid
 /
Water
                           0.00     0.25    0.50    0.75    1.00
                                     Saturation
Fig. 2.14  Fluid saturation distribution at the end of redistribution for Example 2.
Saturation distributions at the end of the infiltration period (0.07 h) and 200 h after the
start of redistribution are shown in Figures 2.13 and 2.14, respectively, for Method 1.
Oil was predicted to reach the lower boundary at about 120 h.  The total number of
time steps and iterations, CPU time and the maximum mass balance error for all five
methods are given in Tables 2.5 - 2.11 for the infiltration and redistribution periods.

In Example 2 as  in Example 1, Methods 1 and 2 both performed well, but the former
yielded much  lower  mass balance  error  (0.34% vs.  3.7%) and required 16% less
computer  time.  Method 3 yielded  mass balance accuracy similar to Method 1 up until
the time oil reached the lower boundary, after which the imposed boundary conditions
permitted oil drainage from the lower  surface.  Method 4 failed to converge from the
start of the simulation for inferred reasons discussed for Example 1. Method 5 behaved
well up until  the time the oil front reached the  water table.   Thereafter, the  mass
balance error increased from only about 0.1% to in excess of 10%  within 50 h.
                                     46

-------
Example 2i  In this example,  we investigate a situation involving a domain in which
water pressure is everywhere negative.   Column dimensions, spatial discretization and
time-integration parameters  are the same as in  Example 1.  Fluid and porous media
properties are similar to those of Example 1, except for a higher organic liquid viscosity
and higher hydraulic conductivity  (Table 2.4).  Initial conditions correspond in this case
to equilibrium with a water table  at the bottom of the column (200 cm depth) and the
water  head is  fixed  throughout  the simulation at  the  lower boundary at  hw=0 for
Methods 1, 2 and 5.   For Methods 3  and 4, the  lower boundary conditions are  as
described for Example 1 — corresponding to the stipulation of zero oil saturation. The
results  of  Method 1  at  the  end  of  infiltration  and  at  200  h after  the start  of
redistribution are  shown in Figures 2.15 and 2.16, respectively.  The  total number of
time-steps  and iterations, CPU time and the maximum mass balance error are given in
Tables 2.5-2.11.

For the present example, Method 1 yielded results with somewhat greater but still
acceptable mass balance error (0.75%) compared with the previous problems reflecting a
higher degree of nonlinearity in this problem associated perhaps  with  the  lower initial
water  content  and/or greater mobility contrast between oil  and water.    Method  2
succumbed entirely to these difficulties  and  eventually failed to converge after the mass
balance during redistribtion grew above 10%.  Method 3  yielded nearly identical results
in terms of accuracy  and computational effort to Method 1.  Methods 4 and 5, which
behaved miserably for  Examples  1  and 2,  now  exhibit  remarkably good  results  with
mass balance errors 7-10 times lower than Method 1.  Method 4 yields the most efficient
solution for this problem both during infiltration and  redistribution,  while Method 5 was
somewhat more computationally intensive.
2.4  Development of an Adaptive Solution Method

In most practical problems, large changes in fluid pressures and saturations do not occur
throughout the spatial domain at a given time step.  Computational effort is thereby
inefficiently spent solving equations in areas where little activity occurs  rather  than
concentrating effort in the  more  active zones.  We  discuss in this  section  a new
"adaptive  solution  domain"  (ASD)  algorithm  designed  to  reduce  unnecessary
computational effort.
                                      47

-------
                200.00 -r
                       j     Water


                 0.00  | I I : I : I I I I | I I i , I I I II I I M I I II I I | i I i , I I i I I
                     0.00    0.25    0.50    0.75    1.00

                               Saturation


 Figure 2.15. Fluid saturation distributions at the end of infiltration foi Example 3.

               200.00  -TT
T
                          \
                           \
                150.00  H
             E
            o
                100.00  -
             CD
                 50.00 -
   \   _  ,  ,  ..   .  .
   "• 'oial  liquid
                  0.00 	I I I I i I I I I | . I i i i I I I i i i .	I | I I M I M I I

                      0.00    0.25    0.50     0.75    1.00

                                Saturation


Figure 2.16. Fluid saturation distributions at the end of redistribution for Example 3.
                               48

-------
 2-4-1  Description of ASD algorithm

 In the ASD method, elements are categorized as either "active" or "inactive" at a given
 iteration.  If the element is classified as active, then it is included in the global matrix
 assembly,  otherwise it is excluded.  The criteria for changing a node from inactive to
 active is based on the changes in water pressure head, hw, oil pressure head,  hoi oil
 saturation, S0, and water saturation,  Sw at connected nodes from the last converged
 time step  to the current iteration.  The criteria for switching an active  element  to an
 inactive  status is based  on the  difference in hw,  h0 Sw and 50 at the connected  nodes
 between the converged values at the previous and the current time step. The specific
 criteria for changing the classification of an element is as follows:

 Inactive -»  active if any nodes meet the conditions:
or
or
or
or
                              Ahw
                              Ah0  > Tj
                               ASW > r2
                              AS0  > T,
where Ahw,  Ah0>  ASW and A50 are the changes  in  hw,  /i0, Sw  and 50, respectively,
between the previous converged time step and the current iteration.
Active -» inactive if all nodes meet the conditions:
                              Ahw
                       and
                              Ah0
                       and
                       and
where Ahw, Ah0i ASW and A50 are the changes in hw, h0, Sw and 50 refer to the changes
between the converged values at the end of the current and the previous time step, and
£ hw is the cumulative change  in hw from the last time the node met the criterion to
the present time.

To ensure  an accurate solution, a bias is imposed towards maintaining active  elements
by making T4 < rl  and rs < r2.  Preliminary simulations for a number of cases indicated
                                      49

-------
good  results  using  r3=2   cm  and   r4=r1/lQO  subject  to  the  constraint   that
0.001 < r4 < 0.0001 cm and r5=r2/10 subject to the constraint that 0.0001 < rs < 0.00001.
The foregoing values will be employed in the present study.  To avoid erratic switching
between active and  inactive subdomains, an  active element is labelled inactive  only
after convergence conditions are satisfied for three consecutive time-steps. This allows a
smooth  transition  between  active  and  inactive elements  while  achieving   high
computational efficiency.

It may be noted that complete elimination of a node from the global matrix only occurs
if all elements associated with the node are classified as inactive.  Equations for nodes
on the active-inactive frontier will  have contributions only from active elements, and
the  boundary will shift  naturally as imbibition  or  drainage "fronts" pass through the
domain.

2.4-2  Numerical Results

Two examples will  be  described to illustrate the  ASD method  and  to compare  its
performance with a conventional  full  domain  finite element  approach.   The  first
example involves infiltration and  redistribution of oil in a one-dimensional soil column
and the second example involves a two-dimensional domain.

Example  1^. This problem involves  vertical oil infiltration and redistribution into a 225
cm long column initially free of hydrocarbon and in  equilibrium with a water table at a
depth of 200 cm. A  dense oil was  added at the top of the column under a constant head
of ha = 0 over a period  of 0.14 h  until the cumulative oil infiltration reached 6.92  cm.
At the end  of the infiltration period, the upper boundary condition for oil was specified
as zero flux.  Fluid  redistribution was permitted for a period of 250 h after the end of
the  infiltration stage.   A  zero flux condition  for water was imposed at the upper
boundary at  all times.   At the  lower  boundary,  water  pressure was  maintained
continuously at hw = 25 cm, and zero oil flux was  imposed.   Lateral boundaries were
zero flux for both fluids. Spatial  discretization for  the problem was achieved using 45
rectangular  elements each 5 cm wide and spaced 5 cm apart in the vertical  direction
yielding a total of 92 nodes.  All simulations  were  started with an initial time-step of
0.0002 h, and a time acceleration  factor of F = 1.03 was used to increase the time-step
size up to a maximum time-step of 0.5  h.  Soil and fluid properties are given in Table
2.12 which correspond to the constitutive model described in Chapter 3.

                                       50

-------
            Table 2.12 Soil and fluid properties for example problems.
Parameter
4>
n
Q, cm
Sm
Ksw, cm h~
Pao
ft ow
Pro
r]ro
1
0.4
3.0
0.05
0.0
20.0
3.2
1.455
1.2
O.S
Example
2
0.4
2.7
0.02
0.0
2.1
3.2
1.455
O.S
2.0
Table 2.13 CPU time (seconds) and mass balance error on IBM 3090 for Example 1.
*
TI TI CPU time (s) MBE
(cm) (-) infiltration redistribution (%)
Conventional FEM 30 127
0.002 0.00005 9 92
0.010 0.00010 6 77
0.050 0.00050 6 76
0.100 0.00050 6 73
0.33
0.66
0.37
0.43
0.46
                                   51

-------
                    200.00 -
                                                    ASD  o
                                                   CFtM 	
                                       Total  liquid
                      o.oo
                          o.oo
0.25    0.50
  Saturation
i.oo
 Figure 2.17 Water and total liquid saturation distributions at the end of redistribution
             for CFEM and ASD method with ^=0.01 cm for Example 1.
Simulations were carried out using the  ASD method with various values of r^  and r2
and using a conventional full-domain finite element solution (CFEM) based on the same
numerical formulation employed for the ASD solution but with no element  exclusion.
Simulations were performed in two stages corresponding to the period of oil infiltration
and the subsequent redistribution period. Table 2.13 shows the CPU time required for
infiltration and redistribution periods using different ASD tolerances and for the CFEM.
The  maximum  percentage  mass balance  error for the oil  phase (MBE) during  the
redistribution  period  is also given  in Table 2.13. Water and total liquid saturation
distributions predicted by the ASD method for the r^= 0.01 cm case and the CFEM are
virtually identical (Figure 2.17) and the other ASD results were visually indiscernible.
Small differences among the various solutions were evident  in the mass balance error
(Table 2.12).  The ASD cases with rl  > 0.01 cm yielded nearly the same mass balance
error as  the CFEM, while the ASD case  with  rl = 0.002  cm exhibited substantially
higher, although still tolerable, mass balance  error.  We attribute the poorer behavior of
                                     52

-------
the latter case to erratic changes in the number of active  elements  for consecutive
iterations of the same time-step which result when imposed  ASD tolerances  undercut
the precision in the numerical solution.   At this point, the ASD criteria are simply
driven by noise in the solution.

Marked savings in CPU time are obtained with the ASD method (Table 2.13).  For the
TI >  0.01 cm cases, the ASD method yields an overall speed-up factor (CPU  time for
CFEM/CPU time for ASD method) of 5 relative to the CFEM for the infiltration
period.   For the entire redistribution period, the overall  speed-up factor  diminishes to
1.7.  The large speed-up factor for the infiltration period reflects the fact that only a
small fraction of the total elements are active early in the simulation. As  redistribution
proceeds, the number  of active elements increases as the oil front advances down the
column.
                   ci I
                 •II-
              6.3S:
        E
          0.00
                                            I	I
                                                                    2.55
                                                                 7.5
                                       X,  m
                        Figure 2.18 Geometry of Example 2.
                                      53

-------
Example  2^  A two-dimensional domain is considered  in  this example involving  a
vertical  slice through an unconfmed aquifer which is 7.50  m wide and 6.38 m in height
(Figure  2.18).  The system initially contains  no oil and is in equilibrium with a water
table at  an elevation of 2.55 m from the bottom boundary. A low density oil is added on
a 1.1 m  strip (A-B) at a constant pressure of  h0 — -0.1 m for a period of 3.43 d yielding
a cumulative oil  infiltration of 0.32 m3 m"1. Thereafter, A-B is a zero-flux boundary for
oil.  All  other boundaries are zero-flux for oil  at all times.  Redistribution  was permitted
out for a period  of 45 days following the end of infiltration. Water pressure is fixed at
the initial condition on C-D and all other boundaries are zero-flux for water at all times.
A total  of 294 rectangular elements  was used  to  discretize the domain yielding  330
nodes.  All simulations were started with an initial time step of 0.002 d, and a factor of
F=1.03 was used to increase the  time step size up to a maximum time-step of 0.2 d.
Information on soil and fluid properties for the problem is given in Table 2.12.

Simulations were carried out with the ASD method using different tolerances and using
CFEM.  Comparisons  of water and total liquid  saturation distributions on  vertical
sections  at x = 0 and 1.2 m are shown in  Figures 2.19 and  2.20.  It is observed that at z
= 0, the CFEM results slightly undershoot the ASD solution and slightly overshoot at x
=  1.2 m.   This reflects a slight oscillation in the  CFEM solution  which could be
observed in the y-direction.  Such behavior was not evident in the ASD results.  The
superior  stability of the  ASD method  may also be evidenced  by  the more rapid
convergence behavior of the nonlinear iterations which averaged 4.0 iterations per time-
step for  the CFEM and 3.5 for the ASD method for the TI = 0.1 cm case.  The greater
accuracy of the ASD method is confirmed by the consistently lower mass balance error
for all ASD solutions (Table 2.14). As in Example 1, the  lowest mass balance error for
the ASD method was  obtained for the case with rl = 0.1 cm with smaller and higher
tolerances resulting in somewhat greater error.

Total CPU times for infiltration and  redistribution periods  showed even more marked
contrasts between CFEM and ASD results than were observed for the one-dimensional
problem. Overall speed-up factors for the present case were 6 for the infiltration period
and  2.5  for the redistribution  period.   This greater  speed-up  reflects  the  larger
proportion of the domain which can on average be excluded from the active domain as
dimensionality increases. As in the first example, as redistribution proceeds, the fraction
of active elements gradually increases. At sufficiently long times, the active elements
will diminish as steady state is  approached.

                                      54

-------
 Table 2.14  CPU time (seconds) and mass balance error on IBM 3090 for Example 2.
T\
(cm)
T2
(-)
Conventional FEM
0.05
0.10
0.15
0.0001
0.0001
0.0001
CPU
infiltration
511
86
84
84
time (s)
redistribution
871
371
349
323
MBE
(%)
0.68
0.39
0.12
0.14
   600.00 -E

   500.00 i

   400.00 i

J= 300.00 ^

   200.00 I

   100.00 -E
  CD
       0.00
                  Water
                    ASD

                   CFEM
                                    Total  liauid
                                                    x=O.Cb
           0.00
                   0.20
0.40
0.60
0.80
1.00
                             Saturation
Figure 2.19 Water and total liquid saturation distributions at the end of redistribution
       at z=0 m for CFEM and ASD method with Ta=0.1 cm for Example 2.
                                55

-------
CD
     600.00  -E


     500.00  '-.


     400.00  \


     300.00  \


     200.00  \


     100.00  \
        o.oo
               Water
                ASD  o

                C'EX 	
           o.oo'    '0.20
                                 otal  liquid
                            0.40      0.60      0.80
                           Saturation
1.00
Figure 2.20 Water and total liquid saturation distributions at the end of redistribution
      at i=1.2 m for CFEM and ASD method with Tj=0.1 cm for Example 2.
                               56

-------
             3.  HYSTERESIS IN THREE-PHASE FLOW RELATIONS

3.1  Evaluation of General Hysteretic Formulation

S.I.I Model description

In order to model three-phase flow, functional relationships between fluid  pressures,
saturations and permeabilities must be known. In most cases, the NAPL is less wetting
to the solid than water, so that water will occupy the pore space in immediate contact
with  the solid, although oil-wet  or  mixed  oil-water-wet  systems may occur  —  for
example, in soils with high organic matter contents or in cases  where mineral surfaces
exhibit  natural organic coatings. We will  assume in the following that water (w) is the
wetting phase, organic liquid (o) is the intermediate wetting phase, and air (a) is the
nonwetting phase. Hysteresis will occur in three-phase systems due to nonwetting fluid
entrapment, contact angle hysteresis, ink bottle effects and other  phenomena.

Fluid entrapment. To model  hysteresis  in three-phase  air-NAPL-water systems,  the
concept of apparent saturation may be used. Apparent water saturation in a two-phase
air-water system is defined by

                               IT  = sr +  5™                             (3.1)

•where 5^"" is the effective water saturation and  Sataw is  the effective  trapped air
saturation of  a two-phase air-water system.  Effective water saturation is defined by
Sw = (Sw — STW}I(\ — 5ru;), and effective nonwetting phase saturations (i.e., air) have the
form 5n = Sn/(l — Srw) where 5n is the actual nonwetting phase saturation.

For three fluid phase systems, apparent water and total  liquid saturations are defined.
Apparent water saturation is defined as the sum of the  effective water saturation and
effective trapped air and oil saturations, and apparent total liquid saturation is defined
as the sum of the effective water and  oil saturations  and the effective trapped air
saturation. The apparent  water saturation can be used to determine the flow channel
radii where NAPL-water interfaces reside. The apparent total liquid saturation can be
used to determine the flow channel radii where air-NAPL interfaces reside. Apparent
water and total liquid saturations are given as
                                      57

-------
                             Sw  = Sw + Sot +  Salu,                          (3.2a)

                            St = Sw  + S0 + Sat                             (3.2b)

where Sgi is the effective trapped oil saturation (i.e., in water),  Saiw is  the  effective
saturation of air occluded by water, and 5B< is the effective total trapped air saturation
(i.e.,  air that is occluded either by oil or water).  The effective trapped saturations are
defined analogous to that for the two-phase case.  Using apparent saturations in lieu  of
effective or actual saturations to model relationships among fluid pressures, saturations,
and permeabilities has physical merit. Apparent  saturations can be used  to determine
fluid  flow channel dimensions that separate continuous fluid phases that are immiscible
with  each other. This  modeling approach will yield a more accurate description of fluid
distributions  in  porous media  for  arbitrary   saturation  paths.  Correlating  fluid
distributions to flow channel radii is important when predicting relative permeabilities
from  S-P relations via a pore size  distribution model.  Note that the model assumes the
entrapped fluids are discontinuous and immobile.

Fluids that wet other fluid surfaces can become trapped by the fluid  phase they  wet.
That is, fluids with  lesser  wettability can become occluded  by fluids  with greater
wettability. For example, air may become trapped by oil or water, and oil can become
trapped by water. The model assumes that water cannot become trapped  by oil or air,
and that oil  cannot  be trapped  by air.  Entrapment  of fluid  occurs as fluid-fluid
interfaces advance into larger flow channels, which implies  that  one fluid  phase  is
displacing another.

Trapping of  a fluid   phase can occur  only in  flow  channels  larger  than those
corresponding to the smallest or narrowest  flow channels where the fluid-fluid interfaces
resided.  For example, air can be trapped by advancing air-water interfaces in a two-
phase fluid system  in flow  channels  larger than those corresponding to the  historic
minimum apparent  water saturation. Historic minimum  saturations  are  the  smallest
fluid  contents for a given saturation  path history.  For  two-phase systems, this value
would be associated  with the lowest saturation obtained on the main water  drainage S-P
branch. Note that apparent and effective saturations are  equal for the main drainage  5-
P branch because this path has no trapped fluids. For  all other saturation paths, the
                                      58

-------
model assumes some trapped  fluid exists  and the  apparent water or  total liquid
saturations  are greater than  effective water or total  liquid saturations for the same
actual water or total liquid content, respectively.

The amount of trapped air in  a two-phase system, therefore, is zero at  the historic
minimum apparent water saturation and is at a maximum when the porous medium is
apparently water saturated, which is at an air-water  capillary pressure of zero on an
imbibition path. An  algorithm proposed by  Land (1968)  is  employed  to  estimate
trapped  fluid  saturations of  two-phase fluid systems that occur at  a  zero  capillary
pressure on  imbibition  saturation paths. Land's algorithm  predicts the  amount  of
nonwetting fluid that becomes trapped in  pore spaces when the wetting fluid saturation
increases from some point along the main drainage  branch to  an apparently saturated
condition (i.e., saturated with respect to wetting fluid  and entrapped nonwetting fluid)
indicated by a capillary pressure of zero. The effective  wetting fluid content at the
saturation path reversal from the main drainage branch to an imbibition scanning path
is given by the symbol  Sv. Land's empirical relationship may be written as

                             V  = 1 + R (1 J\^}                       (3.3a)

in which
                                 Rv =  7TU -  l                            (3-3b)
                                          tr
where 5t>" and ISiji are the maximum effective trapped nonwetting fluid saturations  or
the effective residual fluid  saturations corresponding to the imbibition  scanning path
beginning from ^S»  and to the main imbibition branch, respectively. This algorithm
predicts that as A5';—»0, the effective residual fluid saturation for that  scanning path
(Sjji)  approaches  the  maximum  effective entrapped  nonwetting  fluid saturation
associated with the main imbibition  path (75-r'').  In other words, as more pore space
becomes occupied by a continuous nonwetting phase as wetting fluid drains, the amount
of nonwetting fluid that potentially can become trapped will be greater because of the
larger volume of flow channels that can trap nonwetting fluid. Note that S^  = ISij>
when A5^ =  0 and,  by definition, A5;" = A5;-*'. Note that we use the term irreducible
saturation to refer to the minimum wetting fluid content associated with wetting fluid
drainage,  while residual saturation refers to the  nonwetting fluid saturation at  zero
capillary pressure. The irreducible saturation thus represents a continuous  fluid phase,
                                      59

-------
 albeit that only thin films may be  present, and  the  residual saturation  represents
 trapped, discontinuous nonwetting fluid.

 The algorithm proposed by Land (1968) places bounds  on the amount of nonwetting
 fluid that can become trapped in a porous medium with  two immiscible fluids. That is,
 at a wetting fluid saturation of A5-^, there is no trapped nonwetting fluid, and at an
 apparent wetting fluid saturation of 1 (i.e., at a capillary pressure of zero), the effective
 entrapped  nonwetting fluid saturation is  equal to  SlV«. To interpolate between these two
 end-points, it  is assumed  that all  pore size  classes will entrap nonwetting fluid in
 proportion to their volumes. Accordingly, the amount of entrapped nonwetting fluid is
 predicted to vary linearly as
                             _   / 5..; _ 45 .«; \
                        *       j                                              (3'4)

where 5,^ is given by (3.3). Note that for apparent saturations equal to A5^, the main
drainage path  is  assumed, and  for apparent saturations  greater than A5-^', scanning
saturation paths are assumed. When modeling S-P behavior, A5.v needs to be updated
every time the main drainage path is followed. Implicit in the modeling scheme is that
as interfaces between continuous nonwetting and  wetting fluids advance into larger-
sized flow channels,  nonwetting fluid becomes trapped  (i.e.,  discontinuous) by  the
wetting fluid as it displaces nonwetting fluid into  larger pore cavities. The process is
assumed to be reversible during drainage.

Algorithms to predict  nonwetting fluid entrapment in two-phase systems  are employed
to estimate fluid entrapment in three-phase systems. This is because in three-phase fluid
systems the entrapment of nonwetting fluids are essentially two-phase fluid processes.
Air in a three-phase system will  be trapped by  advancing  air-oil interfaces and oil  will
be trapped by advancing oil- water interfaces.  However, because the fluid system is
initially an air-water system, air entrapment by advancing air-water interfaces must
also be considered.

Distinguishing  between air trapped  by air-water interfaces and air trapped by air-oil
interfaces is done by tracking the  historic minimum apparent  total liquid  saturation
(Stmin) and *SW™.  If 5tnin >  *3W™, then air is trapped by air-water interfaces in flow
channels larger than those corresponding to *SW*W but smaller than those corresponding
                                      60

-------
to  5/71"1,  and air is  trapped by  air-oil interfaces in flow channels  larger than  those
corresponding to 3tmin. The historic minimum apparent total liquid saturation (3t""B) is
used to determine the smallest flow  channel radii where  air-oil interfaces resided.  If
§min < *Swaw,  then all air is  entrapped  by advancing air-oil interfaces.  Once a two-
phase system becomes a  three-phase system, the value of *Swtw becomes fixed (i.e.,
invariant). When the interfaces between the continuous air and oil phases recede into
flow  channels  smaller than  what  air-water  interfaces  historically occupied  (i.e.,
51T7"n
-------
 and the effective entrapped air saturation occluded by water for a given saturation path

 history can be predicted from the following relations:
               ? aw
               Jw
for 5



                   _      _   / C mi'n  	 Ac
                   C   — C   I  °<        ^u
                                                                 _

                                                              -1
for 5 5,, > *Swau'




                                "rt        ~C>   I  *^1H ^~   *^ll
                                               1     A C au'
                                               •I  —   Ou
                                                  _ C tntn
                                   "       ar° V  1  _  C miti
                                              \  J.  —  O<







for 5^ > Stmin




                                 —       —   /  C   _  C mi
                                 C    	  C   ( '•'ui    "t
                                 ^S/fim  ^~  ^/I^A 1 ^^^^^^^^^^^^^^™"
for 5^ >  5,"11'"


                   _       _    / ^min _ A^ ou; \      _   / ^  _ ^ min \

                                                                      -              (3-6b)
for 5,  <*SW™



                                          Stttw =  0                                    (3.6d)



for Stmin <
for 5W < S™n



                                          S*t*  =  0                                   (3.6g)




                                           62

-------
where Sat is the effective total entrapped air saturation, Saiw is the effective entrapped
                                         =     s=
air saturation in the aqueous phase, and  5, and Sw are the current apparent  total liquid
and water saturations, respectively. Note that the effective saturation of air trapped in
oil (5a<0) can be obtained from Saio  =  Sai — 5alu, for any saturation path history.

In a three-phase fluid system, displacement of oil by advancing oil-water interfaces will
result in  discontinuous blobs or ganglia of oil being formed in flow channels  larger than
those  corresponding to the minimum apparent water saturation (Swm>n).  Within the
trapped oil phase, there may be entrapped air  that resulted from air-water or air-oil
interfaces. To account for concomitant entrapment of air and oil by oil-water interfaces.
an effective total trapped oil saturation (Sgii) is defined as

                              Sott  =  Sot + Sotw +  Soio                          (3.7)

where  Soi is the  effective trapped  oil  saturation, Sotw  is the  effective  trapped air
saturation contained within the trapped  oil that resulted from air-water interfaces, and
Soto is the effective entrapped  air saturation  contained  within  the trapped  oil that
resulted from air-oil interfaces.
The maximum amount of trapped oil for a given saturation path history is estimated by
                             -     /      1 - ? "»«'»      \
                             <>  —
                             s  ~
                                        i  D   fl _ C miiA  /                        v    '
                                       T •"•ouA.1   &w   ) /

where 5or is the effective residual total oil  saturation, Swmin is  the historic minimum
apparent three-phase water saturation, and  Rgw is a calibration term  defined in (3.3).
5wTOI'n  represents  the  smallest flow  channel  radii  where  oil-water  interfaces  have
occurred.

Effective total trapped oil saturations  between 5^  =  Swmin and  5^, =  1 are estimated
by
   _  / C  _  C tnin
_  C  f  Jw    &w                              (n Q\
-  \A       fe   .   I                          (3.9)
      \
                                             _   W|.n
                                             —
                                        63

-------
where 5^ references  the current location of the continuous oil-water  interfaces in a
porous medium.  The amount of trapped air  contained in Sott  (i.e., Soio  and 50 S *SU™

                             ~        SaroSor(5w- $r»)
                                 =  a -  S^KI  - 5,m'n)
for §„,*» > *SW™
                                     5~    C  IC min    C min^
                                      5arw  Jor \J1       Jw   I                     /O inu\
                             olw =  	s	'•	A^	                     (O.1UDJ

for I,,,"1"1
  mtn  ^
'U)    i
for 5tt > 51rai'n and Stmin <  ^Sw
                                         5C  / C mtn 	 Ac nu'^
                                      Soru1 Jor \Jt       Jw  )                     I*
                                ^~  • • i i i^•»   .1-^^^^^^^^^^-.                    I
                             0TW     /^    7r   ' \/*    A7r   \                     ^

fnr ^  ^  C min aT1H C min ^ A C ou/
lOl Jw v  O^    ollU »jj   «^  iJ^


                                       50<0  =  0                                (3.10d)


for Swm>n  > A5u,au'


                           ^otw =     ^C°minVl _ Ao aw\                    (3.10e)

for Swmtn  <   Sv/
-------
                                      5otu, =  0                               (3.10h)

for 5,,, <  5,""'" and Stmi'n < A5u,au'
                                      S,to  =  0                               (

                                      Sot*  =  0                               (3.10J)

Concomitant entrapment of air and oil needs to be considered for mass balance purposes
as oil-water interfaces move through a porous medium. The historic saturations need to
be  continually updated so  that  current  apparent  saturations  are  not  less than  their
historic values.

Scaled saturation-pressure function. Three-phase capillary pressure curves are estimated
from two-phase relation using an extension of the scaling proposed by Leverett (1941) as

                                 3£WOUAJ =  S*(fc*)                      (3.Ha)
where 5 £,J/ is the apparent water saturation in the three fluid phase system, S \IJ is the
total liquid  saturation  in  the three-phase system, S*(h*}  is the  S-P function for a
reference two fluid phase system, and flao and (3OW  are fluid-dependent scaling factors.
Capillary heads are defined by

                                   haw =  ha -  hw                        (3.12a)
                                   hao =  ha-  h,                         (3.12b)
                                   h,w =  hc -  hw.                        (3.12c)

If  we  take  a  pristine   two-phase  air-water  system  as  reference   such  that
5*(/i*) = 5 ™(haw), then the scaling factors are given by

                                     &,o  =  TT*                              (3.13a)
                                             v ao                              x      '

                                    Pow  =  -^                              (3.13b)
                                       65

-------
where 0 for the three-phase system,  it may be inferred from
(3.11)  that the  air-water capillary pressure relations for  a  contaminated  system is
described by
                               3 I"(l3'aw haj =  S*(h*)                         (3.14a)

                                     ft* = %                               (3.14b)
where a'aw is the surface tension of water with dissolved components of the oil present.
Organic contaminants  generally decrease the surface tension of water so that 0'aw  > 1,
which indicates that the capillary pressure in an air-water system with dissolved organic
contaminants will be less than that in the pristine system. It may also be shown that
                                      ao
If /3'aw ^ 1, then Sw(hatv) will exhibit a discontinuity when a transition is made from an
oil-free system to one with oil present. The oil pressure at which this transition occurs
may be found by letting St—>Sw, indicating the necessary condition to classify a location
as oil-free is

                                      h0  <  V                              (3.16a)

                                h? =  ^A +  JoA,                         (3 16b)
                                         POW  '  "ao

where hcj is a critical pressure head at the transition from a two-phase air-water system
to a three-phase air-oil-water system.
                                       66

-------
Apparent  saturation-capillary  pressure  hysteresis  is  modeled by describing main
drainage and imbibition S-P relations using the van Genuchten function, while arbitrary
scanning  paths are described by scaling the  main S-P branches through appropriate
reversal points. The main drainage and imbibition branches are parameterized by
                                (A*) =  (l  4- (D«haw)n)'                   (3.17a)

                               *(O =  l  4- (7a/Un~m                  (3.17b)
where ^a, Ja, and n are curve shape parameters with m = 1 - 1/n, and superscript D
and / denote  main  drainage and imbibition curves, respectively.  To  predict  S*(h*)
scanning  paths, the main  drainage and imbibition branches are scaled to pass through
the  appropriate  saturation path  reversal points  such that closure of scanning  5*(/i*)
paths is enforced. The main imbibition branch is scaled to pass through  reversal  points
to give the current imbibition scanning path as
                                                   +  IDS*                    (3.18)
where h* and 5* are the current scaled capillary head (0^) and  apparent saturation
(Swaw, Sw or 5,), (DIS*,DIh*) is the most recent drainage to imbibition reversal point and
(IDS*,IDh*) is the  most recent imbibition  to  drainage reversal, and 75*(/i*),  75*(£l//i*)
and IS*(IDh*) are  apparent saturations of  the main imbibition branch that correspond
to the specified scaled capillary heads. Note that when using this function to predict S-P
relations of a primary imbibition scanning path (i.e., an imbibition  path that originates
from the main drainage branch), IDS* and 7S*(7I>h*) are equal to 1.

