WATER POLLUTION CONTROL RESEARCH SERIES
16060 ELJ 03/72
   DENSITY INDUCED MIXING
     IN COUCINED AQUIFERS
U.S. ENVIRONMENTAL PROTECTION AGENCY

-------
          WATER POLLUTION CONTROL RESEARCH SERIES
The Water Pollution Control Research Series describes the
results and progress in the control and abatement of pollution
in our Nation's waters.  They provide a central source of
information on the research, development, and demonstration
activities in the water research program of the Environmental
Protection Agency, through in-house research and grants and
contracts with Federal, state, and local agencies, research
institutions, and industrial organizations.

Inquiries pertaining to Water Pollution Control Research Reports
should be directed to the Chief, Publications Branch (Water),
Research Information Division, R&M, Environmental Protection
Agency, Washington, D. C.  20460

-------
              DENSITY INDUCED MIXING IN CONFINED AQUIFERS
                                  by
                             L.  W.  Gelhar
                             J.  L.  Wilson
                             J.  S.  Miller
                             J.  M.  Hamrick
   Ralph M. Parsons Laboratory for Water Resources and Hydrodynamics
Department of Civil Engineering, Massachusetts Institute of Technology
                    Cambridge, Massachusetts 02139
                                for the

                  Office of Research and Monitoring

                    ENVIRONMENTAL PROTECTION AGENCY
                          Project #16060 ELJ


                              March 1972

-------
               EPA Review Notice
This report has been reviewed by the Environmental
Protection Agency and approved for publication.
Approval does not signify that the contents neces-
sarily reflect the views and policies of the Environ-
mental Protection Agency, nor does mention of trade
names or commercial products constitute endorsement
or recommendation for use.
                      11

-------
                              ABSTRACT

Analytical techniques are developed to describe the mixing between two
fluids of different density in a confined aquifer,  in which one fluid
is introduced to the aquifer by well recharge.  The immiscible displace-
ment process in both linear and radial flows is analyzed and the effects
of longitudinal and lateral dispersion are included using a boundary
layer approximation.  The theoretical results demonstrate the effect of
hydrodynamic dispersion in retarding gravity segregation due to density
differences.

The theoretical results are compared with observations of aquifer mixing
in linear and radial flow laboratory models.  During recharge excellent
agreement between the theoretical predictions and experimental results
is found and the predicted retarding effects of longitudinal dispersion
are verified.  During withdrawal some systematic differences between the
theory and observation are noted.  Theoretical predictions of recovery
efficiency during a recharge-storage-withdrawal sequence show trends
similar to those observed but are typically 5 to 10% lower than those
observed.

Direct theoretical predictions of recovery efficiency during single or
multiple sequences of recharge-storage-withdrawal are developed for an
immiscible system, and similar developments outlined for miscible dis-
placement .  The results are suggested for use in developing preliminary
estimates of operating conditions in the field.

This report was submitted in fulfillment of Project Number 16060ELJ
under the sponsorship of the Office of Research and Monitoring, Environ-
mental Protection Agency.
                                  iii

-------
                             CONTENTS









Section                                                        Page




           Conclusions                                           1




           Recommendations                                       3




   1       Introduction                                          5




   2       Analysis of Density Induced Mixing in Linear Flow     9




   3       Analysis of Density Induced Mixing in Radial Flow    31




   4       Linear Flow Experiments                              71




   5       Radial Flow Experiments                              79




           Acknoxjledgments                                     111




           References                                          113




           Definition of Symbols                               117




           Appendices                                          121

-------
                               FIGURES


Figure                          Title                           Page

 2.1      Definition Sketch for Linear Flow                      10

 2.2      Comparison of One-Dimensional Concentration Model      10

 2.3      Theoretical Horizontal Interface Projection for
            Linear Flow                                          23

 2.4      6/L as a Function of Time for the Case of No
            Dynamic Effects of Grain Dispersion                  25

 2.5      6/L as a Function of Time for Interaction between
            Grain Dispersion and Density Induced Mixing          26

 2.6      g(L/6) for One-Dimensional Linear Flow                 29

 2.7      Total Dispersion Coefficient, Et, as a Function
            of Time for Linear Flow                              29

 3.1      Definition Sketch of a Radial Confined Aquifer         32

 3.2      Comparison of Interface Shapes for Various Values
            of e  During a Recharge Operation                    37

 3.3      Inverse Interface Slope, 0, as a Function of Time
            During Recharge            "                          56

 3.4      Inverse Rate of Change of Relative Concentration,
            At, as a Function of Time During Recharge            57

 3.5      A/BH as a Function of Time During Recharge for the
            Case of no Dynamic Effects of Grain Dispersion       59

 3.6      A/3H as a Function of Time During Recharge for the
            Case of Interaction betx^een Grain Dispersion and
            Density Induced Mixing                               62

 3.7      Sketch of the Interface at the End of Withdrawal       66

 3.8      Recovery Ratio, R, as a Function of e  and e
            (ts=0)                             r      w          ^

 4.1      View of the Linear Sand Box Model                      72

 4.2      Horizontal Projection of Interface, L, as a
            Function of Time                                     75

                              vii

-------
Figure                          Title                           Page

 4.3      A Typical Linear Flow Experimental Breakthrough
            Curve Compared with Theory                           77

 5.1      View of the Radial Model                               80

 5.2      Photographs of the Radial Model in Operation            81

 5.3      Location of Conductivity Probes                        83

 5.4      Detail of Well Construction                            85

 5.5      Calibration of Averaging Conductivity  Probes            88

 5.6      Media Dispersivity Calculated  from the Averaging
            Probes                                               93

 5.7      Piezometric Head Distribution  in Radial Model           94

 5.8      Comparison of  Visual  and Electronic Measurement
            of  the Interface with  the Immiscible Theory           97

 5.9      Comparison of  Theory  and Experimental  Time
            Dependent Concentration Slopes During Recharge        99

 5.10      Volumetric Length of  the Interface, gH = AV, as a
            Function of  Time                                    100

 5.11      Comparison of  the Coupled Theory and Experimental
            Time  Dependent  Concentration  Slopes  During
            Recharge                                            101

 5.12      Experimental Breakthrough Curves at the Well           103

 5.13      Experimental Breakthrough Curves During Recharge
            and Withdrawal                                       104

 5.14      Comparison of  Experimental  and  Theoretical
            Breakthrough Curves During Recharge                  105

 5.15      Comparison  of  Experimental  and  Theoretical
            Breakthrough Curves During Recharge                  105

 5.16      Comparison  of  Experimental  and  Theoretical
           Breakthrough Curves During Withdrawal                105

5.17     At as a Function  of Time  During  a Recharge-Storage-
           Withdrawal Sequence, Run A-2                         106

5.18     At as a Function  of Time  During  a Recharge-Storage-
           Withdrawal Sequence, Run A-7                         106

                              viii

-------
                               TABLES


Table                           Title                           Page

 3.1      Summary of Equations Applicable to a Recharge,
            Storage, Withdrawal Operation                        69

 4.1      Range of Test Results for the Linear Model             76

 5.1      Radial Model Horizontal Hydraulic Conductivity         89

 5.2      Determination of Dispersivity from Averaging
            Probes                                               92

 5.3      Experimental Conditions for the Radial Model           96
 5.4      Observed Values of -QAt/V  during Withdrawal          108

 5.5      Comparison of Theoretical Recovery Ratios with
            Experiments                                         109
                                  IX

-------
                             CONCLUSIONS

1.  Analytical techniques have been developed to describe the combined
effects of density difference and hydrodynamic dispersion on mixing in
linear and radial flows in confined aquifer systems.   The theoretical
results demonstrate the retarding effects that dispersion may have on
the rate of tilting of the interface.  The roles of both lateral and
longitudinal dispersion in the mixing are established.

2.  Laboratory experiments with linear and radial flow aquifer models
confirm several aspects of the theoretical developments.  During re-
charge operations the predictions are found to be in excellent agree-
ment with the observed interface tilting.  In most of the recharge
experiments, the interface motion is dominated by immiscible dynamics
but some experiments confirm the predicted retarding effects of hydro-
dynamic dispersion.  However, during withdrawal there are significant
differences between the observed and predicted interface motion.  Pre-
dictions of recovery ratios showed trends similar to those observed
although the predicted values were typically 5 to 10 percent lower.

3.  Direct application of the analytical results in the evaluation of
field operations is feasible.  The results of the immiscible theory
(Eq. 3.119) provide conservative estimates of the recovery ratio for a
recharge-storage-withdrawal sequence in terms of aquifer and fluid pro-
perties and flow rates.  Multiple recharge-storage-withdrawal sequences
can be treated and improved recovery efficiency is found after a few
sequences.  Conditions under which dispersive effects become important
can be evaluated (see Section 3.4) and these effects can be included
using the computational procedure developed in Section 3.5.

-------
                           RECOMMENDATIONS

1.  Several additional factors of importance In field situations should
be considered in future analytical and experimental studies;  these in-
clude partially penetrating wells, phreatic aquifers, sloping and multi-
layered aquifers, and ambient flow in the aquifer.
2.  Analytical or numerical solution techniques, should be developed to
evaluate the implicit limitations of the proposed analytical models;
these should include evaluation of the effects of interface curvature
during the storage and withdrawal cycles and influence of no-flux boun-
dary conditions at the aquifer boundaries.
3.  Field experiments are suggested for further evaluation of the pro-
posed analytical techniques.  This involves the development of a series
of experiments in which breakthrough curves are observed at several
monitoring wells and at the recharge-pumping well during a recharge-
storage-withdrawal sequence.  The hydraulic conductivity and dispersivity
of the aquifer should be determined as part of the experimental program.

-------
                           1.  INTRODUCTION

                       1.1  General Discussion

The pressures of population and economic growth lead inevitably to
problems in water resource management.   One method of trying  to optimize
present water resources to keep up with growth involves  the utilization
of ground water aquifers for storage or disposal of surface or waste
waters.  During periods of the year when there is excess surface water
available, it can be pumped into aquifers through wells  for storage and
use during dry periods.  This use of underground aquifers for water sto-
rage is attractive because of the scarcity of natural reservoir sites,
the low initial costs compared to ordinary surface reservoir  construc-
tion, the elimination of evaporative losses, and the possibility of par-
tial purification of the stored water during its passage through the
aquifer.

These storage schemes can be used to either mix and dilute contaminated
surface waters with naturally occuring ground water, or  for storage of
surface water of high quality in an aquifer initially containing water
of lower quality.  This latter case in which fresh water is stored, for
example, in a saline aquifer could find wide application in the United
States where one third of the land area has saline ground water less
than 500 feet below the land surface (see Kohout, 1970).  Moulder (1970)
has reported field studies on the recharge of saline aquifers with fresh
water and has suggested specific geographic areas where  this  scheme may
be feasible.

Ground water aquifers are also finding wide use for disposal  of liquid
wastes and the trend is toward increased use of aquifers for  this pur-
pose.  The number of industrial waste injection wells, for example,
increased from close to 100 in 1967 to nearly 150 in 1969 (see Kohout,
1970).  In addition to the disposal of normal wastes, ground  water
aquifers can also be used to dispose of thermally polluted waters.

When surface water is introduced into an aquifer, mixing occurs with  the
native ground water.  Usually the surface water and native ground water
differ in physical, chemical, and biological characteristics.  In de-
signing and operating recharge or disposal systems, it becomes necessary
to predict and control, as much as possible, the amount  of mixing between
the injected and native waters.

The mixing in the aquifer is caused by the large scale convective motion
of the native ground water as well as the flow from wells during recharge,
disposal, or recovery.  Further mixing is due to hydrodynamic dispersion
which includes the effects of molecular diffusion as well as  what may be
termed grain dispersion.  Grain dispersion is caused by the velocity vari-
ation of the fluid flowing along different paths within the pore structure
of the medium.  A final mixing process comes about because of differences
in the physical properties of the native and injected waters.  Differences
in density may be a significant cause of mixing due to gravity currents

-------
 that develop.  The resulting convective motion  increases mixing  in the
 aquifer.  The storage of fresh surface  water  in saline aquifers  is an
 example where density induced mixing is important.

                     1.2  Review of  Previous Work

 Considerable work has been done on  hydrodynamic dispersion and its
 effects on the mixing of miscible fluids flowing in  porous media.
 Schiedegger (1961) and Bear and Bachmat (1964)  theorized a linear re-
 lationship between the dispersion tensor and  the velocity of flow
 through porous media.  Harleman and Rumer (1963)  and List and Brooks
 (1967)  gave experimental evidence that  some nonlinearity  between the
 velocity and dispersion coefficient may exist.   Studies by Li and Yen
 (1968)  indicated that differences in viscosity  and density have  little
 or no effect on the local mixing due to dispersion at the interface.
 More recently Gelhar and Collins (1971) developed an analytical  solu-
 tion that describes the concentration of a substance at a particular
 radius  for flow from a well.   Numerical solutions of the dispersion
 equation have been developed  by Shamir  and Harleman  (1967), Reddell and
 Sunada  (1971), and Guymon,  Scott, and Hermann (1970).  Rose and  Passioura
 (1971)  showed that current  methods  for  calculating the coefficient of
 hydrodynamic dispersion during lab  tests may  be profoundly affected by
 even seemingly negligible density differences between the original fluid
 and the fluid used as a tracer during tests.  Furthermore Theis  (1967)
 and Mercado (1967), demonstrated that field determinations of dispersion
 coefficients may be several orders  of magnitude greater than those in-
 dicated by laboratory tests because of  the large  scale heterogeneity of
 natural deposits.   These factors should be considered when studying and
 applying results from tests on hydrodynamic dispersion in the laboratory.

 There has also been extensive work  done on salt water intrusion  in
 coastal aquifers.   Dispersion and density induced mixing are both im-
 portant in the study of salt  water  intrusion.   Henry (1960) studied the
 combined effects of gravity and dispersive mixing for a steady saline
 wedge in a confined aquifer.   Rumer and Harleman  (1963) studied  the ef-
 fects of tidal motion on interfacial dispersion.  Bear and Dagan (1964)
 developed an approximate analysis of the unsteady movement of a  salt
 water wedge due to a change in the  fresh water  flow.  They compared their
 results with model experiments.  Shamir and Dagan (1971) developed numer-
 ical  solutions for the  motion of the seawater interface in coastal
 aquifers.

 There have  been a  few investigations  of dispersion and density induced
mixing  more directly  related  to  this  paper.   Esmail  and Kimbler  (1967)
 and Kumer  and Kimbler (1970)  studied the possibility of storing  fresh
water in saline  aquifers.   Their work was  derived from an "ad hoc" as-
 sumption of  extending results  based on  the rate of interfacial tilting
 in a  linear model  with  no net  flow  to radial  flow from a well using a
numerical procedure which was  not explicitly  described in the papers.
Although their  procedure involves this  extrapolation from linear to
radial  flow,  they  did report  good agreement between  their numerical

-------
simulations and data collected in a three dimensional radial flow model.
Perrine and Gay (1964) developed numerical solutions of the problem of
density induced motion during radial flow from a well with density and
viscosity difference between the two fluids.   Their solutions did not
take into account hydrodynamic dispersion at  the interface, and no com-
parison was made between their theory and experimental or field data.

Green and Cox (1968) developed numerical solutions of the equations of
motion and mass conservation based on the method of characteristics.
They analyzed miscible displacement of fluids with different density and
equal viscosity in a linear model using constant dispersion coefficients
in a numerical scheme.  Results of the numerical analysis were compared
with those from a laboratory model.  Keulegan (1954) and Gardner, Downie,
and Kendall (1962) reported experiments on the tilting of the interface
between fluids of different density in linear laboratory models with no
net flow.  De Josselin de Jong (1960) developed the potential flow solu-
tion for the initial motion of a vertical interface in a linear model
with no discharge.  Gardner, Downie, and Kendall (1962) have also used
the immiscible approximation to develop analytical descriptions of inter-
face tilting for small and large times under  conditions of no net flow.

                      1.3  Objectives and Scope

The primary objective of this study is to investigate the effects of hydro-
dynamic dispersion and density stratification on mixing in ground water
systems.  Of particular interest is the mixing process which occurs when
a fluid whose density differs from that of the native fluid is injected
into a confined aquifer.  These mixing phenomena can be important in ap-
plications involving the storage of fresh water in saline aquifers or
deep-well waste disposal.  Methods of evaluating and predicting these
mixing phenomena under field conditions are sought.

The general approach includes the development of analytical models which
describe the effects of density differences and dispersion for two flow
configurations.  A series of laboratory experiments in also developed to
evaluate the analytical predictions.

The analytical methods are first developed for mixing in a linear flow
with a constant net convective velocity, and later extended to mixing in
a radial flow well system.  Initially, the theoretical analyses are based
on immiscible displacement for the case of fluids with different density
and equal viscosity.  The effect of dispersion on the density induced
tilting of the interface and the overall mixing is then analyzed theore-
tically using boundary layer type approximations.  From this analysis it
is possible to determine when dispersion will influence the interface tilt-
ing and the overall mixing.  The analyses also reveal the conditions under
which longitudinal or lateral dispersion will dominate the dispersive mix-
ing and influence the interface tilting.  The relative importance of grain
dispersion and density induced mixing is characterized by a dimensionless
parameter for both flow regimes.

-------
Based on the results of the linear flow analysis., a concept of gross dis-
persion due to density difference is introduced and a total dispersion co-
efficient representing both density and grain dispersion is also derived.

The radial flow system is analyzed for the three processes of recharge,
storage and withdrawal.  The results are used to simulate multiple se-
quences of recharge, storage and withdrawal.

The results of laboratory experiments designed to evaluate the theoretical
developments are presented.  The linear flow case is simulated in a hori-
zontal linear sand model.  In the radial flow case a 15° radial sector
model of consolidated plastic beads was constructed.  Interfacial tilting
and gross mixing were measured by visual and electronic means.

The application of the theoretical results to field operations is consid-
ered and predictions of recovery efficiency in a recharge-storage-withdrawal
operation are developed.

-------
       2.  ANALYSIS OF DENSITY INDUCED MIXING IN LINEAR FLOW

The objective in this section is to develop methods of analysis which
can be used to describe the combined effects of dispersion and gravity
segregation on aquifer mixing.  The flow configuration considered here
is a one-dimensional displacement in a confined system with horizontal
boundaries, i.e., a linear flow.  A theoretical analysis of immiscible
displacement in a model with a net flow is developed for the case of
fluids with different density but equal viscosity.   Based on this
analysis, a new concept of gross dispersion due to  density differences
is introduced.  Using boundary layer approximations, the effects of
grain dispersion on the tilting of the interface and the overall mixing
are analyzed.  The effects of both longitudinal and lateral dispersion
are included in the analysis.  In Section 3, these  same concepts and
methods of analysis are extended to the case of the radial flow system.

2.1 Immiscible Displacement in a Confined Aquifer of Constant Thickness

The problem of mixing of two liquids of different density in a homogen-
eous, confined aquifer may be simplified by neglecting grain dispersion
and molecular diffusion.  In such a case, the mixing process is des-
cribed by an immiscible displacement model.  Consider the displacement
of a heavier liquid with density p.,, by a lighter liquid with density
p,, in a horizontal confined aquifer of constant thickness H (see
Figure 2.1). If the two liquids have approximately the same kinematic
viscosity, v, (i.e., fresh and salt water), the hydraulic conductivity,
K, may be taken as a constant throughout the flow.