For drainage S*(h*)  scanning  paths,  an analogous procedure is  followed  to  scale the
main  drainage branch to pass through the reversal points.  The interpolation formula for
an arbitrary drying S*(h*) scanning path is

                                                        - DiA
                                       67

-------
35*(7ZV) - rS*(c/fc*)j
                                                                            (3.19)
where (IDS*,IDh*) is the most recent imbibition to drainage reversal and (DIS*,DIh*) is
the most recent drainage to imbibition reversal, and other symbols follow the notation
of the imbibition formula.
Relative permeability  relations. Three-phase  relative  permeability functions may be
derived using the theory of Mualem (1976) after making suitable corrections to account
for  effects  of trapped  fluids.  For  the case of the  water phase relative permeability,
trapped oil and air trapped within the water phase will occupy some  of the pore space
which  would otherwise be filled with water, thus displacing water into slightly larger
pores.  For  the  air and oil phases, dispacement  of the fluid interfaces within the pore
space will affect the hydraulic radius of the phase, while the portion of the phase which
is trapped  is assumed to have no contribution to  the phase  permeability. Correcting
Mualem's integral to account for these effects yields for the  water,  oil and air  phase
permeabilities
            k   -
            "Vu>  —
7
J 0

»dSw
h*(S*)

"ot
J 0 h*(S*)
f1 
-------
        • 1
        0
                                                 dS*
                                                                            (3.20c)
Using the van Genuchten model for S*(h*) and  the expressions  for  apparent and
trapped  saturations described  above,  closed form expressions for phase  permeabilities
may be derived. For the case of no trapped fluids, the expressions reduce to the simple
form  derived earlier by Parker et al. (1987)
          krw = sj/2 [ i - (i-

          kro = (5t - 5
 ]2                                     (3.21a)

)m - (l-5t1/rn)m ]2                      (3.21b)

2m.                                     (3.21c)
Results for the  general case in which fluids are trapped following arbitrary saturation
path histories are given by Lenhard and Parker (1987).
3.1.2 Numerical simulation of hysteretic flow

Problem description. The hysteretic three-phase hysteretic k-S-P model described above
was  incorporated into an upstream-weighted finite element multiphase flow code, and
the numerical  model  was employed to simulate a pulse injection of NAPL  into the
vadose  zone under conditions  of  a fluctuating water  table  to  assess the effects  of
hysteresis.

Parameters used in the simulations (Table 3.1) correspond to a sandy porous medium
and  a NAPL less dense and more  viscous than water. A one  dimensional vertical soil
section is modeled which has a depth of 150 cm discretized in 1 cm intervals.  Initially,
the system is assumed to be free of oil and in static equilibrium with a water table  at
the lower boundary. Initial fluid saturations are taken to correspond to main  drainage
branches of the S-P relations. Water phase pressure at the lower boundary as a function
of time is  shown in Figure 3.1. During the first 10 hours, the water table rises as the
                                      69

-------
water pressure head  at the lower boundary increases from 0 to 50 cm. At 10 hours, a
pulse of NAPL is applied under a constant oil pressure head of 3 cm water-equivalent
height at the upper  surface.  NAPL is applied until  7.5  cm3 cm'2 infiltrates into the
porous medium after which a zero flux upper  boundary condition  is imposed for NAPL.
The upper boundary condition  for water and the lower boundary condition for  NAPL
are at all times zero flux conditions (with the water head of 50 cm at the bottom, 5-P
relations preclude NAPL from reaching the lower boundary).  After NAPL infiltration,
a  period of redistribution follows during which  water head at the lower boundary is
fixed at 50 cm (A-B of Figure 3.1).  Subsequently, the lower boundary water head again
increases (B-C), decreases (D-E), increases  (F-G) and finally returns to the 50 cm level
(H-I) with constant pressure periods between period of changing head (Figure 3.1). The
total simulation time is 500 hours.

Three different simulations of the  above physical scenario  were carried  out  using
variations of the constitutive model.  Hysteretic analyses were  carried  out using the full
hysteretic  model described above which accounts for reversible  hysteresis effects and
nonwetting fluid entrapment. Additionally, analyses were conducted using a simplified
model which considers  fluid entrapment  effects  only.   Since  the scaled  saturation
function becomes  single-valued for the latter case, storage requirements axe reduced
(scaled reversal saturations are  not relevant)  and computational  requirements  are
slightly diminished.  Finally, simulations were also  carried out with hysteresis entirely
disregarded, for which case the present model reduces to the model described by Parker
et al. (1987).

Simulation results.  To assess the system behavior for full  hysteretic, trapped fluid only
and nonhysteretic simulation cases, we first consider the temporal system response at a
depth 50 cm below  the  soil surface. Water, total NAPL and free NAPL saturation
histories at the 50 cm depth are illustrated  for the three  cases  in Figures 3.2  - 3.4.
Water saturations predicted for the trapped fluid only case are similar to those of the
full hysteresis analysis, while the nonhysteretic analysis indicates substantially  higher
water saturations since air entrapment is disregarded (Figure 3.2).
                                      70

-------
                 Table 3.1   Parameters Employed in Numerical Simulations
                                      Parameter
                                                               Value
                                      Porous Mciliiini Clmrticierimk's
                                'o                            0.10 cm"1
                                Ja                            0.05 cm"'
                                S,,,                           0
                                n                            2.0
                                •S";                          0.25
                                '5';'                          0.10
                                's;"                          0.25
                                *                            0.4
                                Saturated water phase            50 cm h"'
                                 conductivity

                                            Fluid Properties
                                A,,,                           -25
                                A.,.                           '-80
                                NAPL specific gravity            0.8
                                 (P.,'PJ
                                NAPL to water viscosity          2.0
                                 ratio (TJ./T;,,.)
                                    100     ZOO     300     400     500

                                         TIME   Chours)
Fig. 3.1    Variation in water head at lower boundary for simulations.
                                              71

-------
                    0. 9
                              100     200     300     400



                                  TIME  Choui-s)
                                                          500
Fig. 3.2   Water saturation histories at 50 cm depth.
                              100    200    300    400




                                  TIME  Chour-s)
                                                          soo
Fig. 3.3   Total NAPL saturation histories at 50 cm depth.
                                        72

-------
                                     100     200     300    400


                                         TIME  
-------
                         Q_
                         U
                            25 -
                            50 -
                            75 -
                           100 -
                           125 -
                                                 t. -  100 n
                                         FULL
                                       HYSTERESIS
                                         NDNHYSTERETJC
                                 /   FLUID
                                '/ ENTRAPMENT
                                     ONLY
                           150
                              0. 0  O. 2  0. 4   0. 6   0. 8  1.0

                               FREE  NAPL SATURATION
 Fig. 3.6   Free NAPL saturation distributions at 100 hours.
                           25 -
                           SO -
                           75 -
                        D.
                        O 1DD
                          125 -
                          150
                                                t - 100 h
                                     FLUID
                                   ENTRAPMENT
                                      ONLY
                            O. O   O. 2  0. 4  0. 6   0. 6  1.0

                                 WATER SATURATION
Fig. 3.7   Water saturation distributions at 100 hours.
                                          74

-------
                      Q.
                      LU
                      D
300 -
                        125
                        150
                          0. 0   20. 0   40. O   60. O   80. 0

                         NAPL-WATER  CAPILLARY  HEAD
 Fig. 3.8   NAPL-water capillary head distributions at 100 hours.
                         so -
                         75 -
                     £L

                     D 100
                        150
                                    .  \   NONHYSTERETIC
                              FULL

                             HYSTERESIS
                                       FLUID

                                     ENTRAPMENT
                                        ONLY
                          0. O  O. 2   0. 4   0. 6  0. 6   1.0

                              WATER SATURATION
Fig. 3.9   Water saturation distributions at 200 hours.
                                         75

-------
                      so -
                      75 -
                   Q.

                   D 1DO
                     125 -
                     150
                                  FULL  HYSTERESIS
                                 FLUJD ENTRAPMENT ONLY
                             NONHYSTERETIC
                                           t - 200 h
                        D. D  0. 2  0. 4  0. 6   0. 6   1.0


                         FREE  NAPL SATURATION
 Fig. 3.10  Free NAPL saturation distributions at 200 hours.
                   E
                   0
                  Q.
                  UJ
                  D
100 -
                    125 -
                    150
                                   PULL

                                 HYSTERESIS
                             FLUID

                            ENTRAPMENT

                              ONLY
                                          AT ZOO HRS
                       0. 0  0. 2  0. 4   0. 6   0. 6  1.0


                      TRAPPED  NAPL SATURATION
Fig. 3.11  Trapped NAPL saturation distributions at 200 hours.
                                         76

-------
 Table 3.2
Free and trapped fluid saturations at a depth of 50 cm which correspond to
the time in Figure 3.1 for different simulations
Time

A
B
C
D
E
F
G
H
1
J

A
li
C
D
E
F
C
H
1
J

A
B
C
D
E
F
G
H
I
J
Su-

0.199
0.299
0.663
0.667
0.493
0.472
0.544
0.582
0.385
0.342

0.200
0.215
0.548
0.579
0.355
0.332
0.405
0.425
0.300
0.266

0.199
0.249
0.549
0.574
0.385
0.347
0.408
0.458
0.320
0.274
s.,,
Nonhvxterrlic
o'.o
0.141
0.337
0.333
0.123
0.076
0.403
0.3S9
0.085
0.057
Full Hvxierexix
0.0
0.143
0.191
0.146
0.112
0.086
0.126
0.159
0.093
0.074
Fluid Entrapment Onl\
0.0
0.136
0.182
0.146
0.102
0.082
0.257
0.252
0.094
0.071
s.,,












0.0
0.004
0.144
0.158
0.063
0.054
0.0X4
0.092
0.040
0.026

0.0
0.014
0.143
0.154
0.072
0.056
0.082
0.104
0.044
0.025
s«,












0.002
0.038
0.117
0.117
0.064
0.055
0.077
0.086
0.049
0.039

0.001
0.050
0.126
0.126
0.076
0.064
0.105
0.116
0.060
0.046
For imbibing water saturation paths corresponding to periods of rising water tables,
large differences occur  between total and  free NAPL  saturations predicted by  the
hysteretic  and nonhysteretic analyses (Figures 3.3 and  3.4).  On  the other hand, for
conditions  in which water saturation paths do not induce NAPL  entrapment, good
agreement  occurs between NAPL saturations for the different simulations.

The history of apparent water saturation versus capillary head at the 50 cm depth for
the full  hysteretic simulation is shown in  Figure 3.5. Imposed boundary conditions
generated two closed scanning loops. The first loop is barely  distinguishable from  the
primary  imbibition curve (Figure 3.5a). Greater detail is shown in Figure 3.5b. Since
initial conditions reflect main drainage branch 5-P relations, a  .primary imbibition path
ensued when the  water  table was raised  at time A.  Letter symbols in  Figure  3.5
correspond to the times designated in Figure  3.1.
                                      77

-------
 NAPL  injection  was  initiated  immediately  following  time A.  As NAPL  moved
 downward, water was displaced resulting in a water front preceeding the NAPL front.
 The hollow square symbol in Figure 3.5a designates when the fluid system changed from
 an air-water  to an air-NAPL-water system which is also the origin of the first scanning
 loop.

 As NAPL displaces water, a secondary drying path ensues. A drying water saturation
 path continues until the  NAPL front  moves beyond 50 cm at which time  the NAPL-
 water capillary  heads adjust to the NAPL front passage causing water saturation to
 increase and inducing a saturation path reversal. At time B, which is after  the passage
 of the NAPL front, the water saturation path is on a tertiary wetting scanning path.
 Subsequent rising of the water table to time C produced higher  water saturations  and
 an imbibition  saturation path  continued until  time  D.  The internal  scanning loop
 (Figure 3.5b) closed when the water saturation increased beyond the water saturation at
 the inception of the three-phase fluid system.

 In interval D-F,  the water table is lowered initiating a secondary drying scanning path
 until time F. Later, the water table elevation increases during period  F-H producing a
 saturation path  reversal to a wetting path. During the final simulation period H-J, the
 water table is lowered  causing scanning loop F-H-F to close whereupon the secondary
 drying path D-J is resumed.. The second scanning loop F-H-F is more distinguishable
 than the first because the  range of apparent  saturations  encountered in F-H-F is
 significantly greater than  that encountered in the first  scanning loop. Free and trapped
fluid saturations corresponding to the times in Figure 3.1 are listed in Table 3.2.  For
nonhysteretic and fluid entrapment only  simulations,  scanning S-P loops do not exist
because S-P relations are modeled  to follow the main drainage branch.

To investigate the spatial distribution of fluids for the various cases in greater detail, we
consider saturation distributions at t  =  100 h (time B) and at 200 h (time  D).  At 100
h, injected NAPL has redistributed for a  period of approximately 90 h while boundary
 conditions remained constant leading to nearly static  conditions. Predicted  free NAPL
 distributions at  100 h  are shown  for the three simulation cases  in Figure  3.6.  Nearly
identical free NAPL distributions are predicted  for the three cases reflecting the fact
that  water saturation was nearly  constant over  the period of NAPL- redistribution so
                                      78

-------
that  no NAPL entrapment is predicted.  In the upper part of the profile  where total
liquid saturation is  draining, all cases effectively  use  the same S-P  functions (i.e.,
drainage), so that results are indistinguishable.  Slightly greater deviations arise lower in
the profile where saturation paths are more complicated, but it is clear that effects of
reversible hysteresis phenomena are of secondary importance at this time.

Finally concerning the NAPL predictions, it is worth noting that all simulations predict
a nearly constant NAPL saturation in the upper zone of about 17% which for practical
purposes  is  irreducible  under the  influence of gravity.  This nondrainable  residual
saturation should be clearly distinguished from the  residual oil saturation introduced in
the constitutive  model which represents the maximum amount of trapped  NAPL. The
nondrainable saturation evidenced in the present simulation results is not trapped and it
is  not parametrically defined by the constutitive model. It is simply an outcome of the
relative permeability model which indicates that as NAPL saturation decreases, NAPL
permeability eventually diminishes to the point that flow virtually ceases and precludes
attainment of true equilibrium conditions.

Water saturation distributions versus depth at 100 h are  shown for the three simulation
cases in  Figure  3.7. Full  hysteretic and fluid  entrapment results are  quite similar;
however,  the nonhysteretic case overpredicts  water saturation since  air  entrapment
associated with the rising water table is unaccounted  for.  Comparison of NAPL-water
capillary  head distributions at 100 h shows  very little  difference among  the three cases
(Figure 3.8) indicating that predicted apparent water saturations are nearly identical.

At 200 h  the  water table has been held at its highest elevation for an extended period.
Deviations between hysteretic and nonhysteretic model predictions  of water saturation
exhibit similar trends to those observed at 100 h, but are now more pronounced due to
increases  in air  entrapment associated with the rising water table (Figure 3.9).  Full
hysteresis and  fluid   entrapment   only  simulations  again  exhibit  rather  close
correspondence  except  near  the upper  surface where reversible hysteresis effects
apparently become  more  significant due  to  increasing  separation between  scaled
scanning curves.

In contrast to the results at 100 h, free NAPL saturation distributions at 200 h exhibit
marked differences for the  hysteretic and nonhysteretic simulation cases (Figure 3.10).
                                       79

-------
Full hysteretic  and fluid entrapment only analyses indicate free NAPL volumes which
are only about 1/3 of that predicted by the nonhysteretic case. The discrepancy reflects
the substantial  amount of trapped NAPL for the hysteretic cases (Figure 3.11) which is
by assumption nonexistent in the nonhysteretic analysis.

Mass balance calculations for NAPL were performed during all three simulations by
integrating NAPL contents over the spatial  domain  and normalizing to the injected
NAPL volume. The mass balance  errors ranged from 1.7 to 4.9% using the pressure-
pressure formulation described  in  Section 2.2.1, with  the lowest error for the full
hysteresis  model. Apparently, for this particular problem, inclusion of hysteresis had a
stabilizing effect on  the numerical  solution, not only  reducing  the mass balance error
but  slightly reducing  the  average number of iterations  per time-step for nonlinear
convergence. The latter effect partially offset the additional computational  burden for
the hysteresis model.
3.2 Development and Testing of a Simplied Hysteretic Model

The model described in Section 3.1 distinguishes two sources of hysteresis: (I) hysteresis'
in apparent  water saturation (actual water  saturation  —  trapped fluid saturation in
water)  versus oil-water capillary  pressure  relations  and  apparent liquid saturation
(actual water saturation -f  actual oil saturation — trapped fluid saturations in water
and oil) versus air-oil  capillary pressure relations due to contact angle hysteresis and
differences in pore-scale  fluid entry and drainage pressures, and (II) hysteresis  in
saturation-capillary pressure relations and in relative permeability relation-saturation
relations  due to nonwetting fluid  entrapment associated  with pore bypassing during
wetting fluid imbibition. The results of numerical simulations discussed in the preceding
section indicated that a simplified model that considers  only Type-II hysteresis yielded
nearly identical results to the full model with both Type-I and Type-II hysteresis.

Simplification of the model to consider only Type-II hysteresis is an attractive option as
this will  markedly reduce the computational effort and storage requirements of  the
model.  Furthermore, by invoking this simplification it becomes feasible to treat various
aspects of fluid entrapment with greater rigor than is possible for  the full model.   In
particular, although the detailed model describes maximum trapped fluid saturation for
                                       80

-------
 a given saturation  path as a hyperbolic function of wetting fluid  saturation at the
 drainage-imbibition reversal point, actual fluid entrapment  at  a given wetting fluid
 saturation  is  taken as  a linear function  of  wetting fluid saturation.  The latter
 assumption is in fact  inconsistent with  the former,  but is  invoked for the  sake  of
 expediency.

 In this section, we describe a simple model for three-phase k-S-P relations that accounts
 for effects  of nonwetting  liquid  entrapment  in  a stringent fashion while avoiding
 complications  arising  from  less  important  factors.  A number of one-  and  two-
 dimensional simulations involving low and high density hydrocarbons will be performed
 to demonstrate the behavior of the model.
8.2.1 Model description

Saturation-capillary pressure relations. The proposed constitutive model is based on the
nonhysteretic three-phase k-S-P relationships of Parker et al.  (1987) with extensions  to
consider hysteresis due to nonwetting liquid entrapment. Explicit consideration will not
be given to gas entrapment during periods of increasing total liquid saturation, since gas
entrapment has a relatively minor effect on liquid phase relative permeabilities. Effects
of trapped gas may be accomodated implicitly by employing the maximum liquid-filled
porosity in lieu of actual porosity  and the hydraulic conductivity at "field  saturation"  in
lieu of the fully saturated conductivity (Kool and Parker, 1987).  It will be assumed  in
the following that these  definitions of porosity and saturated conductivity have been
employed where relevant.

Adopting the notation  of the preceding  section, effective  water and total liquid
saturations are defined by

                                     ~SW=  S~m                          (3-22a)
                                     5'=     ^2                          (3-22b)
                                      81

-------
 where St  =  Sw -\- -S0 is the total liquid saturation and 5m is the irreducible water
 saturation. Effective oil saturation is defined such that S0 =  St  - Sw, hence

                                       ~S0 = -y&jr-                           (3.23)

 Furthermore, oil saturation is divided into two fractions  so that 50 = Soi  + Sot where
 subscripts / and t denote "free" and "trapped" oil, respectively.  In terms of effective
 free and trapped oil saturations, defined by analog to (3.23), we have

                                    ~S0 = ~Sot  + ~Soi.                          (3.24)

 Finally,  we define a quantity  referred  to as apparent water  saturation, S w,  which
 corresponds to the sum of effective saturation of water and trapped oil, i.e.,

                                      1 „,  =  ~SW +  50,                       (3.25)

 Following Parker  and Lenhard (1987), apparent  saturation will be assumed a simple
 function of capillary pressure.

 Prior to the occurrence of oil at a given location, the porous medium is treated as a two-
 phase air-water system described by the van Genuchten (1980) function

                                  5, = [ 1  +  (a haw)«  J-                      (3.26)

•where a  and n are porous medium parameters and m =  1-1/n, and haw is the air-water
 capillary head. Following the occurrence of oil at  a  given location,  the system  is
 described by the three-phase relations

                              1 w =  [ 1 + (a $ow how)» ]'m                   (3.27a)

                              3,  = ( 1  +  (a 0ao hao)n ]-»                    (3.27b)

•where how and hao are oil-water and air-oil capillary heads, and /?ao and 0OW are scaling
factors. To avoid numerical oscillations associated with  changes from a two-phase air-
water system to a three-phase air-water-oil system, once a node  is classified as  a three-
                                       82

-------
phase node,  it is not  be allowed  to  revert  to  a two-phase  air-water system.  The
necessary condition to classify a location as a three-phase node is
                                  0   + 0
                                  "ow  '  ~a
                                                                              (3.28)
To compute water  saturation in  the three-phase  system, an  expression  for  Sot is
required. Since gas entrapment is disregarded in the simplified  model, we  assume oil
entrapment in a three-phase air-oil-water system can  be inferred directly from that in a
two-phase oil-water system. Consider an oil-water system with main drainage and main
imbibition curves  described by (3.27a).  On any given  scanning curve,  the  effective
water saturation at how  =  0 is less than 1 due to oil entrapment  given by  1 - Sor where
Sor is the effective residual oil saturation which may be estimated using  an empirical
relation given by Land (1968) as
                             _          •t _  c
                             ST  =
                              OT
                                         R  l-
                                              • 1                             (3.29b)

where 5J^'n is the minimum effective water  saturation corresponding to the reversal
from water drainage to imbibition  and S?" is the effective residual oil  saturation for
the  main  imbibition curve. Residual oil saturation  corresponds to the  trapped oil
saturation at zero capillary pressure.

In Section 3.1,  the trapped  saturation at nonzero capillary pressure was estimated by
linear interpolation with ~Sot  = 0 at f „, = Sj7"" and ~Sot  = 507" at f „, = 0. This
method  has the  advantage of simplicity, but  is inconsistent with  (3.29).   Linear
interpolation can  also lead to numerical problems at capillary heads near  zero  due to
excessive changes  in trapped oil saturation with small head  changes.  A procedure which
is  fully  consistent with  (3.29) involves estimation of trapped oil saturation  as  the
difference between residual  oil saturation for the actual scanning curve and that for a
curve with a reversal point equal to  the saturation on the actual path.
                                      83

-------
 To clarify, consider Point 2 on Figure 3.12 with apparent water  saturation 5 w on  a
 path starting from Point 1 where 5,,, = 5J""1.  The effective trapped oil saturation at
 Point 2 is computed  as  the difference between the  residual saturation on the path
 starting from Point 1 and that starting from Point 3. From (3.29) we have
                     _ {      1 - S™»	1-5,_     \
                   "11+  R(l-S™)      i + tf(l-SJ  /
(3.30)
Equation (3.30) is valid if 5^ >  S „,""", otherwise 50, = 0. For the three-phase air-oil-
water system, an additional  constraint which must be enforced is that 5ot at a given
node is  the  maximum  of Sot computed by (3.30) and  50 at the node —  that is,  the
system cannot trap more oil than is present at the node.   Additionally, for the three-
phase system, 5Jn>n takes on the operational meaning of being the minimum effective
water saturation since the first occurrence of oil at the node.  The  value of SJ711" is
updated at the end of each time step after convergence and stored for each node.

Relative permeability  relations.  At  a given water saturation, oil entrapment affects
water permeability by displacing water into larger pores. The net effect can  be  shown to
be small and experimental studies often report no observable hysteresis in wetting phase
permeability vs. wetting fluid saturation relations (e.g.,  Saraf et al., 1982). Therefore, in
the present analysis, we will assume water relative permeability to be a simple single-
valued  function of  water saturation  as described by Parker et  al. (1987). Since  gas
entrapment is not  considered in the present analysis, we will assume that  gas relative
permeability can be described as a unique function of gas saturation also  as  given by
Parker et al. (1987). To describe oil relative permeability, it  is assumed that  only free
oil is a continuous phase subject to convection while trapped  oil cannot move.  Parker
et al. (1987) derived an expression  for oil  relative permeability in the absence of fluid
entrapment which  is a function of  effective water saturation  and effective total liquid
saturation.  In the presence of trapped oil,  the oil-water interface in the porous medium
will  be  displaced   to  larger pores identified  with the  apparent water  saturation.
Replacing  effective water saturation  by  apparent  saturation  in  the previous model
should therefore yield a good  representation of oil relative permeability.
                                      84

-------
Phase  relative permeabilities in  the  present  model  are  described by the following
relations

                          *™ =  ~SJ'* [ 1 - (1-5^/T ]2                   (3.31a)

                 kro  =  (5,-f J1/2[(l-1^/m)m-(l-5t1/rn)m]2          (3.31b)

                          kra =  (1 - 5t)1/2 [ 1 - St1/m ]2m                   (3.31c)

where fcruj, kro and kra denote  relative permeabilities to water, oil and air, TO  =  1-1/n is
the van Genuchten parameter  and  saturations are as previously defined.
Table 3.3.  Soil and fluid properties used in simulations.
Property
*
sm
n
a [L-1]
K,W[L T'1]
OV71OT
Pao
"ow
Pro
Iro
Example 1
0.4
0
2.5
0.05
40
0-0.4
2.67
1.6
0.8
2
Example 2
0.4
0
3
0.8
3
0.25
1.8
2.25
0.8
2
Example 3
0.4
0
3
0.8
3
0.25
1.8
2.25
1.2
2
Units for Example 1 in cm and hours and for Examples 2 and 3
in meters and days.  Conductivity assumed isotropic.
                                     85

-------
 3.2.2 Example problems

 Example li  One-dimensional infiltration and redistribution of low density NAPL. The
 purpose of this problem  is to investigate model sensitivity to the maximum residual oil
 saturation, SJ£ai.  The flow  domain is  taken to be a 75 cm long soil column which is
 initially free  of oil and at hydrostatic equilibrium with a water table located 10 cm from
 the lower surface. The  soil  hydraulic and fluid  properties used in the  simulation are
 given in Table 3.3.  A  total of seven simulations  were performed  with  S?0ax  values
 ranging from 0 to 0.5 which is the extreme range which is likely to occur in geologic
 media. Note that 5J^OZ =  0  corresponds to no oil entrapment in which case the model
 reduces to that described by Parker et al. (1987). Boundary conditions for each of the
 simulations consist of following stages:

 Stage 1: A  slug of  oil is allowed to infiltrate  from  the upper surface  at  a water
 equivalent  head of 1 cm until the total accumulation  is 5 cm3cm~2 while a zero-flux
 condition was stipulated for the water phase.  At the lower boundary, the water head is
 maintained at its initial value of 10 cm and the oil phase is zero-flux.

 Stage 2:  Oil is allowed to redistribute for 25  h under zero-flux conditions for both water
 and oil at the upper surface,  while the lower boundary conditions remain the same as in
 Stage 1.

 Stage 3:  The water head at  the lower boundary is raised from 10 cm to 35 cm over a 1
h period and maintained at  this value for 175 h.  A zero-flux condition is imposed  for
water at the  upper boundary and for oil at both boundaries.

Simulations were carried out with a finite element mesh consisting of 75 elements using
a time-step size varying  from 0.001 to 1 hour. Predicted oil saturations versus elevation
at the end  of Stage 3 are shown in Figure 3.14 for the cases S%?x = 0 and 0.25.  During
Stage 3, water saturation increases as the  water  table rises resulting  in oil entrapment
when this is  considered by the model. As a result, vertical  displacement of oil is much
less efficient  and the zone of oil saturation is  much more dispersed when oil entrapment
is  considered. At the end of the simulation nearly 40%  of  the oil in  the system  is
trapped for the case with 5^QI =  0.25.
                                      86

-------
 Variations in the percentage of added oil which has become trapped at the end of Stage
 3 in response to variations in 5J£ai are shown in Figure 3.13.  The results indicate that
 the fraction of the 5 cm3cm~2  spill volume which becomes trapped following a 25 cm
 increase  in  water table  elevation increases in a  markedly  nonlinear  manner  with
 increasing S%?x. The entire spill becomes immobilized under these conditions when S?0ax
 « 0.4.  For a given maximum  residual saturation, immobilization of the  spill would be
 greater if the total spill volume  were smaller or the water table rise greater.
            0
            01
            r
            x
            L
            0
            Q_
            D
            U
               Effective  water saturation  
-------
                  100
                     0. 0
D. 1
                                           0. 2
                                         C.   mcx
                                          or
0. 3
Fig.  3.13 Variation  of percentage  trapped  oil  volume in  domain with  maximum
          residual oil saturation for Example 1.
                          75
                                         Total  oil
                                         
-------
Example 2i Two-dimensional infiltration  and redistribution of low density NAPL. The
objective of this example is to demonstrate  the effects of fluid entrapment under field
conditions in which lateral  migration of hydrocarbon is significant. The flow domain is
taken  to be a 4 m deep  x 17 m wide vertical slice through  the  unsaturated  and
saturated zones which is initially free of NAPL (Figure 3.15) and in equilibrium with a
water table along plane BI  with a gradient of 1.6%.  Soil hydraulic and fluid properties
used in the simulation are given in Table 3.3. Simulations were performed with S^x =
0 and  0.25 using a  finite element mesh consisting of 232 elements with the time-step
varying from 0.0001 to 0.4  days. Boundary conditions were changed in three stages as
follows:

Stage 1: A slug of oil is allowed to infiltrate into the system under a  water equivalent
head of 1 cm  along a 0.5 m long strip  (EF)  until the total accumulation is 0.8 n^m"1.
All other boundaries are zero-flux for oil.  Water head is fixed at the  initial conditions
on sides AD and GJ and other boundaries are zero-flux for water.