The following analysis is based on the Dupuit approximation, which
assumes hydrostatic pressure distributions and thus essentially hori-
zontal flow.  The analysis is strictly correct only when the interface
slope is small and hence the vertical velocities are small.

The continuity equations for the two liquids are:
                                      IF                          (2-2>
where £ is the distance to the interface, in liquid 1, measured from  the
top of the aquifer, and r\ is the distance to the interface, in liquid 2,
measured from the bottom of the aquifer.  The seepage velocities of the
two liquids, u.. and u«, are given by the Darcy equations for  the two
liquids,

                                  K 3*1
                           Ul = - --jr-i                           (2.3)
                            1     n 3x

-------
                1
«f»1-H/2
                      u
<}>0-H/2
                                                            H
     //ANN //A\\//A.\s //A\\//A\V/A \V/ANV /A\\/ /' NV/ /\\ V//Qv


                   f-i	 L 	*•]



       Figure 2.1  Definition Sketch for Linear Flow
     1.0
C/C   .6 -
   o
      .4 -
      .2 -
                                                    1	T
                       Eq.  2.33
                           Eq. 2.24
                           I
                                        r	L
             -.6   -.4   -.2   _0    .2     .4    .6

                               x/L



   Figure 2.2  Comparison of One-Dimensional Concentration Models
                              10

-------
                                 K
Vertical velocities are neglected due to the Dupuit approximation.  The
porosity of the aquifer is n, and the piezometric heads of the two liq-
uids are (j>. and~.  Along the interface, the fluid pressure must be
continuous.  The pressure continuity equation is
                       *2 =    *l -     (C-H/2)                   (2.5)

                       Ap = P2 - PI

If the density difference, Ap , is small compared to p9, and the  rela-
tive density difference is identified by Ap/p_ = Ap/p, Eq .  (2.5) becomes


                                                                 (2-6)
The continuity equation for the entire flow is obtained by adding  Eqs,
(2.1) and (2.2) and integrating
                        U;LC + u2n = UH                           (2.7)


The convection of the interface through the aquifer  is  due  to  the  net
seepage velocity, U.

The solution of the problem consists of determining the  interface shape
and location, and determining the velocities in the two  liquids.  Sub-
stituting Eqs. (2.3) and  (2.4) into Eq. (2.7) gives,
                                                                 (2'8)
Differentiating Eq.  (2.6) and substituting into  Eq.  (2.8)  and simpli-
fying the results using C + n = H, gives

                     ^ ,i
                       1     nU _ Ap_ _n !n_                       (•) Q\
                     8x    ~ K    p  H 3x                       ^   '

Substituting Eq.  (2.3) into Eq. (2.1) and eliminating  £  using £ = H - n
gives
                                  11

-------
 The partial differential equation for n is obtained by substituting
 Eq. (2.9) into Eq. (2.10).
                                            .
                                            .
                   3t     8x   n p    2   2  ~ 3H
 Introducing a conyected coordinate x - x - Ut ,  Eq.  (2.11)  is simplified
 tQ
                                         _
                      3t   n p  ~* •  2    3H
 Equation (2.12)  may be reduced to  an ordinary  differential  equation
 using a similarity transformation.
                                      = F(z)                      (2.13)
                                H/F
The resulting ordinary differential equation in z, the similarity
variable, is

                                2     23
                        2 dz ~  , 2   2  ~ 3
                               dz
Equation (2.15) is solved for an initially vertical interface by
assuming a power series in z, to give
                      n =   ( H + — )                            (2.16)
This solution may not apply for small times, when  the interface  slope
is large and vertical velocities are not negligible.

The expression for the velocity in liquid 1 is obtained using  Eqs .
(2.3) and (2.9).
                             u +
                         1       n  p H
                                 12

-------
Using Eqs. (2.4), (2.6) and (2.9) the expression for the velocity in
liquid 2 is obtained.


                     „  - IT   - ^- (1 - Ih 8n                    (7
                     U2 ~ U   n  p (1   H} "3x~                    (

From Eq. (2.16), the horizontal projection of the interface, L, is
                      L = 2H/f & I = 2H/F                      (2.19)


The position of the interface is given by x., the distance to the inter-
face from the x origin.


                           x£ = Ut +|y                         (2.20)


Results of Eq. (2.16) indicate that the interface is a straight line,
of decreasing slope, which is convected through the aquifer with a
velocity U.  The solution is also valid for the case of no through
flow, U = 0.

If the density difference is due to the presence of some material in
solution (i.e., salt), and the density is a linear function of the
concentration of the material, c, then
                          p = P  + Ap --                          (2.21)
The concentration of material in liquid 2 is co.  To obtain a one-
dimensional description of the variation of the liquid density, the
concentration, c, is averaged over the aquifer thickness.
                                  H
                          I = M 2  c dy                         (2.22)
                              H J  H
                                   2
Using the indicated concentration distribution,
                   c = co * - I < y < -  (f -
                   c = 0  , -  (f - n) < y < f
                                                                  (2.23)
                                 13

-------
 The average concentration is

                        —      n    C    —   T
                        c = c  — = — ( x + — 1                    O ")L\
                        *—      IT    T  ^     o  *                    \ *-•*•"/

 which indicates a linear distribution of average liquid density,


                           p" = p   + Ap |-                         (2.25)
                                        o

 The average flux of_dissolved material with respect to the convected
 coordinate  system (x,y)  is
                                 H
                         J  =         pcu dy                        (2.26)
                             H  J   g
                                "  2

                         "u  =  u  - U
which is evaluated, using  Eqs.  (2.18)  and (2.23),  as

                   j^.coKAP./.TK.nlH
                   J     co  P2  n  p  U   H;  H 3x

Using Eq.  (2.24), Eq.  (2.27) is modified  to  the form

                     —       K Ap .    TK   8c                   .
                     J " * P2 nF (1  "I)n  ^                   (2

For a dispersive mixing  process,  the flux is related  to  the concentra
tion gradient through  a  dispersion coefficient,  E,  as  follows
                                                                  (2.29)
The dispersion coefficient can  thus be  evaluated  by  equating Eqs.
(2.28) and (2.29) as
for P2 = p".  The result of Eq.  (2.30)  suggests  that  the  gross character-
istics of density induced mixing may be  approximated by  the one-dimen-
sional dispersion equation of the form
                                 14

-------
 H*'H-

                                                                 (2.31)
where EI is a constant dispersion coefficient.  Although Eq.  (2.30)
implies a parabolic distribution of  the  dispersion coefficient, E,
over the length of the interfacial zone,  satisfactory results can be
obtained by using the average value.
                             _ - _  H K Ap
                          h,  — h —  —	
                           1       6 n  p
                                          (2.32)
This dispersion coefficient,  which characterizes the mixing due to
density differences* will be  identified  as  the  density dispersion co-
efficient.  This expression for  E  is later  used  to estimate the im-
portance of the effect of the density difference on the mixing process.

The longitudinal distribution of average concentration can thus be
determined approximately by solving Eq.  (2.31)  subject to the initial
condition
                        t = 0
                                c =  c
                  x>0
                                c =  0
                  x<0
which corresponds to an initially vertical  interface  at  the  origin
x = 0.  The solution is
c
C
                            1,1   c,
                            - + - erf (
                            2   2
(2.33)
A comparison of this one-dimensional approximation with  the  exact  re-
sult from Eq. (2.24) is shown in Figure 2.2.

In many field situations, it is the average water quality, over  the
aquifer thickness, which is of importance.   If the mixing process  is
assumed to be due only to the density difference between the mixing
liquids, the one-dimensional model represented by Eq.  (2.33) will  ap-
proximately describe the mixing process, and  predict  the depth-averaged
water quality.  The concept of a one-dimensional dispersion  model  to
represent a two-dimensional density induced mixing process is  an im-
portant simplification of this phenomenon.

                   2.2  Grain Dispersion Effects

The immiscible displacement model is an idealization, since  there  is
always some degree of interfacial dispersion.  The  objective here  is  to
estimate the importance of grain dispersion in comparison  to the density
                                  15

-------
 induced mixing.

 The validity of  the assumption of immiscible  displacement  depends  on
 the magnitude of the density difference Ap .   In certain  cases  the
 density difference will be large enough to  completely  govern  the
 dynamics of the  mixing, (i.e., the velocity field is that  predicted by
 the immiscible model);  however,  the resulting model may  not adequately
 describe the variation  of density near the  interface.  To  obtain a
 better description of the density variation,  grain dispersion  must be
 considered.

 The mass conservation equation which describes the distribution of
 solute concentration c  in a porous medium is  [Bear, et al, 1968]


              9c  ,    3c  ,    3c   3  ,_   3cx  ,  3  ,_ 3cx           /„ ~,s
              ^+U 3^+V^ =  H °W  +  ^T (D2 37}           (2'34)

 where  u and v are  the horizontal and vertical seepage  velocities respec-
 tively,  and D-. and D~ are the longitudinal  and lateral dispersion  coef-
 ficients respectively.   To simplify the solution of Eq.  (2.34), the
 following assumptions are made.

 (i)  The dispersion coefficients,  D..  and D~ ,  are generally velocity
 dependent,  but for this analysis the velocxty changes  are  assumed  small
 and  D   and D  are  taken as constants and proportional  to the net seepage
 velocity U.

 (ii)   The  vertical convective transport may be neglected when  the  Dupuit
 approximation, which  assumes  negligible vertical velocities, is
 applicable.

 (iii)  The  convective velocity in  the  dispersed region can be  approxi-
mated  by  the velocity of  the  interface as determined by  Eq. (2.20).

                                     3x.
                            u  = u.  = —=-                         (2.35)
                                 1    O L

The  interface is idealized as  the  line of 50% concentration.

The  three assumptions simplify Eq.  (2.34) to  give,

                                     2         2
                    3c ,      3c       9 c     3  c                  .„  ,,
                    3F + ui  3i "  Di 7T + Dz 72                 (2'36)
                                    3x       3y

and  the  interface  location  is  as given by Eq.  (2.20).  The interface
slope,  S, and its  reciprocal,  3, are defined  as
                                                                  (2-37)
                                 16

-------
Equation (2.20) is rewritten as
                           x. = Ut + 3y
                                                       (2.38)
The following coordinate transformation will be used:
                         Xl =
                                                       (2.39)
Expressing the concentration c as a function of x , y, t, Eq.  (2.36)
becomes
  3c
  3t
     xi>:
                ^
- r>    c  + n   8 c
~ Dl ,  2 + D2I    2
                                                            3x
Since the lines x  = constant are parallel xvith the interface,  it  is
reasonable to assume that
                                    = 0
which can be verified from the results  (Eq. 2.44).  The  convective
dispersion equation then reduces to
                          . 0
                                     o

                                 c  =  0     x<0
                                                        (2.43)
                                 17

-------
 The solution of Eq.  (2.42)  subject to these conditions  is
                         c
                         c
                                              (2.44)
 If the dynamics of the mixing are given by the  immiscible  displacement
 model, 3 is given by Eq.  (2.19),  and hence £  may be  determined.   From
 Eqs.  (2.19) and (2.41),
                       = 2/F
5  = D^ + 2D2T t
                                                                  (2.45)
                           j.     f
                   — =  2  +2  erf
                     o
                                        x -  x.
                                              (2.46)
 The  result  of  Eq.  (2.46)  indicates  that  as  time  increases,  the  impor-
 tance  of  lateral dispersion  increases.   The approximate width of  the
 dispersed zone about  the  interface,  6, is defined  as
            6  =
                  3(c/co)
                         x=x.
                             1-
                                -1
                                   )t
(2.47)
As  the width  of  the dispersed  zone,  6, becomes  large, dynamic  effects
of  grain  dispersion are expected, and the  dynamics of the mixing  are
not given by  the immiscible model.   The ratio 6/L, where L  is  given by
Eq.  (2.19) is proposed as a criteria to estimate  the importance of
grain dispersion.
                                                  1/2
                          7TD
                                                                  (2.48)
The result given by Eq.  (2.46) is expected to be valid if  6«L.  As  6
approaches L it becomes necessary to consider the dynamic  effects of
the grain dispersion.  This analysis will be developed in  the next
section.

2.3 Interaction Between Grain Dispersion and Density Induced Mixing

In Section 2.2 the additional mixing due to grain dispersion was deter-
mined under the assumption that grain dispersion has a negligible in-
fluence upon the dynamics of the process.  As mixing due to grain
dispersion becomes large, the immiscible model of Section  2.1 is no
                                 18

-------
longer adequate to describe the motion.  To determine the influence of
grain dispersion on the fluid motion, the mass and momentum equations
must be solved simultaneously.  Coupling the mass and momentum equations
for the entire flow field results in non-linear partial differential
equations, which are extremely difficult to solve.  However, a useful
approximation can be developed by coupling these equations at the mean
interfacial position, x = x..

The governing equations are dynamic equations (Darcy equation)


                            u=-^f                          (2.49)
                                  np 3x
                                                                 (2'50)

where p is the pressure, k is the permeability, and y is the dynamic
viscosity.  The equation for conservation of total mass is, under the
conditions of small density difference (Ap/p «1)


                           |^4^=0                           (2.51)
                           3x   3y                                    '

Using the same assumptions made in Section 2.2, the mass conservation
equation for a solute of concentration c is

                                    2        2
                   3c      3c      3 c      8 c                  ,0 ,.„.
                   -3T + ui 3x~= Di7T + D27T                  (2-52)
                                   3x       3y

The Dupuit approximation assumes a hydrostatic pressure distribution,
8p/3y = - pg, and thus from Eq. (2.49) it follows that


                        3u _ kg_ 3p _ K 1 3p
                        "^  ~    "^  —     "^                       \^-
                        3y   ny 3x   n p 3x

where K is the hydraulic conductivity.

Equation (2.53) can be evaluated at the mean interfacial position,
x = x
                       3y
                          x=x.
                             i
                                                                  (2.54)
If the width of the dispersed zone, 6 = AirC  , is large  in  comparison
to the horizontal interface projection, L, the velocity  gradient  in
Eq. (2.54) can be approximated by
                                 19

-------
                            _3_u
                            9y
                                      3u.
                                        x
                                                                  (2.55)
                               x=x.
 The range of validity of this approximation is evaluated in detail in
 Appendix A.   Using this approximation, Eq. (2.53) can be rewritten as
                      3u.    3 x.
                                   K
                                          x=x.
                                             i
                     3y    3y3t   n p 3x


Using Eq. (2.38), Eq. (2.56) may be rewritten in terms of 3
                          dg = K 1 3p
                          dt   n p 3x
                                                                  (2.56)
                                                                  (2.57)
                                      x=x.
 The  solution  of  Eq.  (2.52)  for the appropriate boundary conditions is
 that obtained in Section 2.2.
c_
c
                         o
                             11    cf
                           =2+2  6rf
                                      \
                                        x-x,
                                           J
                                                                  (2.58)
                                           dt
                                                                  (2.59)
Differentiating Eqs.  (2.21)  and  (2.58)  and combining gives

                          _   Ap         (X-XJ)2
                                                                  (2.60)
To couple the mass  and momentum equations at the mean interfacial posi-
tion, Eq. (2.60) is evaluated at x = x.,  and substituted into Eq. (2.57)
                          _d_B _  K A
                          dt =  n P
                                                                  (2.61)
Equation  (2.61) must be solved  simultaneously with Eq.  (2.59) in dif-
ferential form.
                                                                  (2.62)
The solutions of Eqs.  (2.61) and  (2.62)  with the initial conditions
                                  20

-------
                      3 = 5 = 0   at   t=0
are, for the reciprocal slope 3,

                       »2 + i££-!l-*                      (2-63)

and for the dispersed zone width, 6,
The ratio of the dispersed zone width to the horizontal  projection of
the mean interface  is,  from Eqs. (2.63) and (2.64),

Equations (2.63)  and  (2.65) may be approximated for both large  and
small interface slopes, noting that D.. is usually an order of magnitude
larger than D .  For  large slopes, the lateral dispersion term  is neg-
ligible and Eqs.  (2.63) and (2.65) reduce to
                              :
If the interface slope is  small  (large 3), longitudinal dispersion is
negligible and Eqs.  (2.63)  and  (2.65) become
                                                                (2.69)
It should be noted that, in the analyses  of  grain dispersion in Sections
2.2 and 2.3, the boundary conditions  of no mass flux at the upper and
lower aquifer limits were not explicitly  applied.  The resulting solu-
tions are therefore strictly valid only for  aquifers which are of infinite
vertical extent.  It can be seen by direct evaluation of the vertical
                                21

-------
 mass flux at the boundaries (y = ±H/2)  that the boundary  conditions  are
 not exactly satisfied.   In Appendix B,  some possible effects  of this
 boundary condition are  explored.   However,  no explicit  evaluation  of the
 effects was obtained,  other than an indication that the boundary effects
 may become important when 6/L is large.

               2.4  Discussion of Theoretical Results

 In the previous sections, theoretical models were presented to describe,
 for a one-dimensional  linear flow in a  confined aquifer,  the  mixing  of
 two fluids due to the  effects of density difference and grain dispersion.
 A_lso introduced was the concept of the  density dispersion coefficient,
 E, the magnitude of which may be used to estimate the relative importance
 of density difference  in the mixing process.

 In the first model (Section 2.1)  only the effect of density difference
 is included.   An immiscible displacement process is assumed and results
 describing the tilting  of an interface  as it is being convected through
 the aquifer at the net  seepage velocity are obtained.   One character-
 istic of this model is  the horizontal projection of the interface, L,
 given by Eq.  (2.19).  Previous investigators have obtained expressions
 for L,  for immiscible displacement in enclosed systems  with no net flow^.
 For the case of no net  flow, Gardner, et al.,  (1962)  obtained for  small'
 and large time respectively
                              L/H =  /3  T
(2.70)
                              L/H =  2/F
(2.71)
The second  result  (Eq. 2.71) was originally  developed by  Dietz  (1953)
and also agrees with  that  found in  this  study  (Eq.  2.19).   Gardner,
et al.,  (1962) also suggested an interpolation  formula which  encom-
passes Eqs.  (2.70) and (2.71) in the  form

                                   12t
                                   4+3T
(2.72)
In the development leading to Eq.  (2,19)  it was noted  that  the  result
was independent of the net seepage velocity and thus the agreement
between Eqs.  (2.19) and  (2.71) is expected.  The  results of  the immis-
cible displacement theory  (Eq. 2.19) and  Gardner's  interpolation form-
ula (Eq. 2.72) are shown in Figure 2.3.