Stage 2:  Oil  is  allowed to redistribute under ambient hydraulic  conditions to  t  =  8
days with zero-flux for oil on all boundaries and water boundary conditions the same as
in Stage 1.

Stage 3:  Water  heads on the segments AD  and  GJ are increased by 0.7 m raising the
water table to points C and H on the  boundaries  over a  0.5 day period and other
conditions are  maintained  as  in  Stage  2  for  another 12  days  bringing the  total
simulation time to 20 days.

The duration  of Stage  1  was determined to be  0.45 days irrespective of the value of
S^". Figure 3.16 shows the variation  in percentage of the  oil volume in the domain
which is trapped over the course of the  simulations for the case with S?*x  =  0.25. The
results indicate that negligible oil entrapment occurs  during Stages 1 and 2 since water
saturation paths are predominantly draining.  As the water table rises during Stage 3,
the percentage of trapped  oil increases  sharply  to  almost  55%  of the  spill volume.
Distributions of total oil saturation (free + trapped)  at the end of Stage 3 are shown in
Figure  3.17  for  simulations with S?™  =  0  and 0.25.   Trapped oil  saturation
distributions are  shown for the case of S?*x —  0.25 (for S™"  = 0 no oil is trapped).
                                      89

-------
The results show that fluid entrapment markedly reduces the lateral migration of the
liquid plume since much of the oil volume is trapped and  thus immobile. When fluid
entrapment is disregarded,  the upward displacement of the oil by the rising water table
is more efficient and lateral migration occurs more readily.
         3.5
                  E   F
   o
  B
                                   17
Fig. 3.15  Two-dimensional flow domain used in Examples 2 and 3.
                                     90

-------
                     100
                      eo
                  a
                  a.
                  o
                  <
                     20 H
                                Water-table  rise
	 Low density
	 	 High density
                                  5         10        IS
                                     Time  (days)
 Fig. 3.16  Variation  of percentage  trapped oil  volume with  simulation time for
           Examples 2 and 3.
 Example 3j.  Two-dimensional infiltration and redistribution of high density NAPL. The
 objective of this  simulation is to investigate the effects  of fluid entrapment on  the
 migration of dense NAPL plumes. The simulation will be identical to that described in
 Example II except that the fluid density ratio, pro, will be taken as 1.2 instead of 0.8
 (Table  3.3).  The  flow  domain,  initial and boundary  conditions  and numerical
 discretization is the same as in the previous example. The first stage of the simulation
 predicted an infiltration time of 0.36  days for a total oil accumulation of 0.8 nAir1.
 The predicted variation of the percentage of oil volume in the domain which is trapped
 with  simulation time is shown in Figure 3.16.  The dense NAPL simulation predicts
 almost 10% of the oil  to be trapped at the end of Stage 2 whereas no  trapped oil is
 predicted for the light  hydrocarbon at the end of Stage 2.  The reason for the greater
 entrapment for the dense fluid is the  more complex saturation history of the medium
 associated with the penetration of dense NAPL into the saturated zone.   As the front
initially  moves downward, high NAPL saturations develop  in the capillary fringe  and
upper saturated zone which  gradually diminish as gravity drainage continues. As water
saturation rebounds, oil entrapment occurs. As the water  table is raised, additional
entrapment occurs in the capillary fringe zone.
                                     91

-------
Predicted  oil saturation  contours  are shown in Figure 3.18 for  simulations with and
without fluid entrapment effects. Marked differences in oil saturation distributions occur
for the two cases. A much larger fraction of the spill volume is predicted to be retained
in the unsaturated zone and capillary fringe when fluid entrapment is  considered and
the  NAPL plume migrates laterally along the aquifer lower  boundary to a far lesser
extent.  The time required for NAPL to first reach the lower boundary was also much
longer when fluid entrapment was considered.
                                                       Total  oil
                                                       S  ™ox-0. 25
                                                      Trapped  oil
                                                      Sromox-0. 25
Figure 3.17 Predicted oil saturation distribution for Example 2 at end of Stage 3.
                                      92

-------
                                                     "Propped oil
                                                        i-ox-0. 25
Fig. 3.18  Predicted oil saturation distribution for Example 3 at end of Stage 3.
3.3 Experimental Verification of Hysteretic Flow Model

S.S.I Static three-phase measurements

Experimental methods and materials. Measurements of fluid contents and pressures as a
function of saturation path history were taken in two different porous media to evaluate
the hysteretic S-P model.  The experimental apparatus used to conduct the two- and
three-phase S-P measurements consists  of  alternating plexiglass sleeves  containing
treated and untreated porous ceramic rings  on their inner surfaces (see Figure 3.19).
There are two  treated and two untreated ceramics rings composing the retention cell.
The  treated  ceramics  were  immersed  in  chlorotrimethysilane  to  render  them
hydrophobic.  This is  done so that NAPL could be  continuous from NAPL-filled flow
channels within a porous medium to a pressure transducer located outside the retention
cell.  In this manner,  the NAPL pressure within a porous  medium could be measured
with a pressure transducer mounted outside the experimental cell.  Pressure transducers
                                     93

-------
  were  manufactured by  Soil  Measurement Systems in Tucson, Arizona, and had  a
  sensitivity of 1  mbar.  The untreated ceramics maintained  a  continuous water phase
  from  water-filled flow channels within a porous medium to the pressure transducer for
  the aqueous  phase  also located outside the retention cell.  Therefore, NAPL  and water
  pressures could  be measured simultaneously  for a three-phase fluid system.   The gas
  phase within the porous media was always in contact with the ambient atmosphere and
  its  pressure  was assumed  to be atmospheric (i.e., zero gage pressure)  when  liquid
  movement within the retention cell ceased.
              SCHEMATIC OF THREE-PHASE FLUID
             PRESSURE-SATURATION APPARATUS
                                    POROUS MEDIA
             FITTINGS
 TO TRANSDUCER.
   BURET AND
PRESSURE/VACUUM
   REGULATOR •*-
                                                                DETAIL OF CELL SEGMENT
                                                                 FITTING
                                                                           CERAMIC RING
 TO TRANSDUCER.
   BURET AND
PRESSURE/VACUUM
   REGULATOR •*
                                                *• TO TRANSDUCER.
                                                   BURET AND
                                                PRESSURE/VACUUM
                                                   REGULATOR
                                                                            PLEXIGLASS  CAVITY
   Figure 3.19.  Experimental saturation-pressure apparatus.
                                         94

-------
Two  vacuum-pressure regulators, one for water  and the other for  NAPL, controlled
movement of liquids into and out of the cell.  Each regulator was connected to a buret
that was connected to the retention cell via tubing.  Fluid contents were determined by
recording the volume  of fluid that drained  from the porous medium or that imbibed
into the porous medium.  The apparatus also was used to measure two-phase air-water.
air-NAPL, and  NAPL-water S-P relations.  For a  more  detailed  description  of the
experimental apparatus, see Lenhard and Parker (1988).

The NAPL was Soltrol 220, which is manufactured by  Phillips Petroleum Company,
Bartlesville,  Oklahoma.  Soltrol  220 is  a mixture  of  branched alkanes  having  low
solubility in water and a liquid density of 0.80 g cm"3.  The aqueous phase was distilled
water.  Air-NAPL  and NAPL-water interfacial  tensions, as measured by  the ring
method, were 0.026 and 0.036 N m" , respectively. The NAPL-water interfacial tension
reflects prolonged contact between NAPL and water.

The porous media employed in the experiments were unconsolidated. Porous medium 1
was a glass bead mixture with  99% of the beads between 0.125  and 0.25  mm, 0.06%
between 0.106 and  0.125 mm, and 0.04% less than 0.106 mm. Porous medium 2 was a
sand  consisting of approximately  97%,  1%,  and 2%  sand, silt and clay sized particles
(USDA), respectively.  A measured amount of air-dried porous media was packed into
the retention cell to achieve a constant bulk  density. Preliminary investigations were
undertaken to determine what air-dry weight would yield a mass  density where porous
media grains would not move significantly during liquid wetting and  drying cycles (i.e.,
a  constant pore  volume). The  resulting bulk densities of the glass bead and  sand
packings were 1.59 and 1.72 g cm"3, respectively.

Selection of the two porous media used in the experiments was based on  satisfying a
major constraint of the model  that  it be rigid  porous media (i.e., no swelling clay
minerals that may induce changes in the pore-size  distribution with  fluid content
changes).  Although both porous  media can  be  classified with  respect to grain-size
distributions  as sands  (USDA), flow  channels that result from packing glass beads will
have  a much different geometry than those that result  from packing natural angular
porous material. The glass beads are significantly  smoother and more rounded than are
the sand grains (porous medium 2). This variation is expected to produce differences in
the amount of nonwetting fluid that  becomes entrapped and in contact angle changes
                                     95

-------
with  saturation path  reversals.  Porous media  were added  in  several stages  to  the
retention cell containing de-aired water. After each stage, the porous media were mixed
to homogenize the packing and  to dislodge any air bubbles that might have become
trapped.  The initial condition for the S-P measurements, therefore, was a completely
water-saturated porous medium.

Two- and three-phase saturation path histories were initiated by desorbing wetting fluid
(i.e., water for two-phase air-water,  two-phase NAPL-water, and three-phase systems,
and NAPL for a two-phase air-NAPL system) from the porous media via adjustments in
the vacuum-pressure regulators.  As  the wetting fluid content decreased, a continuous
nonwetting phase (i.e.,  air for  two-phase  air-water, two-phase  air-NAPL,  and  three-
phase  systems,  and NAPL  for  a two-phase  NAPL-water  system)  entered  the void
spaces. Measurements along this saturation path  corresponded to  the main wetting fluid
drainage branch.  After the wetting fluid content decreased to some low  arbitrary level,
wetting fluid was permitted to imbibe back into the porous media by adjusting  the
vacuum-pressure   regulators.   Measurements  along  this  second  saturation  path
corresponded to a wetting fluid imbibition scanning path.

For  two-phase fluid systems,  only  the main wetting fluid drainage  and imbibition
scanning paths were measured. In addition, measurements for the imbibition scanning
path were conducted until the porous media appeared to be apparently saturated (i.e.,
containing wetting fluid and entrapped nonwetting fluid at  a zero capillary pressure).
This yielded a measurement of the residual nonwetting fluid content (5,r'J) from  which
the maximum amount of entrapped nonwetting fluid (i.e., residual nonwetting fluid
content) for the main imbibition branch (/5,>^) could be calculated by inverting Land's
(1968) algorithm.  The  residual nonwetting fluid contents  associated  with the main
wetting fluid imbibition branch are parameters in the hysteretic S-P model.

For three-phase saturation path histories, NAPL was allowed to imbibe into the porous
media at a point along the imbibition scanning path. When NAPL imbibed, the fluid
system changed from a two-phase to a three-phase system. For subsequent three-phase
S-P measurements,  NAPL and/or water were  permitted to  imbibe into  the  porous
media or were extracted from the  porous media  by adjusting the vacuum-pressure
regulators. Both  NAPL  and water pressures  and saturations were recorded for each
three-phase measurement.  This yielded  simultaneous measurements  of water and total
                                     96

-------
liquid saturation path histories that can be compared to model predictions.  The total
liquid path refers to water saturations during air-water  S-P measurements and to  the
sum  of water  and  NAPL  saturations during three-phase S-P  measurements.  The
saturation path history was different for each porous media packing. During the three-
phase S-P measurements, cyclic changes in fluid contents (i.e., wetting and drying) were
imposed to investigate scanning path behavior.

Experimental results. Figure 3.20 shows the water saturation path  history for porous
medium 1. The open square symbols are two-phase air-water measurements, the closed
diamond  symbols  are  three-phase  measurements,  and  the  dashed  lines  connect
successive measurements. The first series of measurements involved draining water from
an initially water-saturated porous medium to Swaw =  0.42 (i.e., ASu,au/ =  0.42). As
water drained, air entered the porous medium.  This saturation path  corresponds to  the
main drainage branch. An imbibition  scanning  path was then initiated  from Swaw =
0.42 to Swaw  =  0.55. During  this path,  air should theoretically become trapped and
occluded by water as air-water interfaces advance into larger flow channels.

NAPL was then allowed  to imbibe into the  glass  beads  creating a three-phase system
while the  water  saturation was held constant  at  0.55. Thereafter, water was  drained
forming a drying scanning path until  the main drainage path was  rejoined  at Sw =
0.42.  Note that when scaled capillary heads (i.e., /?y\) are employed, a closed 5-P loop
is formed when the main  drainage  branch is rejoined.  Air  trapped by  air-water
interfaces and occluded by water during the imbibition segment  of the 5-P loop should
become occluded by NAPL during  the  drainage segment. This  follows  from  the
assumption that entrapped fluid ganglia axe immobile  as fluid-fluid  interfaces move
within a porous medium. Upon rejoining the main drainage path, air  initially trapped
by air-water interfaces should be occluded by NAPL or be released into the continuous
gas phase.  Whether trapped air remains occluded by NAPL or released into  the  gas
phase depends on the corresponding total liquid saturation path history.
                                      97

-------

I
E 60-
U

O 50 -
^
UJ
1 40 -

>•
tr
< 30 -
— '
— '
£20-
U
10 -
o
UJ
_) n -

\^
\\ \
\\v
t ^ SP-
\ \ h \ ^°"
^ ^ B^ ^^""B-^
\ ^^VV,^ "%
\ \ Nj. S
V v 1
~* \

\
1
1
WATER SATURATION
PATH HISTORY I


1

                      U
  0.0  0.2   0.4  0.6   0.8
      WATER SATURATION
                                                      1 .0
Fig. 3.20  Experimental two- (open squares) and three-phase water (closed diamonds)
          saturations measured for the water saturation history in porous medium 1.

IU J\J
X
E
JJ 25

D
< 20 -
UJ
I
<
M 10 "
OL
U 5-
D
UJ •
-J o -

l *
\ ^-^

\ \ ^

\\\ \
^X^il \
^ Ri a
Vt»l T
]L\ i i
1 1
-I1 '
TOTAL LIOUIO * { '
SATURATION PATH \ '
HISTOBY

                     U
                     tn
 0.0   0.2   0.4   0.6   0.6   1.0
TOTAL  LIQUID  SATURATION
Fig.  3.21  Experimental two-phase water (open squares)  and three-phase total liquid
          (closed diamonds) saturations measured for the total liquid saturation path
          history in porous medium 2.
                                     98

-------
After rejoining  the  main drainage path  at  Sw =  0.42, water contents were further
reduced, yielding a continuation of the main drainage branch.  As water drained, NAPL
filled the water-vacated flow channels. Note that the three scaled main drainage S-P
points for the three-phase fluid system appear to be an extension of the air-water main
drainage data. At 5^  =  0.24, which was the lowest  water saturation attained (i.e.,
Swmtn  _  0.24), another imbibition scanning path was produced by allowing water  to
imbibe into the porous medium.  During this imbibition path,  however, initially only
NAPL is trapped  as NAPL-water interfaces  advance into larger flow channels. This is
because air has never entered these flow channels. The lowest total liquid saturation for
this experiment was 0.37 (i.e., 5
-------
saturations in the three-phase fluid system, and the dashed lines connect successive S-P
measurements. As in Figure 3.20 and for all  S-P experiments, the main drainage path
was initially followed. Water was drained from porous medium 2 until Swaw  =  0.40,
then an imbibition scanning path ensued.

At Swaw = 0.60 on the first imbibition scanning path, NAPL was permitted to imbibe
into the sand creating a three-phase fluid system. For the next three measurements, the
total liquid  saturation continued to increase. These  measurements  should lie on the
same imbibition  scanning  path that  was formed in a  two-phase  fluid system. No
saturation path reversals occurred that would generate a different  saturation path. Note
that when scaled capillary pressures are employed, these three experimental three-phase
S-P points appear to be an extension of the  imbibition path formed in the two-phase
system.  Scaling the capillary pressures  by  best-fitting essentially removes effects that
are due  to differences in interfacial tensions and contact angles  of  the different fluid
systems  (Lenhard and Parker, 1987b).  Theoretically, scaled  S-P main branches should
be  unique for  a given porous  medium.  Scaled  S-P scanning  paths,  however, are
dependent on the saturation path history.

At St = 0.78 on  the  first  imbibition  scanning  path,  the  total liquid  content  was
decreased, resulting in a drying scanning path.  This path was followed until St = 0.53.
At this  point, the total liquid content within  the sand was increased again forming
another imbibition scanning path, which was internal to  the first imbibition scanning
path. Note that the second imbibition scanning path did not  close with the first at
St = 0.78.  This  is because  the saturations that index the historic fluid-fluid interface
locations changed. However, if apparent saturation scanning paths were shown in lieu of
actual saturation paths in Figure 3.21, they would close at the scaled capillary pressure
Lead corresponding to the beginning of the drying scanning path (i.e.,  St = 0.78).

The explanation why  scanning paths would close when apparent saturations are used
but not  when actual saturations are employed follows from  tracking the movement of
5tro". On  the first imbibition scanning path, air was trapped  by air-water interfaces
from 5|  =  0.40, which corresponded to *SW™, to St = 0.60, which was the saturation
where the fluid system changed from a two to a three-phase system (i.e.,  5tmi'w  = 0.60).
For the three-phase segment of the first imbibition scanning path, air was trapped by
advancing  air-NAPL interfaces from St  = 0.60 to St  =  0.78. On the following drying
                                      100

-------
scanning  path,  air was released to the continuous gas phase  as  air-NAPL  interfaces
receded into smaller flow channels. As air-NAPL interfaces receded into flow channels
smaller than those corresponding to 5, =  0.78, but larger than 5, =  0.60, air trapped
by air-NAPL interfaces  was released into the continuous gas  phase.  As  air-NAPL
interfaces receded into flow channels smaller than those corresponding to St =  0.60, but
larger than St  = 0.40 (note that  ^Swaw  =  0.40), air trapped by air-water  interfaces
was released into the continuous gas phase. Therefore, as 5, became less than 0.60, Stmtn
changed. Stmin at the termination of the drying scanning path was 0.53 (see Figure 3.21).
For the ensuing second imbibition scanning path that  originated at St  =  0.53, all air is
trapped  by air-NAPL  interfaces  as they move into larger flow  channels. The total
amount of air trapped on the second imbibition scanning path,  however, will be less
than the amount of air released on the preceding drying scanning path because ISarao is
less than  ISa,raw. Hence, for the same scaled capillary pressure head corresponding to the
initiation of the drying scanning path, there should be a higher water saturation for the
second imbibition scanning  path  than for the first imbibition scanning path. This is
because less air is entrapped on the second imbibition scanning path, which agrees with
what  was experimentally observed in  Figure 3.21.  When apparent saturations  are
modeled,  these  differences are taken into account by definition.  Apparent saturations
track the movement of fluid-fluid interfaces that separate immiscible fluid phases.

As the total liquid content was increased beyond 5, = 0.78, the measurement point at
St = 0.84 appears to lie on an extension of the first imbibition path, which began in the
two-phase system. It is assumed  that once a drying or imbibition scanning  path has
closed, the previous drying or imbibition  saturation path  is respectively followed. The
experimental measurements in  Figures 3.20 and 3.21 confirm  this behavior.  The final
total liquid saturation  path for the sand was another drying scanning path.  However,
the latter drying scanning path  originated  from a saturation reversal of 5^, =  0.84, and
the former drying scanning path was  at a reversal of Sw — 0.78. Note how  the latter
drying scanning path does not cross  over the former drying  scanning  path and the
second imbibition scanning path does not cross over the first imbibition scanning path.
This  suggests that higher order imbibition and drying scanning paths are internal to
their corresponding lower order paths. From  the data shown in Figures 3.20  and 3.21,
there  appears  to be orderly progression  of  saturation paths as  the  saturation path
history develops. Hence, an  accurate representation of history dependent S-P relations
may be obtained with a model that incorporates the major observed features.
                                     101

-------
 Modeling results.  Model parameters were obtained by best-fitting  the  model, via  a
 nonlinear regression algorithm, to experimental water and total liquid  saturation path
 histories. Results of these regression analyses are given in Table 3.4. Also shown are the
 residual nonwetting fluid saturations of the main imbibition branch, *5,>';'.

 Figure  3.22a shows a comparison between predicted and experimental water saturation
 path histories  for porous  medium  1.  The  dashed  lines  connect  successive  S-P
 measurements and the  solid line is the model predictions.  There is close  agreement
 between  predicted and experimental  water  saturation  path  histories.  Significant
 discrepancies  between predicted and experimental S-P relations are found only for the
 larger scanning loop. Parameters used for predicting the water saturation path history
 in Figure 3.22a were obtained by best-fitting the model to the experimental two- and
 three-phase water saturation path history.

 An alternative method for calibrating the model is to best-fit the van  Genuchten (1980)
 retention function to only hysteretic air-water S-P data to obtain the parameters, ca, ;a,
 n and Sm. Ratios of air-water to NAPL-water and air-NAPL  interfacial  tensions can be
 used to predict the scaling parameters /3OW and /?ao, respectively (Lenhard and Parker,
 1987). The residual nonwetting fluid saturations of  the main  imbibition S-P branch for
 two-phase fluid systems (/5irt^), however, will have to be experimentally determined by
 imbibing wetting  fluid into  the porous  medium that  is  initially  saturated  with
 nonwetting  fluid until the porous medium is apparently saturated. Using  this approach
 to obtain model parameters, Figure  3.22b  was constructed.  Note  that  this alternative
 calibration method, which  can be used in the absence of three-phase S-P data, predicted
 the measured two  and three-phase water  saturation path history  in porous  medium 1
just  as accurately as did the best-fit calibration method.

 Figure  3.23 shows  the total liquid saturation path history measured in porous medium
 1. Again, there is relatively close agreement between predicted and experimental total
 liquid saturation path histories. Model parameters were obtained from  best-fitting to the
experimental  data  (see Table 3.4). Figure 3.23 shows  two  closed  scanning paths;
 however, they are  difficult to discern. The coefficient of rmiltiple determination (R2) in
the regression analysis for the total liquid saturation path history was 0.96.
                                      102

-------
If the alternative method  of  calibrating  the  model to a  porous medium is used for
predicting the total liquid saturation path history in porous medium 1, the results are
not as accurate as those for the water saturation path  history. This can be deduced by
noting that the ratio of air-water  to NAPL-water interfacial tensions (i.e., 0OW) required
for predicting  water saturation paths is 2.0. The best-fit  Pow was 2.05 (see Table  3.4)
when the model was regressed to  the experimental data. These values  are very close to
each other; hence, model predictions by both calibration methods should be similar, as
shown in Figures 3.22a and 3.22b. However, when predicting the  total  liquid saturation
path history by using the alternative calibration method, the ratio of air-water to air-
NAPL interfacial tensions is 2.77,  whereas  the best-fit /?ao was 2.39.

A noteworthy observation is the agreement between predicted and measured total liquid
saturations for the imbibition scanning path at a scaled capillary pressure of zero.  For
this point, the porous medium was apparently saturated with NAPL and water. That is,
the sum of the NAPL, water  and entrapped air contents equaled the porosity or flow
channel  volume.  At this  5-P point, air is predicted to be trapped by advancing air-
water interfaces in a two-phase fluid system and by advancing air-NAPL interfaces  in a
three-phase fluid  system.   The entrapped air saturation  that  resulted from  air-water
interfaces was predicted to be 0.015, and the entrapped air saturation that resulted from
air-NAPL interfaces was predicted to be 0.137.  The sum of these  amounts differed from
the measured amount of trapped air in the glass beads by a saturation of 0.002.

A similar water saturation path history, as measured in porous medium 1, is shown for
porous medium 2 in Figure 3.24.  A closed scanning path that developed from the main
drainage 5-P branch was measured.  Agreement between experimental data and  model
predictions  is  very  close,  particularly for  the  scanning paths. Figure 3.25 shows the
corresponding experimental and predicted  total liquid saturation path histories.  Again,
very  close  agreement  was  found between predicted and experimental  total  liquid
saturation path histories.  Model predictions in Figures 3.24 and 3.25 were generated by
best-fitting the model to the experimental data.

Although the same  three-phase fluid system was used  for porous medium 1 and  2, the
best-fit scaling factors are different. These differences  may reflect contact angle effects
or experimental uncertainty.   Theoretically, contact angles in dissimilar porous  media
could be different.   The observed  differences in the  best-fit scaling factors,  0OW and /?
                                                                                 00'
                                      103

-------
for the two porous  media  suggest  that  estimating  the scaling  factors via ratios of
interfacial tensions may not  always yield satisfactory results.  Previous investigations by
Lenhard and  Parker (1987,  1988) for drainage saturation paths, however, have  shown
that interfacial tension ratios provides a good estimate of the best-fit scaling factors.

Table 3.4.   Model parameters used to predict  experimental water  and total  liquid
saturation path histories for porous media 1 and 2.
Parameters Porous Medium 1 Porous Medium 2
Da 0.022 cirf; 0
*a 0.032 cm-' 0
n 10.70 5
5m 0.22 0
P.. 2.39 1
/*,„ 2.05 2
/5arai" 0.06 0
/5arao 0.16 0
75erei" 0.15 0
o o _
»„ /\, —
I

S60~

D 50 -
^
UJ
1 40 -

EC
< 30 -
_J
M 20 -
a
CJ
10 -
D
UJ
_J n

«
N
*\ »t
tV^t*^
VV L^^*^*^^
V \ »4c\ ^'^''^k-.
N. \ V''Sv\* ^^XB
>^ l^i ^^n
•>«»4fc '



WATER SATURATION
PATH HISTORY



I

j>-

D 50 -
^
LU
1 40 -
-^
EC
< 30 -
_)
fT 20 -
CL
i U
10 -
O
> Lb
i — 1 0 -













.042 cirf;
.072 an1
.86
.10
.93
.30
.19
.14
.09


\

\x\
\\w^_
\ NV v\r^^c^^
\ ^ ^A ^t&^afc^.
\/. "SLv >1>:^*
X. \V
-------
                              o
                               (u 50
                              E
                              U 50
                                40 -
LJ
I

£30-



_J 20 -

a

u 10 -|

D  '
111
d  0
                                       TOTAL LIQUID

                                     SATURATION PATH

                                        HISTORY
                              CJ
                              en
                                 0.0  0.2  0.4
                                                0.6  O.B
                                                          1.0
                                 TOTAL LIQUID  SATURATION
 Fig.  3.23  Comparison of experimental (dashed lines) and predicted (solid line) total

           liquid saturation  path histories for  porous medium  1  when the model  was

           calibrated by best-fitting the model to the experimental data.
 cv 4B
I

E
U 40
                              LU
                              I
                                32 -
                                24 -
                                16
D.

CJ

D
UJ
_l

U
                                 a -
                                           WATER SATURATION

                                             PATH HISTORY
                                 0.0  0.2  0.4   0.6   O.B   1.0

                                     WATER SATURATION
Fig. 3.24 Comparison of experimental (dashed lines)  and predicted (solid line) water

          saturation path histories for porous medium  2 when the model was calibrated

          by best-fitting the model to the experimental data.
                                      105

-------
                           a- 30
                          I
                          E
                          U 25
                            20 -
tu
I


I1""
_J
-J 10 -
M
D.
0  5 H
Q
ID
_J
                                   TOTAL LIQUID
                                 SATURATION PATH
                                    HISTORY
                          U
                          in
                             0.0  0.2  0.<
                                            0.6  O.B
                                                     1.0
                             TOTAL LIQUID SATURATION'
Fig. 3.25  Comparison of experimental  (dashed  lines)  and predicted  (solid line)  total
          liquid saturation path histories for porous medium 2  when the model was
          calibrated by best-fitting the model to  the experimental data.
8.8.2 Two-phase dynamic measurements

Experimental methods.  A one dimensional air-water flow experiment was conducted to
measure water  content  and pressure distributions over time  that  resulted  from a
fluctuating water table scenario. The objective of this exercise was to compare measured
hysteretic fluid  distributions that included effects  of nonwetting fluid entrapment to
results of numerical simulations of the experiments that incorporated the hysteretic k-S-
P model for the two-phase case.

The porous medium employed in the experiment was an unconsolidated sandy material
comprising approximately 97.5,  0.8 and 1.7 percent sand, silt and  clay sized particles
(USDA), respectively. The sand was packed in a glass flow cell with a cross sectional
area of approximately 39 cm2 (6 cm x 6.5 cm) and 1 m in height.  A 10-cm-thick gravel
layer underlaid  72  cm of  packed sand.   Porous  ceramic tensiometers connected to
pressure transducers were inserted into the packed column through access ports located
                                     106

-------
 10  cm  apart along  the  flow  cell.   Ceramic  tensiometers  were cylindrical  with
 approximate dimensions of 3 cm long by  5  mm in diameter. Pressure  transducers,
 manufactured  by Omega Engineering (model 240),  were  constructed  with  silicon
 diaphragms and possessed a range of ±  17 kPa or ± 175 cm of water equivalent height.
 Transducers were accurate to within 1.5% and  the response time was 1 ms.  Seven access
 ports were outfitted  with tensiometers. Drainage and imbibition of water from and into
 the flow column were accommodated by an outlet at the  cell base. The  sand-packed
 flow  column  was flushed with C02 before saturating with water to minimize  gas
 entrapment on the  inital water wetting of  the dry  sand.  Initial condition for  the
 experiment, therefore, was  assumed  to be a water saturated porous medium.