In the second model (Section 2.2), the effects of grain dispersion were
considered under the assumption that grain dispersion  has no  influence
upon the dynamics of the flow.  Under this assumption, the horizontal
projection of the interface, now idealized as the line of 50  per cent
concentration, is the same as the immiscible displacement result,
                                 22

-------
                                                                                                 E/D,
         A  Upper^ Limit of Immiscible Displacement Model
            for E/D  = 87f/3  (5/1X1/2)
         B  Lowejr Limit of the Coupled Analysis
            for E/D  = 8ir/3  (6/L=2)
  10.0
         Immiscible Displacement
            Eq. 2.19
            Eq. 2.72
L/H
Coupled Theory (Eq. 2,63)
   2   	
   1.0
            6/L <2   —
                                                                          Increasing Effect of
                                                                            Grain Dispersion
   0.1
     0.1
                  1.0                  10.0                    100.0
                                             T  =  KApt/npH
           Figure 2.3  Linear Theoretical Horizontal  Interface Projection
                                                                                                  1000.0

-------
 Eq. (2.19).  An estimate of the width of the dispersed zone, 6, about
 the interface is given by Eq. (2.47),  As indicated in Section 2.2,  the
 immiscible model (Eq. 2.19) is valid when 6/L is small compared _tp unity.
 The ratio 6/L, as evaluated from Eq. (2.48) for three values of ₯/D,
 is shown in Figure 2.4 as a function of the dimensionless time, T.  From
 Figure 2.4, it is seen that even for large values of E/D ,  mixing due to
 grain dispersion will eventually become as important and influence the
 rate of interface tilting.  The exact limits of applicability of the
 immiscible model can not be specified but reasonable estimates are
 obtained by assuming that the limiting value of 5/L is a fixed number
 less than one, say 6/L - 1/2.  For a given value of E/D , the indicated
 maximum value of dimensionless time, T, for which the immiscible solu-
 tion applies, is found from Figure 2,4 or Eq.  (2.48).  For  example,  when
 E/D..  = 8rr/3 the maximum T = 15,  as shown in Figure 2.3.

 In the third model (Section 2.3),  effects of density difference and
 grain  dispersion are analyzed by an approximate coupling of the mass
 and momentum equations based on large 6/L.   The solution is given by
 Eq.  (2.63),  which represents a family of curves relating g, the inverse
 interface slope (g^L/H), and T for various values of E/D and D /D .
 Equation (2.63) has  two asymptotes, Eqs.  (2.66) and (2.68), for large
 and small interfacial slopes or small and large times ,_respectively.
 In Figure 2.3,  Eq.  (2.63) is shown for four values of E/D.. , with D /D_
 constant and equal to 10,  As noted in Section 2.3, the coupled analysis
 (Eq. 2.63)  is valid  for 6/L»l.   The ratio 6/L from Eq.  (2.65) is shown
 in Figure 2.5 for four values of E/D .   An approximate limit on the  range
 of application  of Eq.  (2,63)  can be^ found by taking 5/L = 2 in Eq. (2.65)
 or Figure 2.5.   For  example,  when  E/D.  =  877/3, the limiting value is
 T  = 200  and  the coupled  analysis (Eq.  2.63)  should be valid for times
 greater  thin this value  as indicated in Figure 2.3.  It  can be seen  that
 the regions  of  validity  of the immiscible theory (Eq. 2.19) and the
£pupled  analysis  (Eq.  2.63)  are  mutually  exclusive for a given value of
 E/D  .  In general, there will be an intermediate time range in which
neither  theory  is applicable.   In  the example  E/D..  = 8ir/3,  the range is
 15  < T <  200.   The actual behavior in this  range is not  known, but a
smooth transition between the models is expected.   The portions of the
curves representing  Eq.  (2.63)  for 5/L <  2  are shown as_dashed lines in
Figure 2.3.   It  is seen  that,  even for large values of E/D  , lateral grain
dispersion will  eventually influence the  interface tilting.  From Figure
2.3, it  is seen  that  as  E/D  decreases, interface  tilting becomes less
and as lateral  dispersion becomes  important, the rate of tilting decreases.

The ratio  6/L indicates  the  relative importance of the two  mixing mech-
anisms,  grain dispersion and  density difference.   When 6/L  is small
 (Section  2.2),  the grain dispersion has no  effect  on the interface
tilting and mixing dynamics  are  governed  by  the immiscible  solution.
When 6/L  is  large the  combined effects  of the  two  mixing mechanisms  are
important and the coupled analysis  governs  the mixing.   The coupled
solution  does not reduce  to  the  immiscible  solution as D ,D« -> 0,
because the  regions of validity  for the two  analyses do  not overlap.
                                 24

-------
Ul
              10.0
           6/L
               1.0 -
                 0.1
10.0
                                                            =  K _Ap_ t_
                                                          T    n p   H
100.0
1000.0
                Figure 2.4  6/L as a Function of Time  for  the  Case of No Dynamic Effects of Grain Dispersion

-------
    10.0 —
6/L
     0.1
        0.1
1000.0
              Figure  2.5   6/L as  a Function of Time for Interaction Between Grain Dispersion  and
                                           Density Induced Mixing

-------
A numerical example for some typical field conditions will illustrate
the importance of the different mixing phenomena.   Consider a sandy
aquifer 50 feet thick having a hydraulic conductivity, K, of 0.03 ft/sec
and a porosity, n, of 0.35.  For a_one percent   density  difference, the
density dispersion coefficient is E = 7.15 x 10"^  ft /sec.  For a hy-
draulic gradient of 0.01, the net seepage velocity from Darcy's law is
U = 8.6 x 10~^ ft/sec.  The longitudinal dispersion coefficient is esti-
mated to be DI = 2.8 x 10~° ft/sec from results by Harleman, Melhorn
and Rumer (19o3) and D.. /D? = 10 is assumed.  The ratio E/D  is approxi-
mately 26, indicating that the mixing process is initially one of immis-
cible displacement.  The immiscible model  (Eq.  2.19) will be valid up to
a dimensionless time T = 57 which is determined from Eq.  (2.48).  Thus a
period of somewhat over one month would be required before the interface
tilting would be influenced by grain dispersion.  After one month, the
horizontal projection of the interface, L, is 670 feet and the width of
the dispersion zone about the interface, as given by Eq.  (2.48), is
300 feet.

The results of the analyses of the combined effects of grain dispersion
and density- induced motion can also be interpreted in terms of a total
dispersion coefficient, E  , defined by
                                                                  (2'73)
where j  is the mass flux relative to the mean flow U, and c is the
depth-averaged concentration given by
                                 fH/2
                          -   1
                          C = H
               c  dy                         (2.74)

          -H/2
The total dispersion coefficient, E  , can be evaluated  directly  from
the results of  Sections 2.2 or 2,3 in the general form

                             ..   1     x-x.

                        ^"2 + 2«f~                       <2'75>


where x. = Ut + 3y  (Eq. 2.38), 3 is  given by Eqs.  (2.19)  or  (2.63)  and
C  by Eqs.  (2.45) or (2.64),  The average concentration is determined
using Eq. (2.75) in Eq. (2.74) with  the result

                     2    ?
                        -Q
[0erf9-6erf9+(e   -e  i)/v^T]           (2.76)
         -    ,    n-=-                    -6    -Q
                         2211
where  0  =  (x+3H/2)//4T ,   9   =  (x-3H/2)//4T and x = x - Ut.   The gradient
of  the depth-averaged  concentration  is  then
                                 27

-------
         23H
                                                                  (2.77)
 The mass flux, j  ,  is expressed as
                      H
                        H/2
                           PD
                               3c
                   rH/2
                       -H/2
                      pcu dy
                                                                  (2.78)
                  -H/2
 where u = u - U.   Assuming small density changes,  Eq.  (2.78)  is adequately
 approximated by
                                       rH/2
                                           cu dy
                                            (2.79)
                                       -H/2
Using  Eqs.  (2.73)  and (2.79),  the  total dispersion coefficient becomes
                                  rH/2
Et -
                                     cu dy)/ -r—
                                                                  (2.80)
                                  -H/2
The  convective  flux integral  in  Eq.  (2,80)  was  evaluated at x=Ut (x=0)
in Appendix B  (Eq.  B.6)  as
                        rH/2
       — ,      u x jj
      cu dy =  OQU  ir-
                      _!
                      H

                       -H/2

in which L = 3H,_5 = ATT? and $ is defined in Eq.  (B.7).
Eq. (2.77) with x = 0 and Eq. (2.81) in Eq.  (2.80),


                         VD1   3 F     T
                          L.  J. _ J ,£j •.  f"\
                                                                  (2.81)
                                                           Introducing
                                                                  (2.82)
where g(L/6) = $(L/6)/erf (/JrL/26).   The  right  hand side  of Eq.  (2.82)
represents the relative increase of  the  total  dispersion coefficient
over the longitudinal dispersion coefficient for  the medium.   The func-
tion g(L/<5) was evaluated numerically  and is shown in Figure  2.6.  For
a given value of the parameter E/D.. , L/6 is determined by the appropriate
model from Section 2.2 or 2.3  (Eqs.  2.48 or 2.65)  as discussed earlier_
in this section.  The results of this  computation for three values of
are shown in Figure 2.7.  The dashed portion of the curves represent an
estimated interpolation in the region  where neither of the two theories
for <5/L is strictly valid.  It is seen from Figure 2.7 that the total
                                 28

-------
g(L/6)
        o.i
                                                                 100.0
            Figure 2.6  g(L/6)  for  One  Dimensional Linear Flow
        100.or	r-
    I!
    Di
                                           i MI
              E/D  =  8^/3
	Immiscible Thy
	Transition zone
	 Coupled Thy
                      TT/6
           Figure 2.7   Total  Dispersion Coefficient, E  , as  a  Function
                              of Time for Linear Flow
                                29

-------
dispersion coefficient, E , decreases with increasing time and that
eventually the additional dispersion due to density differences becomes
negligible.

Rose and Passioura (1971) reported experiments in which the additional
dispersion associated with density differences was determined from break-
through curves.  These experiments were done with columns of circular
cross-section and therefore are not directly comparable with the present
results.  However, the increase in dispersion coefficient indicated from
Eq. (2.82) for these experimental conditions is found to be of the same
order of magnitude as that observed.
                                 30

-------
      3.  ANALYSIS OF DENSITY INDUCED MIXING IN RADIAL FLOW

In this section the methods of analysis developed for linear flow in
Section 2 are extended to the more practical case of radial flow from
a fully penetrating well in a horizontal confined aquifer.   The approach
generally parallels that for linear flow including first an analysis of
immiscible displacement, next a superposition of dispersive effects on
that result and finally a consideration of the combined effects of den-
sity differences and grain dispersion.  The analysis is generalized to
treat a recharge-storage-pumping sequence that would be typical of field
well operations.

                   3.1  Immiscible Displacement

As in Section 2.1, the problem of mixing two liquids of different density
in a homogeneous, confined aquifer may be simplified by neglecting grain
dispersion and molecular diffusion.  In such a case, the mixing process
is described by an immiscible displacement model.  Figure 3.1 illustrates
the displacement of a heavier fluid of density, p~, by a lighter fluid
of density, p7 , in a horizontal confined aquifer of constant thickness H.
If the two liquids have approximately the same kinematic viscosity, v,
the hydraulic conductivity, K, may be taken as a constant throughout the
flow.

The following analysis is based on the Dupuit approximation, which as-
sumes hydrostatic pressure distributions and thus essentially horizontal
flow.  The analysis is strictly correct only when the interface slope is
small and hence the vertical velocities are small.

The continuity equations for the two fluids are
where t, is the vertical distance to the interface, in liquid 1, measured
from the top of the aquifer, and n is the vertical distance to the in-
terface, in liquid 2, measured from the bottom of the aquifer.  The
seepage velocities of the two fluids are given by the Darcy equation,
                                 31

-------
                                                           Fluid 2
                                        / \\y/L\\v/,\v//, \\v/.
Figure 3.1  Definition Sketch of a  Radial Confined Aquifer
                           32

-------
Vertical velocities are neglected  due  to  the Dupuit  approximation.
The porosity of the aquifer is n,  and  the piezometric  heads  of the  two
liquids are .. and -.  The continuity expression  for  the entire flow
field is obtained by adding Eqs.  (3,1) and  integrating,


                       ?un + nu_ =  -r-*— = —  H                      (3.4)
                        1     2    2irnr   r

where A = Q/2TrnH.  The convection  of the  interface through the aquifer
is due to the net seepage velocity, A/r.

Along the interface, the fluid pressure must be  continuous (p1 = p?),
or
where
                           Ap = P2 -  PI                           (3.6)


For small density differences Eq.  (3.5) becomes


                           4>2 = 4>-,	(C-H)                       (3.7)


in which the relative density difference  is  identified by Ap/p~= Ap/p.

The problem involves the  determination of the  interface shape and
location.  Substituting Eqs.  (3.2) and (3.3) into Eq.  (3.4) gives

                      S^-i     9o         n
                        -L^!^_       U                       f n n \
Differentiating Eq.  (3.5) and substituting  into Eq.  (3.8) with H = n +
results in the expression

                        ^ J%
                         1 _   An  _  A_p  n_ 9n.                      n QN
                        3r      rK    p   H 8r                      ^   ;

Substituting Eq.  (3.2)  into Eq.  (3.la)  and  eliminating t, using £ = H -
gives
                                  33

-------
 Replacing the derivative of   with Eq. (3.9), the partial differential
 equation for r) is found to be
 where D = - ^- H.
           n p

 Recharge Cycle  Because of the nature of the flow geometry, it is helpful
 to express Eq. (3.11) in a volumetric coordinate system.  During recharge,
 with a positive flow rate Q  into the aquifer, the transformations are,


                             ₯ = (TrnH)r2                          (3.12a)
                            ^* = Qrt                              (3.12b)

 which represent, respectively, transformed space and time coordinates.
 Using these transformations, Eq.  (3.11) becomes


                    3n ,  9n _ 2D.-r) ,..   aw -§ai                    f-\ i -i\
                    o v*.   J V   A  11     n   o~V~
                      *

 Note that  this  substitution transforms the radially dependent coef-
 ficient of the  convective term into a constant.  The equation can be
 simplified further by transforming to a convective coordinate system:

                                ₯ = ₯ - *a                        (3.14)

 Eq.  (3.13)  now  becomes
This  recharge  equation can be reduced to an ordinary differential equa-
tion  using  a similarity transformation

                           f = n/H = F(z)

                                                                  (3.16)

                           z - Wer₯*


where
                                                                   (3.17)

-------
                            Ar = Qr/27mH                         (3.18)


and the subscript "r" refers to the recharge cycle.  The resulting
ordinary differential equation in z, the similarity variable, is
The parameter e  indicates the relative importance of the density dif-
ference, represented by D, as compared to convection, represented by A.
For increasing e , the relative rate of tilting of the interface should
increase.

For small e , a perturbation solution for Eq. (3.19) of the form


                     f = fQ + E/-L + erf2+'"                    (3'20)

is assumed.  For equations of 0(1) this results in a nonlinear equation
in f  and z
    o
This equation is solved for an initially vertical interface by assuming
a power series in z , to give
                                                                  (3.22)


The equations for higher orders, e , are of the general linear form
                  A
                  dz
                                                                  (3.23)
                          n = 1,2,3,'

For n = 1 the equation is
                  2  d         dfl           1     2
                  Z) —± - 2-z~^ - 2f1 = - y(l-3zZ)             (3.24)
                     dz
and the particular solution is
                          fl = I6(1"3z)                          (3<25)
                                   35

-------
 Similarly,  the equation for n  =  2  is
                               df
                          -  2z
              '     '    2        dz      2    3_
                     dz

 for which  the particular  solution  is


                                  1- ~ T^z2)                        (3.26)
 Therefore,  during recharge, the interface behaves as
                                     2
                                     £ 2
                                  -     (l -  i Z2)+- • •            (3.27)
The  solutions of Eq.  (3.23) will, in general, include two solutions of
the  linear homogeneous equation and a particular solution which depends
on F.  However, the homogeneous equation, which is closely related to
the  Legendre equation,  can be shown to have mathematical singularities
at z = ±1 and therefore  the solutions of the homogeneous equation must
be excluded on physical  grounds.  The result represented by Eq. (3.27)
is shown in Figure 3.2 for two values of e  .  It can be seen that,
even for the relatively  large value e  = .65 , the departure from the
linear interface shape of the zero order solution is not large.  Note
that the upper and lower limits of the interface (f = 1 and f = 0) ad-
vance as e increases, but that the arrival  time of f = 1/2 has been
delayed relative to the  average displacement.

The length of the interface in volumetric coordinates is expressed as

                   _.      —.            21      21


which can be determined  from Eq. (3.27) by  evaluating the end points of
the interface,
                                                                  (3-30)
Thus from Eqs.  (3.16),  (3.28),  (3.29) and  (3.30)
VJ1 -
                     = 2E V1 -     ^r  + •'•                    (3.31)
                                  36

-------
f =
          -1.0
+1.0
                              z = V/e ₯.
                                     r
f = n/H
          -1.0
0


z =
+1.0
            Figure 3.2  Comparison of Interface Shapes for

                Various Values of e  During a Recharge Operation
                             37

-------
 The interface location can be described by a coordinate -V-. , the volu-
 metric distance from the origin.  For the zero order solution (Eq. 3.27
 with e  =0) the interface position is given by
                                                                  (3.32)


 where
                               „
                           g = 2er

 and y is  the vertical distance from the bottom of the aquifer.  The
 radius to the interface is expressed as r. and the interfacial velocity,
 u,  is
                              3r.    A
                                                                  (3-33)
where
                       r± = y^/TmH = r*/l+g                      (3.34)
                                      /2A^F                       (3.35)

Storage  Cycle   During  storage  of liquid 1 in the aquifer, there is no
net flow (Q=0)  and  the convective term of Eq.  (3.11) becomes zero,  thus


                     111 -  n i  — m  - -^1-2- r  9nl
                     3t ~  D r  8rU1    H;H r  3^J

This equation may be transformed to  volumetric form using Eq. (3.12a)


                  |2. = 2D(2TrnH)  |^ [§(1 - £) V |j]                (3.36)


and may  be further  simplified  by introducing the coordinate system


                            £  =•₯•-₯•
                                       o

where ₯•  is the volume of  liquid of  density  p.. which was injected
during recharge.  Equation (3.36)  then transforms to


                  = 2D(2,nH)    - [d -  )(^ W-]            (3.37)
                                  38

-------
When  the length of the wedge is small compared to the distance from
the well, ₯  « •₯•  , Eq.  (3.37) is approximated by


                 |a - 2D<2»«*  ^[iU - j}> ffr-l               (3.38)


This  equation may be reduced to an ordinary differential equation using
the similarity transformation
                                    ,
                         z  =           —                        (3.39)
                                      t+Cn
                                     o   1

where C.. is an arbitrary constant added in order to fit the initial
interface shape at the start of the storage period.  The resulting or-
'dinary differential equation is
                          dz    dz        dz
                                                                  (3.40)
which has the  solution  (see Eqs. 3.21 and 3.22)
                            f = |(l+z')                           (3.41)


The result  of  Eq.  (3.41) represents a linear interface in volumetric
coordinates and  is implicitly limited to cases in which the volumetric
length of the  interface is small compared to the initial volume ₯• .
The explicit features of this limitation will be discussed in  sections
3.4 and  3.5, but it can be seen from Eqs. (3.39) and  (3.41) that  this
solution will  become invalid for large times.  It is  also implied that
the initial condition for the storage cycle should correspond  to  a
linear interface at the end of the recharge cycle and thus that e be
small.

It would be desirable to obtain solutions during the  storage  cycle
which are not  limited by the approximation in Eq.  (3.38), i.e., small
e   during  recharge.  Although the equation governing the storage cycle
(Eq. 3.37)  is  similar to that for the recharge cycle  (Eq. 3.15),  the
method of solution used for the recharge cycle does not apply because
a similarity form  of Eq.  (3.37) can not be obtained.

Withdrawal  Cycle  During withdrawal, fluid is pumped  from the  aquifer
at a rate 0 and the governing equation becomes, from Eq.  (3.11),
                                  39

-------
                                                                  »•«>
 where A  = Q /2imE.   This equation may be  transformed  to a volumetric
 convective coordinate system using Eq .  (3.12a)  and
                               o     w
                        ₯• =

 where
                t   =  time  at  the  start  of withdrawal
                o

               ₯-   =  net injected volume at  t
                o         J                 o
 Equation  (3.42)  then  reduces  to
                     = -  2e2 ?-  [f(l-f )($ + *) M]               (3.43)
with  e  =  /D/A   .  When  the length of  the interface is relatively small
compared to  the  distance from the origin to the  front, V;,. » V, Eq .  (3.43)
is approximated  by


                   - 51— = e2 1-  [f (1-f) %                 (3.44)
                   3(C2-₯* )      3^         3*

This  equation has the solution


                                                                 (3.45)
where C« is a constant to be determined from the initial conditions for
the withdrawal cycle.  The length of the interface in volumetric coordi-
nates will be given by
                                                                  (3.46)


and the location of the interface is expressed as

                             2e
                                j /     /
                                                                  (3.47)
                                  40

-------
The solution for the withdrawal cycle, as given in Eq. (3.45), represents
a linear interface in volumetric coordinates.   The range of applicability
of this solution will, as discussed in Sections 3.4 and 3.5, depend on
the initial condition from a storage or recharge cycle.  It can be seen
from Eq. (3.46) that as ₯A approaches zero (i.e., the entire stored vol-
ume is withdrawn) , the ratio AV/V^ becomes large and the approximation
implied in Eq. (3.44) becomes invalid.  A perturbation solution of the
exact equation (Eq. 3.43) was attempted using the method employed for
the recharge cycle.  A similarity form of Eq.  (3.43) can be found, but
in the withdrawal case, the initial conditions can not be satisfied by
the similarity form.

Matching Between Cycles  The solutions for the storage and withdrawal
cycles  (Eqs. 3.41 and 3.45) can be combined with the zero order solution
for recharge (Eq. 3.22) to represent general recharge-storage-pumping
sequences.  The general approach in matching between the cycles will be
outlined here and explicit applications to aquifer operations will be
developed in Section 3.5.  Because each of the three solutions repre-
sents a linear interface in volumetric coordinates, the matching between
the cycles can be expressed conveniently in terms of the inverse slope
of the interface in volumetric coordinates, 3, which is given by
                                                                  (3-48)
The constants C.. and C_ appearing in the storage and withdrawal solutions
can be expressed in terms of the initial inverse slope, 3  , at the start
of the cycle.  From Eq . (3.41) for the storage cycle
                       3 = 2/4TTnHDV t+CL/H
                                   o   1

and if the time t is measured from the start of the cycle,  the constant
C  can be determined in terms of 3Q.  The resulting expression for  the
inverse slope during storage is therefore

                                 16irnDV t   .
                      3 = g  (1 + - =— — )1/Z                     (3.49)
                                   32H
                                    o

Likewise for the withdrawal cycle the inverse slope is, using Eq.  (3.46),
where 3  is the initial inverse slope when V. = V  ,  the  initial  net
       q   .,                                *    o
injected volume.

The zero order solution for recharge may also be generalized to  include
                                  41

-------
 initial  conditions  corresponding to an interface of inverse slope g
 and  initial  injected volume V  .  Noting that the zero order solution
 during recharge  is  represented by Eq.  (3.15) with V « V , and defining
 VA=Q t + V   with t  measured from the  start of  this recharge cycle,
the solution can be written
                       f = f [1 +
(3.51)
The constant ,C_ ,is again evaluated in terms of 8  and •₯•  to yield
              3                                 o      o
                                                                 (3.52)
The inverse slope, B, determined from any of the three cases can provide
the initial value B  for any subsequent cycle.

Using the above notation, the solutions in all  three  cycles  (Eqs.  (3.22),
(3.41) and  (3.45)) can be represented in the general  form
                                    ••
                            f = 2+ BH                            (3'53)

where 3 is given by Eq.  (3.49), (3.50) or  (3.52).

                 3.2  Effects of Grain Dispersion

The immiscible displacement models of section 3.1 are an idealization
of the actual situation, since there will  always be some degree of in-
terfacial dispersion.  The objective of this section is to estimate the
importance of grain dispersion, under the  assumption that the dispersion
does not attenuate the rate of interfacial tilting.  This corresponds
to the case in x^hich the density difference, Ap , is large enough  to
completely govern the dynamics of mixing (i.e., the velocity field is
predicted by the immiscible model) , but in which the resulting model
does not adequately describe the variation of density near the inter-
face.  To obtain a better description of the density variation, grain
dispersion must be considered.  The analysis of the recharge cycle will
be developed in detail, and the results for the other cycles will be
obtained as a simple extension of this case.

Recharge Cycle  The convective-dispersion  equation, neglecting molecular
diffusion, is, during the recharge cycle,
                                  42

-------
                 +'r-7?+7
-------
 and 3 = 2e  ^A/H,  the  concentration is  expressed as  a  function  of -V^,
 •V-  and y.
   9c  .   i  3c      i    3c     1   r_  a  c
Ar/F
+ »2~
,2
r 8 c
1 2
3y
V'
                               -  28
                             (3.59)
                                            V'
 Since  the  coordinates, •₯•.  =  constant,  are parallel  with the  interface,
 it  is  reasonable  to  assume that
                            _§£
                            3y
= 0
The  convective  dispersion  equation  then  reduces  to
                   3c
       ,2,  9 c
                                                                  C3.60)
When the dispersed zone around  the  interface  is  small  compared with  the
distance, r^, from the origin,  the  two  convective  terms  on  the left  hand
side of Eq.  (3.59) will cancel  near the interface,  r = r..   The  second
term in brackets in the longitudinal dispersion  term will also be  neg-
ligible under these conditions  because  it  involves  a lower  order of  dif-
ferentiation than the first.  This  type of approximation can be  formally
justified, as done by Gelhar and Collins (1970), using perturbation
methods and can be verified from the result of the  present  analysis.  It
is also consistent to replace the coefficients (A/r) in  the dispersive
terms by A/r.,..  The coefficient in  this equation can be  eliminated by
the' introduction of a new variable,  £.
                                                                  (3.61)
The function, £, is evaluated from the known flow  field, based on  the
immiscible displacement model, Eqs.  (3.35) and  (3.31).
                                 2
                          al
                         [rri
                          3r
                                H
                             (3.62)
                                  44

-------
With this new variable, Eq.  (3.60)  transforms  to
                                                                 (3-63>
The initial condition corresponding to a vertical interface  at  the
origin (V=0) is
                     t = 0    c = c     ₯ > 0
                                   o


                              c = 0     ₯ < 0

and the boundary conditions are


                     t>0    c=c     •₯--><»
                                   o


                              c = 0     ₯ < -°°

The solution of Eq. (3.63), subject to these conditions, is


                        c_ , 1 + 1 erf !L                       (3.64)
                        co   2   2     /45

The dispersed zone width, in volumetric coordinates A, may be defined as
                                   -Tr   e 2 r^
                                   i- + -f-r-a2*               (3.65)
                                    K   H


and the approximate radial width of the dispersed zone, 6, may be defined
as
                                 fa., ir   e 2 r^
                          = 2**\hr- + -^ -- ^                <3-66)
                              * V 3r*   H2  5

As the width of the dispersed zone, 5, becomes large, dynamic effects
of grain dispersion are expected, and the dynamics of the mixing  are
not given by the immiscible displacement model.  The ratio, A/AV, where
AV is given by Eq.  (3.31) and A is given by Eq.  (3.65), is proposed as a
criterion to estimate the importance of grain dispersion in the radial
flow problem during recharge
                                   2   2
                      A a  IT       3e  r .
                                                                  (3.67)
                                  45

-------
 Equation (3.67)  indicates  that  the lateral  dispersion  becomes  important
 for
                           _  2   2
                           3c  r*  a
                          — -2-->l                          (3.68)


 The concentration distribution  described  by Eq.  (3.64) is  expected to
 be accurate when A«A₯.  However,  as A approaches  AV,  it is necessary
 to consider the  interaction  of  grain dispersion  and density differences
 in the mixing dynamics.

 Withdrawal Cycle  Under  the  assumptions outlined at the beginning of  this
 section, the convective  dispersion equation (Eq. 3.55) during  withdrawal
 for A = - A  is
            w
                        .,         A  .2       A  -2
               9c .     9c          w 8 c       w  8  c              ,
where  the  subscript  refers  to withdrawal.   Using the  same  method of
analysis described for  the  recharge  period, this equation  can be trans-
formed to  the volumetric  coordinate  system, V  ,  measured from the inter
face.
          -__      _   _      _       .
          Vr  9V£  Vri 3V£ " Vr    3V "     3r
                  2  0       3V02
                      w        X.
For withdrawal, VA is given by  (see  Section  3.1)
while V. and V. are defined as before,  in  terms  of  VA  (Eqs.  (3.57)  and
(3.58)).  Using the same approximations used  to  obtain  Eq.  (3.60)  from
Eq.  (3.59), Eq. (3.70) reduces to

                                   2

                             If-'luS-                            (3'72)
                             at,    d₯  /
where
                                                                  (3-73)
                                  46

-------
and C  is an arbitrary constant to be determined by matching with the
concentration field at t=t  .  The appropriate solution of Eq. (3.72) is
of the form
where £ can be evaluated from the known value of 3 found for the with-
drawal period using immiscible displacement theory (Eq. (3.46) with
A₯=gH or Eq. (3.50)).  Substituting this value of g into Eq . (3.73) yields


      r      "4    r TrnH      3/2   £W2    „ 1/2 ,_    ** M  . „  /o -7^
        = -    t — ot  ₯A    + — a  V^ '  (C  - — )] + C  (3.75)
If the form of C_ implied in Eq. (3.50) is used, and C^ is evaluated in
terms of the initial volumetric width of the dispersed zone, A  = ATT? ,
when t=t  and VA=₯  , the variable £ is given by
             2
       (irnH)
                + 1 (v 5/2_₯    )}] +T-                         (3.76)
                  j   O     ^         A-TT

As in recharge, the ratio A to AV represents the relative importance of
grain dispersion.  The volumetric width of the dispersed zone A = /47r^
is given by Eqs. (3.75) or (3.76), and AV is obtained from Eqs. (3.46)
or (3.50).  Thus the ratio A/AV can be expressed as
                              =  r _   _  1/2
                                 l                      J
In this solution it has been assumed that the mixing dynamics are given
by the immiscible displacement theory.  Therefore, A/AV must be  small
for the solution to be valid.

Storage Cycle  When the net convective motion in the aquifer is  zero,
the additional mixing due to grain dispersion is negligible.  Physically,
this is implied from the fact that the volumetric length of the  inter-
face, AV, must be small compared with the mean volumetric distance to
the interface, V , for the solution during storage (Eq. 3.41) to be
valid.  Since the width of the dispersed zone generally is proportional
to the square root of the distance traveled by the interface, and the
length of the interface is , under these implied limitations , small com-
pared with the distance from the origin, the additional dispersion during
                                  47

-------
 storage  will generally  be  small.   Thus  it  is  consistent  to  assume  that
 the volumetric thickness of  the  dispersed  zone  around  the interface  does
 not change  during  storage.   For  extremely  large storage  periods  (e.g.
 several  years) it  may,  however,  be necessary  to include  the broadening
 of the dispersed zone as a result  of  molecular  diffusion.

 3.3  Interaction Between Grain Dispersion  and Density  Induced Mixing

 In Section  3.2, the  additional mixing due  to  grain dispersion was
 analyzed, assuming that the  dynamics  of the mixing is  described  by the
 immiscible  displacement model  (Section  3.1).  As  the mixing due  to grain
 dispersion  becomes large,  the immiscible displacement  model no longer
 describes the  fluid  motion.  To  determine  the influence  of  the grain
 dispersion  on  the  motion,  the mass and  momentum equations must be  solved
 simultaneously.  When the  mass and momentum equations  are coupled  for
 the entire  flow field,  a practically  intractable  system  of  nonlinear
 partial  differential equations is  obtained.   However,  as in the  linear
 case, a  useful approximation can be developed by  coupling these  equations
 at the mean interfacial position,  ₯=₯.,

 The governing  dynamic equations  (Darcy  equations) are
                            v  =                                  (3.78)
                             r   ny  3r
in which v  and v  refer to radial and vertical seepage velocities re-
          r      y
spectively, p is the pressure, k is the permeability and p is  the dynamic
viscosity.  The governing mass conservation equation is


           3c ,    3c  ,    3c   1 3 /rv   3c,.  , 9  ,_.  8cN          /0  on^
           K + Vr 3? + Vy 3? = 7 ^(Dlr 3?> + ^2 3^          (3'80)

which is simplified by the same assumptions appearing in Section 2.2.


                 3c .     3c      IA| 32c  ,    IA! 32C                 ,,  _..
                 3F + Ui 3? = al r -2 + a2 r 7T                 (3'81)
                                   9r         3y

The interface is assumed to be linear in the volumetric coordinate system
(Eq. 3.58) and 3 is an unknown time-dependent function which represents
the inverse volumetric interfacial slope.

The dynamic equations are manipulated in the same fashion as their linear
counterparts (Section 2.3).  The Dupuit approximation implies  a hydro-
static pressure distribution,
                                  48

-------
                             9y
                                = -pg
and thus from Eq.  (3.78) it follows that
                         r = kg_ _3p_ = K 1 _3p_
                       3y    ny 3r   n p 3r
                                                     (3.82)
where K is hydraulic conductivity.

Equation (3.82) can be applied at the interface
                          3v
                          3u.
                            i
                              r=r.
                                   =  3y
when the thickness of the dispersed zone, A, is much larger  than  the
horizontal projected length of the wedge, AV.  The analogous approximation
for the linear case is examined in Appendix A.  The result of this  ap-
proximation is
3u.
  i
                  32r.
                     i
            3y  ~ 3y3t
3V.
                  3yvr.  3t
       K _! ^p_
       n p 3r
                                                                  (3.83)
                                                  r=r.
Using the linearity between density and concentration  changes  (Eq.  2.21),this
becomes
                      3u.   _,  .  3c/c
                      	i _ K A_p_,    QS
                      3y    n p   8r   r=r.
                                                      (3.84)
Recharge Cycle  The solution to Eq.  (3.81) during recharge  is  obtained
by the approximate method used in Section 3.2, and is given by Eq.  (3.64)
in which 5 is given by Eq.  (3.61).   However, in  this case 3 is an unknown
function of time.  Differentiating c/c  with respect to  r and  evaluating
the result at the interface gives
                      3(c/cQ)
                         3r
                         2TtnH r.
                               i
                              r=r.
This result is used to evaluate the right hand  side  of  Eq.  (3.84),  and
the left hand side of Eq.  (3.84) is, after multiplying  by r.   ,  written
as
                  3r-
            r. 3t3y
                       3r2
             r. 3t"2r. 3t
              i      i
                                   49

-------
where 3 = 9₯./3y has been introduced.  It is consistent with the approxi-
mation of the convective dispersion equation (see Section 3.2, Eq. 3.60)
to  approximate r.  by rA in the last expression.  Equation (3.84) then
reduces to

                                            2

                            JIT (  V /o' ~   IZHT                    (3.85)
 Equation  (3.85)  must  be solved simultaneously with Eq. (3.61) in differ-
 ential  form,
                                                                  (3-86)
 In  order  to  simplify the  simultaneous solution of these two equations ,
 a new variable,  9  =  3/₯,j.V2,  is  introduced.   Equations (3.85) and (3.86)
 then become
                                *-\
                                 r     -                         (3.87)
                           d*A
                  =  [2a  (7rnH)1/2  + - 2— r-   02]₯                 (3.88)
                *                   2(7rnH)

which can be  solved  simultaneously to give


              51/2 =  [2ai(TrnH)1/2  6 + -   "2     G3]               (3.89)
                        1              6(TrnH)17/     2e^


This result is -substituted  into Eq . (3.87) in order to find the inverse
slope, 6 = VA  6, for  the  initial condition 3=0 at t=0,

                                         A       4« 5/2
               /  TTxl/2  .2  ,     a2     3     4 Er V*              ., ...
            a , OmH)     0  + - T7T w~ = T - 9 -           (3.90)
The volumetric width of  the  dispersed zone,  A = /4ir£, can be found from
Eq. (3.89)

                           2a    a0r.   „
                     A.  [-1 + ^-lB2]^                     (3.91)
                           •>-    AII       ^>
                                 ov .       e
                                   *        r

and the ratio of A to AV  is
                                  50

-------
                  AV - 3H
Equation (3.90) can be approximated for both large and small interfacial
slopes, noting that a., is usually an order of magnitude greater than a .
For larger slopes (3 small), lateral dispersion is negligible and Eq. fs.90)
reduces to
                                                                 (3.93)
                                         )-,/.


The volumetric width of the mixed zone is

                                    /"~t   ,1/2

                                    ^-3	—                  (3.94)


and the ratio A to AV is,


                        A
                        A₯     2      1/2
                             er     V*
For small interfacial slopes (large 0), longitudinal dispersion is negli
gible and Eq. (3.90) simplifies to
                                                                  (3'96)
                                     2    7TH

In this case the ratio of A to AV is given by
                    AV ~ 3U2,  U.1/2
                           H (irnH)
                                      1/2   1/2
                                     J    V                       (3'97)
Storage Cycle  As explained in Section 3.2, the additional mixing during
storage due to grain dispersion is negligible.  Suppose that at the start
of the storage period the volumetric length of the interface, AV, is
much smaller than the width of the dispersed zone, A.  Then the rate of
tilting of the wedge will be governed by a coupled analysis which accounts
for the interaction of the density difference and the given dispersed
zone.

In this case it is assumed that the volumetric width of the dispersed
zone remains constant and equal to its initial value, A ,  throughout the
storage cycle.  From the definition of A,
                                  51

-------
                    A -  = -t--\  =        .(^\
                     o     3₯vc  '    27rnHr 3rvc '
                              o              o

 and from this expression the right hand side of  Eq.  (3.84)  is  evaluated
 to yield


                         8Ui  _ K Ap_ 2rmH
                         3y  ~ n p   AQ  ri                        (3'98)


 The velocity gradient is represented by

                     3u.    _   3r.      ,    ,,   _
                     3y     3t3y      27rnH 3tV.


where  B =  3V±/3y  and ₯£ =  VQ + 3(y-H/2).   Equation  (3.98)  then becomes


                                                                  (3.99)
                        rrnHr.  3t  r.      A
                             ix      o
and by  introducing the approximation  analogous  to  that preceding  Eq.
(3.85), i.e., r. £ r  or ₯. = ₯  ,this reduces to

                            d3
                             dt ~   A
                                    o

If 3  is the inverse volumetric slope  of  the interface  at  the  start  of
the storage cycle , t=t  , then

                               4imW
                      6 = S0 +   A  °(t-t  )                       (3.100)
which remains valid as long as 0H = A₯- « ₯  .