 A gamma radiation  attenuation system was  employed  to nondestructively measure
 water saturations. Gamma radiation measurements were taken within 1 to 2 cm of the
                                                                              *
 ceramic  tensiometers.  This was done  to avoid gamma radiation attenuation by the
 ceramic  tensiometers and  to avoid  conducting measurements in areas that might be
 disturbed (i.e., possessing a different soil particle arrangement) as a result of inserting
 the tensiometer tips  into the packed column. The radiation collimator was circular with
 a 6.6-mm diameter.  By  moving two interconnected parallel platforms, one supporting
 the radiation sources and the other supporting  the gamma detector, to desired locations
 with  stepper  motors under computer control, the gamma radiation beam  could be
 returned repeatedly  to any elevation. The seven  measurement elevations  were 70, 60,
 50, 40, 30, 20 and 10 cm, where 0 cm corresponds to the base of the packed  sand.

 The gamma system was  calibrated with a procedure reported by Lenhard et  al. (1988).
 A sectioned small Plexiglas cell and repeated  measurements  of 10-min durations  were
 used  to  determine  water  and  soil  attenuation  coefficients.  In  all  calibration
 measurements, 10 min counting times were  employed. Any gamma  attenuation caused
 by  air was neglected. Flow cell path lengths  for each measurement position  were
 determined  via empty and water filled gamma counts of the flow  cell and by direct
measurements  with  calipers.  Close agreement was  found between  path  lengths
 determined  from the gamma attenuation measurements and those that  were  physically
measured with calipers.

Water contents were determined from the relationship
                                     107

-------
                                            nwpwewx)                       (3.32a)
where
                                V = I0exP(-n,P,x)                        (3.32b)

in which  //, and nw are soil and water gamma attenuation coefficients, respectively, pt
and pw are bulk soil and aqueous phase densities, 6^ is the volumetric water content, i
is the path length, and 70 is the average empty cell gamma count of each measurement
location.  Bulk  densities employed  in  (3.32)  for  calculating water contents  were
physically measured by  sectioning the soil  column at  the  end  of  the  experiment.
Differences between soil bulk densities as determined by gamma counts of the dry  soil
column prior to the experiment and  those physically measured after  the experiment
were less than  0.04 g cm'3,  except for the uppermost two  positions where differences
were larger.  The average soil bulk density of the packed column was  1.70  g  cm"3.
Because of the sandy nature of the porous medium, changes in bulk densities  resulting
from water  saturation path  reversals were not expected  (i.e., porous medium was
assumed to be rigid).  Gamma counting periods during the transient experiments were
90 s and  all  gamma counts were corrected for resolving time and  Compton scattering
following methods reported by Stillwater and Klute (1988) and  Lenhard et al. (1988).

After initially wetting the dry soil column, the water table was positioned at 72 cm,
which corresponded with the soil surface, for  approximately 12 h.  At the  onset of the
experiment  (i.e., t  =  Oh), the water table was lowered 5 cm,  and every 10 min
thereafter it was lowered an additional 5 cm until the water table  reached  an elevation
of 7 cm at t = 2 h, where it remained stationary for an hour. At  t = 3 h, the water
table was raised  5 cm  every 10 min to a  final  elevation  of 42 cm and remained
stationary there for an  hour until t  =   5 h.  All changes in  water table elevations
proceeded at the rate of 5 cm every 10 min. The first lowering and raising of the water
table should  produce the main water  drainage and primary water  imbibition  scanning
paths, respectively. An  internal water drying scanning path was then generated by
lowering the  water table at the prescribed rate from 42 cm at  t  =  5 h to 17 cm where
again the water table remained  stationary for an  hour until t  = 6.67  h. The final
saturation path, a water imbibition scanning  path, was produced  by raising the water
table elevation from 17 cm back to the soil surface at 72 cm, which was attained at t
=  8.33 h. In  this last path as the water table was raised past  42 cm elevation, all
internal scanning  S-P  loops  should  have  closed and, thereafter,  primary water
                                     108

-------
imbibition scanning paths should have been followed. A distinct S-P path history should
have developed for each measurement elevation. The rate of lowering and raising the
water table was arbitrary and was imposed, in lieu of more abrupt elevational changes,
to minimize potential numerical oscillations in the simulations that could be induced by
large, instantaneous  changes in boundary conditions.  Water  content  and pressure
measurements were conducted until t  =  10 h.

The upper water boundary condition  as well as the lower air phase boundary condition
were no-flow. Evaporation from the  upper  boundary was  minimized by  covering the
flow  cell  with plastic that  contained numerous small  holes to  ensure  that the  air
pressure within the flow cell would  be atmospheric.  The  upper air phase  boundary
condition,  therefore, was constant  at  atmospheric pressure.  The lower water  boundary
conditions  were  constant  pressures  that corresponded  to  the  changing water  table
elevation scenario described above.

The experiment was simulated using the numerical model described in Chapter 2. A one
dimensional 72 cm vertical homogeneous soil section was discretized in 1 cm intervals to
model numerically. Although the multiphase flow code was developed to simulate fluid
flow in air-oil-water systems, two-phase air-water flow can be simulated  by regulating
the  oil phase boundary  conditions such that the system  remains oil free.  The code
assumes that the air phase, when present, is at atmospheric pressure.

To calibrate the k-S-P model, the drainage van Genuchten parameters, do, n and Sm,
were fit via a nonlinear regression  algorithm  to transient  S-P measurements taken
during the first drainage sequence  (i.e., main drainage). Two sets of parameters were
generated and five  different  simulations of the  experiment were conducted. Parameter
Set  1 was  obtained  by fitting da,  n and  Sm  to  the main S-P drainage  data and
Parameter Set 2 was obtained  by fitting  only  da and  n  to the main drainage data
assuming  Sm  = 0. In Simulation 1, the full hysteretic model was employed  using
Parameter Set 1. In Simulation 2, a simplified  hysteretic  model was employed that
accounts only  for  nonwetting  fluid  entrapment  effects in k-S-P relations (i.e.,  no
hysteresis in apparent saturation-capillary pressure  relations). We refer to the latter
analyses as  "fluid entrapment  only" simulations.  In  Simulation  3, hysteresis  was
disregarded altogether (i.e., no apparent saturation  hysteresis and no fluid entrapment
effects).  Simulation 4 was a variant of Simulation 1 in that Parameter Set 2 was used
                                     109

-------
and Simluation 5 was identical to Simulation 4, except the tortuosity term in Mualem's
(1976) relative permeability integrals was changed from Sw°-5 to Sw2, which conforms to
experimental work conducted by Burdine (1953) and Corey (1954).

Figure 3.27 shows the resulting  best fit main drainage van Genuchten function with
experimental data of the measurement elevations.  The solid line is for Set 1 parameters
and the broken line  is for Set 2 parameters. Although there is scatter in the  data, the
soil  column was  assumed to  be  homogeneous and isotropic for the simulations. The
residual air saturation of the main water imbibition  branch used in simulating the
experiment  was estimated to be 0.25, which was based on results  observed  in earlier
studies  of similar porous  media (Parker  and Lenhard, 1987). The  saturated  water
conductivity was measured  to be 119 cm h"1  in the packed flow column by the falling
pressure head method before  initiating the experiment. The porosity  (i.e.,   =  0.36)
corresponded  to the average  bulk  density of all  measurement elevations.  Parameters
used in the  5  simulations are  listed in  Table 3.5.   The hysteretic k-S-P routines can be
employed  to simulate nonhysteretic flow (i.e., Simulation 3) by equating the a for main
imbibition S-P relations  with  the a for main  drainage relations ('a  =  da) and using a
negligible  quantity for '£,„. (i.e., 10'^). A nonzero number is needed for *Sm or (4) will be
undefined.
Table 3.5.  Model parameters for simulations 1-5.
Parameters
"a
n
sn
X
*
Kftu
Parameter Set 1.
0.042 cm'1
5.25
0.17
0.25
0.36
119 cm h*1
Parameter Set 2
0.039 cm'1
4.02
0.0
0.25
0.36
119 cm h'1
                                     110

-------
                           ID
                           cr
                           D
                           cn
                           w
                           UJ
                           a:
                           a
                          a
                          <
                          u
                              o                          i

                                APPARENT SATURATION
Fig. 3.26  Apparent saturation-capillary pressure relations showing main drainage and

          imbibition branches and a scanning path scenario.
                                      Ill

-------
                           E
                           u
                          UJ
                          I

                          LJ
                          o:
                          D
                          C/)
                          tn
                          UJ
                          DC
                          a.

                          oc
                          LU
                          H-
-10 -
                                0.0   0.2  0.4   O.B  0.8  1.0

                                   WATER  SATURATION
Fig.  3.27  Comparison of measured  water saturations (symbols) as a function of air-

          water capillary head  and best fit main drainage S-P relations for 5m > 0

          (solid line) and for  5m  =  0 (broken line) where the closed diamond, closed

          square, closed triangle, open inverted triangle, open circle, open triangle, and

          open squares are the symbols used for S-P relations measured at the 70, 60,

          50, 40, 30, 20 and 10 cm measurement locations, respectively.
                                     112

-------
   1 .0
  1 .0
z
2 o.e H
I-
in 0.4 -
DC
  0.0
                  4      6
                TIME  (h)
                         i .0
                       z
                        '
                         o . e H
                       <
                       en 0.4 -
                       cr
                       UJ
                         0.2 -
                         0.0
                                    10
  1 .0
z
2 O.B -j
                                             50.
                                             <
                                             U5 0 . 4 -
                                             o:
                                               0.0
                                               1.0
z
2 O.B 1
  0.4 -
                                               « -, J
                                               0.2 -
                                               0.0
                                       TIME  (h)
                                                               4      B
                                                             TIME  (h)
                  4     6
                 TIME  (h)
                                                           10
                                                                                 10
                                                                                 :o
 Fig. 3.28  Comparison of measured (solid symbols) and simulated full hysteretic (solid
           lines), fluid entrapment  only  (broken lines with equal sized segments)  and
           nonhysteretic (broken lines with alternating long and short segments) water
           saturation distributions at measurement locations of a) 70 cm, b) 60 cm, c) 50
           cm, d) 40 cm and e) 30 cm.
                                      113

-------
  -55
 E
 u
  -45 -
HI
X

HI
cr
D
in
en
UJ
tr
D.

oc
UJ
i-
-35 i
-25 -
-15 -
 -5 -
                    MEASUREMENT

                     ELEVATION

                      70 em
     0.0  0.2  0.4   0.6   O.B   1.0


         WATER  SATURATION
                                                     E -45

                                                     U
                                                     HI
                                                       -30 -
UJ
EC
D -15
in
in
UJ
cc

a   o
EC
UJ
                                                        15
                      MEASUREMENT

                       ELEVATION

                         50 cm
                                                        0.0   0.2   0.4  O.B  O.B   1.0


                                                            WATER SATURATION
  -40
E
U
  -30 -
UJ -20


ID
CC -10

in
in
UJ   o
cc
a
cc
UJ
 10 -
   20
                    MEASUREMENT

                     ELEVATION

                      50 cm
                                                       -40
     0.0  0.2  0.4  0.6   O.B


         WATER  SATURATION
                               i .0
E
u
                                                     D -20
                                                     <
                                                     UJ
                                                     I
UJ
DC

in
in
UJ
cr
a.

cc
UJ
                                                         0 --
                                                        20 -
                                                        40
                                                              MEASUREMENT

                                                               ELEVATION

                                                                 40 cm
                                                       0.0  0.2   0.4   0.5  O.B  i.D


                                                           WATER SATURATION
  Fig. 3.29  Comparison of measured (solid symbols) and simulated hysteretic (solid lines)

            S-P histories at measurement locations of a) 70 cm, b) 60 cm, c) 50 cm  and d)

            40 cm.
                                          114

-------
— «JU -
"E
u
— -30 •
D
UJ -20 -
X
UJ
H -10 -
D
cn
c/i
UJ o .
DC °
a
oc 10 -
UJ
i-
S pn -
.
*V \ >x
^^t N
*N 1
^ ^

t I
*
I *



t
                              0.0   0.2  0.4  0.6  0.6   1.0
                                  WATER  SATURATION
Fig.  3.30  Measured  5-P history of the 50 cm measurement location showing a closed
          internal scanning loop where broken lines are employed to connect successive
          5-P measurements.
z
2 o.e -
i-
U) 0.4 -
cc
t- °-2 "
2
0

*L
V MEASUREMENT
I ELEVATION
\ 70 en
\
V
w .


2 4 6 E
TIME (h)
['£"*'•"?•
*
•

*


) 1C

                                              1.0
                                            go.aH
                                            U) 0.4 -
                                            UJ A _ ,
                                            ,_ 0.2 -
                                              0.0
MEASUREMENT
 ELEVATION
  40 cm
                                                              4      6
                                                             TIME  (h)
                                                                                1C
Fig.  3.31  Comparison of measured (solid symbols) and simulated full hysteretic water
          saturation distributions  using Parameter  Set 1  (solid line), Parameter  Set 2
          (broken line with  equal sized  segments)  and  Parameter  Set 2  with a
          tortuosity modification in the k-S relations (broken line with alternating long
          and short segments) for measurement locations of a) 70 cm and b) 40 cm.
                                     115

-------
 Results  and discussion. Measured and predicted  water saturations are compared in
 Figure 3.28 for the different measurement locations. Experimental water saturations are
 shown as diamond symbols. Full hysteretic, fluid entrapment  only  and nonhysteretic
 simulations are shown. Results of Simulation 1, the full hysteresis model, are shown as
 solid lines, results of Simulation 2, the fluid entrapment only option (FT), are shown as
 equal spaced broken lines, and results  of Simulation  3, the nonhysteretic  simulation
 (NH) that employs only k-S-P relations of the main  drainage branch, are shown as
 broken  lines with alternating long and short  segments. In these three simulations,
 Parameter Set 1 was employed. Very close agreement was found between experimental
 and predicted water saturations at all  measurement locations when the full hysteretic
 constitutive relations were employed (i.e., solid lines - Simulation 1). Predicted water
 saturations using the fluid entrapment only and nonhysteretic k-S-P relations, however,
 deviated significantly from measured water saturations of the scanning paths.

 Note in Figure 3.28 how poorly nonhysteretic predictions compare with water saturation
 measurements of the scanning paths. At the  70- and 60-cm positions (Figures 3.28a and
 b), measured water pressure heads decreased approximately 10 and 20 cm, respectively,
 during the first raising of the water table to an elevation of 42 cm, yet measured and
 predicted full hysteretic water saturations increased only slightly. The  reason for this
 phenomena will be explained later. Predicted saturations in the fluid entrapment only
 and nonhysteretic simulations, however, increased significantly during the same period.
 For measurement locations closer to the water table, there were only relatively minor
 differences among experimental water saturations and predicted  full hysteretic and fluid
 entrapment only simulations. The nonhysteretic simulations,  however, poorly matched
 the  measured water saturations  at all elevations. Furthermore,  the early  increase in
 water saturations predicted by the nonhysteretic k-S-P relations for the final wetting
scanning path  (i.e., t  = 6.7 to 8.3 h)  in Figures 3.28a-c may suggest that effects of
nonwetting  fluid entrapment on wetting  phase relative  permeabilities  may  have
 significant consequences. For the remaining  two  measurement locations that are not
 shown in Figure 3.28  (i.e., at elevations of 10 and 20 cm), differences between measured
 and simulated water saturations are generally less than 0.04. At these positions the soil
was nearly water saturated at all times.
                                      116

-------
Finally, note that for the 40-cm-measurement elevation (Figure 3.28d) experimental and
predicted water saturations are constant when the water table is positioned above this
location. During these periods, constant water saturations infer constant entrapped air
saturations. This occurred twice since the maximum water table elevations attained in
the first and second water imbibition paths were 42 and 72 cm, respectively, and the
minimum water table elevations attained for the first and second drainage sequences
were  7 and  17 cm, respectively. All measurement  locations  between 17 and 42 cm
should, therefore, exhibit plateaus in water and entrapped air saturations during periods
in which the water table is positioned above those elevations.  Unfortunately, however,
only the measurement elevation at 40 cm (Figure 3.28d) drained sufficiently during the
second drainage sequence to exhibit recurring plateaus. The pore size distribution of the
porous medium (i.e., height of water saturated fringe above the water table)  was such
that no water  drained during the second drainage sequence at the 20-cm-measurement
location and only an insignificant amount drained at the 30-cm location. These distinct
and  equal plateaus  observed at the 40-cm elevation suggest that  the entrapped air
saturation corresponding to an apparent saturation of unity may be a  constant  for a
given saturation path history, and can be  reproduced following cycles of wetting and
drying provided the saturation path reversal from  the  main water drainage  branch
remains  unchanged. Similar results  have  been reported recently by Stonestrom and
Rubin (1989) where they observed a reproducible functional dependence of entrapped air
content on water saturations.

The constant water saturations measured  during periods when  the  water pressure is
greater than atmospheric infers that there is negligible movement of trapped air out of
the porous  medium  for conditions  imposed during the experiment.  This  is  to be
expected considering that water was not continually flowing and was probably saturated
with respect to air at atmospheric pressure.  Water-occluded air bubbles in porous media
may move in response to increasing the viscous forces that tend to push the air bubble
through the porous medium (i.e., water pressure gradient), to decreasing the interfacial
forces that hold the air bubbles stationary,  and to dissolution of the air in the aqueous
phase and subsequent transport by convective and dispersive mechanisms. Researchers
in the petroleum industry investigating various techniques to enhance oil recovery in
petroleum reservoirs  have reported  a  correlation  between the amount of  entrapped
nonwetting  fluid that remains in  porous  media  and a  dimensionless  group called  a
capillary number, which is defined  as a ratio of viscous forces to interfacial forces, and
                                     117

-------
 can be expressed by

                                    Nc = fy^f                              (3.33)

 where qw is the Darcian water flux,  rjw is the water absolute viscosity, and   10~5, the amount of mobilized
 entrapped nonwetting fluid increased  rapidly with increasing -Afc. For air-water fluid
 systems in porous media, the Darcian water velocity needs to exceed  approximately 2.6
 m  h   , which is a substantial  flow  rate, in order to mobilize trapped  air bubbles
 according to  (3.33). For  boundary conditions imposed during our experiment such a
 large  flow rate would not be attained.

 Dissolution of air bubbles in  water has been  studied by Bloomsburg and Corey (1964)
 and others based on theoretical considerations. The time required for  a nonwetting fluid
 ganglion to  dissolve in water is a  function  of  the pressure difference between the
 nonwetting fluid and water and the nonwetting fluid solute concentration in  the aqueous
 phase. For small ganglia,  the dissolution time is orders  of magnitude faster  than for
 significantly  larger ganglia, however, as shown by  Bloomsburg  and Corey (1964) the
 time  period for significant  amounts of entrapped air to dissolve in the aqueous phase
 may be on the order of weeks to months. Considering the time required for entrapped
 air bubbles to dissolve in  water and the Darcy  water velocities required to dislodge
 entrapped  bubbles, the modeling assumption that entrapped nonwetting ganglia (i.e.,
 entrapped  air bubbles) are immobile is reasonable, particularly for a fluctuating water
 table  scenario where  the  duration between  fluctuations is less than several weeks.
 Incorporating dissolution of entrapped air in k-S-P constitutive models  may yield more
 accurate fluid distribution predictions. However, markedly  increased computational
 effort will be required since a transport equation for dissolved gas must be solved.

 Measured and simulated water S-P histories are compared in Figure 3.29. Experimental
 S-P relations are shown as diamond symbols and simulated S-P relations employing the
full hysteretic constitutive model (i.e., Simulation 1) are shown as solid  lines. There are
 several experimental features that  are worthy of attention in  Figure 3.29. First, note
 that scanning  S-P loops formed  by the first  rising of the water table  and subsequent
water  table lowering (i.e., t =  3 to 6.67 h)  become more pronounced as the reversal
                                      118

-------
water saturation from the main drainage branch increases. At the 70-cm-measurement
location (Figure 3.29a), the lowest water pressure head measured was approximately -45
cm, which  occurred  during the first  lowering of the water table (i.e., main  drainage
branch). During the first rising of the water table from  7 to 42 cm, the water pressure
heads increased to approximately -34 cm; however, the slope of the S-P scanning curve
is quite steep in this region and a negligible increase in water saturation is predicted,
which matched the observed behavior. This is why the predicted full hysteretic water
saturation distributions  in  Figures 3.29a and b show little response to the first table
elevation increase at t = 3 to 5 h. Slopes of water imbibition 5-P curves will always be
steeper than the drainage curves at lower water saturations (i.e.,  fo > *a).  As the water
table was lowered for a second time  from 42 to  17 cm, the measured water pressure
heads approached -45 cm and water contents very closely followed the saturation path
as for the initial wetting. During the  final rising of the  water  table from  17 cm to  the
soil  surface, the water saturation  path mirrored the path measured  during  the first
increase in water table elevations suggesting that the scanning S-P loops close.

Forming closed 5-P loops for cyclic changes  in boundary conditions can be better
observed in Figure 3.30 where the experimental 5-P history  is plotted for the 50-cm-
measurement location. Successive experimental  measurements  are connected by broken
lines in  Figure 3.30.  Note that a closed internal  5-P loop is formed during the cyclic
raising and lowering  of the water table. Also note that  during the second water table
rise  (i.e., to the soil surface), the  same 5-P path is followed as that  during  the first
increase in  water table  elevations. This closed loop behavior can be observed  for all
measurement locations shown in Figure 3.29.

Predicted  5-P relations  employing the full hysteretic model  and Parameter Set 1
matched fairly well the experimental 5-P paths. Main features  of the 5-P histories were
captured by the hysteretic constitutive relations. This  is  very encourging in  that  the
hysteretic model was calibrated only from main water drainage 5-P measurements and
an estimate of the maximum amount  of entrapped air corresponding to the main water
imbibition 5-P branch. Water imbibition 5-P relations were predicted by employing an
a (i.e., *a) of the van Genuchten retention function to be  twice  that of the drainage o
(i.e.,  a) and using the same parameter n for imbibition  as for drainage. This approach
greatly  simplifies  calibrating  the  hysteretic flow  model  and  appears  to  perform
satisfactorily, at least for the experimental conditions employed in this study.
                                      119

-------
Two  additional simulations were conducted to evaluate  consequences  of assuming  an
irreducible wetting fluid saturation, 5m, of zero. For  Simulation 4, the full hysteretic
constitutive  model  was employed. However, it  was  calibrated to the  main  water
drainage  S-P  experimental  data  assuming  STO  =   0  (i.e., Parameter Set  2). For
Simulation  5,  the  full hysteretic constitutive model  and Parameter Set  2  were
employed. However, the  tortuosity  term  in Mualem's  (1976) relative permeability
integrals was changed from Sw°-s to  5^*. That is,  in  Simulation 5  we used a slightly
different k-S model. Water distribution results from these two additional simulations are
shown in Figure 3.31 for  the  70 and 40 cm measurement  locations. Simulation 4 is
shown as broken lines with equal sized segments and  Simulation 5 is shown as broken
lines with alternating long and short segments. For comparison purposes, Simulation 1,
which is the full hysteretic model with Parameter Set  1 (i.e., Sm > 0), is also shown as
solid lines. The parameter sets employed in the simulations are listed in Table 3.5.

In Figure 3.31a, it can be observed that Simulation 4 predicts lower water saturations
during drier periods than were measured or predicted by Simulation 1. This is primarily
attributed to imposed k-S relations. In Simulation 1,  as  water  saturations approached
0.17 (i.e., Sm =  0.17), the water relative permeabilities were negligible. Hence, water
drainage  essentially  ceased  as  predicted  water saturations   approached  0.17.  In
Simulation  4,  k-S relations were  such that  water relative permeabilities would not
approach a  negligible  value  until the water saturations were close  to  zero,  which
permitted water to continually drain to lower water contents. It is clear from comparing
Simulation  1  and 4  results   to the experimental data that  an irreducible  water
saturation, or a residual water saturation  as is commonly refered to in soil physics and
hydrology literature, may be required to accurately model subsurface water behavior or
that the k-S relations of the van Genuchten-Mualem model may not be appropriate for
this soil. To further explore this supposition, we conducted Simulation 5, which used the
same parameters  as Simulation 4, but with  modified k-S relations. We changed the
tortuosity  term in Mualem's  (1976) relative permeability integrals to conform with
experimental work completed by Burdine (1953) and  Corey  (1954).  The  results  of
Simulation 5 are shown in Figure  3.31a for the 70 cm measurement location where it
can  be observed  that the predicted water  distribution compares  favorably  to the
experimental  measurements.   Furthermore,  hysteretic  S-P histories  predicted  by
                                     120

-------
 Simulation 5 compared favorably to the experimental data shown  in Figure 3.29.  At
 some measurement locations, Simulation 5 predictions yielded better agreement with
 the experimental  data than did Simulation 1 predictions. However, at other locations
 the reverse was true.

 These results  seem to suggest that employing an  irreducible water  saturation in
 numerical modeling  may  not  be necessary  provided  k-S  relations  axe described
 adequately. Considering that Mualem (1976) obtained the tortuosity  term in his relative
 permeability integrals by best fitting to data from 45 soils in which recorded S-P and k-
 S relations were available, it is not unreasonable to assume that this tortuosity factor
 may  depend on  flow channel geometry  and may  be  empirically correlated to other
 retention parameters. Mualem  (1976) noted that the tortuosity  term may be porous
 medium specific.

 In Figure 3.31b there is better  agreement among Simulations  1,  4  and 5. This is
 primarily because at the 40 cm measurement location, measured and simulated water
 saturations are significantly greater than 5m of Parameter Set 1 (i.e.,  0.17) and predicted
 k-S relations are similar for the  different simulations. For higher water contents, i.e.,  30,
 20 and 10 cm  measurement locations, Simulations 1, 4 and 5 predict almost identical
 water saturation paths. Finally, we note that the coefficient of determination (R2 value)
 obtained  from  Parameter Set 1 was 0.966 and that obtained for  Parameter Set 2 was
 0.963 (see Figure 3.27). Although this difference in R2 values is small, substantially
 different predicted water saturation paths may be obtained as observed in Figure 3.31a.
 It is  not  sufficient to accurately  describe S-P relations; k-S  relations  must also be
 accurately described.
3.8.8 Three-phase dynamic measurements

Experimental methods. A three-phase column experiment  was conducted  to evaluate
the hysteretic constitutive model with fluid entrapment effects. The experimetal setup
consisted of a 71 cm long soil column (6 cm x 6.5 cm) initially free of oil and the water
table located at the upper surface. Fluid saturations along the column were measured
•using a dual-energy gamma radiation apparatus and liquid phase pressure measurements
were taken  using  ceramic  tensiometers  connected  to   pressure  transducers.   The
                                     121

-------
experiment was  conducted  with time  dependent boundary conditions  that  can be
summarized as follows:

Stage 1: At t = 0, the water table was lowered at a rate of 1 cm per minute until it
reached an elevation of 7 cm at t = 1  h. The flow system therafter  was allowed to
redistribute for a period of 12 h.

Stage 2: At the end of stage 1,  a slug of oil containing 90% Soltrol  and  10% 1-
iodoheptane was allowed  to infiltrate from the upper  surface  under a  zero  water
equivalent  oil  head until the accumulation was 250 cm3. At the  end of the infiltration
period, the system was allowed to redistribute under natural hydraulic  conditions until
t = 17.45 h.

Stage 5: The lower surface  water table at the end of stage 1 was varied  as shown in
Figure 3.33 and the experiment was conducted until t = 22.1 h.

Soil hydraulic properties of  the medium  and the fluid  properties  of the oil mixture are
given in Table 3.6. The experimental procedure described  above was simulated using
the simplified  hysteresis model described in Section 3.2. For the simulation, oil head at
the upper and lower surfaces was fixed at zero flux condition except during infiltration.
The water  boundary  conditions at the upper surface water head  was fixed at zero flux
condition during the entire simulation.
                                     122

-------
                         Table 3.6.  Soil and fluid properties
                        Parameter
   Value
                         n
                        Pro

                        Iro

                        POO

                        &
                          OW
                           ax
0.4
2.48
0
0.05 cm'1
121 cm h'1
0.83
3.62
2.85
1.7
0.25
                        75. O
                       6O. O
                     T3 45. O
                     0
                     0
                       3O. O
                     L
                     0
                        0. O
                                         10     IS     20
                                       Time  Chrs)
Fig. 3.33  Water table elevation during column experiment.
                                      123

-------
' • U
D. a -
C
0
; O.B-,
0
L
D 0. 4 -
JJ
0
t/1
0. 2 -
0. O -
o
\
\ w"°" •«"'-«'•>"
j Oopth - 14 cm
\
•\ r—
" \ /* !• " •
• \ I*/
" * * i * /•
m \ IB g ' *^ *
IW^L-^JM " I • ^** * • '. * S •
~* ^ — — ^^f • \
v 	 J
i i 	 1 	 1 	
5 10 IS 20 25
Tima 
i . U

o. e -
C
0

~ 0. 6 -
0
L
3 C. 4 -
4J
0
l/l
0. 2 -
0. 0 -
\
\ Vota,- ...u,nr ,
\ °t0" "*"•-«">«
\ DoptM - 34 cm __
\ /'""^ '»""•"•
\ 1 I /

,_\ ^ /"""''.'"
>_ \«_ _ . /^ ^ ^
"

™ "• * " J*
* • «r •"•
*


                                                            10
                                                                   15
Fig.  3.34  Predicted and observed water saturations  for the column experiment. Solid
          line and symbols indicate predicted and observed saturations, respectively.
1. U -

0. B •
C
0
•" 0. 6 -
0
L
3 0. 4 -

0
in
D. 2 -
0. 0 -

Oil •otvi-otlon











C>pth - 14 cm

1
\ m
if fv —
v r\. .
"^\^ v^-

"•••TT"" — — "• .— '
• *
1 1
J. U -

o. e -
c
0
- 0.6-
0
L
D 0. 4 -
*J
0
in
o.z-
v.o-

Ol 1 •ct.wi-otlon
Ooptn - 34 cm









• »
\ , *
'K • •"
•\^_^^^ /^"
"•• . «^~~^
i
i i
1O :S 2O 25 10 15 20 2!
Timo Chrs) • Timo 
Fig.  3.35  Predicted and observed oil saturations for the column experiment. Solid line
          and symbols indicate predicted and observed saturations, respectively.
                                      124

-------
Predicted water and oil saturations at  14 and 34 cm depths are shown in Figures 3.34
and 3.35, respectively. Predicted water saturations at the 14 cm depth indicate a drop
in saturation at approximately  12  h due to the advancing oil front. During the same
period, water saturation at the 34 cm depth increased due to the downward movement
of water associated with the advancing oil front and decreased thereafter for a short
period as the oil front reached this depth. Fluctuations in the water table caused smaller
changes in the water saturation  at the 34 cm depth compared to the 14 cm depth since
during the most period of the simulation, the 34 cm depth remained liquid saturated.
Predicted oil saturations after 17 h at  the  14 cm depth  indicate less oil entrapment
compared to the 34 cm depth.