Withdrawal Cycle  The coupled solution for the withdrawal  period  utilizes
the convective dispersion equation found in  Section  3.2  (Eq.  3.69)  and
its solution  (Eq. 3.74).  The analysis parallels  that  described for the
recharge period, and yields two equations analogous  to Eqs.  (3.85)  and
(3.86) for the recharge period.  The governing equations differ only by
signs, and the details of the development are omitted.  The  result  for
the equation describing the flow dynamics  (the equivalent  of Eq.  (3.85))
is
                                          2
                       /9
                       '
                                  52

-------
and the equation expressing dispersion  (equivalent to Eq .  (3.86)) becomes

                   j>-    a-iQ_      i     A  /r,
                 -irrfe-SV

In these equations VA is given by
                        V* = V  -  Q  (t-t  )
                         *    o    xw   o

The simultaneous solution to these equations is  found as in the recharge
case and
                                                 c.              0.103)
                                                  6

where C, is a constant used to  match with  the previous cycle.  The in-
verse slope, 3 , is found by substituting Eq.  (3.103) into Eq .  (3.101)
and integrating

                                      2                 4
        1/2 32       a2       84    2ew    3          4ew     3/2
 a (irnH)    — H        r-r     - + — 7—    ~- C, = -- =- V     + C
                                                                 (3.104)
The constant, C7, is dependent  on the  initial  slope,  3  •
When considering multiple cycles,  it  may be  convenient to evaluate the
constants C, and C7 in terms of  the initial  dispersed zone  thickness,
A , and the initial inverse slope, 3  .  Using V  = irnHr 2}  the  dispersed
zone thickness becomes
                                            3
                                                    3
    A - (*.«    -    2V- -   ,  +          -       „ *.A<>      (3.105)
                    w      *    o        ₯A      V

and the general expression for the inverse interfacial slope, 3, is
                  o     24(irnH)      V.    V
                                     K     o
                                           ir !/2'
                   O    "   V      ₯., •     V
                             0*0
                                  53

-------
                        3TTH
                              IV*3/2 - V]                     (3.106)
 The  concentration  field is obtained by using Eq.  (3.105) in the general
 solution  as  given  by Eq.  (3.74).  The interpretation and application of
 these  results  are  discussed in  the following sections.

              3.4   Discussion  of the Theoretical Results

 In the previous portions  of this  Section,  theoretical models were pre-
 sented to describe the mixing of  two fluids in a  radial flow system, due
 to density difference and grain dispersion.  The  first model presented
 in Section 3.1 was based  on an  immiscible  displacement process in which
 the  only  mixing mechanism was the density  difference.  A   nonlinear
 partial differential equation (Eq. 3,11) describing the interfacial shape
 and  position for either recharge, storage  or withdrawal was derived.  An
 equivalent equation was used  by Perrine and Gay  (1964) in  their investiga-
 tion of secondary  recovery of oil.  However, they were primarily inter-
 ested  in  the effects of large viscosity differences on the interface and
 did  little analysis of those  situations in which  the density difference
 is the predominant effect.

 A  solution of  Eq.  (3,11)  was  first developed for  the recharge process.
 It was found that  the interface is described by a straight line in volu-
 metric coordinates when the volumetric length of  the interface, A₯=3H,
 is small  in  comparison to the volumetric distance of the convected  front
 from the  origin, V^.  This is the case when e  (Eq. 3.17)  is small  com-
 pared  to  unity.  For an initially vertical interface at the well this
 ratio,  AV/₯^,  remains constant  throughout  recharge  (Eq. 3.31) and the
 interface is adequately described by the straight line interface at all
 times  during the recharge period if e  is  sufficiently small.  When there
 is some amount of  tilting at  the start of  recharge  (Eq. 3.52), such as
 would  occur  after  a recharge, storage, withdrawal sequence, the ratio
 AV/V^   decreases as VA increases,  Thus the description of the interface
 by a straight  line will improve with time.

 In order  to  improve upon  the  linear interfacial solution found for  re-
 charge, perturbation techniques were used  to investigate the curvature
 of the  interface.   The results  of this perturbation analysis, as shown
 in Figure 3.2, indicate that  there is only a small departure from the
 linear  interface,  even for relatively large values of E .  It is seen
 that for  large e   the lower end of the interface  has advanced into  the
 aquifer farther than it would for a linear interface.  In  the subsequent
 analysis  of  the recharge  process, it will  be assumed that  the interface
 is linear in the volumetric coordinate system.

 One  of  the primary  characteristics of the  immiscible model is the volu-
metric  length of the interface.   For the linear interface  (lowest order
 form of Eq.  (3.31)),
                          AV = SH = 2e V,                         (3.107)
                                      r *
                                  54

-------
                                                                2
is shown in Figure 3.3 as a dimensionless plot of gD/Q H=AVD/Q H  versus
2e T.  The parameter x is a dimensionless time,

                             = K Ap_ _t = Dt
                               n p  H   H2
It is also useful to consider the interface as it appears to an observer
in a single observation well at a radius r.  This is often the type of
information which may be gathered in the field.  The observer_would
record depth averaged relative concentration in the aquifer, c/c , as a
function of time
                           c_ = I
                           c    H
                                   H  c
                                      7- dy                     (3.108)
                                      c
                                       o
Because the fluids a_re considered immiscible for the purpose of this
model, the value of c/c  is identical to the dimensionless height of
the interface, n/H_-_  A simple method of representing these results is
the time slope at c/c =n/H=0.5 (i.e., V=VA)  which is, from Eq. (3.51),

                    9n/H
                                 = (At)"1 = -Q /3H
                         n/H=0.5              r
where At is the inverse of the rate of change of relative concentration,
It is convenient to present this result in terms of the dimensionless
variable
as obtained from Eq. (3.107).  This result is shown graphically in Fig-
ure 3.4.

Equation (3.11) was also solved for storage and withdrawal processes.  Dur-
ing storage the interface is described by a straight line in volumetric
coordinates when AV is small compared to the initial volume, V  (Eq. 3.41).
The requirement of small AV/V  must be met for the initial condition at
the start of storage.  However, unlike the recharge process examined
earlier, the ratio AV/V  is a function of time, V  is a constant and A^=0H
is increasing with time (Eq. 3.49).  If storage is continued long enough,
AV will eventually become large compared to V  and the approximation of a
linear interface may become inappropriate.  A reasonable limit on the ap-
plication of the linear interface solution (Eq. 3.41) is the condition
AV/V <1.  From the results for recharge, it would be expected that inter-
face curvature will become important for AV/V  near unity.  Analysis of
these curvature effects is suggested as a topic for future study.

The withdrawal phase also yielded a linear volumetric interfacial solution
when the volumetric interfacial length, AV (Eq. 3.50), is small compared to
the volumetric distance of the convected front from the origin, VA.  A
necessary, but not sufficient, condition for this solution is that e  be
small compared to unity (e «1).  The initial condition inherited from the
previous recharge or storage process must also be considered.  Thus  P H/V
                                 55

-------
O"
o"
CO.
     10
     10
     10
              H/a  =  75,   a/ct
Immiscible Thy, Eq. 3.107


Coupled Thy, Eq. 3.90, 6/L>2
                    Coupled Thy, Eq. 3.90,

                      6/L<2
                                               A  Upper Limit of
                                                  Coupled Thy, 6/L=2
                                               B  Lower Limit of Im
                                                  Thy, 6/L=.5
                                               C  Upper Limit of Im
                                                  Thy, 6/L=.5
                                               D  Lower Limit of
                                                  Coupled Thy,  S/L=2
       10
               Figure 3.3  Inverse Interface Slope, 3, as a
                           Function of Time during Recharge
                                   56

-------
                                       Immiscible Theory,
                                          Eq. 3.109
                                   	 Grain Dispersion
                                       Effects, Eq.  3.113,  -
                                       6/L > 1
                                        Coupled Theory,
                                        Eq.  3.114,  6/L >  2
                                        Coupled  Theory,
                                        Eq.  3.114,  6/L  <  2
10
              10
10
10
10'
                                                     10
                           2e T
                             r
Figure 3.4  Inverse Rate of Change of Relative Concentration,
            At, as a Function of Time during  Recharge
                       57

-------
 must be small  compared  to  unity at  the  start of withdrawal.  Finally, it
 is important  to  examine the  temporal  conditions under which the linear
 interfacial solution  is valid.  The ratio A₯/₯^ can be expressed as
 (Eq. 3.50)
 As  VA decreases  to  zero  during withdrawal , the ratio A₯/₯^ increases and
 the linear  interface  solution eventually becomes invalid.  As in the sto-
 rage case,  the limit  for this solution may be taken as A₯/₯^<1.  Using
 this condition in Eq.  (3.110) there is the implied lower limit on the ap-
 plication of  the linear  interface result during withdrawal,

                              (6 H/V )2 + 4e
                                        2
                                       w
When VA approaches this limit, it would be expected that interface curva-
ture effects may become significant.  If a linear interface is assumed,
the lower end of the interface reaches the well when A₯/2=₯A.  For this
condition A₯/₯A>1 and thus significant errors may be introduced when the
linear interface theory is applied at the well.  The possible effects of
interface curvature as the interface approaches the well should be ana-
lyzed in future studies.

The effects of grain dispersion were analyzed in Section 3.2 by assuming
the grain dispersion to have no influence upon the flow dynamics.  The dy-
namics are given by the immiscible model (Section 3.1) where the inter-
face is now idealized as the line of 50% relative concentration.  The
volumetric interfacial length, A₯, was assumed to be small compared to the
distance from the origin, ₯A or ₯ , and a linear volumetric interface was
obtained.  An estimate for the volumetric width of the dispersed zone, A,
was found for recharge (Eq. 3.65) and withdrawal (Eq. 3.76).  The ratio of
the dispersed zone width to the volumetric interfacial length, A/8H, was
proposed as the criterion to estimate the importance of grain dispersion
during radial flow.  For recharge this is given by Eq. (3.67) and for with-
drawal by Eq. (3.77).

As described in Section 3.2, this model is valid when A/3H is small com-
pared to unity.  Figure 3.5 shows the ratio, A/gH, versus dimensionless
time, T, for several values of e  during a recharge operation with an in-
itially vertical interface.  This graph was constructed by assuming a
typical ratio of dispersivities (a1/a«=10) and by taking the ratio of
aquifer height, H,  to longitudinal dispersivity, ct, , as a constant equal
to 75, a typical field value.  A/$H is large for small time indicating
that interaction between density difference and longitudinal grain disper-
sion may be important near the well.  A/3H decreases at first, but it
                                 58

-------
Ui
VD
                  10.0
                   1.0
            A/BH
                   0.1 i-
                  0.01
                      0.01
0.1
1.0
10.0
100.0
1000.0
                                 Tlgure 3.5  A/3H as a  Hinction  of  Time During Recharge for the Case of No
                                             Dynamic Effects  of Grain  Dispersion

-------
 eventually increases as lateral dispersion becomes important (Eq. 3.67).
 It can be expected that even for large values of e ,  mixing due to grain
 dispersion will eventually become as important as density induced mixing.
 The exact limits of applicability of the immiscible model cannot be spe-
 cified, but as was demonstrated in Section 2.4 reasonable estimates may be
 obtained by assuming the limiting values of A/BH at a fixed number less
 than one.  A/8H=l/2 will be chosen here.  For a given value of e , the in-
 dicated range of dimensionless time, T, for which the immiscible analysis
 is valid is found from Figure 3.5 or Eq. (3.67).  For example, when e =0.1
 the minimum 2e T is shown in Figure 3.5 as 0.35 and the maximum is 25.  Be-
 tween these two values the immiscible theory can be expected to apply.

 The dimensionless length of time,AT, as defined in Eq. (3.109), is found
 by differentiating the depth averaged concentration (Eq. 3.108) with re-
 spect to time.  The depth averaged concentration is obtained by integrat-
 ing Eq.  (3.64),

       _                                         2  _  2
       p    1     v/ZT                        1    ^9   ^1
       ^*    -*-r-it"~3r      c        c    i   -^~ /   ^~    J- \ i i       /1 -i i i \
       — = -^[H	;rrr 1  y ~erf U9~u ..erf u  H	(e   -e   ) ) J       (3.111)
        O                                  /IT

 where
                         n
                           =  [V+(-l)nBH/2]//4T
The  derivative  of  (Eq.  3.111)  with respect to time is evaluated at ₯j.=V
(V=0)-
               3c/c
                  o
                3t
_   = (At)-A = --i_lerf(^fi)           (3.112)
v=o              3H ot       2 A
For  the  recharge  case  (9V.fc/8t=Q  ) with an initially vertical interface,  BH
is given by Eq.  (3.107)  and  A/$S by  Eq.  (3.67).   Thus AT  becomes

                                      J~       \1^
         AT = -^--|  =  -  2e T[erf	r^	r-   ]~X      (3.113)
              n p   H        r           ^ ct2  (2e^T)
where
                               _
                              a   3?r
This result is sho;ra graphically  in Figure  3.4.   Note that  for some mid
range values of time there is effectively no  increase in AT due to the
addition of grain dispersion over the value predicted by the immiscible
model.

During the storage cycle, the additional mixing  due  to grain dispersion
was neglected.  This assumption is consistent with the conditions neces
sary to derive the immiscible model with a  linear volumetric interface.
                                  60

-------
The model requires that the dispersed zone width,  A,  be small compared to
3H at the beginning of any process to which it is  applied.   During storage
3H must also be small compared to V  for the model to apply.   The dispersed
zone width at the start of storage,  A, is proportional to  the distance
traveled by the interface, V .  During storage the interface will tilt about
its midpoint, with the result that the interfacial endpoints would travel
the most and undergo the greatest amount of di^v-r=ion.  But this travel
must still be small compared to V  in order fo<      A₯/V  ratio to remain
small.  Thus it is consistent to neglect this additional dispersion.

In the third model, presented in Section 3.3, the  combined effects of den-
sity difference and grain dispersion were analyzed by an approximate coup-
ling of the mass and momentum equations based on large A/3H.   It was as-
sumed that the interface remained linear for this  development.  The inter-
facial slope for a recharge process with an initially vertical interface
is given by Eq. (3.90) and can be represented by a family of curves relat-
ing the volumetric inverse slope,  3, and V^ for various values of e ,  a, /a-
and H/ex. .  Equation (3.90) is shown in Figure 3.3  as a dimensionless plot
for several values of  e  with a./a=10, and the ratio H/o, = 75.  As noted
in Section 3.3, the coupled analysis is valid when A/3H»I.  The ratio
A/&H (Eq.(3.92) with Eq.  (3.90) substituted for 3) is shown in Figure 3.6
for a number of values of e ,  The range of applicability of the coupled
model (Eq. 3.90) can be analyzed by setting an approximate limit on the
value of A/gH.  For this discussion the condition A/gH>2 is assumed for
the coupled model to apply.  For example, when e =0.1 the coupled model is
applicable up to 2e T=0.018 during which longitudinal dispersion is impor-
tant.  Later when 2e T exceeds 1000, the coupled analysis again becomes
applicable with lateral dispersion as the predominant mixing mechanism.
Figure 3.3 indicates these regions of validity and if closely examined,
it can be seen that the regions of validity for the immiscible model  (Eq.
3.107) and the coupled model (Eq. 3.90) are mutually exclusive for a given
value of e .  In general, there will be intermediate time ranges during
which neither model is applicable.  In the example,  e ™.l,  these inter-
mediate ranges are 0.018<2e r<0.35 and 25<2e  i<1000.  The actual behavior
of the wedge  in these  ranges is not known,but a smooth  transition between
the models is expected.

The  curves representing Eq.  (3.90) in Figure  3.3  are shown as  dashed  lines
for  A/8H<2.   It can be  seen that,  even for large  values of e  ,  longitu-
dinal dispersion will  influence  the  tilting  near  the well, and if recharge
continues long enough,  lateral  dispersion will influence the  tilting.   As
e  decreases, interfacial tilting  becomes less and as  lateral  dispersion
becomes important, the  rate of  tilting decreases.  The  point  of  intersec-
tion of the  immiscible  theory and  the  coupled theory can be  found by  solv-
ing  equations  (3.90) and  (3.107)  simultaneously.  That  expression  can in
turn be solved for the  value of  e  for which the  coupled theory is  tangen-
tial to the  immiscible  theory at  the point of intersection,  e  =(2T,a1/H)
(3a2/2a )1'2=3.26xlO~2  for Figure  3.3.   For  values of  e  greater than
this, there will be an  intersection  of  the  two theories; for  values of e
less than this, the immiscible  theory  is never valid during  recharge.
                                  61

-------
  100.Or	1	r
   10.0
A/BH
    1.0 -
    0.1
       10
10.0
100.0
                Figure  3.6   A/BH as  a Function of Time During Recharge for the Case of
                             Interaction Between Grain Dispersion and Density Induced Mixing

-------
The ratio A/3H indicates the relative importance of the two mixing mech-
anisms, grain dispersion and density difference.  When A/8H is small (Sec-
tion 3.2), grain dispersion has no effect on the interface and the dynamics
are given by the immiscible model.  When A/gH is large, the combined ef-
fects of the two mechanisms are important (see Appendix A for analogous
development in plane flow) and the coupled analysis governs the mixing.
The coupled solution does not reduce to the immiscible solution when
a  ,a~-*Q because the regions of validity of the two solutions do not overlap.

The length of time, At, defined by the inverse time derivative of c/c
(Eqs. (3.108) and  (3.111))at V=0 is given by Eq. (3.112) for both the°coup-
led and immiscible dynamic theories.  For recharge (3V./9t=Q ) in which
interaction between grain dispersion and density difference is significant
3H is given by Eq. (3.90) and A is given by Eq. (3.91).  Then the dimen-
sionless form of At can be written as
                                      3/rT k
           AT = KAp_At = _|^. [erf	:	ri       (3ell4)
                n p  H      Q H             7          J
                                          2k^ (3D/Q Hr
                                     4(1+ —£	^	)
                                           k j (2£rT)

where

                                2TT  ^ 1 1 / /,   ^ / Q
                           _    n   .I... -L/ 4   o/ o
                        ^7 ~~ ^rr" 7v  7»^"T


This expression is presented graphically in Figure 3.4 for several values
of e .  For values of A/3H less than 2, the curves are shown by dashed
lines.  Note that for large values of e  there is effectively no increase
in AT, for some midrange values of T, over that predicted by the immis-
cible theory, Eq. (3.109).

The combined effects of density difference and grain dispersion were also
investigated for the storage and withdrawal cycles.  During storage the
change in dispersed zone thickness, A, was assumed to be negligible and
the rate of interfacial tilting was described by Eq. (3.100).  For with-
drawal, expressions were presented which described the inverse slope of
the interface (Eq. 3.106) and the volumetric width of the dispersed zone
(Eq. 3.105).  Some general features of these results may be noted.  If
lateral dispersion predominates during withdrawal (Eq. 3.106 with a =0),
then it can be shown that A/3H always increases with time and the coupled
analysis can definitely become applicable.  The same trend is noted for
the immiscible theory (Eq. 3.77), thus indicating that a transition from
immiscible dynamics to coupled dynamics can occur.  If longitudinal dis-
persion is of primary importance during withdrawal, A/3H may increase or
decrease depending on magnitude of V  ,3  and  A  (Eq. 3.106 with a =0).
If V »3 H, the indication is that A?3H°increases.
    o   o

The analysis of the interaction between density difference and grain

                                   63

-------
 dispersion,  as presented in Section  3.3,  has  certain implicit limitations.
 One of these is that  A/3H be large compared to unity.  This restriction is
 explained in Appendix A for the  case of linear two-dimensional flow and
 the same restriction  is implied  for  radial flow.  As pointed out in Appen-
 dix B, there is implied,  from these  approximate solutions, a relaxation of
 the no-flux  boundary  conditions  at the upper  and  lower boundaries of the
 aquifer when A/3H is  large.   The possible effects of these boundary condi-
 tions should be investigated in  future studies.

                 3.5   Applications of the  Radial Theory

 One of the objectives of  this research is to  make it possible to forecast
 the amounts  and quality of fresh water which  can be removed after storage
 in  a saline  aquifer under various operational schemes.  Procedures by which
 the theoretical results in the previous sections can be combined to simu-
 late complex operational  conditions  will  be outlined.  These procedures
 should also  be applicable to deep-well waste  disposal operations in which
 there is a density difference between the injected wastes and the native
 water.

 The immiscible theory (Section 3.1)  can be used to construct simple, ap-
 proximate results  for many types of  applications.  To illustrate the use
 of  this  theory,  a  one sequence operation  of recharge, storage, and with-
 drawal cycles  will be analyzed.  Additional sequences may be analyzed by
 matching the solutions between cycles described in Section 3.1.

 Consider a recharge operation with a flowrate, Q  , for time, t , storage
 for time,  t  ,  and  withdrawal at  a rate Q  .  For these conditions, we want
 to  find  the  interface position as a  function  of time and the amount of
 water recovered.   Utilizing  the  set  of solutions described by a linear
 volumetric interface,  the  volumetric length of the interface during re-
 charge, 6H, is  described by Eq. (3.52) with g  =₯ =0.  At the end of re-
 charge when  a  volume  V =Q  t   has been injected, the volumetric length of
 interface is 3H=2e V   which  provides the  initial value of 3 for the stor-
 age cycle.   During storage the net flow is zero and the interface is des-
 cribed by Eq.  (3.49),  with  3  =2e V  /H and after storage for a period t.
                              o   r o                                  s

                         (3H/VQ)2 = 4£
Using this value of 3 as the initial value, 3  , in Eq. (3.50) for with-
drawal, we find

               (3H/₯Q)2  = 4e2(l+2ts/tr)  + 4E2[l-(₯^/₯o)2]          (3.115)


which can be used to evaluate the volume of water recovered.

For this analysis it will be assumed that withdrawal ends when the depth
averaged relative concentration, c/c , reaches some specific level c/c =b,
0
-------
based on geometric considerations of the interface as it reaches the well,
From Figure 3.7 it can be seen that when c/c =b at the well,


                            VA = gH(l-2b)/2                      (3.116)

The total volume of water recovered at that time is
                                                                 (3.117)

Note that this recovered volume includes some contaminated water with
"c7c =b.  The recovery ratio, R=₯_/₯ ,  is, from Eqs.  (3.116) and (3.117),
   o                            K  o

                        R = 1 - (l-2b)(BH/2V )                   (3.118)
                                            o

The value of (3H when withdrawal ends is found by solving Eqs. (3.115) and
(3.116) simultaneously and with that result the recovery ratio becomes


                                  e 2+e 2(l+2t /t )    .
                 R = 1 - (l-2b) ( —	 S /  )1/Z         (3.119)
                                     l+(l-2b) e
                                               w

This simple result can form the basis for approximate comparisons of al-
ternative recharge, storage and withdrawal schemes.   As  indicated in Sec-
tion 3.4, some error may be introduced because of interface curvature
near the well.  However, the comparisons with experiments in Section 5
indicate that Eq. (3.119) provides reasonable estimates  of R which are
somewhat less than those observed.

Some characteristics of the result in Eq. (3.119) can be seen by consid-
ering t =0 and b=0, in which case the recovery ratio can be represented
graphically as shown in Figure 3.8.  Note that lines of  constant R are
roughly circular in the e , e  plane, hence, a given value of recovery
ratio can be attained approximately by specifying a constant distance
from the origin in this plane.  The graph also illustrates the importance
of the parameters e  and e  in determining the recovery  ratio.  Favor-
able recovery ratios are generally obtained with low values of e  and e  .

Some other features can be analyzed using the immiscible solutions.  For
multiple sequences of recharge-storage-withdrawal, interface movements
and recovery ratios can be found by repeated application of Eqs. (3.49),
(3.50) and (3.52) with appropriate matching between cycles.  As a spe-
cific example of multiple recharge-withdrawal sequences, consider a well
operation in which e  and e  are unchanged (i.e., the same flox^ rates)
for several sequences and t =0.  In each recharge cycle  the same volume
of water is injected and during withdrawal, water is pumped until the
wedge first reaches the well (b=0).  Under these conditions we find,
using Eqs. (3.49), (3.50) and  (3.52)with matching between  cycles, the
following recovery ratios for each sequence:


                                  65

-------
      1.0
e     0.5
 r
I    I    i	1	1	1	1
                            0.5
                                1.0
                              w
            Figure 3.8  Recovery Ratio, R, as a

                  Function of e  and e  (t =0)
                               r      w   s
        Figure 3.7  Sketch of the Interface at

                    the End of Withdrawal
                         66

-------
                    - V          *1 " 
             R  = 1 + £  T  - £ ,    £2 = £2 ..  + £2(l+2£  .)
              n        n-1    n'     n    n-1     1     n-1

where recovery ratio R  for the n'  sequence  of cycles is defined as the
ratio of the volume of water withdrawn during that sequence to the volume
injected during recharge in that sequence. The following examples il-
lustrate how the numerical values of the recharge ratio increase during
repeated sequences:
      Sequence no.     12345

      !Ll = 0.141:    0.859     0.928     0.942     0.949     0.953

      !L1 = 0.526:    0.475     0.608     0.643     0.661     0.672


It can be seen that there is a significant increase in recovery ratio
during the first few sequences but that an upper limit on the recovery
ratio is approached asymptotically after several sequences.   However, it
should be recognized that dispersive effects are cumulative and thus are
likely to alter these immiscible results after several sequences.

When the effects of dispersion are included, prediction procedures are
more complex and general analytical representations are not possible.
As discussed in Section 3.4, there are two different theories - immiscible
dynamics with superimposed dispersion (Section 3.2) or coupled dynamics
(Section 3.3) - which may be applicable depending on whether the ratio of
dispersed zone thickness to interface length, A/8H, is respectively much
less than or much greater than unity.  In addition, there will be some
transition region in which neither of the theories is strictly correct
and transition from one theory to the other could concievably occur sev-
eral times during a recharge-storage-withdrawal sequence.

For purposes of specific prediction during a sequence, it will be assumed
that the immiscible model applies when A/$H<1 and the coupled model when
A/0H>1.  Thus, the values of A and $H are computed from the appropriate
equations in Sections 3.2 or 3.3 and final values from the previous cycle
provide the initial values of the dispersed zone width, A , and  the in-
verse slope, 3 , for the current cycle.  The ratio A/3H is monitored
throughout the computation and the appropriate model  (immiscible or coupled)
is selected using the current value of that variable.


                                  67

-------
Table 3.1 summarizes the various equations  that  are  used  in the  computa-
tional Trooe-.luro for a recharge-storage-withdrawal sequence.   The table
is esse  .: 1 :.y complete, but those  recharge  equations marked with an as-
terisk       ••-.-.j.o.iad for the particular initial  condition of a vertical
interface ^  \..= 0 with A =0.  More  general  forms would  be required in
some cases.  For example, when an initial dispersed  zone  width,  A >  is
specified during recharge with an initial radius, r  , and volume, V ,
Eq. (3.65) is generalized to
               16Tm
    A
    A  =
2
16i;e a~
r 2
H2

2
₯ r -₯
62H2
U 4* 2
r
2
r
                                                                  (3.120)
The computational procedure outlined above is used  in  Section  5.6  to
develop comparisons between the miscible theories and  the  laboratory
experiments.
                                   68

-------
     TABLE 3.1  SUMMARY OF EQUATIONS APPLICABLE TO A RECHARGE, STORAGE, WITHDRAWAL OPERATION
                       A/6H1

e
(AV/H)
A
(V^TO
A
6H
Immiscible Model (Sections 3.1 & 3.2) Coupled Model (Section 3.3)
Recharge
Eq. 3.52
Eq. 3.65*
or
Eq. 3.120
Eq. 3.67*
Storage
Eq. 3.49
A
o
A /Eq. 3.49
J
Withdrawal
Eq. 3.50
Eq. 3.76
Eq. 3.77
Recharge
Eq. 3.90
Eq. 3.91*
Eq. 3.92
Storage
Eq. 3.100
A
o
A /Eq. 3.100
Withdrawal
Eq. 3.106
Eq. 3.105
Eq.3.105/Eq.3.106
For the case V  = A  = 0; a more general form (e.g., Eq. 3.120) is required in some cases.

-------
                    4.   LINEAR FLOW EXPERIMENTS

The objectives of this  part of the experimental  investigation were  to
evaluate the theoretical results for linear flow and  determine any
limits of their validity.  Experiments designed  to  evaluate  the immis-
cible displacement model were successfully conducted.   Due to model
capabilities and the availability of materials for  the  porous media,
experiments to evaluate the theory for the interaction  between grain
dispersion and density  induced mixing were not conducted.  Addition-
ally, experiments were  run to obtain data on the behavior  of the inter-
face for small T.

                    4.1  Experimental Equipment

The linear experiments  were conducted in a lucite box which  contained
the porous media in a section 120 inches long, 3 inches wide and
6 inches deep (Figure 4.1),The porous medium was placed in the test
section through the top of the box.  A sponge rubber  strip was placed
between the medium and the top plate.  The sponge rubber could expand
to prevent short circuiting along the top in the event  that  the medium
settled with time.  The medium was kept in place by brass  screens at
the ends of the box.  Beyond the screens, the inlet and outlet reser-
voirs, which were free of porous media, provided constant  heads over
the cross section.  Inflow and outflow of the inlet and outlet reser-
voirs, respectively, were connected to constant  head  devices in order
to maintain steady flow rates during experiments.

Piezometer taps were located in the side wall at distances of 4.75 cm,
149 cm and 296 cm from the upstream or inlet screen.   The  upstream
screen was taken as the origin of the coordinate axes,  with  the x-axis
horizontal and at the mid-height of the medium.   Three  electrical con-
ductivity probes were located at x-^ = 62.4 cm,  X2 = 182.9  cm and
Xo = 274.2 cm.  The probes consisted of two parallel  stainless steel
rods, 1/8 inches in diameter, spaced 3/8 inches  on centers.   The probes
averaged the conductivity over the depth of the  medium to  give a one-
dimensional record of concentration,

                          4.2  Procedure

After a careful packing, the porous medium was saturated with water,
before the rubber strip and top plate were installed, to prevent trap-
ping of air in the medium.  In place measurements of  hydraulic conduc-
tivity were made using fresh and salt water before and after each  series
of runs.

Preliminary experiments were conducted to determine the longitudinal
grain dispersion coefficient for each medium used.  Salt solutions of
specific gravity 1.0005, were used as tracers.  For the initial
condition
                      t = 0,  c = 0  x>0

                              c = c  x < 0
                                   o

                                  71

-------
N3
              I
                                 Constant Head Devices
                       >J     tfl
                   Fresh
                   Water
                                       .1 Inle

                                 ife
                                      "*"
              Inlet Valves
                   x
                       Probe  1
                      Probe  2
                                    01
                                    
-------
the dispersion coefficient could be determined using the solution
                                                                 (4.1)
where U is the steady seepage velocity.  The slope of the concentration
versus time curve, for a constant x, at t = x/U = t,-n, is used to deter-
mine D.. from
                           Dl =
                                47T
where
50
                               c  3t
                                o
           t=t
                                        50
To satisfy the initial condition, the experimental model was initially
filled with fresh water, c = 0, and the flow rate set by observing the
manometer and adjusting the constant head devices.  Then, the fresh
water flow was shut off and the salt water inlet valve and the reservoir
outlet valve were opened simultaneously.  The fresh water in the inlet
reservoir was replaced by the salt tracer solution.  Next, the reservoir
outlet valve was closed and the flow rate valve was opened to start the
experiment.  This procedure was successful in introducing an initially
vertical interface, corresponding to the initial condition.

The conductivity probes were connected through an AC Wheatstone bridge
to a recording oscillograph to produce a continuous time record of
conductivity.  Since conductivity was approximately a linear function
of tracer concentration, the conductivity records were interpreted
directly as concentration records and the necessary information was
obtained from them.

For the actual experiments, the model initially contained salt water
of known density p« and salt concentration c .  Again the flow rate was
set by observing the manometers and adjusting the constant head devices.
The initially vertical interface was introduced by the procedure used
for the grain dispersion experiments, except in this case dyed fresh
water replaced the salt water initially in the inlet reservoir.  Records
of conductivity were obtained in the same manner as for the dispersion
experiments.  For runs with a density difference, the interface posi-
was observed visually and recorded periodically on the front of the box
with a china marker.  These visual markings were recorded and the hori-
zontal projection of the interface, L, was determined for each marking.
                                 73

-------
 A series of experiments  was  conducted with the  box containing  a  crushed
 quartz sand having a mean grain diameter,  d,_n,  of  0.069  cm,  and  a  uni-
 formity coefficient of 1.37.   The  inplace  hydraulic conductivity was
 0.10 cm/sec and the porosity was 0.35.

              4.3  Experimental Results for Linear  Flow

 Preliminary experiments  to determine  the longitudinal grain  dispersion
 coefficient produced results  indicating that local heterogeneities were
 present in  the  medium.   The  dispersion coefficients determined from
 conductivity records at  probe 3 were  larger than those determined  from
 records at  probe 2  which were in turn larger than  those  determined from
 records at  probe 1,   Visual  observations indicated thin  layers having
 higher hydraulic conductivity than  the surrounding media.  Thus, the
 measured dispersion coefficients reflected the  effects of this layering.
 The  largest indicated dispersion coefficient was used to characterize
 the  longitudinal dispersion  coefficient, D , for the medium.

 A number of experimental runs were  made for specified values of  the
 density difference,  Ap,  and  the seepage velocity,  U.  The parameters
 for  these_experiments are given in  Table 4.1 along with  the  computed
 value  of E/D for each run.   Smaller  values of  E/D  could not be ob-
 tained because  of limiting model capabilities.  The horizontal inter-
 face projections,  as determined by  visual  observations,  are  shown  as a
 function of time on the  dimensionless plot in Figure 4.2.  Also  shown
 are  the theoretical expressions for L/H versus  T given by Eq.  (2.19)
 and  the interpolation equation (Eq, 2.72)  given by Gardner,  et al.,
 (1962).   The result  given by  Keulegan (1954)


                              L/H =  /2r                          (4.2)

 to represent  his  experimental results for  an unconfined  system with no
 net flow is  also  shown in Figure 4,2,   It  appears  that the presence
 of a free surface reduces the rate  of interface tilting.  Gardner,
 et al.,  (1962)   also reported experimental results for models with no
 net flow which  confirmed  their theoretical result  (Eq. 2.72),

 The_experimental  results  of the current study show that, in  this range
 of E/D ,  the  rate of  interface tilting is  not affected by the net  seep-
 age velocity  in  these experiments.  The current experimental results
 do not  provide  a  direct  confirmation  of the coupled analysis in  Section
 2.3  (Eq.  2.63) but  are ^consistent with that theory.  That is, for  each
 value  of  the  parameter E/D.. the range in dimensionless time  T is such
 that the  immiscible  solution  (Eq. 2.19) should  be  valid  as discussed in
 Section  2,4  and  indicated in  Figure 2.3,   It was found that, using this
 porous medium, it is not  possible to  operate this  experimental equip-
ment in  the  range where grain dispersion becomes important.  A model of
 greater  length or larger  dispersivity would be  required  to observe these
 effects.
                                  74

-------
                                          1     I—I   I  II
0.1
   0.1
                   1.0                           10                          100
Figure 4.2   Horizontal Projection of Interface,  L, as a Function of Time

-------
u
(cm/sec)
.008
.0101
.0085
.0309
Ap_
P
0,030
0.030
0.030
0,030
E
,2, .
(cm /sec)
0,0218
0,0218
0.0218
0,0218
"E7D
£./ JJ,
13.7
10.8
12.7
3.5
                a = 0.2 cm,  D. = Ua , E = - &- f
                 1         '1     1'     n p  6

      Table 4.1:  Range of Test Results for the Linear Model
One-dimensional concentration, c, was recorded as a function of time at
the three conductivity_j>robes, for each experimental run.  A typical
experimental curve of c/c  versus T is shown, compared with the approxi-
mate one-dimensional concentration model given by Eq. (2.33) in Figure
4.3.  A comparison of the results indicates a reasonable degree of agree-
ment between theory and experiment especially when it is noted that
Eq. (2.33) was based on an average dispersion coefficient over the mixed
zone.
                                  76

-------
      1.0
      0.9
       .7

       .6

C/CQ   .5

       .4

       .3

       .2

      0.1
                                               Probe  1 (x=63.4 cm)
                                               E/I>1 =12.7
                                               Ap/p = 0.03
                                               Experimental  Curve
                                               Theoretical Curve
                                                 Eq.  (2.33)
        2.0
3,0
4.0
5.0
6.0
7.8
8.0
9.0
                Figure 4.3  A Typical Linear Flow, Experimental Breakthrough Curve Compared
                            with Theory

-------
                    5.  RADIAL FLOW EXPERIMENTS

There were two principal objectives of the radial flow experiments:
(1) to construct a model with radial flow characteristics  capable of
modeling real aquifer field situations; and (2) to evaluate the theo-
retical results and to determine the limits of their validity.   The
theoretical results initially were to be evaluated for the immiscible
displacement model including evaluation of the separate topics  of in-
jection, storage and withdrawal.  Other tests were needed  to evaluate
the effects of grain dispersion on the immiscible model and the combined
effects of density and grain dispersion on the total mixing process.

               5.1  Model Construction and Equipment

The construction and operation of the radial model was more complex
than the linear model because radial flow from a line source is more
difficult to achieve than uniform flow in a constant area  model.  Also,
experience during operation of the linear model indicated  that  the choice
of media and the method of packing were important for measurement of  dis-
persive effects.

Considerable care was taken to determine the model dimensions and media
to be used.  Several porous media including porous foam rubber, crushed
quartz sand, and a consolidated media formed from plastic  beads mixed
with epoxy resin were tested in a small linear model.  The plastic beads
mixed with epoxy resin forming a rigid porous media were found  to be
most suitable considering homogeneity and ease of construction.  The
dimensions of the model were determined from practical considerations.
The necessity of packing the model uniformly while still leaving a model
thickness large enough to allow accurate interface measurements indi-
cated that a model thickness of three inches was desirable.  Using this
depth and considering estimated interfacial tilting rates, it was deter-
mined that the model should have an eight foot radius.  To keep the
model manageable with this radius, it was necessary to construct a sec-
tor model with a 15 degree included angle.  These dimensions enabled
the modeling of a variety of real aquifer situations by varying the
density difference of the fluids and the flowrate.