The results show good agreement between the  observed and predicted fluid saturations.
The  more  pronounced peak  in  oil saturation  as oil  moves  downward relative to the
observed measurements  probably reflects the  simplifications involved in treating the
saturation-capillary pressure  relations. In particular, disregarding hysteresis in apparent
water saturation produces  slightly sharper fronts than are observed.  However, the
discrepencies diminish markedly as the distance from the source  increases, indicating
that ignoring such effects will be of little consequence if predictions very near the  source
are not of primary concern.
                                     125

-------
          4.  EQUILIBRIUM-CONTROLLED MULTIPHASE TRANSPORT

4.1 Mathematical Model for Multiphase Transport

4-1.1  Governing equations

The mass conservation equations for water (w), organic liquid (o) and air (a), assuming
an  incompressible porous medium, incompressible liquid phases  and compressible gas
phase, may  be written in summation  convention for  a two-dimensional  Cartesian
domain as
                              dpaSa_         >
                            * —§r =  —      +    a                      (    J

where tf is  porosity, 5p is  the p-phase saturation,  x,  (and  !_,) are Cartesian spatial
coordinates  (i,j = 1, 2), qp. is the Darcy velocity of phase p in the i-direction, pp is the
density of phase p, Rp is the net mass transfer per unit porous media volume into ( + )
or out of (-) phase p, and t is the time.  Darcy velocities in the p-phase are defined by

                                                                             <4-2>
                                                                              »
where Kp..  is  the p-phase  conductivity tensor,  hp  =  Pp/
-------
 Boundary conditions may be stipulated as type-1 or type-2 as follows

                    hp(xt> t)  =  hpi(Xi, t)        on Sl for  t >  0               (4.4a)

                       -n    =    .  *, *       on $  f°r   t> 0
Relationships between  phase permeabilities, saturations  and pressures  and numerical
solution procedures for the flow equations have been discussed in Chapter 3.

To  model  component  transport,  continuity  and  mass  flux  equations  for  each
partitionable component in each phase must -be specified.  Mass conservation of species
a in the p-phase requires that
                          dcnnsn      dJOD.
                                              + Rap +  7OP                    (4.5)
where Cap is the concentration of the noninert a component in p-phase expressed as the
mass of a per phase volume [M L"3], Jop, is the mass flux density of a in p-phase per
porous  media cross section in the i-direction [M  L"2 T"1],  Rap is the net mass transfer
rate per porous  medium volume of species o into (-I-) or out  of (-) the p-phase [M L"3
T"1], and  7op is the net production (-f) or decay (-) of  a within phase p per porous
medium volume due to reactions within the p-phase [M L"3 T'1] described by

                                 fop =   ~ "op Cap                            (4-6)

where nap is an apparent first-order decay coefficient.

The mass  flux  density of component o  in phase p due  to  convection, diffusion and
mechanical dispersion is described by

                                                    dC
                         Jopi =  Cop qpi - t Sp Dapij -gg                     (4.7)


where Dopij is a dispersion tensor given by
                                    127

-------
                                                                               (4-8)

in which  D^ is the molecular diffusion  coefficient of o in the p-phase of the porous
medium,  Dptyd is a  mechanical dispersion coefficient, and gap is a nondilute  solution
correction factor.  For the case of transport of low solubility organic components, small
volume fractions of organic components in water and gaseous phases will occur  and gQW
 =  gaa =  1 is assumed.  For nonaqueous  phase liquids, the phase composition may
reach 100% (e.g., single component organic liquid) at which point  dispersive transport
becomes  nonexistent.  We  accommodate nondilute  solution  diffusion-dispersion  by
approximating the oil phase nonideal solution factor by

                                 900 = 1 - C00/.Po                             (4.9)

Employing the tortuosity model of Millington and Quirk (1959) yields the expression for
*^op
                              J~> d\1  _  jl/3  C 7/3  r>o                         1 A  in\
                              Vap   ~  '   Jp    ISap                        (4-lUJ

where D^p is the diffusion coefficient of o in bulk p-phase.  The  mechanical dispersion
coefficient is shown by Bear (1972) to have the form
                                                                              (4.n)
where AL and AT are longitudinal and transverse dispersivities [L], qpi and qpj are p-
phase  Darcy  velocities  in the  t  and  j  directions,  ~qp =  /E^,2/1/2  is  the absolute
magnitude of the p-phase velocity, and  £,- •  is Kronecker's delta.

Combining the phase continuity equation and the mass flux equation for transport of
component o in the p-phase yields
Expanding the first and third terms in (4.12), employing the bulk p-phase continuity
equation (4.1), and assuming density derivative terms to be of second order importance
                                     128

-------
within a given time-step yields
                              C       „  dC<>p  ,
                                    ~ 9" ~~
                                                     op
where the total  phase mass transfer rate,  Rp, it  may be noted, is related  to  the
individual component mass transfer rates by

                                  RP =  £  Rap                             (4-14)
                                       0 = 1
where n, denotes the number of "noninert" or partitionable species.  In the present
context, we will use the term "inert" to refer to components of the NAPL phase which
are for  practical purposes insoluble and nonvolatile.  For the three fluid phase system,
(4.13) constitutes a system of three equations which we may write for the water phase
(p = w), the organic liquid phase (p = o) and the gas phase (p = a) as
                                                                           (4.15b)
                                               +  Raa - („„  +     ) Coa     (4.15c)
To accommodate adsorption of a by the solid phase, an additional continuity equation
is required which may be written as

                              dCg,  _  p        r                           u -\R\
                              -0T  ~  R°' ~ "« C<"                         (4-16)
where  Cat is the solid phase concentration expressed as mass of adsorbed component a
per porous medium volume [M L"3], nat is the first-order decay coefficient (T"1) in the
solid phase and Rai is the mass transfer rate per porous medium volume to (+ ) or from
(-) the solid phase [M I/3 T1].
                                     129

-------
 4-1.2  Phase-summed equations for local equilibrium transport

 Coupling between the phase transport equations arises due to the interphase transfer
 terms.   Explicit consideration  of interphase transfer  kinetics may often be justifiably
 avoided  by assuming  phase transfer to be equilibrium controlled.  We consider first the
 case of linear partitioning and introduce the approximate thermodynamic relations

                                   Cao  =  T00 Cow                             (4.17a)
                                   Caa  =  roa Cow                             (4.17b)
                                   Ca.  =  ra. Cow                             (4.i7c)

 where roo  is the  equilibrium  partition coefficient for species o  between  water  and
 organic liquid  (Raoult's  constant), Toa is the  equilibrium partition coefficient between
 water  and  gas (Henry's constant), and Tat  is a dimensionless equilibrium partition
 coefficient between water  and  solid phase.  Using the equilibrium relations, we may
 rewrite the phase transport  equations  in  terms of a single-phase concentration.   For
 water-wet systems, it  is logical  to retain the water phase concentration, since water will
 always be present in  the system.  Using  (4.17) to eliminate oil,  gas  and solid phase
 concentrations from (4.15) and summing the equations noting that

                           Raw + Rao  +  Ra. +  Raa =  0                      (4-18)

leads to the a-component  phase-summed  transport equation in terms of water phase
 concentrations

                 *:

where
                      *« =  *$«,  + * SOTOO  + * 50  roa  +  ra.                (4.i9b)

                Dafj  =   Sw Dawi; + 4 S0 Dooij roo  + * 5a Doaij Taa          (4.19c)

                           9«*  =  9«i + 9oi roo  + qai roa                     (4.19d)
                                                      D      /? r      /? r
          *  —       1     T»    I     T»    I      T*   I    W  i  ^O CrO  I    fl   QQ   / A i r\ \
         it  ~  u   •+•  u  1   -T- ii  1   -T- 11   1    -I-     U-   .-"v -L      "u   (A 1 Qo i
         '^^k     ~^t*»  '  ~f*n  nn  i  ™*v*i  rt/i i  ~f*» * /v« i   /i    I    /)     f    /)      1~*X^C f
                                       130

-------
Note that (4.19)  has  the same form as the  simple single-phase transport  equation.
However, the coefficients represent pooled effects of transport in all phases.  Note also
that interphase mass transfer terms occur in the phase summed  equation only as a sum
over all components.

An alternative form of the phase-summed transport equation may be written in terms
of the oil phase  species concentrations.  The a-component phase-summed  transport
equation in terms  of oil phase concentrations has the form
                                    **
                                                             ** r>           (A t)n.\
                                                          " "«  GO°         (4.20a)
where
                                  i   V'-'w   i  v-'o x oo   i  •* as                i A onv^
                                 +  -P—  -(-  —f	  -I- TT—                (4.20b)
                                      oo        oo         oo
                                                   i   T^a^aaij  a a           i/, nr\\
                                                  -I-   	f	           (4.20c)
                                                          ^00
            **
                                                                            (4.20d)
                                                                            (4.20e)
The form of the oil-based equation is identical to that of the water-based phase-summed
equation.   The water-based  equation  has the conceptual advantage that  Caw  is
physically meaningful at all  locations in  the  porous medium, whereas Cao only has
physical meaning at locations where S0>0.  The latter constraint, however, imposes no
mathematical difficulty since (4.17) can in  any  event be employed as a definition of Cao
as well  as  Caa whether or not the phase exists.   Since most of the species mass
commonly resides in the oil  phase for  organic  liquid  spill  simulations, the oil-based
equation has the advantage that mass  balance accuracy  of the solution may be more
readily controlled.
                                     131

-------
 When simulating two-phase liquid flow  consisting of water and oil with  constant gas
 pressure, only diffusive transport is  considered in the  gas phase — i.e., qa  =  0  is
 assumed for this case.  It should  be noted that even  if gas pressure  gradients have
 negligible effect on liquid phase flow,  this does not strictly imply zero gas flow.  If fluid
 saturations change in time, gas  flow  will occur.  Furthermore, phase partitioning may
 result in density gradients in the gas phase which can induce significant flow.

 4-1.3  Initial and boundary conditions

 Since the transport equation is  written  in the phase-summed form, it  is necessary  to
 specify  initial  and boundary  conditions  in terms of  a single concentration.   In the
 following, as well as  in the program itself, we will specify initial conditions and  type-1
 boundary conditions  in terms of water phase concentrations.  If the oil concentration
 form of the phase-summed transport equation is employed, oil phase concentrations are
 internally computed  from  the  specified water  phase  values  and  the  equilibrium
 relations.   Type-3 or flux-type boundary conditions  are defined in terms of  phase-
 summed fluxes.  Initial conditions for  component a are given as

          Cow(Xi, t = 0} = COWo(x,)              on  R for t =  0             (4.21a)

where Caw (i,) represents  the initial water phase concentration of component a  at
           O
location  i,.  Note that the concentrations of o in all other phases are implicit in (4.21a)
via the phase  partition relations (4.17).  Boundary conditions for  the  phase-summed
equation are stipulated by

          CUM) =  Cawi(xi,t)                  on  5:  for  t > 0           (4.21b)

                 =   0                           on  52 for t> 0              (4.21c)

                                            /„    on  S3   for O 0             (4.21d)
                                         3

Equation (4.21b) describes a type-1  boundary condition on  boundary  region  5j with
specified o-species concentration Cawi(xitt) within the porous medium at the boundary.
Note again  that stipulation of the  water phase  concentration simultaneously fixes the
organic liquid,  gaseous and solid  phase concentrations via the partition relations.  Type-
1 boundary conditions for transport will  be used rarely  since direct control over porous

                                      132

-------
media concentration seldom occurs.   One case which  may  be relevant would  be the
occurrence of a source of known concentration where diffusion is the dominant transport
mechanism (a type-3 condition would be used if significant convection occurred at the
boundary).

Equation (4.21c) describes type-2 boundaries with zero normal concentration gradient
specified on region S2. The zero gradient condition indicates zero normal dispersive flux
at the boundary.  If normal fluid velocities are nonzero, transport across the boundary
will be permitted by convection.  This boundary  condition is commonly applied at
boundaries  which   have  fluid efflux  to  permit  constituent transport  across  the
boundaries.  In  the  event that fluid fluxes are zero  normal to  the boundary, condition
(4.21c) stipulates zero component flux.  Caution needs  to be exercised in applying the
type-2 condition on  boundaries with inward fluid velocities of uncontaminated fluid. If
the  concentration at the boundary is  nonzero due to presence of chemical  near the
surface,  the  type-2  condition will stipulate an inward  mass flux density equal to the
water phase concentration at the boundary times the phase-summed velocity, g*, which
is  incorrect.  In  such  a  case,  a  type-3  boundary  condition  with zero  influent
concentration should be specified to  force the desired zero-flux condition.   The type-2
zero gradient condition is the default condition.

Equation (4.21d) is a type-3 boundary condition for  use on boundary regions, 53, which
have inward NAPL  phase fluxes predicated on the assumption  that contaminant enters
the system solely via inflow of nonaqueous phase liquid. Here, qon is the normal Darcy
velocity  of the  oil  phase at  the boundary  and  C£{,fe  is  the  equivalent water phase
concentration of species a in the influent oil  phase  defined by C£^e = C^/T^ where
Cof/ is the  actual  concentration of  a  in the influent oil.   Note  that applying the
condition C^e  = 0  corresponds to imposing zero mass flux of component a  on the
boundary segment.  In applying the type-3 boundary  condition during periods of oil
infiltration,  it  is   important  to ensure  consistency in  the  specified  oil  phase
concentrations such  that
                                                                             (4.22)
where Cao (a = 1, . . . , n,) is the  oil  phase concentration expressed  as  mass  of
component per phase  volume  for partitionable component a and Cj is the oil phase

                                      133

-------
 concentration of  "inert" or nonpartitionable components.  Equivalent!}-, in terms of
 mass fractions the condition may be expressed as

                                 // + £  /00  =  1                            (4-23)
                                     0 = 1
 where fao and fj are mass fractions  of partitionable and inert components, respectively,
 in the oil phase.   Mass fractions and phase concentrations are related by

                                   Cap  =  fap Pp                             (4.24)

 where pp is  the  bulk  phase density.  Simple  mixing theory permits component  mass
 fractions and volume fractions to be related as
Cr      n»
77  +   £
                                                    *                         (4-25)
                                     0=1
where pa  and  pj  are  the  densities  of pure  noninert  components  and inert  oil,
respectively.  The oil phase density may be related to phase composition by

                                                                              (4-26)
                                    fl/Pj + £
                                            0 = 1
In  applying  the  type-3  oil  infiltration  boundary  condition,  influent  oil  phase
concentrations, C£^e must be specified such that (4.22) - (4.26) are obeyed.
4.2  Numerical Model Description

4-2.1 Central solution approach

Fluid flow equations are highly coupled with each other due to interdependence of fluid
permeabilities, saturations and pressures mandating a simultaneous solution approach as
discussed in the  previous  chapter.   The flow equations are also coupled  with the
transport equations via  the occurrence of interphase mass transfer terms and through
the dependence of various coefficients in the flow equations on fluid composition — e.0.,
phase density, viscosity, scaling factors, etc.  Since mass transfer rates are necessarily

                                      134

-------
small for the case of low solubility organic fluids and changes in fluid properties over
short time periods will accordingly also be small, the dependence of the flow equations
on transport  will be  very  mild over  short  time  spans permitting  computational
decoupling.  Over extended periods of time, cumulative phase mass transfer may have
significant effects on  flow as dissolution and volatilization deplete the nonaqueous phase
liquid requiring some means of updating phase transfer rates as well as compositionally
sensitive fluid  properties.  We will consider fluid compositional effects on phase density
but disregards effects on other coefficients. Interphase mass  transfer rates and phase
densities are time-lagged in the flow  solution and updated at the end of each time-step
after solution of the transport equations.

The transport  equations are highly dependent on  the solution  of the flow equations due
to the occurrence of  fluid velocity and phase saturation terms directly in the transport
equations and  in the functional forms for the dispersion coefficients.   Thus, the flow
equations must be solved concurrently with or prior to evaluating transport.  Due to the
weak back-coupling, the most opportunistic approach is to solve the transport equations
serially with the  flow equations.  Furthermore, since the individual component transport
equations  axe  weakly  coupled  with each  other  in  the  present  model  due  to the
assumption that  components do  not  interact,  the transport equations for  different
components may be solved serially in arbitrary sequence.

The solution approach for solving the coupled  multiphase flow and  multicomponent
transport problem with equilibrium mass transfer is as follows:

1.   Solve the fluid flow equations simultaneously for the current time-step using time-
     lagged phase densities and interphase mass transfer rates.

2.   Solve the phase-summed transport equation using the previous time-step phase
     densities  and interphase mass transfer rates.

3.   Back-calculate  new interphase mass transfer rates and update  phase densities for
     the current time-step.

4.   Proceed to  the next time-step.
                                      135

-------
4-2.2  Finite, dement formulation

The phase-summed  transport equations  given  by (4.19) or (4.20) are solved  using an
upstream-weighted Galerkin finite  element method with linear  rectangular elements.
We  will  present  the solution  for  (4.19).   The  formulation in  terms of  (4.20)  is
fundamentally  identical.   Employing linear shape functions, Nj,  the concentration
C*OU)(ij,<) can be approximated by
                                      4
                         Caw(xt,t)  = £  Nfa) CowJ(t)                   (4.27)

where  CQWj denotes the  concentration at node  J and c and rj  are local coordinates.
Detailed  information  pertaining  to the  upstream  weighting  functions  have been
discussed   by   Kaluarachchi  and  Parker  (1989).    The  Galerkin   finite  element
approximation of (4.19) over the entire domain R  can be written as

                               J NjL(Caw)dR  = 0                         (4.28)
                               R
which leads to the set of equations given by
                       [Pi iCkw}  +  M {     *}   =   {R}                 (4.29)
Using the type-3 boundary condition given by (4.21d) and disregarding cross derivative
terms, the matrices [P], [M\ and {R} are defined by
 fj                     N
£  [NI»ZNJdR -   £   <9*>  / WjN
 = lJZe                c = 1        S
                +         NI»NJdR  -        <9>    WjNjdS         (4.30a)
                         MIJ =   E       ^/ *Z NJ dR                   (4-30b)
                                  « = 1   Re

                               N>                    r
                    RI =  -  E     rao C^i/e /  W7 dS              (4.30c)
                              e = l
                                     136

-------
where N is the total number of elements, Nt is the total number of boundary elements,
 WT is the asymmetric  upstream weighting function,  is the average normal oil
flux along the surface 5 where contaminant enters the system and  is the average
normal phase-summed velocity on surface 5.  If an evaporative boundary is present, the
term  in  (4.30c)  will have an additional term (-.D^1/ roo) to account for the
atmospheric boundary layer and the terms containing C^J,' in (4.30c) will vanish.

Treatment of  the integrals in  (4.30)  employs the method of influence  coefficients
described in the flow analysis and will not be repeated here. Handling of the convective
term in (4.30a) using the influence coefficient method has  not been previously described
and will be discussed below. Phase-summed velocities within an element are represented
by
                                        4
                          fcf (**-,«) =   E  *M foSW                     (4-31)
                                      9=1

where qQf*p is the phase-summed velocity at the evaluation points in the element,  which
may be at the nodes, at Gauss points or elsewhere as controlled by the parameter GP
(see discussion  in flow analysis).  Using equation  (4.31) on the  convective  term of
(4.30a) and using linear upstream weighting shape functions (Kaluarachchi and Parker,
1989), it is possible to show that
                                                 dR =
                              	   J   l ""  "*!
                             e = lR*
}
/
                                                            +  [Hv9}a}        (4-32)
where m  and d are the width and height of a rectangular element, x and y axe the
coordinate axes and matrices [H\ with superscripts xg and yg refer to the contributions
from standard linear shape functions while matrices with superscripts xu and yu refer to
contributions  from  upstream  weighting coefficients.   Details  of matrices  [H\  are
described  in detail by Kaluarachchi and Parker (1989) and will not be repeated here.

The surface integrals in (4.30a) and (4.30c) can be described by
                                     137

-------
                                    =     (4±3w)   if I=J               (4.33a)

                              =  ^ (2  ± 3w)  if  7?fc 7

and
                                     dS  =   (1 ± w)                      (4.33b)
where £ is  the  length  of the side and  w is the upstream weighting coefficient.  The
appropriate sign of the integrals given in (4.33) will be determined by the side of the
element exposed to the boundary condition.

Time integration of (4.29) is performed using a standard finite difference approximation.
The resulting final set of equations can be written as
                                                                              (4<34)
where k and k + 1 refer to previous  and current  time-steps and 6 is a time weighting
factor.    In  the  present  work,  we  use  a  Crank- Nicolson  time-stepping  scheme
corresponding to  6 = 0.5 to provide an unconditionally  stable second-order accurate
temporal approximation.  Time-step size is  varied under program  control by a time
incremental factor which increases or decreases time-step size depending on the number
of iterations required for solution of the flow problem.  The time-step size for solution of
the transport problem is taken as identical to that for the flow problem.  If equilibrium
mass transfer is assumed the transport equation is linear and no iteration of the solution
is required. For the case of nonequilibrium mass transfer, nonlinearities are handled by
a  simple Picard iteration scheme with apparent  partition coefficients, mass  transfer
rates and phase densities updated at each iteration.

4-2.S  Interphase mass transfer and density updating

After solution of the phase-summed transport equations, the interphase mass  transfer
terms Rap must be updated.  This  is performed  after  each iteration of the transport
equations in the case of nonequilibrium mass transfer.   For the equilibrium case, the

                                      138

-------
transport model is  linear and  no iteration is  required.  Hence,  for the latter  case,
updating is performed only once at the end of  each time-step.  To compute the  mass
transfer rates, the concentrations of each component in the oil, gas  and solid  phases are
first calculated from the partition coefficients and from the control phase concentrations
(Caw for the case of eq.  4.19 or Cao for eq. 4.20).   The mass transfer rate terms, Rap.
are computed  by back-substitution into the component transport equations for each
phase (eqs.  4.15  and 4.16) using a finite difference approximation.  Summing over the
components  yields the  desired Rp terms for each phase  for  use in the  subsequent
solution of the flow and phase-summed transport equations.

For equilibrium mass transfer problems, effects  of mass  transfer terms on the flow and
transport equation solutions at  any given time-step  will be quite  small.  However,
cumulative  effects can be very important  since these terms will control the long  term
removal  of  NAPL associated with dissolution and/or volatilization.   The accuracy of
interphase mass transfer calculations will generally be most troublesome during periods
of highly transient NAPL flow. During such periods,  it may  be preferable to suppress
mass transfer updating to avoid numerical instability as well as to reduce computational
effort and improve accuracy.

With increasing simulation time, the density of each phase will change depending on the
net  mass  of each  component leaving or  entering  the phase.     Although density
derivatives were  neglected in the equation development, on the assumption  that  these
terms are small over a given time-step, cumulative changes in density  cannot  be ignored
since over long times these can accumulate and become quite large in magnitude. To
accommodate these effects,  liquid phase densities are updated at the  end of each time-
step as follows
                                       nt   n        n>
                               O  f 1    ^"^  ^QtU 1  i   V™*  /t                   /1 n v \
                       Pw -  Pw I * - 2^ ~f~ J  +  Z-  C™                 (4.35a)
                                     a=l        a=l

                                       n,   Q        n,
                        P0  =  Pj [ 1 ~ ^_- -p22 ]  "I"  2_s  ^oo                 (4.35b)
                                     o=l         o=l

where nt denotes the number of "noninert" or partitionable species, p°w is the density of
uncontaminated  water,  pa  is the density of pure a-component,  pr  is the  density of
"inert" oil components, and Ma is the molecular weight of component  a. Density ratios,

                                     139

-------
 pru), pro and pra, are computed as the ratio of pw, p0 and pa to p°.

 For the gas  phase,  density  depends on  both  pressure and composition.  Pressure
 dependence is treated as a nonlinear effect in the gas flow equation according to

                                Pa = *ha  + Pa" + Pa'                          (4'36)

 where A is the gas compressibility taken to be 1.17 x 10~6 g cm'4, pa° is the density of
 native soil air taken to be 1.12 x 10~3 g cm"3, and pac is the density of contaminants in
 the gas phase computed as
                                  Pac=  £  Caa                             (4.37)
                                       0 = 1
where Coa is the gas phase concentration of species a and  na is the number of noninert
organic species.
4.3 Model Applications

4-3.1 Example 1.  1-D Spill of Two-Component Hydrocarbon

The first example problem involves a spill of a mixture of toluene and o-xylene at the
soil surface followed by gradual leaching by rainfall.  The domain of the problem is a
200 cm vertical section with a water table at a depth of 150 cm described by a single
column of forty 5 x 5 cm elements using 82 nodes. Properties of the porous medium are
given in Table 4.1. Note that since the problem is one dimensional, Kfw alone controls
the flow.   However, since the solution employs two-dimensional elements Ktw   is still
employed.   To minimize the possibility of oscillations in 1-D vertical problems, it  is
advisable to input Ktw > Klw .  Here, we use Ktw  — 2 Ktw  . The nonaqueous liquid
entering the system is assumed to be composed of 50% mass fractions of toluene and o-
xylene.  The  oil specific gravity,  relative viscosity  and scaling factors of the mixture
estimated as mass fraction-weighted  averages of the individual  component values are
given in Table 4.1.  Component values of air, oil and water phase diffusion coefficients,
and oil-water and air-water partition  coefficients are also  summarized in  Table 4.1.
Solid-water partition coefficients and decay coefficients were  assumed to  be zero for
both components.  The problem is simulated in three stages involving one initial run

                                      140

-------
and two restart runs as follows:

          Stage 1: Constant head oil infiltration
          Stage 2: Redistribution with transport for 25 days
          Stage 3: Constant flux water infiltration for 100 days

In  Stage 1, oil  infiltration is considered  into soil which is initially oil-free  and in
equilibrium with a water  table at a depth of 150 cm.  Oil is added under a constant
head of  h0 — 0 cm at  the  top  two nodes until the cumulative oil infiltration is  21 cm3
( = 4.05  cm3cm"2). Other  boundaries for  oil are  no flow.  The  bottom  nodes are
maintained at a constant  water head  of hw = 50  cm corresponding  to the  initial
conditions throughout Stage 1  and other boundaries are no  flow for water.  Transport is
not considered during  Stage 1  since the total duration of the first stage is only 0.0074  d.
At the end of infiltration, the oil front is  at a depth of approximately  20 cm  (Figure
4.1).

Stage  2  involves a 25 day period during which oil redistribution occurs with no flow
conditions at the upper boundary.  Initial conditions  for flow are read from the restart
file. Transport is considered during Stage 2 with  initial conditions computed internally.
Initial oil phase concentrations of toluene and  o-xylene  corresponding to  50%  mass
fractions  are 436.7 mg cm"3.  Equilibrium  water  phase concentrations for toluene and
xylene are 0.2583 and 0.0762 mg cm"3, respectively.  Stage 2 is carried out for a period
of 25 days with no flow boundary conditions for  oil and with boundary conditions for
water  as in Stage 1.   Type-3 boundary  conditions for transport  are  employed with the
influent  concentration, €%£> equal to zero  corresponding to zero component mass flux
on all  boundaries.  The water  and total liquid saturation profiles at the end of Stage 2
are shown in Figure 4.1.   Note that a residual oil saturation of about  7% occurs in the
upper  unsaturated zone due to a limitation on  drainage  under gravity  when  the oil
permeability is low.   Oil permeability  increases near the water  table  as  the water
saturation increases resulting in faster oil drainage near the  water table.
                                      141

-------
 Table 4.1  Input parameters for Example 1.
Soil properties:                       Bulk fluid properties:
     KIWi =  800 cm cf1               /3ao  =  2.1           pow  =  1.83
     K.w'z =  400 cm fl               r,ro  =  0.695         pro  =  0.873
         =  0.4                     Component properties:
     Sm   =  0.05                           o-Xylene      Toluene
     5or  =  0.2                     Da°w  =  0.620       0.821
     o    = 0.05  cm"1                 Da°0  =  0.70        0.987  cm2^1
     n    = 2.5                      D0°a  =  6099.       6765.  cm2^1
     AL   = 2.0 cm                  roo   =  5729.       1683.
     AT   =0.2 cm                  roo   =  0.22        0.28
                                      pa.   =  880.        862. mg cm'3
                                     142

-------
                               JOO.O -i
                                                                    Stoje 1
150.0 •
'•
-g- :
u, :
JS 100.0 -
.Q :
•
•
•
•
«
\
v
X.
^•^
»••«) St
*-»-*-*-*s*




                                50.0 -
                                 0.0'
                                   0.03
                                           'olio'
0.4O     0.60
  Seturction
                                                                     aoo
                                200.0 -,
                                150.0 -
                              £100.0
                              0.
                                 50.0 -
                                  0.0
                                    0.00
                                            0.20
                                                    0.4O     0.60
                                                      Soturetien
                                                                    Stc«e 2
                                                                     0.60
                                                                             1.03
                                200.0 -i
                                 150.0
                               5*100.0

                               I



                                  50.0
                                   0.0
                                                                     Stoje 3
                                     o.oo     0.20     0.40     0.60	blab	"i!6o
                                                       Saturation
Figure 4.1  Water and total liquid saturation vs depth at end of Stages  1, 2 and 3 for Example 1.
                                                     143

-------
  At the beginning of Stage 3, water infiltration is initiated at the upper surface at a
  constant flux of qw =  10 cm d~l via a type-2 flow boundary condition.  The  water head
  was maintained as in  the previous stages at 50 cm at the bottom of the column.  Zero
  oil flux was  assumed on all boundaries.  Initial conditions for flow and transport were
  read from a restart file from the end of Stage 2. Assuming infiltration of  clean water, a
  type-3 transport  boundary condition with C^w' — 0 was imposed at the  upper surface.
  At the lower surface and elsewhere a zero concentration gradient (type-2  condition) was
  employed to  allow  convective mass  loss in the water phase  from the bottom of the
  column.

  Concentrations of xylene and toluene at the lower boundary are shown as a function of
  time in Figure 4.2.  Due to its higher  solubility, toluene is lost from the system more
  rapidly than  xylene leading to an increase in the  mass  fraction of xylene in  the  oil
  phase, and hence an increase in aqueous phase  concentration at short times, followed by
  a reduction as extraction in the water phase continues.
                0.3
                Q 25
              E 0.2-
               : 0.15-
                a as-
	Toluene
	Xylene
                        ZQOO    40.00    eaao    earn    taaao   120.00
                                    Time
Figure 4.2  Aqueous xylene and toluene concentrations at outflow boundary for Example 1.
                                       144

-------
4.3.2 Example 2.  2-D Planar Spill with Sloping Water Table

This problem involves a light hydrocarbon spill in a two-dimensional Cartesian section
through saturated and unsaturated zones with a water table gradient.  The problem is
simulated in two stages involving one initial run and one restart run as follows:

          Stage 1: Constant head oil infiltration on strip
          Stage 2: Redistribution with transport for 25 days

The domain is 8 m in the vertical and 11 m in the horizontal with a water table initially
located 4 m from the bottom on the left  boundary and 3.5 m on the right.  The lower
surface is an aquitard  and the top is  the soil surface. An oil leak occurs on a strip 2 m
in width centered 5 m from the left boundary along the top.  Spatial  discretization is
achieved with 88 elements of 1 x 1  m spacing for a total of 108 nodes.  Initially  the
water table is assumed to vary linearly  from the left to right boundaries  and water
pressures are hydrostatic with no oil  in the system. Boundary conditions for the water
phase in Stage 1 are no flow on the top and  bottom boundaries and prescribed head on
the side corresponding to the initial conditions allowing  water  flow in the aquifer from
the left to the right. Boundary  conditions for oil are no flow except on  the strip source.
(Note: this presumes the physical domain is large enough that free oil never reaches the
boundaries.) At the beginning of Stage 1, oil infiltration  commences on the strip source
under a constant head of h0  =  -0.1 m and is continued until the total  accumulation of
oil is 1.0 m3 (per m in the third dimension).  The organic liquid is assumed to be 89.5%
inert oil and 10.5% benzene.  Soil properties and bulk fluid properties are given in Table
4.2.  Transport is not simulated  during Stage 1 which lasts for only 4.17  d.

During Stage 2, oil redistribution  and benzene transport is simulated subject  to  the
same boundary conditions for the water phase as in Stage 1 and with no flow for oil on
all  boundaries.   A zero concentration  gradient was  imposed  on the downstream
boundary to allow  mass loss  with  the water phase by convection.  The zero gradient
condition was employed  on  the upper and  lower surface which have zero fluid flux
resulting in  the equivalent of a zero mass flux condition for transport. On the upstream
boundary, a type-3 condition with zero influent concentration was imposed.  Stage 2
was continued for a period of 25 days during which time the solution indicated a mass
loss of benzene due to  dissolution and transport of about  2%. Contours  of oil saturation
and water and gas concentrations at the end of Stage 2 are shown in Figures 4.3 - 4.5.

                                      145

-------
Table 4.2  Input parameters for Example 2.
Soil properties:
Ktw = 10.0 m d'1
Ktw = 5.0 m fl
$ — 0.35
5m = 0.05
Sor = 0.2
o = 5.0 m"1
n = 2.8
AL = 0.2 m
AT = 0.04 m
Bulk fluid properties:
Pao = 2.69 /3OW
Iro = 2-0 Pro
Component properties
D0°w = 0.94x10-" m'
D°0 = 1.13xlO'4 m'
^^ O f\ 7CO MVk* J"l
^/ =^ U. /OO T7Z O
roo = 493.
Toa = 0.24
Pa = 0.87 xlO6 5

= 1.59
= 0.832
(benzene):
!(Ta
lfl



m-3
                  1.0   2.0    3.0   4.0
                                        5.0
                                             6.0
                                                   7.0
                                                        B.O
                                                             9.0
                                                                  10.0
                                                                        11.0
                                                                          o
  o

11.£P
                  1.0
                        2.0
                             3.0
                                   4.0
S.O    6.0

 X(m)
                                                   7.0
                                                        6.0
                                                              9.0   10.0
        Figure 4.3  Oil saturation contours at the end of Stage 2 for Example 2.
                                        146

-------
                        1.0   20    3.0    4.0   5.0   6.0    7.0    6.0    9.0   10.0  11.0
                              2.0    3.0    4.0   5.0    6.0    7.0    6.0   9.0    10.0   11.0
                                                                                    0.0
                   0.0
Figure 4.4 Water phase concentration contours (g m~3) at the end of Stage 2 for Example 2.
                    00    1.0    2.0   3.0    4.0    S.O    6.0    7.0   6.0    9.0   10.0   11.0
                                     3.0    4.0    S.O    6.0   7.0    6.0    9.0   10.0   11.0
                  0.0
 Figure 4.5  Gas phase concentration contours (g m"3) at the end of Stage 2 for Example 2.
                                             147

-------
 4-8.8  Example S. 2-D Radial Dense Solvent Spill with Vacuum Extraction

 This problem involves  a spill of tetrachloroethylene (PER)  in  a radial domain and
 remediation using vacuum extraction.  The simulation is performed in three  stages with
 two restarts as follows:

          Stage 1: Oil infiltration event
          Stage 2: Redistribution and transport under natural gradients
          Stage 3: Remediation using gas vacuum extraction

 The first stage of the problem involves  infiltration  of PER on a circular area with a
 radius of 2.0  m at the soil surface.  A total of 3 m'3 of NAPL is assumed to infiltrate
 under a water-equivalent oil head of -0.25 m.  A water table occurs at a depth of 3.2 m
 and an impermeable layer occurs 5 m below the surface.   The soil is uniform  and has
 properties given in Table 4.3.  Fluid properties for PER are also given in the table. The
 problem is  analyzed as a 2-D  radial section with an inner radius of 0.1 m (in order to
 facilitate  subsequent analysis of pumping from a well of the same radius) and an outer
 radius of 8 m. A  mesh  with  12 nodes in the vertical direction and  16  nodes in the
 horizontal direction is employed.

 Initial conditions for the water phase are assumed to  correspond to equilibrium with the
 water  table with no oil.  Boundary  conditions for  the water phase involve a type-1
 condition on the right side below the water table corresponding to vertical  hydrostatic
 conditions (i.e., same as the initial conditions).  All other boundaries are no flow for the
 water phase.  Boundary conditions for the oil phase are a type-1 condition  with h0 = -
 0.25 TO on the 5 nodes on the upper surface between r = 0 and r = 2.0 m.  All other
nodes are no flow for oil. Gas flow was not considered during infiltration.  Termination
of Stage 1 occurred at t = 6.48 d after infiltration of 3.128 m3 of oil.

 Stage 2 of the problem involves a continuation of the Stage 1 using the restart  option.
 Redistribution of the PER is  permitted  for a period of 25 d under no flow boundary
 conditions for the oil phase on all boundaries.  Boundary conditions for the water phase
 are maintained as  in the infiltration stage.    Gas   phase flow  is  again disregarded.
 Transport is considered with the initial aqueous concentration  of PER at nodes with oil
phase set to the solubility of 150 g m'3.  Boundary  conditions for transport are type-2
 (zero dispersive flux) on  all boundaries which is equivalent to zero flux since there is

                                      148

-------
essentially no flow on the boundaries during Stage 2.  By the end of the redistribution
period, the  NAPL plume has  reached the  lower  aquifer  boundary.   However,  a
substantial volume of the spill is still retained in the zone above the water table as
"residual saturation"  retained by capillary forces.  The oil saturation distribution at the
end of Stage  2 is shown in Figure 4.6.

In Stage 3, remediation is simulated with vacuum pumping in the unsaturated zone. A
vacuum well with a 0.1 m radius is assumed to be placed at the left boundary screened
over a 2.5 m interval from a depth of 1.25 m to 3.25 m and regulated at a pressure head
of ha = -1.5 m.  The gas boundary at the well is treated as a type-1 boundary condition.
Gas inflow is permitted along the upper surface from r = 3.5  m to the outer perimeter
under  a type-1 condition for gas with a constant pressure of ha = 0.  The inner 3.5 m of
the surface, which is  assumed covered, and other boundaries are treated as no flow for
the gas phase. Water head is prescribed on the right boundary below the water table as
in the earlier simulations  and all other boundaries are treated as no flow for water. No
flow conditions are imposed for the oil phase on all boundaries.  Boundary conditions for
transport are type-3 with zero  influent  concentration on  the top  and  right  side
boundaries and zero dispersive flux  elsewhere.  Gas  pumping results in water  table
upwelling as  shown in Figure 4.7.  The gas flow rate stabilizes at 319 m3 d~l after 2 days
and the PER recovery rate reaches a corresponding steady rate of  16.77 kg d'1.  PER
mass in the soil vs venting time (Figure 4.8) shows a small increase ( ~ 1.5%) during the
first  day due to numerical error before the solution  stabilizes and mass  removal  rate
becomes nearly constant.
Table 4.3 Input parameters for Example 3.
Soil properties:
K»w^
K.w*
* *
sm
SOT
a
n
AL
A f
=
=
=
=
=
=
=
zz
:=
5.0 m a
2.0 m 
-------
     £00
.10     100     190     260    370
     403 H
                                           450
                                                  550     640
                                                                750
 10     100    190     280    370     460     550     640
                            R(cm)
                                                                       503
                                                                       400
                                                                       300
                                                                       200
                                                                       100
                                                                713
  Figure 4.6  Oil saturation contours at the end of Stage 2 for Example 3.
       10.0    100.0   190.0    260.0    370.0   460.0   550.0   640.0   7300
500.0
403.0
300.0
—N
E

N 200.0
100.0

1C
II 1 1 1 1 1 1
°"s
" \>^r 	 °35 	 -ass
^^^~ 	 —0.55 	 0.55 	
0^ ^-0.95 	

^ , ^~^"-93
.0 100.0 190.0 260.0 370.0 460.0 SSO.O 640.0 730.0
                                                                       5OO.O
                                                                     H 400.0
                                                                      -\ 300.0
                                                                     ^200.0
                                                                      •4 100.0
                                                                        0.0
                                     R(cm)

Figure 4.7  Water saturation contours at the end of Stage 3 for Example 3.
                                   150

-------
  4.75E+C06 •=


  4.70E4006 •!


  4.65E+006 \
  4.60E-f006
           0.00
           *  I I  I [  I tI I  I I I  i
5.00          10.00          15.00
     Time(days)
Figure 4.8 PER mass in system vs time during Stage 3 of Example 3.
                           151

-------
   5. TRANSPORT WITH NONEQUILIBRIUM INTERPHASE MASS TRANSFER

5.1 Model for nonequilibrium intetphase mass transport

5.1.1 Governing phase transport equations

The transport equations for a component  developed in Chapter 4 can be expressed for
the water phase (w), the organic liquid phase (o) and the gas phase (a) as

To accommodate adsorption of Q by the solid phase, an additional continuity equation
is
                                   =  R0, - vas Cat                          (5.2)

where Cat is the solid phase concentration expressed as mass of adsorbed component Q
per porous medium volume [M L"3], »ot is the first-order decay coefficient [T"1] in the
solid phase and Ras is the mass transfer rate per porous medium volume to (+) or from
(-) the solid phase [M I/3 I"1].

5.1.2  Consideration of nonequilibrium interphase mass transfer

The phase-summed formulation  of the  transport  equations may be  generalized  to
account for nonequilibrium phase partitioning by introducing the concept of apparent
partition coefficients. Under transient field  conditions, at a given location in time and

                                    152

-------
 space, actual phase  concentration  ratios may differ from  the true equilibrium  ratios
 defined by [4.17]. With this in mind, we define apparent partition coefficients  -  which
 will vary in time and space  — by analogs of [4.17] as

                                   Cao = T0'0 Cow                              (5.3a)

                                   Caa = Ta'0 Cow                              (5.3b)

                                   C0. = C CQW                              (5.3c)

 For any two phases which are  in  physical contact,  the rate  of mass transfer will be
 described by first order mass transfer functions of the form

                              *al2 =  *.12 (Cal  ~  Co\)                         (5.4a)

                                   =  -kol,(Ca2 - C0\)                       (5.4b)

 where Ran is the rate of mass transfer of a per porous medium volume from phase  1 to
 phase 2, (7ol is the actual concentration of o in phase 1, C^i is the concentration of a
 which would occur in phase 1 if it  were in equilibrium with phase  2, Co2 is the  actual
 concentration of a in phase 2, C^2 is the concentration of a that would occur in phase 2
 if it were in equilibrium with phase  1, and  A;ol2 is a mass transfer rate coefficient [T"1].

 Consider first,  the case of mass transfer between oil and water  phases  when both phases
 exist at a point in time and space.  Employing [5.4a]  along with [5.3a] for the actual oil
 phase concentration and [4.17a] for the equilibrium concentration gives

                               =  *

•which may be solved in terms of ra'0 as
                                               D
                                Ft   	  r   _L   •"•pom                            (f £\
                               ao  -  Fao T   L    /nr                           (5.6)
•which indicates that the apparent partition coefficient may be expressed in terms of the
actual partition coefficient  plus a correction  term, which  depends  on the  sign and
magnitude of the actual mass transfer rate and on the concentration in the water phase.

                                      153

-------
If free oil is present in the system at a given point in time and space, so that air and oil
are physically in contact, mass transfer between oil and gas phases may be described in
a similar fashion. Employing in this case [5.4b] we obtain
oa
                                        (ro

which yields
                                               "000
                                               -
In the absence of free oil in the porous medium at a given location, mass transfer will
occur between water and gas phases rather than between oil and gas phases.  Proceeding
in the same manner as for oil-gas mass transfer yields

                                               n
                              r '   —  r   —    •aau>a
                              1    —  *
                               era  —   oa     L
                                              "'awa
Note that either [5.8] or [5.9] will be relevant at a given location and time, depending on
whether free oil occurs. Finally, mass transfer between water and solid  phase may be
considered by employing [5.4b] to obtain
                              r.f.  =  r0,  -   ."•!,                           (5.10)
                                              .
                                              atvf
In order to describe transport with nonequilibrium phase partitioning, the  individual
phase transport equations, [5.1a-c]  and [5.2], may be summed as described  in  section
3.1.2 but using apparent rather than equilibrium partition coefficients. Thus  [4.19] will
hold for the nonequilibrium case with apparent partition coefficients used  in lieu of the
equilibrium coefficients.  Nonlinearity is introduced into  the  resulting phase-summed
transport  equation  due  to the  dependence of  apparent partition  coefficients  on
concentration as well as  mass transfer rates. Thus, an iterative solution procedure will
be required for the nonequilibrium problem as will be discussed later.

If the transport equations  are written in  terms of oil phase concentrations,  similar
expressions to those given above may be derived to define apparent partition coefficients
as functions of mass transfer rates and oil phase concentrations.
                                      154

-------
Using the nonequilibrium relations, we may rewrite the phase  transport equations in
terms of a oil phase  concentration.  Using [5.3] to eliminate oil,  gas and solid  phase
concentrations from [5.1] and summing the equations noting that



leads to the a-component phase-summed transport equation  in terms of water  phase
concentrations

                   *  dCaw _
where
                        *Z = 4SW + * S0T'ao + 4>Sa roo + T'a,                 (5.12b)

                 £Q* =Sw Dawij + S0 DQoij T'ao + 4Sa Doaij T'aa          (5.12c)

                            ««,* = Qwi + 9oi r^0 + 9ai rL«                     (5.12d)
                                                   p    P r1     7? r'
             * 	      i     rii   i     r'  _L    r'  _L   w _L  o  oo _L   a  QQ     /c 10Q^
            ^o - "aw ^ ''ao rao + "aa ^0 ^ "o« T<» + ~P^ + —p^~ +    pa       (5.12eJ

Note that [5.12] has  the same  form as the  simple single phase transport equation.
However, the coefficients represent pooled effects of transport  in all phases. Note also
that interphase mass transfer terms occur in the phase summed equation  only as a sum
over all components.

An alternative form of the phase-summed transport equation may be written in terms
of the  oil phase  species  concentrations.  The a-component phase-summed transport
equation in terms of oil phase concentrations has the form
                 **     00     d f r) **    >o 1   ., **   oo    **
                                                                ri           /K 10 \
                                                                 °°         (     ^
where
                                                      ,   Q,                  /K1•^k^
                                                      + p-                  (5.13b)
                                     1 00      * 00       * 00
                                      155

-------
                      **
                                          W             W
                                          1 oo            *• ao
                                                  /T
                             ** 	  _    i  "mi  ,   "oi
                             i  -  9o.  +  p-  +  —p
                                          1 oo       *
                                                                            (5.13e)
                                                                 'a1 oo
The form of the oil-based equation is identical to that of the water-based phase-summed
equation. The water-based equation has the conceptual advantage that Caw is physically
meaningful at all  locations in the  porous medium, whereas  Cao only has  physical
meaning  at  locations where  S0>0.  The  latter  constraint,  however,  imposes  no
mathematical difficulty since [5.3] can in any event be employed as a definition of CQO
as well as  Coa  whether or not the  phase  exists.  Since most of the  component  mass
commonly resides in the oil phase, the oil-based equation has the advantage that mass
balance accuracy of the solution may be more readily controlled.

When simulating two-phase liquid flow consisting of water and oil with constant gas
pressure,  only diffusive transport is considered in the gas  phase — i.e.,  qa  = 0  is
assumed for this case.  It should be noted that even if  gas pressure gradients have  a
negligible effect on liquid phase flow, this does not strictly imply zero gas flow. If fluid
saturations change in  time, gas flow will occur. Furthermore, phase  partitioning may
result in density gradients in the gas phase which can induce significant flow.

5.1.S Solution approach

The solution  approach for solving the coupled  multiphase flow and multicomponent
transport problem with nonequilibrium mass transfer is as follows

1.    Solve the fluid flow equations simultaneously for the current time-step using  time-
      lagged phase densities and interphase mass transfer rates.

2.    Solve the phase-summed transport  equation using current  values  of apparent
      partition coefficients, interphase mass transfer rates and phase densities.

                                      156

-------
3.    Back-calculate interphase  mass  transfer  rates,  update  phase densities  and
      apparent partition coefficients and repeat (2) until transport solution converges.

4.    Proceed to the next time-step.
5.1.4 Numerical verification

Example 1^.  1-D Spill of Two-Component Hydrocarbon. Details of the input data and
the results for this example using equilibrium partition coefficients were presented in
Chapter 4 (Section 4.3.1). In this section, stage 3 of example 1 has been repeated using
nonequilibrium  interphase  partition coefficients. Several simulations  of stage 3 were
performed using nonequilibrium interphase  mass  transfer analyses  and results were
compared with those for  the equilibrium model described in  Chapter 4. To  provide
greater mass balance  stability for the nonequilibrium  analyses, the oil-based phase-
summed transport  formulation was employed. In addition to an equilibrium simulation,
analyses were performed using rate constants for all phase pairs of 1.0, 0.01 and 0.005
d "1.  Concentrations of xylene  and toluene at the lower  boundary  are shown as  a
function of time for the various runs in Figure 5.1 and 5.2,  respectively. Due to its
higher solubility, toluene is lost from the system more rapidly than xylene leading to an
increase in the mass fraction  of xylene in  the oil phase, and hence  an increase in
aqueous phase concentration at short times, followed by a reduction as extraction in the
water phase continues. As the phase transfer rate constants increase to 1.0  d"1,  the
nonequilibrium  model  is seen to give results identical to the equilibrium model. This
degeneration to the equilibrium case provides verification of the nonequilibrium model
formulation. As the rate coefficients  decrease, the exit  concentrations, and  rate of
leaching of contaminants from the oil phase, both diminish as expected.
                                      157

-------
           0.08
              0.00     20.00     40.00     60.00     80.00    100.00    120.00
                                    Time  (Days)

Fig. 5.1   Xylene effluent concentration  in water vs time for Stage 3 of Example 1.
             0.3-r
 E
 to
 e
^"
 c
 o
             0.2H
          a
         i 0.1 5-\
          C
          u
          u
          C
         U  0.1-
          V
          C
         jp o.osH
                                               K=0.005
                                                •4-
                                               K=o.or
                                                »K
                                               K=T.O
                                                        EQUILIBRIUM
              0.00     20.00     40.00      60.00     80.00    100.00    120.CO
                                     Time (Days)

Fig. 5.2   Toluene effluent concentration in water vs time for Stage 3   of Example  1.
                                          158

-------
5.2 Laboratory Investigations

A  series  of  laboratory  column  experiments were conducted to  investigate  the
nonequilibrium interface mass transfer  of soluble constituents from  a  multicomponent
nonaqueous phase liquid (NAPL) at residual saturations during steady-state water flow.
Breakthrough curves obtained from these experiments were inverted to  determine the
mass transfer rate coefficients. The objectives of the laboratory investigations were: (i)
to determine the effects of  aqueous phase pore velocity, NAPL saturation, and NAPL
composition on oil-water mass transfer  rate coefficients, (ii) to test the experimentally
derived rate coefficients against that  generated by existing models for predicting mass
transfer rate coefficients, and  (iii) to develop a new empirical model for nonequilibrium
mass transfer if results from (ii) are not  satisfactory.
5.2.1 Experimental methods

NAPL and porous medium properties. The NAPLs used in experiments were composed
of soluble and insoluble (inert) components. The inert component of NAPL was allowed
to comprise more than 50% of the total NAPL volume in order to minimize volume
changes  during the  course  of experiments,  and  to keep the  NAPL-aqueous phase
interfacial  area nearly constant. Two sets of experiments were conducted using two
different inert components, namely soltrol and hexadecane. For purposes of distinction,
the NAPLs with soltrol and hexadecane as their inert components will be referred to as
NAPL-1 and NAPL-2, respectively. The soluble constituents of NAPL were benzene,
toluene,  ethylbenzene and xylene in varying amounts.  All constituents used with  the
exception of soltrol are well defined compounds. Soltrol is a light hydrocarbon composed
of straight and branched  chained alkanes with carbon numbers  ranging from  10 to 16,
hence it has a low aqueous phase solubility. Table 5.1 lists the relevant properties of the
NAPL constituents.

The porous medium  consisted of a well graded  sand with fractions of 3.4% very fine
sand, 13.2% fine sand, 38.7% medium sand,  29.4% coarse  sand  and 14.6% very coarse
sand.
                                      159

-------
Table 5.1. Properties of NAPL constituents used in the experiments.
Constituents  Mol. Weight (g)  Solubility (mg/1) Spec, gravity Diff. Coef (m2/d)
Benzene
Toluene
Ethylbenzene
Xylene
Hexadecane
Soltrol
78.11
92.10
106.14
106.17
226.44
160.00
1780
515
152
192
0.0009
0
0.878
0.867
0.867
0.880
0.775
0.785
9.42 x 10'5
8.21 x 10'5
6.21 x lO'5
6.21 x 10'5
-
-
    Experimental Setup  1-D Column  Studies
           Soil/Oil Mixture
Constant head
            Sampling Port
        Sand gravel mixture
Fig. 5.3  A schematic of the soil column used for nonequlibrium mass transfer studies.

                                160

-------
Experimental procedure.  Experiments were conducted  in a 45  cm long glass column
with  a  cross section of 7.7 by 7.7 cm (Figure 5.3). The column was furnished with a
Marriotte bottle attachment to achieve steady state flow conditions. A 5 cm thick layer
of pea sized gravels and a wire screen were placed at the bottom of a dry, clean column.
Clean sand was added to the gravel layer leaving the sampling port just above this sand
mixed gravel layer (Figure 5.3).  In a separate container, 1500 g of well graded sand was
mixed uniformly with  a known volume of NAPL to produce a desired NAPL saturation.
The column was packed  uniformly  using NAPL-sand mixture as quickly as possible.
During   the  mixing  and  packing  processes,  care  was  undertaken  to   minimize
volatilization  losses. The  top  of the column was sealed with a plastic wrap to reduce
further  volatilization. The column was slowly saturated from the  bottom, to prevent air
entrapment. The duration of the saturation period ranged from 12 to 15 minutes. Upon
completion  of saturation, the water source was moved to  the  top  of the column to
maintain a  constant  head at  the  top  of  the column with  the  Marriotte  bottle
attachment. Flow was initiated  through the column adjusting the elevation of the flow
outlet until the desired flow rate was achieved. Outflow samples were collected for a
duration of 20 s periodically at 3  to 10 minute intervals during the course of the
experiment  to analyze for aqueous  phase concentrations. Flow  rates  were constantly
monitored and adjusted as required.  Experiments were conducted at discharge rates of
50 and 100 ml/minute for  a duration of 3 and 4 hours, respectively.

The experimental  variables  were  NAPL  saturation of the porous  medium,  NAPL
composition and  aqueous flux.  Table 5.2 summarizes the experiments performed  for
NAPL-  1 and NAPL-2, respectively. Each experiment was replicated at least twice to
ascertain the reproducibility of observed results.

Sample   collection and handling. At each sampling  interval, a syringe was used to
extract  approximately 5 cm3 of solution after the dead volume of the sampling tube had
been withdrawn and discarded. The sample was directly filtered into a 3.7 cm3 vial and
tightly capped, ensuring no head space was created in the process. Samples were  stored
at 8 "Cand were analyzed within 24 hours.

Samples from NAPL-1 were analyzed  using high  pressure liquid chromatography
(HPLC), while those  from NAPL-2 were analyzed using  a packed-column  type gas
chromatography (GC), utilizing standard solutions prepared in methylene chloride.
                                     161

-------
Table 5.2      Specifics of the nonequilibrium mass transfer column experiments (NAPL composition, pore
              water velocity, NAPL content), estimated mass transfer rate coefficients, and W values for
              estimations.
Exp.
*
1
2
3
4
5
6
7
6
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
Chem.
Comp.
Index
1
1
1
1
1
1
1
1
1
1
1
2
2
2
2
2
2
2
2
2
2
2
2
2
3
3
3
4
4
5
5
5
V (cm/s)
0.0303
0.0321
0.0308
0.0324
0.0336
0.0665
0.0653
0.0659
0.0642
0.0631
0.0642
0.030
0.033
0.035
0.033
0.034
0.036
0.033
0.033
0.032
0.069
0.065
0.067
0.063
0.034
0.034
0.033
0.061
0.061
0.061
0.062
0.060
©oil
0.0193
0.0395
0.0397
0.0287
0.0283
0.0388
0.0373
0.0194
0.0284
0.0276
0.0284
0.029
0.030
0.029
0.030
0.029
0.029
0.029
0.020
0.019
0.029
0.030
0.0198
0.0197
0.029
0.0297
0.019
0.0294
0.0294
0.0291
0.0290
0.0296
K (1/sec)
0.3005
0.3120
0.2402
0.4442
0.4533
0.4305
0.5872
0.4233
0.5569
0.3672
0.3791
0.0067
0.2662
0.2475
0.1537
0.3553
0.2487
0.2587
0.2943
0.2735
0.3701
0.3143
0.4234
0.3713
0.2407
0.2373
0.1594
0.3394
0.3618
0.2580
0.2590
0.2575
R'
0.945
0.957
0.941
0.965
0.880
0.985
0.983
0.970
0.902
0.955
0.959
0.830
0.781
0.732
0.742
0.676
0.970
0.964
0.874
0.860
0.956
0.980
0.972
0.984
0.984
0.978
0.902
0.107
0.975
0.990
0.993
0.991
       1:     10% Benzene + 10% Toluene + 10% Ethylbenzene + 10% Xylene + 60%
             Hexadecane (NAPL-1)
       2:     10% Benzene + 10% Toluene + 80% Soltrot (NAPL-2)
       3:     10% Benzene + 3% Toluene + 87% Soltrol (NAPL-2)
       4:     10% Toluene + 90% Soltrol (NAPL-2)
       5:     10% Benzene + 90% Soltrol (NAPL-2)
                                       162

-------
5.2.2 Data analysis methods

Aqueous transport of constituents through the experimental soil column, considering
nonequilibrium mass transfer from a multi- component  NAPL at a residual saturation,
can be described by one- dimensional ad vective- dispersive- reactive equation

                                                                             (5.14)
                                           ... n
subject to the following initial and boundary conditions

                                    C(x,0) = C0'                             (5.15a)

                                     C(0,t)  =0                              (5.15b)

                                    dC(L,t)   n
                                      dx
                                                                            (5.15c)
where Rld is the retardation factor for species t; D is the dispersion coefficient [cm2 s"1];
Vis the aqueous phase pore velocity [cm s"1]; C" is the concentration of species i in
the aqueous  phase [mg L"1]; K is the mass transfer rate coefficient [t~l\; Ce'q is the
equilibrium solubility of component i [mg L~1]; C0' is  the initial concentration of species
in the aqueous phase [mg L~1}; x is the spatial coordinate [cm]; t is time [s]; and n is the
number of soluble components in NAPL.  Rj were estimated for the porous media from
batch  experiments  outlined  by  Abdul  and Gibson  (1986).  C0'  was taken  as the
concentration  of  species  determined just  before flow  was  initiated.   Assumptions
associated with [5.14] and [5.15] are (i) first  order nonequilibrium partitioning between
NAPL and aqueous phase, (ii) linear equilibrium partitioning between aqueous and solid
phases, (iii)  no partitioning  between  NAPL and solid phase,  and (iv) single rate
coefficient describing nonequilibrium mass transport for all species.

The mass transfer rate coefficient (K) is defined as

                                     K =  k a                                (5.16)

where k is the mass transfer coefficient  [cm s'1] and a is the aqueous  phase-NAPL

                                     163

-------
 interfacial area per unit porous media volume [cm"1]. It is more convenient to determine
 K rather than k since there is no rational and independent method for estimating a.

 Since  Ce'g  is dependent on the concentration  of all other species in  the mixture, [5.14]
 was solved numerically using a Crank-Nicholson finite difference scheme to simulate the
 experimental breakthrough curves. The dependence of Ceg on the concentration of other
 species at the current time step was eliminated by using  C e'q  calculated from the
 previous time  step.  Ce'g  at a  given  node was  calculated  as a linear function of the
 component mole fraction at the same node and its aqueous phase solubility limit  from
 the previous time step as
                             t) = C^x, t-st) = jrir- •  C*                  (5.17)
where Ceq(x,t) is the equilibrium solubility at the current time step;  Ceq(x, t — 6t) is the
equilibrium solubility at the previous time step; Cel is the aqueous  phase solubility of
component i;  Af is the number of moles of component «; n+1 is the number of soluble
plus inert components. The denominator is the total  number of moles present in the
NAPL.