Once the dimensions and choice of media were determined, the model was
constructed as shown in Figures 5,1 and 5.2,  The model consisted of a
1/8 inch thick plexiglass sheet fastened to a 1 inch thick piece of
plywood, 2-1/2 feet by 8 feet, which acted as the base.  The two radial
sides of the model were 1/2 inch thick plexiglass sheets,  three inches
high, that allowed visual observation of dyed fluids in the model.  The
sheets were placed at a 15 degree angle with the apex having a 1/8 inch
opening, simulating the well.  Piezometer taps, 1/8 inch in diameter,
were drilled at ten locations in the model, including one within the
well itself (see Figure 5,1).  The piezometer  taps were connected with
plastic tubing to a piezometer board and a manometer.  The piezometer
board consisted of open-ended plastic  tubes  (1/4"ID) fastened  to a board
with paper marked at 1 mm intervals as a background.  The manometer was
                                  79

-------
                            Constant Head Tanks
00
o
               To
           Fresh-Water
            Reservoir

                Rotating Tube
            Valve  to Initiate  lij
           Vertical Interface  [Jj
                             Wastewater
                    Well radius = 0.5 inches
                    Media radius = 7 ft 10 in
                    Media angle = 15 degrees
                                                                                   Piezometer Board
                                                                  Collector
                                                                   Manifold
To Salt-Water
 Reservoir
                                  Well
                                  Flushing
                                  Valve
                                                                             Wastewater
                                                 Note:
                                                                     Outlet
                                                                     Valve
                                 Total model length = approx. 10 ft
                                 Model thickness = 3 inches
                                ,lje.g., Location of piezometer number 1
            Piezometer
              Number
              Radius
          I     (cm)
 123456       78       9       10

1.27   31.3   61.8   123.1   123.1   123.1   199.5   199.5   199.5   239.0
                                       Figure 5.1  View of the Radial Model

-------
Figure 5.2  Photographs of the Radial Model
            During Operation
                    SI

-------
 graduated in hundredths  of  a  foot  and  could be read to a thousandth of
 a foot  accuracy with  a vernier  scale.   Since  it was more accurate than
 the  piezometer board, the manometer was used  in all tests where piezom-
 eter difference was measured.   The apparatus was designed so that the
 manometer could be connected  between any  two  piezometer taps.  The
 piezometer board was  used to  give  an overall measurement of the piezome-
 tric head and to simultaneously compare the head at different locations.

 With the  piezometer taps drilled,  the  model was ready for packing.
 Glass beads about one millimeter in diameter were sieved through a U.S.
 Standard  Sieve Series.   Only  those beads  held by the no. 20 sieve after
 passing the no. 18 sieve (1000  micron  mesh) were used.  This insured
 uniformity of the media  and gave a mean diameter (d,-n) of about 1 mm
 and  a uniformity coefficient  of approximately 1.0.  The beads were mixed
 with epoxy resin (1 part resin  to  25 parts beads by volume) and hand-
 packed  into the model.  While the  epoxy was hardening, 21 conductivity
 probes  were inserted vertically into the  media (see Figure 5.3).  The
 probes  consisted of two hypodermic needles (24 gauge stainless steel
 tubing) located 3/4 inch apart  on  a circumferential line at various
 radii from the well.  A  resistance measuring  circuit, consisting of a
 carrier amplifier system (Sanborn  Co.  Model 350-1100C), was connected
 to the  probes and then to a recording  oscillograph  (Sanborn Co. Model
 296).   The probes were inserted through the entire thickness of media;
 hence,  the recorder gave the  depth averaged change in conductance as a
 function  of time at the  location of each  probe.  These probes are here-
 after referred to as  "averaging probes".  One of these averaging probes
 was  inserted inside the well.   This probe served to indicate if the
 start of  a test had a step  input with  a vertical interface.  During with-
 drawal  tests this-probe  indicated  the  time at x^hich saltwater first
 reached the well.  This  information was used  later in determining the
 percentage of water recovered during the  test.

 What  shall be referred to as  "point" probes were also inserted into the
 top  of  the model.  These probes  consisted of  two very short hypodermic
 needles (16 gauge) inserted just through  the  top coat of the model,
 1/16  inch  apart.  They were located at the same radius and adjacent to
 averaging  probes #6 and #10.  These point probes were also connected to
 a resistance measuring circuit  and then to a recorder.

 The  3/4 inch spacing of the averaging  probes x^as originally chosen from
 consideration of two factors.   First,  it  was felt that they should be
 this  far  apart so that during insertion of the probes the media between
 them wouldn't  be highly disturbed.  Secondly, this spacing yielded sat-
 isfactory  current between the probes for  anticipated salt concentrations
 (1%  to  2.5%).  Experimentation  showed  that it would have been better if
 the averaging  probes were moved  closer together.   There was apparently
 little  disturbance to the media  during probe insertion.  In addition,
 salt  concentrations much lower  than anticipated (less than 0.1% by
weight) were  used for the tests.   This meant that the current was less
 than expected.  Finally, a  third factor was the radial width of the
electric field generated between the probes.   With 3/4 inch spacing
                                  82

-------
00
OJ
                                     —G>-©
              Well
                              1) e.g., Location of Probe  Number  1
                                                                                                 Reservoir
                 Probe Number    12345678      9      10      11
                  Radius (cm)  1.27  19.6   34.9   45.7  50.0  67.7  87.0  104.5  104.2  122.9   142.0
                 Probe Number   12     13     14      15     16     17     18     19     20     21
                |  Radius (cm) 142.0  165.0   165.0   165.0  167.0  193.5  193.5  218.0  218.0   218.0
                                   Figure 5,3  Location  of  Conductivity Probes

-------
 between the probes, the electric field width (area that  is  sensitive  to
 the approaching interface)  became significant  and  the  probes  read  not as
 points but as zones.   This  caused what can be  termed electronic  disper-
 sion of the recorder output.   For example, a vertical  immiscible inter-
 face approaching and passing  the probes would  yield a  characteristic
 I!S" shaped dispersion curve output rather  than a step  change  as  desired.
 If the probes x^ere moved closer  together,  the  thickness  of  this  zone  of
 influence would be lessened and  electronic dispersion  decreased.   Prob-
 ably 3/8 inch spacing of the  probes would  have been preferable.  A po-
 tential theory analysis of  the electronic  dispersion for the  3/4 inch
 spacing, however,  showed it to be approximately an order of magnitude
 less than hydrodynamic dispersion and  it was disregarded during  most
 data analysis.

 With the probes in place, the top surface  of the model could  be  con-
 structed.   It x^as  made by spreading a  1/8  inch thick epoxy  coat  onto
 the exposed media.   Prior to  application the epoxy was allowed to  become
 very viscous so that  it wouldn't flow  into the media.  It hardened into
 a  translucent top  surface,  bound to the media,  eliminating  short cir-
 cuiting of the fluid  flow along  the top.   The  model was  to  be 7.62 cm high,
 but due to seepage of the epoxy  top coat in certain small areas, the
 height was somewhat less.   Visual observation  of the seepage  through
 the side plates made  it possible to estimate an average  height of  7.46 cm
 for the model.

 The well end of the model consisted of  concentric  plastic tubes with
 slots cut  into them that acted as a valve  during tests (see Figure 5.4).
 This system was designed to insure a vertical  interface  between  the
 fresh and  saltwater at the  start of a  test.

 The well tube was  connected through a  "T"  fitting  and  valves  to  saltwater
 and freshwater constant head  tanks (see Figure  5.1).   The constant head
 tanks  could be raised above or below the well  level in order  to  simulate
 injection  or withdrawal from  the well.  The  valves  and "T"  connection
 allowed a  rapid change from freshwater  to  saltwater feeding of the well,
 or  vice versa.

 The  end of  the model  opposite  from the  well  consisted  of  the  media end-
 ing  in  a circular  arc  of 7  foot  10 inch radius  followed  by  a  large
 reservoir  (see Figures.1).  The reservoir was connected through a mani-
 fold to  a  constant  head tank.  During an injection  cycle, water was
 drained  from this  tank, but during withdrawal  fluid was  constantly sup-
 plied  in order to maintain  the head.  The  system of constant  head  tanks
 at either end  of the model  produced hydrostatic conditions  within  the
well and the  reservoir.

                    5.2  Experimental Procedure

Preparation  of  the Model for Operation  Once the model was  constructed,
it was  tilted  at an angle to the  floor  and saturated slowly from the  well.
By tilting  the model,  air within  the media had  a chance  to  escape  through
                                  84

-------
00
Ul
                                     To Constant Head Tanks
                                            Rubber 0-Ring
                                            1/2" I.D.
                                            3/4" O.D.
                                            Rubber 0-Ring
                                            Flow into the model occurs only when
                                            the slots line up


3/4" I.D.
1" O.D. — •}


To Exit Valve
and
Wastewater Tank
x^— ~^ — ^—








— 	 	 — 	
Slot
3" high
1/8" wide
	 	



-*•—
-V




N



— 	 '
— --"
~PH.
X.




^


                                             Figure  5.4  Detail  of  Well Construction

-------
 the reservoir end as the water entering slowly displaced it.   After the
 model was saturated, it was leveled and prepared for experiments.

 Deaerated water was used to saturate the model and run tests.   This was
 done to minimize the likelihood of gas coming out of solution  in the
 model forming air pockets.  Deaeration was accomplished by pumping the
 water through a chamber under a vacuum until the dissolved oxygen level
 was around 3 mg/1.  The water was also sand filtered and passed through
 a demineralizer that removed anions and cations from the water.  Finally,
 to avoid the growth of either bacteria or algae in the model,  a disin-
 fectant, zephiran chloride, was added to the water.

 Model Operation during Tests  To start a test, a solution of sugar, salt
 and water that had the desired density and conductivity was pumped into
 the model.  The freshwater constant head tank was set to the proper
 height for the desired flowrate.  About 30 seconds before the  start of
 the test, valves were opened allowing the freshwater to go from the con-
 stant head tank into the well, flushing the saltwater out through  an
 exit valve (see FigureS.l). At time t = 0 the valve within the well was
 turned so that the slots lined up and flow into the model could occur.
 At the same time, the well exit valve was closed and a valve at the
 reservoir end of the model was opened to allow flow through the model.
 Flow was continued as long as desired, stopped for a period if neces-
 sary,  and could be reversed by changing the relative position  of the
 constant head tanks.  This procedure was used to model recharge, storage
 and withdrawal cycles.

 Methods of Data Collection  There were three types of measurements made
 during tests:  (1) measurement of flowrate; (2) visual measurement of
 the interface shape and location; and (3) measurements made with the
 conductivity probes.  Flow measurements were made by using a 1000  ml
 graduated cylinder and a stopwatch.   Several readings were taken during
 a  test and the average of the readings was used as the flowrate.

 Visual measurement of the interface  consisted of periodic marking  of
 the interface location with a china  marker on the clear plexiglass
 sidewalls.   The time of each marking was recorded.   It was found that
 as long as  the density  difference was in the range of principal interest
 (from  1% to  2.5%), the  dyed fluid made a distinct interface that could
be observed and whose location  could  easily be  marked.   When the den-
sity difference between  the  fluids was  less than 1% by weight,  the inter-
face was no longer sharply  defined.   Apparently,  as the  density  difference
between  the two fluids  decreased,  the  relative  importance  of dispersion
and the  effects of slight inhomogeneities  in  the model  construction
increased.  Since  the experiments normally  had  higher density  differences
than 1%, visual observation  of  the interface was possible.

In addition to the visual observation of  the  interface position, mea-
surement of the interface was made with  the  averaging conductivity
probes.  In order to obtain  a linear  relationship between  salt  concen-
tration and output from the  recorder, salt  concentrations  less  than 0.1%
                                  86

-------
by weight had to be used (see Figure 5.5).   In order to achieve the
desired density difference of 1% to 2.5%, sugar was also added to the
salted water.  The sugar was a nonelectrolyte and hence increased the
water density without affecting its conductivity.  The sugar,  however,
did have some affect on the viscosity of the denser fluid.   Viscosity
increases of 2.5 to 7 percent are found for sugar concentrations of
from 1% to 2.5% by weight.

              5.4  Determination of Model Parameters

The porosity of the model was determined in two ways.  First,  a value
of 0.337 was determined using a small test cylinder that had been packed
at the time of model construction.  Its porosity was determined from
comparison of the volume of water required to saturate the cylinder
versus the gross volume of the cylinder.  Since the test specimen was
packed in a similar manner to the model, with the same bead and epoxy
mixture, its porosity should be the same as that of the model.  The
second method for determining the porosity was done using the in situ
model media.  By measuring the travel time of fluid flow between con-
ductivity probes at different radii, an average porosity of 0.34 x^as
determined.  This value is used in subsequent calculations.

The hydraulic conductivity of the media was also determined in two ways.
First, a cylindrical permeameter  (1-1/2" diameter, 10" long) which had
been packed at the time of model construction was tested.  Hydraulic
conductivity tests were made using this cylinder with the flow perpen-
dicular to the packed surface.  The hydraulic conductivity from these
tests approximated the results that would be obtained by vertical flow
through the actual model.  Consequently, the hydraulic conductivity of
the test cylinder was an approximation of the vertical conductivity of
the radial model.  The average vertical conductivity from these tests
was 0.48 cm/sec at 60°F.

The horizontal conductivity of the model was determined from tests run
on the in situ media.  The tests  consisted of measuring piezometric
head and flow rate in order to calculate the hydraulic conductivity
from the well equation for a confined aquifer between two constant head
reservoirs.  The equation is:
                                                                  (5.D
     where:  K = horizontal hydraulic conductivity  (cm/sec)
             Q = flowrate  (cm /gee)
             s ^ piezometric difference between the radii
                 r  and r2  (cm),  (r,>r?)
             H ^ thickness of the aquifer  (cm)

The average horizontal conductivity derived from the use of  this  equation
was 0,43 cm/sec at  60°F  (see Table 5.1).   The fact  that the  conductivity
                                  87

-------
CO
CD
           V-i
           QJ
           4-1
           n)
           13
           cn
           •H
           p

           C
           •H
           ra
           to
           C
           cu
           CJ
           t-l
           0)
                .10
                .08   ~
.06   —
                .04   —
                .02   —
                             100       200      300       400      500


                                                Recorder  Deflection
                                                             600
700
                             Figure  5.5   Calibration  of  Averaging Conductivity Probes

-------
                      TABLE 5.1

          HORIZONTAL HYDRAULIC CONDUCTIVITY

Note:  The governing equation for the model:
                         r\ '
                 K =

Temperature of the water for all runs = 75°F

From the table below, the average K = 0.52

Average K = 0.43
                                           sec
                                               @ 75°F
                 sec
                     @ 60°F
Location
(Number refers
to piezometer)
1 to 10
1 to 8
1 to 5
1 to 3
1 to 2
2 to 8
2 to 5
3 to 8
Discharge*
Q' (cm-Ysec)
0.569
2.4
3.63
6.14
2.23
3.93
1.56
4.80
1.18
4,18
2.03
4.92
2.34
3.36
4.82
2.34
3.36
4.82
2.34
3.36
4.82
K
(cm/sec)
0.502
0.514
0.500
0.49
0.509
0.501
0.515
0.495
0.504
0.488
0.496
0.474
0.563
0.606
0.531
0.572
0.628
0.532
0.563
0.569
0.548
*This is the sector model discharge; i.e., Q = 24Q'
                            89

-------
 measured in the test  cylinder was  very  close  to  the  in  situ horizontal
 conductivity of the model indicated  that  the  media was  reasonably
 isotropic.

 Since the model was to  be used to  study the effects  of  dispersion on
 the mixing  process  between the two fluids  at  the interface, a measure-
 ment of the media's dispersivity was  required.   The  dispersivity was
 determined  through  the  use of the  conductivity probes already mentioned.
 As  the interface between the  two fluids passed a probe,  the reading from
 the recorder plotted  the change in salt concentration versus time at  that
 point.   There are several theories that mathematically  describe the con-
 centration  c/c  at  a  particular radius  for flow  from a well.  One recent
 study (see  Gelnar and Collins,  1971)  can be used to  estimate the longi-
 tudinal dispersivity  of the media  by  considering the slope of the curve
 c/c  versus time at the point c/c  =0.5 for  the case of no density dif-
 ference between the two fluids.  Gelhar and Collins  derived the following
 equation assuming that  the dispersivity, a ,  is  a constant:
lrl
2u
1 r-.~.f /_
T err\

,16_

^ 3-,
2
r - i
3v ]
2
) r 4— r 4
m , * co . , 1/2
                                                       }]
                                                                  (5.2)
     where;  c = concentration of tracer at radius r
            c  ^ original concentration of the tracer in the model
            ou ^ longitudinal dispersivity
            t  = time
            r = radius to point in question	
           rA = radius to the 50% point (=/2At )
           r  = radius of the well
           D  = coefficient of molecular diffusion in the porous medium
            A =  ^   where n = porosity, H = depth of aquifer, Q = flowrate
            "   /TTnH
The effects of molecular diffusion and the radius of the well are small
in magnitude compared to other terms and Eq. (5.2) reduces to

                                   r2 - r 2
                 C    1                  *^
                 co   2         ,16     3.1/2
                                (-^ v*}
                                                                  (5.3)
Taking the derivative of Eq. (5.3) with_respect to time and evaluating
at c/c  = 0.5 remembering that  r  - r.  = 0 at c/c  = 0.5,
      o                               *            o
                   3(c/c
                           c/cQ=.5
                                         1677  a,
                                                  = S
                                                     50
                                                                 (5.4)
Since t = rA /2A, we can solve for a  in terms of the slope at the
                                  90

-------
point c/c  = 0.5 and r.
         o
Experimentally, rA, is the radius to the probe in question and the slope
at c/c  = 0.5 is determined from the recorder output.

Using this method of analysis, the dispersivity was determined from
several of the averaging probes as summarized in Table 5.2 and Figure
5.6.  The longitudinal dispersivity generally falls in the range between
0.1 and 0.2 cm although at the larger radii, a few observations indicate
higher values.  These large values of the dispersivity may be associated
with inhomogenieties of the media.  The very small density difference
(see Table 5.2) may also have affected the results; however, for these
values of e this effect is indicated to be negligible  using the results
of the coupled analysis in Section 3.3.  The so-called electronic dis-
persion associated with the averaging probes may have  introduced some
error in the dispersivity determination for some probes near the origin,
but, as discussed in Appendix C, this effect is unimportant over most
of the model.

A few determinations of dispersivity made using the more closely spaced
point probes were found to be of the same magnitude (0.075 cm < 04 < 0.18 cm)
as those found with the averaging probes.  In all subsequent calculations,
the longitudinal dispersivity is taken to be constant  throughout the
medium and equal to 0.15 cm.

                      5.5  Model Verification

The model was tested to see if radial flow was in fact occurring.  Obser-
vations from the piezometers were used to obtain the distribution of draw-
down versus radius from the well.  The results are plotted in Figure 5.7
and compared with the solution of the well equation, Eq.  (5.1).  Com-
parisons of piezometric head for the taps at the same radius showed no
measurable differences thus indicating a purely radial flow.  The obser-
vations of piezometric head distribution in the model generally indicate
that the desired radial flow was attained and that the model was homo-
geneous with respect to hydraulic conductivity.