Soltrol used as the inert component of the NAPL is not a pure compound but a mixture
of several compounds.  Hence,  the  determination  of an effective molecular weight for
soltrol,  which is needed to determine  the number of moles of soltrol  in  NAPL-1,
required special  attention.  A batch experiment was conducted,  in  which excess of a
mixture of 80%  soltrol, 10% benzene and 10% toluene  was allowed to equilibrate with
water. The equilibrium  solubilities of benzene and toluene in the mixture determined by
combining the excess of the mixture with water in a volumetric flask and gently stirring
over night with a magnetic stirrer, maintaining the interfacial contact. Aqueous samples
were  analyzed by the  GC for  benzene and toluene equilibrium concentrations. The
number of moles of soltrol in the mixture was calculated from  [5.17] using  the measured
values of equilibrium solubilities for benzene and toluene.  Subsequently, the effective
molecular weight of soltrol  was estimated from the relationship

                                     "    m'                                (5-18)
                                     164

-------
where M, is the number of moles of soltrol ;  mt is the mass of soltrol  [g] used  in the
experiment; and Wt is the effective molecular weight of soltrol [g]. The estimated  values
of Wt corresponding to experimentally determined equilibrium  solubilities of benzene
and  toluene was 156.0 and 163.8 g, respectively. An average molecular weight of 160 g
was  adopted for the analysis,  which corresponds to that of hexane. Since soltrol is a
mixture of  c!0-c!6  alkane,  the estimated molecular weight for soltrol appears to be
reasonable.

The finite difference solution  of  [5.14] was  incorporated in a nonlinear least squares
routine to  estimate the  parameters  Rld, D  and K through  a numerical  inversion
procedure. The  nonlinear least squares routine was implemented using the Levenberg-
Marquardt algorithm to minimize the objective function <£
                                              / - cy                         (5.19)

where m is the number of measured data points, each corresponding to a specified time;
n is the number of soluble constituents; Cj' is the predicted concentration of component
                                A .
i for observation number ;; and  Cj is the measured concentration of component i for
observation number  j.  Hence, the  estimated values of parameters  Rj,  D  and  K
correspond to  the best-fit between the predicted  and observed breakthrough  curves.
Figures  5.4 through 5.7  show some of these fits between predicted  and observed
breakthrough  curves. The estimation  of parameter Rld is  included in the least  square
routine in order to account for the adsorption of the dissolved NAPL components by a
piece of cotton at the  sampling port, which is mistakenly used during the  column
experiments to prevent  the sampling port from  clogging.  The values of Rld estimated
through  optimization serve for the purpose  of explaining the apparent retardation  of
observed peak concentrations, in turn, improving  the fits between the observed  and
predicted breakthrough curves. These values  of R'd  are different from those used for the
soil in that they are limited to the last node (representing the sampling  port) where no
mass transfer is allowed.
                                      165

-------
                         400
                                                      R squore  - 0.993
                                                      Simulated  (Benzene)
                                                      Obnrved (Benzen*)
                                         5000          10000
                                            Time  (seconds)
                                                          150CC
Fig. 5.4
Comparison of observed and predicted breakthrough curves for experiment 31
(10% benzene with pore water velocity of 0.062 cm/sec and NAPL content of
0.029)
                         300-q
                                                      R square • 0.984
                                                      Simulated (Benzene)
                                                      Ob9«rv«d (Benzene)
                                                      Simulated (Toluene)
                                                      Observed (Tolutne)
                                         5000          10000
                                            Time  (seconds)
                                                          15000
Fig. 5.5    Comparison of observed and predicted breakthrough curves for experiment 25
           (10%  benzene,  3% toluene with pore  water velocity  of 0.034 cm/sec and
           NAPL content of 0.029)
                                        166

-------
                       400'
                       300-
                     I 200-
                     c
                     V
                     u
                     o
                     O
                       100-
                                                    R «quore  -' 0.980
                                                    Simuloted (Benjene)
                                                    Obterved (Beniene)
                                                    Simuloled (Toluene)
                                                    Observed (Toluene)
                                        5000          10000
                                          Time (seconds)
                             15COO
Fig. 5.6   Comparison of observed and predicted breakthrough curves for experiment 22
           (10% benzene,  10% toluene  with  pore water velocity of 0.065  cm/sec and
           NAPL content of 0.030)
                       300
                                                    R squore  • 0.674
                                                    Simulated (Benzene)
                                                    Objerved (Benzene)
                                                    Simulated (Toluene)
                                                    Observed (Toluene)
                ,1 I  I
5000          10000
   Time  (seconds)
                                                                    15000
Fig. 5.7   Comparison of observed and predicted breakthrough curves for experiment 19
           (10% benzene,  10% toluene with  pore water velocity of 0.033  cm/sec and
           NAPL content of 0.020)
                                         167

-------
5.2.3 Experimental Results

Mass transfer  rate coefficients  K estimated using the numerical inversion model are
given in Table 5.2 for all  experiments consisting of  combinations  of  different NAPL
composition, NAPL saturation, and aqueous phase pore velocity. Increasing  the aqueous
pore  velocity  generally  increased the mass  transfer rate  coefficient.  However,  no
relationship could be discerned between the rate coefficient and NAPL saturation. The
species  concentration in effluent increased with aqueous pore velocity (Figures 5.4 and
5.7), indicating that  increases in mass transfer rate coefficient contributed more than
compensating for the effects of reduction in detention time. With only one exception, all
the estimated K values  are within  the same order  of magnitude. The  majority  of
experiments  produced high values for the coefficient of determination  R2 indicating a
satisfactory model performance. Coefficients of variation CV of estimated K range from
1.5% to 92% with a mean CV of 18%  and a standard deviation of 21%. These CV values
correspond to a reasonably narrow confidence interval for the vast majority of estimated
K values emphasizing the reliability of estimated K values. In addition, reasonably low
correlation coefficients obtained between the estimated K and D indicate that estimates
of K are insensitive to the estimated D values.

To check the consistency of results presented  in Table 5.2 with those reported in the
literature,  a  comparison was  performed between experimental K values  and K values
obtained from an empirical relationship reported by Miller et al. (1990). This empirical
relationship is of the form
                             Sh = /30Re ^0/zSc1/2                           (5.20)

where /30 = 12,  /^ =  0.75, /32  = 0.6,  Sh is a  Sherwood number, Re is  a Reynolds
number,  5c is  a Schmidt  number and 0n is  the NAPL  volume fraction.  The
dimensionless Sh, Re and Sc numbers are defined  as
                                          .. _ j
                                                                            (5.21a)

                                                                            (5.21b)

                                                                            (5.21c)
                                      168

-------
where v is the mean pore water velocity, p is density of solute, d is the mean diameter
of the porous media grains, /z is  dynamic viscosity of solute,  K is aqueous-NAPL mass
transfer rate coefficient and  D0 is the molecular diffusion  of the component  in  the
aqueous phase.

Based on  grain size distribution data for the sand material used as the porous media in
column experiments,  d10 was determined to be 0.012 cm and dsp was  determined to be
0.029 cm. Using p =  878 mg/cm3, p = 6.028 mg/cm.stc, D0  = 1.0903xlO"5 cm2/sec for
benzene,  and d10  =  0.012  cm with the experimental  v, and 0n values in the above
empirical  relationship, corresponding K values were calculated. Results are presented in
Table 5.3 along with the experimental K values. Empirical  K values  are very close to
experimental K values. The same calculations were repeated using dso = 0.029 cm. The
results, which are also presented in  Table 5.3, produced  K values that are  consistently
lower  than  experimental K values.  These  results indicate that although empirical K
values  are  sensitive  to particle  diameter, use  of  d10  in  [5.20] seems to  be more
appropriate than  d50 as the representative porous media particle diameter to estimate K
values consistent with the experimental results.
                                      169

-------
Table 5.3 Comparison of experimental (K *) and empirical (K } mass transfer rate
          coefficients for effective porous media particle diameter corresponding
          to d10 and dso.
                                  d10 = 0.012 cm.

«.
0.019
0.029
0.039
tP=0.033
K*
0.3005
0.4442
0.3120
v=0.033
K
0.2430
0.3131
0.3740
v=0.066
K*
0.4233
0.3791
0.4305
v=0.066
K
0.4086
0.5266
0.6291
v=0.033
Sh
3.3278
4.2888
5.1231
v=0.066
Sh
5.5967
7.2128
8.6159
                                     = 0.029 cm.
0.019
0.029
0.039
0.3005
0.4442
0.3120
0.0826
0.1064
0.1272
0.4233
0.3791
0.4305
0.1389
0.1790
0.2138
6.3707
8.2104
9.8076
10.7142
13.8081
16.4943
                                     170

-------
              6.  TWO-DIMENSIONAL LABORATORY STUDIES OF
                    MULTIPHASE FLOW AND TRANSPORT

Two experiments were conducted to simulate flow and transport of a  light nonaqueous
phase liquid (LNAPL) and a dense nonaqueous phase liquid (DNAPL) in an unconfmed
aquifer with two-dimensional planar symmetry. Liquid pressures were monitored using
pressure transducers  and a  dual gamma attenuation  system was employed  to measure
fluid saturations.  Aqueous phase samples were collected  at specific locations over time
to monitor aqueous  phase  mass transport of soluble components. The experimental
apparatus and procedures  common to  both  experiments are described below  first,
followed by a separate discussion of the LNAPL and DNAPL experiments.
6.1 Experimental Setup

6.1.1 Cell design

A 2-D steel reinforced tank (Figure 6.1) with amber transparent plastic sides  (Ultem)
resistant to the organic chemicals used in  this  study was used  to  perform these
experiments. The flume was 1.01 m tall x 1.5 m long x 0.085 m thick.  Gamma  readings
were taken at 66 locations in an incomplete rectangular matrix. At 16 of these locations,
water and oil pressures were measured using  hydrophilic and hydrophobic ceramic cups,
respectively, instrumented with pressure transducers tied to a data acquisition system.
Twenty-four Teflon tube inserts served as aqueous phase sampling ports  in the lower
sections of the cell. Aqueous phase samples were analyzed on a gas chromatograph using
a packed column.

6.1.2 Gamma attenuation measurements

Beer's law can  be  used to  describe the attenuation of a radioactive beam passing
through a porous medium with three fluid phases (water,  NAPL and air).

                    / = I0 ezp (-/v,z - nwpjwx - »npnOnx - fgPg8gx)              (6.1)
                                    171

-------
                           Oil Sou re*
           ,        « -       ,   	       oSurtte9
           \\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\
                                             Fnttun Trtntduotr*
                                                                        1.1
                                                                           tn
                                                                    outflow
1*0
                        Figure 6.1 Experimental cell layout.
 Table 6.1   Properties of NAPL constituents used in experiments.
Constituent
Benzene
Toluene
Soltrol
PER
1-Iodoheptane
Specific Water Viscosity
Gravity Solubility Ratio
(ppm)
0.877
0.862
0.785
1.60
1.375
1,780
515
0
150
0
0.65
0.58
4.32
0.90
N.A
Surface
Tension
(dyne cm~1)
27.9
29.8
26.0
31.3
N.A
Interfacial
Tension with
Water (dyne cm~l)
35.0
36.1
43.3
44.4
N.A
                                     172

-------
where / is the exciting radiation intensity in counts per second (cps), I0 is empty cell
count (counts s"1), p,- is the mass attenuation coefficient (cm2 g~l) for phase t, pi mass
density of phase t (g cm"3), and x is the path length (cm) over which attenuation takes
place. Assuming the gas attenuation to be negligible and the porous media bulk density
is invariant in time, (6.1) reduces to

                           / = T0 exp (»wpjwx - »npnenx)                       (6.2a)

where
                                /. = I0 exp (-ntptx]                            (6.2b)

Equation (6.2)  was used to determine  water and  NAPL saturations (Lenhard  et al.
1988) from dual gamma measurements from a radiation source which utilizes low energy
Am-241  and high energy Cs-137.  To accomplish this, ^,, p{ and x at every measurement
location, needs to be determined for Am-241 and Cs-137. The product of /j, and p{ !>,?,,)
for oil and water, pa for soil and i at each measurement location were determined for
Am-241  and Cs-137 according to the  protocols of Lenhard et  al. (1988), for each
experiment.  During the calibration exercise, all  counting times  were  300 seconds in
duration, however, during the experiment due to the transient nature of flow processes
and the number of nodes to be scanned, the counting times were reduced to 90 seconds.
Counts were also adjusted for the high energy Cs-137 detected in the Am-241 window
according to the method of Nofziger and Swartzendruber  (1974) and for the resolving
time of both sources (Fritton, 1969).
6.1.3 Pressure measurements

Hydrophobic and hydrophilic ceramic tips connected to pressure transducers were used
to measure NAPL and  water pressures.  Hydrophobic tips were obtained by treating
ceramic  tips with chlorotrimethylsilane as  outlined  by  Lenhard  et. al.  (1988).
Calibration of individual transducers involved determining voltage responses to known
but  varying pressure heads from  which simple linear  regression coefficients  were
determined. In all cases, the square  of the coefficient of determination (R2) was greater
than 0.99.
                                      173

-------
 6.1.4  Sample collection and analysis

 Aqueous solution samples of  about  5.0 m3 were collected 2 or 3 times a day using a
 syringe. Sample were directly filtered into a 3.7 cm3 vials and tightly capped, ensuring
 no head space was created in  the process. Samples were stored in a walk in refrigerator
 at 8 °C and  analyzed within 24 hours.


 6.2  LNAPL Experiment

 6.2.1  Experimental methods

 The LNAPL used in this study consisted of a mixture of soltrol, benzene, toluene and 1-
 iodoheptane in a 7:1:1:1 ratio by volume. Soltrol is a mixture of straight and branched
 chained alkane with carbon numbers ranging from 10 - 16, hence, relatively insoluble in
 water. Benzene and toluene are much more  soluble than the other constituents and
 their transport in the aqueous phase was monitored during the experiment. lodoheptane
 was  added to the mixture to increase the gamma attenuation of the oil phase enabling
 better resolution of fluid saturations from gamma measurements. Relevant properties of
 the  constituents of the  LNAPL are given in Table 6.1. The specific gravity of the
 LNAPL mixture was 0.826. The soil used in the experiment consisted of 3% very fine
 sand, 13% fine sand, 39% medium sand 29% coarse sand and 16% very coarse sand.

 After filling the experimental  cell with soil, bulk density measurements were performed
•with the gamma system at each observation location for the air dry soil and the water
 saturated soil.  At  the  end  of the  experiment,  bulk  density  measurements were
 determined  again via direct measurements of core samples taken from the disassembled
 cell.  The bulk density measurements determined from the saturated cell were used  in
 calculations  as  these  yielded the  lowest  overall  variance and  were deemed most
accurate.  The average bulk density of  the packed cell  was 1.50  g cm"3, yielding a
porosity of 0.43 assuming a particle density of 1.50 g cm"3.

Table  6.2 lists the values of /i,p, and pt determined for both energy sources.  Variation
in x values between  energy  sources were less than  1.0% for both  sources at  all
measurement locations, which will  yield  a precision in  computed NAPL  and  water
saturations of <1%.

                                     174

-------
Table 6.2   Gamma attenuation parameters of porous medium and liquid  for the
            LNAPL Experiment
Coefficient
^ (cm'1)
*Vn (cm~1)
nt (gm.cm )
Am-241
0.196189
0.755835
0.253297
Cs-137
0.083462
0.071227
0.07534
The LNAPL experiment was conducted over a period of 3 weeks following calibration of
the  gamma system.  The experiment was performed in  "stages" with time  varying
boundary conditions as described below.

Stage  1 ^ Water Drainage.   A two-phase  air-water  experiment was performed  to
calibrate the k-S-P (permeability-saturation-pressure) model. The soil  in the cell was
initially water saturated and was allowed to drain for 9 hours after the water table was
lowered to 36.5 cm and maintained at this level. In this experiment, the origin for the z
(horizontal) and z (vertical)  coordinates was  the lower left  corner.  The boundary
conditions employed during this stage were as follows.
Top and bottom boundaries:
                   water and oil flux = 0

Left Vertical Boundary:
                   water and oil flux = 0

Right Vertical Boundary:
                    hw -r z= 36.5 cm
                       water flux = 0
                        oil flux = 0
        everywhere
         everywhere
(0 < z <  36.5 cm)
  (z > 36.5 cm)
   everywhere
                                     175

-------
 During this stage of the experiment, water saturations and pressures  as well as water
 outflow were monitored. The total outflow for the period was  15,000 cm3. After  nine
 hours, a horizontal gradient  of 1.35% was  established by imposing a constant head of
 36.5 cm on the left-hand side of the cell (Figure 6.1) while the water table was dropped
 to  34.5 cm on the right boundary. At steady state, the aqueous  flux  density was
 measured and was used to compute the saturated hydraulic conductivity of the soil.

 Stage 2 i LNAPL Infiltration. Over a period of 11.0 hours 1,500 cm3 of NAPL was
 added as strip source 50 cm from the top left-hand corner (average rate = 136  cm3 h~l).
 Initial conditions for this stage are the final conditions for Stage 1 of the experiment.
 Using the coordinate system  as previously defined, the boundary conditions for Stage 1
 are as follows.
 Top Horizontal Boundary:
                     water flux = 0
                   oil flux = 136 cm3 h~l
                       oil flux = 0

 Bottom Horizontal Boundary:
                   oil and water flux = 0

Left-Hand Vertical Boundary:
                   hw + z = 36.5 cm
                     water flux = 0
                      oil flux = 0

Right-Hand  Vertical Boundary:
                   hw + z = 34.5 cm
                      water flux = 0
                       oil flux = 0
     everywhere
      at x = 50 cm
     elsewhere
       everywhere
 0 <  z < 36.5 cm
    z  > 36.5 cm
   everywhere
0 <  z <  34.5 cm
  z  > 34.5 cm
  everywhere
Stage 21 LNAPL Redistribution.   At the end of LNAPL infiltration, oil was allowed to
redistribute for 170 hours. Boundary conditions during this phase of the experiment
were the same as  those during Stage 2 except that the oil flux is zero everywhere on the
top boundary.
                                     176

-------
Stage  4 ;. Water  Flushing. To simulate the  effects  of  water flushing on  LNAPL
dissolution and distribution, nine intermittent pulses of water at  constant flux were
applied to the soil  surface over a period of 76 hours.  Each pulse has a volume loading
rate of 2.84 cm3/s per unit width for a duration of 78 seconds. Pulses were applied at 0,
8, 14,  22, 30,  39, 48,  56, 63 hours, respectively. All other boundary conditions were the
same as Stage 3.

Stage £ - LNAPL Entrapment.   To simulate LNAPL entrapment by the aqueous  phase,
the water table elevation was instantaneously raised at inlet and outlet by 11.43 cm and
this constant head  condition was maintained for 60 hours. The  boundary conditions on
the  top and bottom of the cell were the same as  during  Stage 3. However, on the
vertical sides, the conditions for water phase were:
Left-Hand Vertical Boundary:
                   hw + z = 47.8 cm            0  < z <  47.8 cm
                     water flux =0                 47.8cm < z

Right-Hand Vertical Boundary:
                    hw + z - 45.8 cm           0 < z < 45.8 cm
                     water flux = 0                45.8 cm  < z
6.2.2   Model calibration and numerical analysis

The experimental results in the  LNAPL  experiment were compared with simulations
performed using the multiphase flow and  transport model described in Chapter 4.  The
length  (148  cm)  and  height  (101 cm) of  the  experimental cell  were discretized
numerically by a  mesh with 16  columns  and 21 rows resulting in 336 nodes and 300
elements. The cell width (8.5 cm) was taken as an axis of symmetry (i.e., 2-D  Cartesian
analysis). Oil-water viscosity ratio,  oil specific gravity, oil surface tension and oil-water
interfacial tension for the NAPL mixture were calculated as volume-weighted averages
of the  component values,  using the data  in Table 6.1. Fluid scaling  factors  were
estimated as
                                     177

-------
                                 O   ^  U Q  i V OW
                                 Pow ~    O™

where ff0 and  aow are the volume fraction-weighted average surface tension and oil-
water interfacial  tensions,  respectively.  For  purposes  of  modeling,  soltrol  and
iodoheptane were considered as a single pseudo-species of negligible solubility referred to
as "inert." Tabulated pure component densities and diffusion coefficients were obtained
from  the literature. Equilibrium oil-water partition coefficients  were computed as the
ratio  of  water  to oil phase concentrations determined from batch experiments  on the
NAPL mixture,  air-water partition coefficients were obtained from tabulated Henry's
constants for  the  compounds, and solid-water partition  coefficients were  assumed
negligible since the organic content of the sand  used in the experiments was very low.
The true (average)  density  for the inert components was  specified  to enable  correct
calculation  of the  bulk  oil  phase  density  by  the  code  which  is  corrected for
compositional changes,  and the oil-water partition coefficient was assigned a large value
to prevent partitioning. Other parameters for the inert components are arbitrary.
      Table 6.3  Soil and fluid properties used for LNAPL simulation.
                  Soil properties:        Bulk fluid properties:
KIW
£U/«
^n
sm
a
n
AL
AT
= 231 cm h-1
r = 231 cm h'1
' = 0.42
= 0.00
= 0.05 cm"1
= 3.50
= 10.0 cm
= 0.01 cm
0ao = 2.84
/?ao= 1.543
r,ro = 4.620
Pro = 0.826




                             Component properties:
          cm2 /T1
    D~/a,  cm2 h'1
    p0, g cm'3
Benzene
0.035
317.0
260.8
0.24
0
0.877
Toluene
0.027
281.0
1058.5
0.28
0
0.862
"Inert"
0.0001
0.001
106
0.2
0
0.785
                                      178

-------
 The saturated hydraulic conductivity of the soil to water was computed from horizontal
 flow rates through  the cell under  a  fixed gradient and the  soil  was assumed to be
 hydraulically isotropic.  Soil  porosity was  determined  from  gamma  attenuation
 measurements and van Genuchten parameters (a, n, Sm) were estimated by matching
 measured and simulated water outflow and vertical water saturation distributions at the
 end of the Stage 1 drainage period.  Dispersivities were  assigned small enough values to
 avoid oscillations in the solution  with  the selected mesh. A summary of the soil and the
 fluid properties used in the simulation  of the LNAPL experiment are given in Table 6.3.

 Stage 2 involved infiltration of oil in  the  domain. During this stage a flux of 7  cm/hr
 was applied on the top of element 291. Simulation was terminated when the cumulative
 oil  infiltration was 165 cm3/cm.  Initially,  the water table was allowed to vary linearly
 from the left boundary to the right boundary. Water pressures were specified on the left
 and right boundaries  as hydrostatic through Stage 2  corresponding  to  a  water table
 elevation  of  36.5 and  34.5  cm at the  left  and right sides, respectively. The rest  of the
 domain had no flow boundary conditions for all fluid phases.

 Stage 3  involved redistribution  for 31 hours after the end  of the  infiltration.  The
 boundary conditions for water phase during this stage were unchanged, while a zero flux
 was imposed for the oil phase  on all boundaries. The fluxes in the oil phase during the
 early part of the redistribution period  were relatively high. It was observed during some
 earlier simulations that a more stable transport solution is obtained when liquid phase
 fluxes are small. Therefore, simulation of transport was delayed until the next stage.

 Nine cycles of water application  and drainage were imposed during Stage 4. Each cycle
 consisted of an application of 2,000 cm3 of  water applied at the top surface of the cell in
 78  seconds,  followed  by  8 hours  of redistribution, while all other flow  boundary
 conditions were same as in Stage 3. During Stage 4, simultaneous flow and  transport of
 the mixture comprised of benzene, toluene  and "inert"  components was simulated.
 Initial oil phase concentrations for benzene, toluene and "inert" species were assigned to
 be 0.070, 0.075 and  0.68 g/cm3, respectively. Initial  water and gas phase concentrations
were assumed to be in equilibrium with the oil phase and to be zero at locations where
no  oil  phase occurred.  Since  transport prior to Stage 4  occurred primarily  due to
 convection of the oil phase, this assumption for initial conditions should lead to little
 error. For transport,  a type-2 boundary  condition (zero concentration  gradient) was
 applied at the right vertical boundary to allow  mass efflux, and  a  type-3  boundary

                                      179

-------
condition with zero influent concentration was  applied elsewhere.The initial time step
during  the simulation was 0.0001  hours and  the time  incremental factor  and the
maximum time step were 1.03 and 0.05 hours, respectively.

During Stage 5, there was no flow of water through the top boundary. The water table
was raised in the cell by 11.4 cm on the left and  right sides equally, while all other
boundary  conditions  for  all phases  remained unchanged.  Flow  and  transport  was
simulated.

6.2.8   Results

Stage 1 - Water Drainage. A comparison of observed and simulated water saturations at
the end of Stage 1 is shown in Figure 6.2.  Good correspondence is observed.  The van
Genuchten  parameters,  which describe the  air-water capillary pressure curve, were
estimated by matching the drainage data.

Stage 2 i LNAPL Infiltration. Water and oil saturation distributions at the end of Stage
2 are shown in Figures 6.3 and 6.4. The oil plume at this time is  largely confined to the
unsaturated zone and capillary fringe, with some vertical  penetration into the aquifer.
Correspondence between observed and predicted water and oil saturations  is good.

Stage 2 i LNAPL Redistribution. Water and oil saturation  distributions at the end of
Stage 3 are shown in Figures 6.5 and 6.6. Water saturation contours differ only slightly
from those of Stage 2. Oil has drained from the unsaturated zone  and spread laterally
on  the  water table.  Measured oil  saturations in the unsaturated zone  exhibit much
greater  variability than predicted values, perhaps reflecting  some heterogeneity  in the
sand packing.  In  general  the degree  of oil  drainage, vertical penetration and lateral
spreading for the  simulations is  greater than was observed.  Our suspicion is that this
may reflect simplifications invoked in the handling of terms in the hysteresis model, but
we  have yet  to confirm this by performing more rigorous model  simulations. Measured
and predicted aqueous concentrations of benzene and  toluene at  the end  of Stage 3 are
shown in Figures 6.7 and 6.8, respectively. Since aqueous samples could only be taken in
the saturated zone, the number of measured locations is rather small. The simulations
show much more vertical penetration of dissolved species into the aquifer than  observed.
This reflects the greater predicted oil phase penetration as  well as possible numerical
dispersion effects. Overshoot in the transport  solution occurred at  the downstream face.

                                      180

-------
Stage 4 i Water Flushing.  Water and oil saturation distributions at the end of Stage 4
are shown in Figures 6.9 and 6.10. Trends observed at the end of Stage 3 towards too
rapid oil  movement are continued and  accentuated  at  this time. Deviations between
measured and predicted aqueous concentrations  of benzene  and toluene at the end of
Stage 4 (Figures 6.11 and 6.12) are likewise similar to those observed for Stage 3.

Stage £ - LNAPL Entrapment.  At the end of Stage 5, the water saturation contours
have  been  clearly  displaced  upwards  (Figure 6.13).  Observed and simulated  oil
saturations (Figure 6.14) exhibit small upward displacements, reflecting the fact that oil
has already distributed to close to a "residual"  state at which its permeability is very
low  so that movement  is  impeded. Additional  aqueous  sampling  ports  at higher
elevations show high  concentrations corroborated  by the model, but the  problem of
excessive  vertical penetration of the simulated aqueous plume remains  (Figures 6.15 and
6.16).

Saturation-time curves at  selected locations.  To  evaluate  the  model predictions in
greater detail,  several  locations  in the  cell were  selected to compare water and oil
saturation histories  over  the course of the experiment. The 6 locations are shown in
Figure 6.17.  Measured and predicted water and oil  saturations versus time for the
selected locations  are  shown in Figures 6.18 -  6.23. The results show generally good
agreement between observed and predicted data, with a tendency to  underpredict the
oil  saturation within  the oil plume as time passes  due to vertical  penetration  and
horizontal spreading which is somewhat overpredicted by the model.
                                      181

-------
                                              100      IB
     loo F   r
   N
        -04-
        -OA-
                    -OJ-
                    -0.4-
                                                     -0.1 •
                            *—M-
        •OJ
                                                                   10
                                  X (cm)
                                              100      120
                                                              140
Figure 6.2 Water saturation at the end of Stage 1 for LNAPL experiment.
                                               100      IB
                                                                  •=1 100
                                                              140
Figure 6.3 Water saturation at the end of Stage 2 for LNAPL experiment.
                               182

-------
                                                      1»      HO
                                                             140
 Figure 6.4 Oil saturation at the end of Stage 2 for LNAPL experiment.
                                                             140
Figure 6.5 Water saturation at the end of Stage 3 for LNAPL experiment.




                              183

-------
     4i
     M
                                                            140
    Figure 6.6 Oil saturation at the end of Stage 3 for LNAPL experiment.
            II—rn—n	1	1	1	n  Q
                                                                20
Figure 6.7  Benzene concentration at the end of Stage 3 for LNAPL experiment.


                                 184

-------
                                               too      m
Figure 6.8  Toluene concentration at the end of Stage 3 for LNAPL experiment.
                                                               140
        ioo p
      U
      s^


      N
                                         i    I    I   I    i
I   ^ 100

                      -fcj-
                                 -u-
                                           -OJ-
                                                      -M.
                                                                     40
                                 *>       to

                                    X (cm)
                                                100      1»      140
   Figure 6.9  Water saturation at the end of Stage 4 for LNAPL experiment.
                                  185

-------
                                                                   - M
                                                                   - M
                                                               140
    Figure 6.10  Oil saturation at the end of Stage 4 for LNAPL experiment.
                                 tO       *)      100      120      140
                                                                   - to
                                                                   - 40
                                                                   - 10
                                                               140
                                    X (cm)
Figure 6.11  Benzene concentration at the end of Stage 4 for LNAPL experiment.




                                  186

-------
                                               100     IB)      140
Figure 6.12  Toluene concentration at the end of Stage 4 for LNAPL experiment.
        too F
                                                                 - »
  Figure 6.13  Water saturation at the end of Stage 5 for LNAPL experiment.
                                 187

-------
                                                 100      IS
                                                                    -  10
                                                                140
    Figure 6.14 Oil saturation at the end of Stage 5 for LNAPL experiment.
                                                 100      12D      140
       U

       M
                                                                   10-
                                                                     100
                          40       «      «       100
                                     X (cm)
                                                                140
Figure 6.15 Benzene concentration at the end of Stage 5 for LNAPL experiment.

                                   188

-------
                                               100      1JD     140
Figure 6.16 Toluene concentration at the end of Stage 5 ior LNAPL experiment.
                       OII Sourct
             0.6 m
     \\\\\\\\\\\\
                                                 Sell Surface
           \\x\\\\\\\\\\\\\\\\\\\\\\\\\\\
                          x  1.
x 2.   x 3.

x 6.
                                                  X 4.
                                                 M44.44)
                                                        (114,••)
                                                                          m
                                     t.em-
                                                                  outflow
  Figure 6.17 Measurement positions for plots of concentration versus time.

                                189

-------
   1.(
   0.80 -
 „ 0.60 ^
 c
 o
"p
 3


-------





irations
"o




1.00 -
-
0.80 -_
~
:
0.60 '-
0.40 -_

0.20 '-

.00 -
o.c
I
I .
Oil Infiltration
I
i£T ^~> *A *

i A 1"*' * AA>* " *
i
i
i
i
i Oil Redistribution
1
1
1
1
L, g • _
J <*-\v*5vT v ^jy-V* "
,i ' - - - - .
1
^•r — i — i — i — i — i — i — i — i — | — i — i — i — i — i — i — r~
)0 100.00
Time (hr
.
4
.
X1
^ * ' .
r* v.*/0 100.00 	 200.00 " 300.00
                                                                 jkAA^A Water saturation (observed
                                                                 • •••• Oil saturation  (observed)
                                                                 — — Water saturation (simulated)
                                                                 — — Oil saturation  (simulated)
                            Time  (hrs)


Figure 6.21  Water and oil saturation versus time for position
                                 191

-------
1 .UW

0.80 -_
-
g 0.60 -_
.2 -
"o ;
-4_< -
(8 0.40 -_
0.20 i
o no

>






!
'
i
\» . * * *
* "* • / "** *
Oil Infiltration

Oil Redistribution
•
«* -n"" «VrV * Xv*v*> "
•
i i 1 1 1 1 > i i i 1 1 1 1 1 1

* V ^^
» * * *
****** *
*
A A

Water
Infiltration

"W^VVrf*
—
1 1 1 I 1 1 I

4t
*' "V^»


Oil
Entrapment
^^* -BB_. •

1 1 1 | 1
                                                                      AAAAA Water saturation (observed
                                                                      • •••• Oil saturation (observed)
                                                                      — — Water saturation (simuloU ••"
                                                                      — — Oil saturation (simulated)
       0.00             100.00           200.00
                                 Time  (hrs)
                                                      300.00
         Figure 6.22  Water and oil saturation versus time for position
   1.00
   0.80 -
   0.50 ^
 o
'•*;
 o
   0.40 ^
   0.20 -
   0.00
               Oil Infiltration
                    Oil Redistribution
                                         Water
                                      Infiltration
                                                                Oil
                                                            Entrapment
                                                               IAHA Water saturation (observed
                                                               immmm oil saturation (observed)
                                                               —  — Water saturation (simulated)
                                                               —  — Oil saturation (simulated)
     '^^^*"|  I  I	1  I  I I  I  I	1  I  I I  I  I  I  I  I  I I  I  I	1—1	1—]	1"
0.00             100.00           200.00           300.00
                          Time  (hrs)

 Figure 6.23  Water and oil saturation versus time for position
                                         19:

-------
6.3 DNAPL Experiment

6.S.I Experimental methods

The second experiment was  conducted to  study  the movement of dense nonaqueous
phase liquid (DNAPL). The experimental setup used in this study is the same as that
employed  for  the  LNAPL  experiment.  The  DNAPL  consisted  of  a mixture  of
tetrachloroethylene  (also  known  as  tetrachloroethene, perchloroethylene  or  PER),
soltrol,  toluene and  1-iodoheptane, mixed in a 4:4:1:1  ratio by volume,  respectively,
yielding  an overall specific gravity of 1.16.  Properties  of  PER, soltrol,  toluene and
iodoheptane are given in Table 6.1.

As  in the  LNAPL experiment, iodoheptane was included  to provide a larger gamma
attenuation contrast  between the NAPL and water to provide better resolution of fluid
saturations. The material  used in this experiment was a medium sand,  consisting of a
mixture of 2% very fine sand, 17% fine sand, 39% medium  sand, 28% coarse sand and
14% very coarse sand. The gamma attenuation coefficients of various components are
shown in Table 6.4.
Table 6.4   Gamma attenuation coefficients of porous medium and liquids
            for DNAPL experiment.
Coefficient
WW(™-1)
iv.r™-1;
PI (sm-cm^)
Am-241
0.19577
0.94179
0.24712
Cs-137
0.08361
0.09118
0.07466
The cell was filled with soil to a depth of 99 cm and saturated with water from the
bottom.  The average bulk density, evaluated under saturated conditions, was  1.46 g
cm'3. The experiment was performed in three stages  as described below.

                                     193

-------
 Stage  1 - Water Drainage.   A  two-phase air-water  experiment was  performed  to
 calibrate the k-S-P model.  The soil in the cell was initially water saturated and was
 allowed  to drain for 9 hours  after the  water table  was lowered  to  36.5 cm and
 maintained at  this level. In  this experiment the origin  for the  x (horizontal) and z
 (vertical) coordinates was the lower left  corner. The  boundary  conditions  employed
 during this stage were as follows.

 Top and bottom boundaries:
                          qw = 0              everywhere

 Left Vertical Boundary:
                          qw = 0              everywhere

 Right Vertical Boundary:
                     hw  + z = 36.5 cm        (0 < z < 36.5 cm)
                           qw = 0          (z > 36.5 cm)

 Water saturations and pressures as well as  water outflow were monitored. Total outflow
 over a 24 hour period was 15,500  cm3. After the drainage period, a horizontal gradient
 of 1.35% was established by imposing a constant head of  36.5 cm  on the left-hand side
 of the cell (Figure 6.1)  while the water table  was  dropped to 34.5 cm on the right
 boundary.  At steady state, the aqueous flux density was measured  and was  used  to
 compute the saturated hydraulic conductivity of the soil.

 Stage 2 r DNAPL Infiltration. Over a period of 3 hours,  500 cm3 of NAPL was added  on
 a strip 50 cm from top left-hand corner (Figure 6.1) yielding an average application rate
of 167 cm3 h~l. Using the coordinate  system previously defined, the boundary conditions
can be expressed as follows.

 Top Horizontal Boundary:
                     water flux = 0                everywhere
                   oil flux = 167 cm3 h~l            at i = 50 cm
                       oil flux = 0                 elsewhere

Bottom Horizontal Boundary:
                   water and oil flux = 0             everywhere

                                     194

-------
 Left-Hand Vertical Boundary:
                    hw + z = 36.5 cm            0 < z < 36.5 cm
                       water flux = 0                z > 36.5 cm
                        oil flux = 0                everywhere

 Right-Hand  Vertical Boundary:
                    hw + z = 34.5 cm           0 < z <   34.5 cm
                       water flux = 0             z > 34.5 cm
                         oil flux = 0               everywhere

 Stage 2. 2. DNAPL Redistribution.   At the end of NAPL infiltration, oil was allowed to
 redistribute  for 122  hours. Boundary conditions during this stage of the experiment were
 the same as those during Stage  2 except  that the oil flux was zero everywhere on the
 top boundary.

 6.S.2   Model calibration and numerical analysis

 Experimental  results  in  the DNAPL  experiment  were  compared  with simulations
 performed using the multiphase  flow and  transport model described in Chapter 4. The
 length and height of the experimental cell were discretized numerically by a mesh with
 19 columns  and 19  rows resulting in 361  nodes and 324 elements. The cell width (8.5
 cm) was taken as an axis of symmetry (i.e., 2-D Cartesian analysis). Oil-water viscosity
 ratio, oil specific gravity, oil surface tension and oil-water interfacial tension for the
 NAPL mixture were calculated as volume-weighted  averages of the component values,
 using the data in Table 6.1. Fluid scaling factors were estimated as
                                 000 =

                                 V™ =
                                            'ow
•where a0  and aow  are the volume fraction-weighted average surface tension  and oil-
 •water  interfacial  tensions,  respectively.   For purposes  of  modeling,  soltrol  and
 iodoheptane were considered as a single pseudo-species of negligible solubility referred to
 as  "inert." Tabulated pure component densities and diffusion coefficients were obtained
 from the  literature.  Equilibrium oil-water  partition coefficients were computed as the

                                      195

-------
ratio of water to oil phase concentrations determined from batch experiments on the
NAPL mixture,  air-water partition coefficients were obtained from tabulated Henry's
constants for  the compounds, and solid-water  partition coefficients  were assumed
negligible since the organic content of the sand used in  the experiments was very low.
The true (average)  density for the inert  components was specified to enable  correct
calculation  of the  bulk  oil  phase  density  by  the  code which is corrected for
compositional changes, and the oil-water partition coefficient was assigned a large value
to prevent partitioning. Other parameters for the inert components are arbitrary.

The saturated hydraulic conductivity of the soil to water was computed from horizontal
flow rates through the  cell  under a fixed gradient  and the soil was assumed to be
hydraulically  isotropic.   Soil  porosity  was   determined  from  gamma  attenuation
measurements and van Genuchten parameters (a, n, Sm) were  estimated by matching
measured and simulated water outflow and vertical water saturation distributions at the
end of the Stage  1 drainage period. Dispersivities were assigned values small enough to
avoid numerical oscillations with the selected mesh. A summary of the soil and the fluid
properties used in the simulation of the DNAPL experiment are given in Table 6.5.
      Table 6.5  Soil and fluid properties used for DNAPL simulation.
                  Soil properties:        Bulk fluid properties:
                  K,w  =226 cm /f1    /?  = 2.587
                  K  '= 226 cm /f1    0ao= 2.587
                  ^  w = 0.43          f,ro = 2.700
                  5m   = 0.01          /jro = 1.158
                  a    =0.15 cm'1
                  n    =2.0
                  AL   = 5.0 cm
                  AT   = 0.2 cm
                     Component properties:
                     PER
    D0°w, cm2 h'1     0.022
    Da°a, cm2 h~l     266.0
                                 Toluene         "Inert"
                                 0.027            0.00001
_aa, _	      	            283.0            0.001
roo             6920.            1110.            106
raa             0.35             0.28             0.2
rM             o                o                o
pa,gcm-3       1.60             0.862            0.785
                                     196

-------
 Stage 2 involved infiltration of oil in the domain on a strip source 8.5 cm wide, centered
 50 cm from the top left corner of the cell. Stage 2 simulation was terminated after 3
 hours when the cumulative oil infiltration reached  500 cm3. Initially, the  water table
 was allowed to vary linearly from the left boundary to the right boundary. The water
 pressure distribution was hydrostatic and  no oil was  present initially in  the system.
 Water pressures were specified on the left and right  boundaries as hydrostatic through
 Stage 2  corresponding to a water table elevation of 36.5 and 34.5 cm at the left and
 right sides, respectively. The rest  of the domain had no flow boundary conditions for all
 fluid phases. Transport was not simulated during Stage 2 since the duration was  only 3
 hours and transport was dominated by oil phase convection.

 Stage 3 involved redistribution for a period of 122 hours after the  end of Stage 2. The
 boundary conditions for  water phase during this stage were unchanged, while a zero flux
 was imposed for the oil phase on all  boundaries. Initial oil phase  concentrations for
 PER, toluene and  "inert"  species were  assigned the values  of  the initial  DNAPL
 mixture. Initial water and gas phase concentrations were assumed to be in equilibrium
 with the oil phase and to be zero at locations where no oil phase occurred. Boundary
 conditions  for transport  were a type-2 boundary condition (zero concentration gradient)
 on the right vertical boundary to allow mass efflux, and a type-3 boundary condition
 with zero  influent concentration  on the upstream boundary to account for inflow of
 clean water.

 6.S. S   Results

 A comparison of observed and simulated water saturations at the end of Stage 1, used
 to estimate van Genuchten air-water capillary pressure curve parameter,  is shown in
 Figure 6.24. Observed water saturations below the  water table ranged  from 82-95%,
 indicating significant quantities of trapped air. Since the constitutive relations used for
the  simulations did not consider  air entrapment,  the model predicts full saturation of
water below the water table.

 Observed and predicted water  and oil saturation  distributions  at the  end of the
redistribution period (Stage 3)  are compared in Figures  6.25 and 6.26. Water saturations
 are generally slightly underestimated by the model, except below the  water table  due to
trapped  air. The model predicts an oil phase plume  which is largely confined to a
 deformed spheroidal region above the  water table, although some oil is predicted to

                                      197

-------
have  sunk through the aquifer and spread  over the tank bottom. Between  the water
table and the bottom pool, oil saturations after drainage are predicted to be less than
1%.  Measurements could not be made near the tank  bottom, so confirmation of oil
spreading along the tank bottom could not be definitely confirmed. However, nonzero
oil  saturations at measurement  locations 15 cm from the tank  bottom suggest oil
penetration below the water table.  The predicted and measured oil distributions in the
unsaturated zone agree with one interesting exception. The observed data show a small
amount  of oil has  apparently  spread laterally  downgradient  along  the  top of  the
capillary fringe at oil saturations in the vicinity of 1%. The simulation does not predict
such  lateral migration along  the water table.  This is  an  interesting  and unexpected
phenomena. However, whether this is due to physical effects disregarded by  the model
(e.g.,  nonzero spreading  pressures at oil-water interfaces) or to subtle heterogeneities
due to packing anomalies cannot be ascertained.

Observed and predicted aqueous concentrations of PER and  toluene at  the end of Stage
3 are compared in Figures 6.27 and 6.28.  Substantial concentrations along the capillary
fringe downgradient of the source confirm the lateral migration phenomena noted based
on  oil  saturation  measurements.  Sampling  locations  under  the water  table  and
immediately beneath the  source indicate no  dissolved phase contaminants, although
gamma  measurements  indicate  DNAPL  at  or below these depths.  This suggests
downward migration of  DNAPL through the saturated zone occurs through isolated
fingers rather than as a uniform front.
                                     198

-------
              0   10   20  30   40   50  60  70   80   90  100  110  120  130  140
1 1 1 I 1 1
90 - S *?
.? / «^
8U 0.2 • U2 i
• *
70 - , S
* •
60 T,-, /-, . J
u.j 	 t,j s
	 « 	


40 -0.0 • t7.U » .
«?«?«?
30 - **
20 -
10 -
Q 1 1 1 1 1 1
0 10 20 30 40 50 60
i i
• *
» *
* •
4*J£-
— IL— ft? —


.' *«
* *
.*  J* ^ ^ - 7°
* * • •
S 3 - tfj "* - 60

	 • 	 • 	 Jli 	 « 	 — 	 bU
V « ri*^ 4^ ~
• "" » "" * — J" U.ti - 	
£ / .»" .^ 	
• * • *
.* .< - 30
- 20
f .*
- 10
1 1 1 1 1 1 Q
90 100 110 120 130 140
Figure 6.24 Water saturation distributions at the end of Stage 1 for DNAPL experiment.
             0.00    18.50    37.00    55.50    74.00    92.50   111.00   129.50   148.00
33. UU
86.63

74.25
61.88
49.50
37.13
24.75
12.38
n nn
i i i i i i i i i i i i i i i i i i i i i i i
A •$ N* & N1* •£*•
o> o- *' * O' O-
— nt- as "•*' -'*' ' '" -*" n-

O' O O* O O' O
• •••• ~

°-55 "' ~f 	 S? 	 rf" 	 °-5^«-~# 	 ^r5 	 ^# 	 0.55— 	
	 ft 05 	 £ 	 ._, 	 £f 	 03^' J _ J" f J 03—^
. . . °, *.
i i I i t i I i i i i i i i i i I i i l i i i
86.63

74.25
61.88
49.50
37.13
24.75
12.38
nnn
            0.00    18.50    37.00    55.50    74.00    92.50   111.00   129.50   148.00
       Figure 6.25 Water saturation at the end of Stage 3 for DNAPL experiment.
                                          199

-------
       0.00
                20.00     40.00
                                  60.00
                                           80.00     100.00     120.00     140.00
   80.00
   60.00
   40.00
   20.00
    0.00
                                                           I	I     I    I
       0.00
               20.00
                         40.00     60.00     80.00    100.00     120.00     140.00
                                                                             80.00
                                                                             60.00
                                                                             40.00
                                                                             20.00
                                                                             0.00
   Figure 6.26  Oil saturation at the end of Stage 3 for DNAPL experiment.
       0.00     18.50    37.00     55.50    74.00    92.50    111.00    129.50   148.00
    0.00
       0.00
                                                                             0.00
18.50    37.00     55.50    74.00    92.50    111.00    129.50   148.00
Figure 6.27  PER concentration at the end of Stage 3 for DNAPL experiment.
                                       200

-------
          0   10   20  30  40  50  60  70  80  90  100  110  120  130  HO
        90



        80



        70



        60



        50



        40



        30



        2C



        10



         0
                                \\r
o
                                          i    i    i    i    i    i    r
                         •s
                         •

                         o
              10  20  30  40  50  60  70  80  90  100  110  120  130  140
90



80



70



60



50



40



30



20



10



0
Figure 6.28 Toluene concentration at the end of Stage 3 for DNAPL experiment.
                                  201

-------
                              REFERENCES CITED

Abdul, A. S. and T. L. Gibson, Equilibrium batch experiments  with six  polycylic
aromatic  hydrocarbons and two  aquifer materials, Hazardous Waste & Hazardous
Materials, 3,  125-137, 1986.

Abriola, L. M.,  Multiphase migration of organic  compounds in a  porous  medium,
Lecture Notes in Engineering, Brebbia and Orszag (ed.)  No. 8, Springer-Verlag, Berlin,
1984.

Abriola, L. M. and G. F. Finder, A multiphase approach to the modeling of porous
media contamination by organic compounds, 1. Equation development, Water Resour.
Res., 21, 11-18, 1985a.

Abriola, L. M. and G. F. Finder, A multiphase approach to the modeling of porous
media contamination by organic  compounds, 2. Numerical simulation, Water Resour.
Res., 21, 19-26, 1985b.

Aienkiewicz,  0. C., The finite element method, McGraw-Hill, NY, 1986.

Bear, J., Dynamics of fluids  in porous media, 764p., Elsevier NY, 1972.

Bloomsburg,  G. L. and A.  T. Corey, Diffusion of entrapped air from  porous media,
Hydrology Papers No. 5, Colorado State University, 27 p., 1964.

Burdine, N.  T.,  Relative permeability calculations from  pore size distribution data,
Trans. AIME, 198, 71-77,1953.

Cooley, R. L., A finite difference method for unsteady flow in variably saturated porous
media: Applications to a single pumping well, Water Resour. Res., 7, 1607-1625, 1971.

Corey, A. T., The interrelation between  gas and oil relative permeabilities, Producer's
Monthly, 19,  38-41,1954.
                                     202

-------
Falta, R. W. Jr. and I. Javandel, A numerical method for multiphase multicomponent
contaminant transport in groundwater systems, Trans. Amer. Geophys. Union, 68, 1284,
1987.

Faust, C. R., Transport of immiscible fluids within and below the unsaturated zone: A
numerical model, Water Resour. Res., 21, 587-596, 1985.

Forsyth,  P.  A., Simulation of nonaqueous  phase groundwater  contamination,  Adv.
Water Resour., 11, 74-83, 1988.

Forsyth,  P.  A. and P. H. Sammon,  Practical considerations for adaptive implicit
methods in reservoir simulation, J. Comp. Phys., 62, 265-281, 1986.

Fritton,  D. D., Resolving time, mass  absorbtion coefficient and water content with
gamma-ray attenuation, Soil Sci. Soc. Am. J.,33,651-655, 1969.

Huyakorn, P. S.  and K. Nilkuha, Solution  of transient  transport equation using an
upstream finite element scheme, Appl. Math. Modeling, 3,  7-17, 1979.

Huyakorn, P. S. and G. F.  Pinder, Computational method in subsurface flow, Academic
Press, Inc., NY, 1983.

Huyakorn, P. S.,  S. D. Thomas and B. M.  Thompson, Techniques for making  finite
elements  competitive in modeling flow in variably saturated porous media, Water Res.
Research, 20, 1099-1115, 1984.

Kaluarachchi, J.  J.  and  J. C.  Parker,  Effects of hysteresis on  water flow in  the
unsaturated zone,  Water Resour. Res., 23, 1967-1976, 1987.

Kaluarachchi, J. J., and J.  C. Parker, An efficient finite element method for modeling
multiphase flow, Water Resour. Res., 25, 43-54, 1989.

Kool, J. B. and  J. C. Parker, Development and evaluation  of closed-form expressions for
hysteretic soil hydraulic properties, Water Resour. Res., 23, 105-114, 1987.
                                     203

-------
Kuppusamy, T., J. Sheng, J.  C.  Parker and R. J. Lenhard,  Finite element analysis of
multiphase immiscible flow through soils,  Water Resour. Res., 23, 625-632, 1987.

Land, C. S., Calculation of imbibition relative permeability for two- and three-phase
flow from rock properties, Soc. Petro. Engr. J., 6, 149-156, 1968.

Lenhard, R.  J., J.  H. Dane, J.  C. Parker and  J.  J.  Kaluarachchi,  Measurement and
simulation of transient three-phase flow in porous media, Water Resour. Res., 24, 853-
863, 1988.

Lenhard, R.  J.  and J. C. Parker, Measurement and prediction of saturation-pressure
relationships in three-phase  porous media  systems,  J.  Contain. Hydrol., 1,  407-424,
1987.

Lenhard, R.  J. and J. C. Parker,  Experimental validation of the theory  of extending
two-phase  saturation-pressure relations to  three  fluid  phase  systems for monotonic
saturation paths, Water Resour.  Res., 24, 373-380, 1988.

Leverett, M.  C.,  Capillary behavior in porous solids.  Trans. AIME, 142, 152-169, 1941.

Miller,  C.  T.   M. M.  Poirier-McNeill  and  A.  S.  Mayer,  Dissolution of  trapped
nonaqueous phase  liquids: mass  trasfer characteristics,  Water  Resour. Res, 26, 2783-
2796, 1990.

Millington, R. J. and J.  P. Quirk, Permeability of porous media, Nature (London) 183,
387-388, 1959.

Mualem, Y., A new  model for  predicting  the hydraulic  conductivity of unsaturated
porous media, Water Resour. Res., 12, 513-522, 1976.

Nofziger, D. L.,  and D. Swartzendruber, Material content of binary physical mixtures as
measured with a dual-energy beam of gamma rays, J.  Appl. Phys. 45, 5443-5449, 1974.

Osborne M. and J. Sykes, Numerical modeling  of immiscible organics transport at the
Hyde Park landfill, Water Resour. Res., 22, 25-33, 1986.
                                     204

-------
Parker, J. C. and R. J. Lenhard, A model for hysteretic constitutive relations governing
multiphase  flow, 1. Saturation-pressure relations, Water Resour. Res.,  23, 2187-2196,
1987.

Parker, J. C., R. J. Lenhard and T. Kuppusamy, A parametric model for constitutive
properties governing multiphase flow in porous media, Water Resour. Res., 23, 618-624,
1987.

Saraf,  D. N.,  J.  P. Batycky, C.  H.  Jackson  and D.  B. Fisher, An experimental
investigation  of  three-phase   flow  of  water-oil-gas  mixtures   through  water-wet
sandstones,  paper  presented  at  Califronia Regional Meeting,  Soc. Pet. Eng.,  San
Francisco, Calif., SPE paper #10761, March 24-26, 1982.

Stillwater, R. and A. Klute, Improved  methodology for a collinear dual-energy gamma
radiation system, Water Resour. Res., 24, 1411-1422, 1988.

Stonestrom, D. A. and  J. Rubin, Water content dependence of trapped air in  two soils,
Water Resour. Res., 25, 1947-1958, 1989.

Thomas, G. W. and D. H. Thurnau, Reservoir  simulation using an adaptive implicit
method, Soc. Pet. Eng.  J., 23, 759-768, 1983.

van Genuchten, M. T.,  A closed form equation for predicting the hydraulic conductivity
of unsaturated soils, Soil Sci. Soc. Am. J., 44, 892-898, 1980.

Willhite, G. P., Waterflooding, SPE  Textbook Series  Vol.  3,  Society  of Petroleum
Engineers, Richardson, TX, 326 p., 1986.
                                     205

-------
         APPENDIX: PUBLICATIONS RESULTING FROM THIS PROJECT

 Anderson, M. A.  and J. C. Parker, Sensitivity of organic contaminant transport and
 persistence models to  Henry's law constants: case of polychlorinated biphenyls, Water,
 Air and  Soil Pollution, 50, 1-18, 1990.

 Kaluarachchi, J. J.  and J. C. Parker. An efficient finite element method of modeling
 multiphase flow in porous media. Water Resour. Res., 25, 43-54, 1989.

 Kaluarachchi, J. J.  and J. C. Parker. Modeling multicomponent organic  transport in
 three fluid phase porous media, J. Contaminant Hydrol., 5, 349-375, 1990.

 Kaluarachchi. J.  J., J.  C. Parker  and R.  J. Lenhard, A numerical model for areal
 migration of water and light hydrocarbon in unconfined aquifers, Adv. Water Resour.
 Res., 13, 29-40, 1990.

 Kaluarachchi, J. J.,  J. C. Parker and R. J. Lenhard, Modeling flow in three fluid phase
 porous   media  with  non wetting  fluid  entrapment,  Proc.   Conf,   on  Subsurface
 Contamination by Immiscible Fluids. Calgary. Canada, April 17-20, 1990.

 Kaluarachchi, J.  J., J.  C. Parker.  Multiphase flow with  a simplified model for oil
 entrapment, Transport in Porous Media, (in press).

Katyal,  A.  K., J.  J.  Kaluarachchi  and J. C. Parker,  Evaluation  of methods  for
improving the efficieny and robustness of  multiphase  flow models, Proc.  Conf. on
 Subsurface Contamination by Immiscible Fluids, Calgary, Canada, April 17-20, 1990.

Lenhard, R. J., J. H.  Dane, J. C. Parker and J. J.  Kaluarachchi,  Measurement and
simulation of transient three-phase flow in porous media,  Water Resour. Res., 24,  853-
863,  1988.

Lenhard, R. J. and J. C. Parker, A model for hysteretic constitutive relations governing
multiphase fluid flow, 2. Permeability-saturation  relations, Water Resour.  Res.,  23,
2197-2206, 1987.
                                    206

-------
 Lenhard, R. J. and J.  C.  Parker,  Experimental validation of the theory of extending
 two-phase saturation-pressure relations to three-fluid phase  systems for  monotonic
 drainage, Water Resour. Res., 24, 373-380, 19SS.

 Lenhard, R. J., J. C. Parker and  S. Mishra, On the  correspondence between  Brooks-
 Corey and van Genuchten  models, J. Irrig. and Drainage, ASCE, 115, 744-751, 1989.

 Lenhard, R.  J. and J. C. Parker, A model for hysteretic constitutive relations governing
 multiphase fluid flow, 3. Refinements and numerical simulations, Water Resour. Res.,
 25, 1727-1736,  1989.

 Lenhard. R.  J. and J. C. Parker. Modeling and laboratory validation of multiphase fluid
 hysteresis, Proc. Workshop in Indirect Methods for Estimating Hydraulic Properties of
 Unsaturated Soils, Riverside CA, 1989.

 Lenhard, R.  J. and J.  C.  Parker, Modeling multiphase fluid hysteresis and comparing
 results  to laboratory investigations, Proc.  Indirect Methods for  Estimating  Hydraulic
 Properties of Unsaturated Soils, Riverside, CA, Oct. 11-13, 1989.

 Lenhard, R. J. and  J.  C. Parker, Estimation of free hydrocarbon volume from fluid
 levels in observation wells,  Ground Water, 28, 57-67, 1990.

 Parker, J. C.,  Multiphase  flow and transport  in porous media, Reviews of Geophysics,
 27, 311-328, 1989.

 Parker, J. C.  and J. J. Kaluarachchi, A numerical model for design of free product
 recovery  systems  at  hydrocarbon spill sites,  Proc. 4th Int.  Conf. on Solving  Ground
 Water Problems with Models, Indianapolis, IGWMC/NWWA, Feb 7-9, 1989.

 Parker,  J. C.  and  J.  J.  Kaluarachchi, Modeling multicomponent  organic chemical
 transport, Proc. Conf.  on Subsurface Contamination  by Immiscible Fluids, Calgary,
 Canada, April 17-20,  1990.

 Parker, J. C.,  J. J. Kaluarachchi and A. K.  Katyal. Areal simulation of free product
recovery from  a gasoline storage tank leak site.  Proc. Petroleum Hydrocarbons and
 Organic Chemical in Groundwater. National Water Well Assoc.  Houston, 1988.

                                     207

-------
 Parker, J. C., J. J. Kaluarachchi, V. J. Kremesec  and E.  L. Hockman,  Modeling free
 product recovery at hydrocarbon spill sites, Proc. Petroleum Hydrocarbons and Organic
 Chemical in Groundwater, NWWA, Houston, 1990.

 Parker, J. C., J.  J. Kaluarachchi and R. J. Lenhard, Experimental and numerical
 investigations  of  constitutive  relations governing  multiphase flow, Proc.  Conf.  on
 Validation of Flow and  Transport Models for  the Unsaturated  Zone, Ruidoso, New
 Mexico, 1988.

 Parker, J. C., A. K. Katyal, J. L. Zhu  and S. Mishra. Estimation of spill volume from
 monitoring well networks, In Proc.  4th Nat. Outdoor Action Conf., Las Vegas, May
 1990.

 Parker, J. C., T. Kuppusamy, and B. H. Lien,  Modeling multiphase organic chemical
 transport  in soils and groundwater, Proc. Groundwater Contamination: Use of Models in
 Decision Making, Amsterdam, Martinus Nijhoff, October 1987.

 Parker, J.  C. and R. J. Lenhard, A model for hysteretic constitutive relations governing
 multiphase fluid flow, 1.  Saturation-pressure relations, Water Resour. Res., 23, 2187-
 2196, 1987.

 Parker, J.  C.  and  R. J. Lenhard, Vertical integration of three-phase flow equations  for
 analysis of light hydrocarbon plume movement,  Transport in Porous Media, 5:187-206,
 1989.

Parker, J. C.  and R. J.  Lenhard,  Determining three-phase  permeability-saturation-
pressure relations from two-phase system measurements,  J.  Petrol.  Sci. Eng., 4, 57-65,
1990.
                                     208

-------