To further check the flow situation, observations of travel times were
made with practically zero density difference between the two fluids.  A
small amount of salt was added to one fluid  (0.02% salt by weight) in
order to obtain a measurable output from the conductivity probes.  By
comparing the readings from probes located at the same radius, it could
be seen that the saltwater was  reaching radial points at approximately
the same time.  Also visual observation of the  semiclear  top coat of
the model during density tests showed that the  injected fluid moved
                                   91

-------
                            TABLE 5.2

        DETERMINATION OF DISPERSIVITY FROM AVERAGING PROBES
Run Number
Discharge Q'cm-Vsec**
Density difference Ap/p
e
Probe
Number
2
3
5
6
7
8
9
10
11
12
14
17
18
21
Radius
cm
19.6
34.9
50.5
67.7
87.0
104.5
104.2
122.9
142.0
142.0
165.0
193.5
193.5
218.0
C-l
2.27
0*
0
1-2
4.66
0
0
1
4.94
0.001
0.039
3
2.96
0
0
5
3.40
0.001
0
Longitudinal Dispersivity a - cm
0.153
0.108
0.110
0.151
0.150
0.293

0.318







0.105

0.144
0.148
0.293









0.095

0.163

0.172



0.14


0.175
0.355


0.115
0.142
0.125


0.106
0.162
0.152
0.173
0.2
0.217



0.0927

0.146


0.157


0.405



 The  condition  Ap/p  =  0  indicates  cases  in which  the  density  of
 the  liquids was matched using  sugar  and salt  solutions  to within
 ±0.0002  relative  density  difference.

**This is the sector model discharge; i.e.,  Q = 24Q'
                                 92

-------
                0,4
                0,3
Run Number
A  i
V  3
O  5
D 1-2
O c-i
LO
•H
w

Q)
ex
tfl
•H
O
                no
                0.2
                0.1
      I
                                                             1
                            0
              0
                                                             V
                                                                      1
                                20
           60          80

               Radius  (cm)
100
                                                                                 120
140
                                Figure 5.6  Media Dispersivity  Calculated From the Averaging Probes

-------
6 -
                         2.0
3.0
4.0
5.0
6.0
7.0
                   Figure 5.7  Piezometric Head Distribution in the Model

-------
radially into the model with a velocity independent of angular position.

To correlate the effectiveness of visual observation of the  interface
location with the depth-averaging conductivity probe measurements,  the
following comparison was made.  The interface position was marked on
both of the side walls as the interface passed a conductivity probe.  A
comparison was made of n/H as determined from the visual markings of the
interface with the relative concentration,  c/c , from the probe output.
These values were recorded at several different times as the interface
passed the probe.  The results for a run with a relatively large density
difference (Ap/p=0.020) are shown in Figure 5.8.  This figure indicates
a good correlation between the probe reading and the visual  recordings.

             5.6  Experimental Results for  Radial Flow

A series of experiments was undertaken to investigate the behavior of
density induced mixing in aquifers using the previously described visual
and electronic methods of data collection.   Table 5.3 summarizes the ex-
perimental parameters along with the computed values of e and e  for
each run.  In some experiments, only recharge tests were conducted, while
in other experiments complete recharge, storage and withdrawal cycles
were investigated.  When an item is not given in the table,  it was not
measured during that particular run.  A summary of model characteristics,
as used in data reduction, is given for convenience:

             Model Height                 H = 7.46 cm
             Porosity                     n = 0.34

             Longitudinal Dispersivity    a. = 0.15 cm

The hydraulic conductivity, K, was determined for each run using the
average viscosity of the two fluids (see Table 5.3).

The parameters e  and e  were found in Section 3 to characterize the re-
charge and withdrawal processes.  In Figure 5.8 the results  of a recharge
run with a relatively large value of e  (=0.422) using both  visual and
electronic interface measurements are compared with the predicted inter-
face position of the immiscible theory with a linear interface in volu-
metric coordinates (Eq. 3.107).  The theory, which is based  on small e ,
compares well with the experimental data.  The indication is that the
interface tilting is governed predominantly by immiscible displacement
under these conditions and that the curvature of the interface is unim-
portant even for this relatively large value of e  .

Electronic data were obtained in the form of breakthrough curves (see
Figure 5.12) for various conductivity probes during recharge and with-
drawal.  A method was developed for interpreting these data_based on the
measured rate of change of depth averaged concentration at  c/c =0.5.
This information was used to characterize the various breakthrough
curves and was expressed in terms of the parameter At:
                                 95

-------
TABLE 5.3  EXPERIMENTAL CONDITIONS FOR THE RADIAL MODEL
Run
1
2
1-1
A-l
A- 2
A- 3
A- 4
A- 5
A- 6
A- 7
A- 8
A- 9
A-10
B-l
B-2
B-3
B-4
C-l
Ap/p1*
0.001
0.001
0.005
0.005
0.005
0.005
0.0215
0.0215
0.021
0.021
0.021
0.020
0.020
0.018
0.018
0.018
0.018
0.00014
Q '*
r
3
cm
sec
4.94
3.87
2.49
3.70
3.70
0.76
4.79
1.32
0.48
2.43
2.30
0.78
0.26
2.55
1.64
5.04
1.86
2.28
Q '*
w
3
cm
sec



5.74
5.55



6.07
4.70



2.70
1.86
5.39
1.96

Fluid
Temp.
°F
-71
=71
75
70
71
69.5
^70
^70
= 70
71
69
70
67
69.5
-69
^69
68
67
Average
V
centi-
stokes
0.995
0.995
0.940
0.985
0.977
0.985
1.012
1.012
1.012
0.995
1.020
1.012
1.050
1.012
1.010
1.010
1.020
1.050
Average
K
era
sec
0.49
0.49
0.52
0.49
0.50
0.49
0.48
0.48
0.48
0.49
0.47
0.48
0.46
0.48
0.48
0.48
0.47
0.46
e
r
0.020
0.043
0.123
0.098
0.099
0,216
0.177
0.337
0.556
0.248
0.250
0.422
0.726
0.222
0.277
0.158
0.257
0.020
e
w



0.079
0.080



0.155
0.179



0.216
0.260
0.153
0.251

Recharge
Time
sec



1500
1500



6000
2423



1440
1440
1440
2100

Storage
Time
sec



75
49



240
157



480
720
1200
10900

=1.000 gm/cm  for all runs; - denotes estimated
                                   temperature; Q  ',Q  =model
                                                r   w
flowrates,  i.e.

-------




1.0

0.9
n/H 0.8
or
_. 0.7
c/c0
0.6
0.5
0.3
0.2
0.1

. | . | I | . | !
x .1 	 • t_ i — nni 	 / T* _ O oo\
Immiscible lay • (hq . J . Zz.)


~ t \ *\ \ ^
i * \ \ • \
_ i x \
i \ \
\ v \ 1 \
; i ^ \ \ ^x

1 \ A V
k \ " \
- ^ \ \ X.
" \ \ X \
\ ^ v
)> Y \
, .A i v\ i , ix,
2.0 4.0 6.0 8.0
i • i •
O Probe 5
A ii ii
V Probe 6
T
A Probe 7
A II II
D Probe 9
O Probe 10
^k ii ii
Run A-9:
X
\
\x
^8
\
i i ^ ^ i
10.0 12.0
1 ' 1 '
Visual
Electronic
Visual
Electronic _
Visual
Electronic
Visual
Electronic
Visual -
Electronic
e = 0.422
r
-
-
X
e ^
! , I r^
14.0 1 h . 0 J 8 . 0
                               (KApt/npH)
                      L      i.


Figure 5.8  Comparison of Visual and Electronic

            Measurement of the Interface with

            Immiscible Displacement Theory

-------
(At) 1 =
3(c/co)
at
c/c =0.
o
5
 For a linear interface in volumetric  coordinates  c/c  =0.5  coincides with
 ₯ =3s?* (r =rA) ,  where the subscript  "p"  refers  to  the°probe at  which the
 measurement is  being made.   The dimensionless  form of At  (Eq.  3.109)
 found from the  recharge experiments is  shown in Figure 5.9 as  a  function
 of dimensionless time, T (Eq.  2.14) and is  compared with  the theoretical
 results presented in Section 3 (Eq. 3.109).  The  range of  e and T is
 such that the ratio of the  dispersed  zone width to interface length in
 volumetric coordinates, A/$H,  is small.  Therefore only the immiscible
 theory should apply for the most part.   At  low 2er however, there may
 be some effects of dispersion  as noted  by the  dashed  line  for  e  =0.1.
 Figure 5.9 shows excellent  agreement  between theory and experiment,
 although some systematic differences  are noted in the early part of those
 runs with larger values of  e .   Figure  5.10 shows a comparable visual
 measurement of  the length of the interface  in  volumetric  coordinates,
 A₯=j3H (Eq.  3.28), versus time  for three runs with large e  .  Only the im-
 miscible theory (Eq.  3.31)  is  shown,  as e   and T  were in  the range in
 itfhich only density difference  governs the mixing  dynamics.  Excellent
 agreement between theory and experiment is again  found, although some
 systematic decrease in gH for  smaller values of 2e T  is indicated.  A
 glance at Figure 3.3  or Figure  3.4  indicates that this difference in At
 or  H for lower values of 2c T  during runs of  relatively high  e   cannot
 be attributed to dispersive effects.  Similar  differences  were observed
 for the  linear  experiments  presented  in Section 4.3 and are attributed
 to the influence of vertical velocities  for small times, and the limited
 applicability of the  Dupuit assumption  under those conditions.   Figures
 5.9 and  5.10 also show that, for this range of e   and  T,  the  rate of
 interface tilting is  not affected by  dispersion and thus provide a di-
 rect  confirmation of  immiscible displacement theory (Sections  3.1 and
 3.2)  and are consistent with the coupled theory (Sections  3.3).

 Figures  3.3 and  3.4 of Section  3.4  indicate that  density difference and
 grain  dispersion will interact  to govern the mixing dynamics for small
 time when longitudinal dispersion will  be important or for very  large
 time when lateral dispersion will be  important.   It was found  that it is
 not possible to  operate this experimental equipment in the range of 2e T
 where  lateral grain dispersion  becomes  important  for  recharge.   A model
 of  greater  length or  larger dispersivity would be required.  However,
 several  experimental  runs were  conducted for small e   over the range of
 times  (2e T) in  which the coupled theory should apply.  The results are
 presented in Figure 5.11; good  agreement between  theory and experiment
 is  indicated. The value  a =0,15  cm was used for the plots  of the coupled
 theory.   It is interesting  to note that  these  runs were made with a very
small density difference  (Table  5.3),  and that several of  them were also
analyzed  to determine  longitudinal dispersivity by  ignoring this small
density difference  (Table 5.2).   In light of the  importance of density
difference as indicated  in Figure 5.11,   it is possible that several of
the larger dispersivities found  in Table 5.2 may be due in part  to the
density effect.


                                  98

-------
 40  -
 10  -
  1	
0.1  _
0.01
          O A-l.A-2   0.098,0.099
                       0.216
                       0.177
                       0.337
                       0.556
                       0.248
                       0.250
                       0.422
                                                 Eq. 3.109
                  Eq. 3.90,  e  =0.1
.01
.1
                                                     10
                       2e T = 2e   - ^- -
                         r      r  n p  H
    Figure 5.9  Comparison of Theory and Experimental Time
                Dependent Concentration Slopes During Recharge
                        99

-------
    100
Ci
<
     10   _
 Cf
 Q
 CQ.
      1   -
                                           O  A-
                                           A  A-9
                                           D  A-10
        .1
             1                10               100

                0      0    K Ap t
                2e T = 2e	 —
                  r      r  n p  H


Figure 5.10  Volumetric Length of the Interface, BH=AV,

             As a Function of Time
                           100

-------
 1.0
 0.1
0.01
10"
10
-4
 10
                 •Immiscible Theory,  Eq.  3.109

                 -Coupled Theory.  Eq.  3.114
             /
     -4
10" J              0.01
    2e T= 2e   K Ap  t
      r     r  n~H
                                                          0.1
      Figure 5.11  Comparison of the Coupled Theory and
                   Experimental Time Dependent Concentration
                   Slopes During Recharge
                         101

-------
 It has been shown that simple direct comparisons between experiment  and
 theory are possible for the recharge cycle with an initially vertical  in-
 terface and that excellent agreement is found.   However, during  storage
 and withdrawal direct comparisons are more difficult because of  the  com-
 plex dependence on conditions in previous cycles of the  recharge-storage-
 withdrawal sequence.   The depth averaged concentration was monitored at
 the various probes and at the well during withdrawal. Visual observa-
 tions proved to be inadequate for these runs, which generally involved a
 low value of e .   In  Figure 5.12 some typical breakthrough curves  at the
 well are shown in dimensionless form where ₯ is the net injected  volume.
 The curves are skewed, and the 50 percent   point occurs  earlier  than
 would be expected with a linear interface in volumetric  coordinates.
 These features may reflect some effects of interface curvature and local
 mixing in the well.

 A series of breakthrough curves for several probe locations  during a
 complete recharge-storage-withdrawal sequence^ (Run B-2)  is shown in  Fig-
 ure 5.13.   The depth  average concentration,  c/c , is plotted in  terms  of
 a dimensionless time  coordinate,  (^n~^*)/^0> where V  is the pore  volume
 between the origin and the probe.   The increasing length of  the  mixed
 zone is obvious for the recharge portion of the sequence, but during the
 withdrawal process this trend is not clear.

 Figures 5.14,  5.15 and 5.16 present a comparison of several  observed re-
 charge and withdrawal breakthrough curves with  theoretical results from
 Section 3  at various  radii.   The theoretical curves in Figures 5.14  and
 5.15 represent recharge with an initially vertical interface.  The in-
 itial period during which the coupled analysis  applies (A/8H>1)  is short
 and was neglected in  the calculation.   Therefore, the theoretical  curve
 represents the immiscible displacement theory with superposed dispersion
 (Eq.  3.111 with 0H given by Eq.  3.107 and v^ given by Eq. 3.65).  There
 is  excellent comparison between the theory and  experiment at Probe 7
 (Figure 5.15),  especially for the  leading  edge  of the curve  (V^
-------
o
U)
        c/c
                -1.0     -0.8     -0.6
0.4      0.6      0.8     1.0
                                  Figure 5.12  Experimental Breakthrough Curves at the Well

-------
c/c
1.0



0.9 h




0.8




0.7




0.6




0.5




0.4




0.3




0.2




0.1
        -0.4
              -0.3

                                                                                  Run  B-2
                                                          // /   	Probe  3  Recharge


                                                                   	 	  Probe  6  Recharge
                                                                             Probe  3 Withdrawal


                                                                             Probe  2 Withdrawal
-0.2
-0.1
0.1
0.2
                                                                                       0.3
                               (V -V.)/V     (V  = rrr  nH,  the pore volume to the probe)
                                 p  *   o      p     p
                                                                                                       0.4
                  Figure 5.13  Experimental Breakthrough Curves During Recharge and Withdrawal

-------
             c/c
                      -.0]   -.«   -.01
                                          Km *-7 rtou 1
                                           • o.jw -* . o.wai
                                              I  .  I
Figure 5.14   Comparison of Experimental and  Theoretical Breakthrough
              Curves During Recharge
             c/c
Figure 5.15   Comparison of Experimental and Theoretical Breakthrough
              Curves During Recharge
             c/c
                       I - 1 - 1 - 1
                              /
                             /
                       I  • -*^ I / I	L.
                                          •HI A-7  F10BI 7


                                         H« t • O.U9 s1 • 0.*IS
                                       1  .  I   1  11.
Figure 5.16   Comparison of Experimental  and Theoretical Breakthrough
              Curves During Withdrawal
                                  105

-------
             1.0
       ₯
                           	THEORETICAL CUKVE NEGLECTING DISPERSION
                                                           EXPERIMENTAL POINTS

                                                            O RECHARGE

                                                            A WITHDRAWAL
                0     r.l    0.2    0.3    0.4     0.5     0.6    0.7    0.8    0.9    1.0
             0,4 .—
0.2 —
 Figure 5.18  At as a Function  of Time During  a Recharge-Storage-
                Withdrawal  Sequence, Run A-7
            0.4
                  n—i—i—i—i—i—i—i—i—r
                                     —I	'	1	1	1	1	i	r

                                  ^0.15 cm    Theoretical
                                       	a, - 0.30 cm    CurV6S
            0.3 —
     -QAt
           0.2  —
           0.1
                                                          EXPERIMENTAL POINTS

                                                            O RECHARGE

                                                            A WITHDRAWAL
               0     0.1     0.2     0.3   0.4     0.5    0.6    0.7    0.8     0.9   1.0
Figure 5.17   At as  a Function of  Time  During a Recharge-Storage-
                Withdrawal  Sequence,  Run  A-2
                                     106

-------
of the dispersed zone.  However,  it is  observed  that  At decreases  system-
atically.  A possible explanation for this  decrease is  the  interaction  of
density difference and grain dispersion.  As shown in Section  3.3,  this
interaction could decrease the rate of  interface tilting,  thus lower At.
This possibility was examined for Run A-2 in Figure 5.17.   The longitu-
dinal dispersivity, a,, was doubled to  a  value of 0.3 cm and the theo-
retical curve recalculated.  For this value of a,  and for the  given
parameters of Run A-2, fj gH remained at a value  greater than one for the
entire recharge, storage and withdrawal process  reflecting  the importance
of longitudinal dispersion.  The theoretical analysis is now based on
the coupled theory (Eq. 3.114) and is shown as a dashed line in Figure
5.17.  It appears that the additional dispersion makes  little  difference
in the general trend of the theory, and that the decrease in  At during
withdrawal must be due to some other factor.

It is of interest to evaluate the theoretical results of the  immiscible
displacement dynamics with and without superimposed grain dynamics.  Fig-
ure 5.18 presents two theoretical curves, one accounting for  the superpo-
sition of dispersion  (solid line) and one neglecting  dispersion (dashed
line) for Run A-7.  In fact, since both predict  the same interface length,
it should be expected that the superposed curve  would give higher values
of At.  In both cases, predicted values of  At during  withdrawal are sub-
stantially larger than those observed.  Other experimental values of At
are tabulated in Table 5.4.

These discrepancies between theoretical and experimental data  during the
withdrawal cycle may be related to several  factors.   However,  it is un-
likely that dispersive effects can account  for the difference, for those
effects were evaluated in Figure 5.17 and were shown  to be relatively
small.  Most likely,  the primary factor is  the increasing importance of
the curvature of the  interface as withdrawal proceeds.   This  is reflected
in Figures 5.17 and 5.18 in which the difference between theory and ex-
periment increases with time.  The potential weakness of the  theory with
a linear interface in volumetric coordinates was identified and discussed
in Section 3.4.

One of the ultimate purposes of this research is to predict the amount of
freshwater  that could be recovered in a  recharge-storage-withdrawal
sequence.  The observed recovery ratios obtained by analyzing the break-
through curves at the well will be compared with the  theoretical results
from Section 3.5.  For an analysis of various breakthrough curves, it
was found that 10% relative concentration at the well was a reasonable
choice for the value  determining the end_of the withdrawal cycle.  The
observed recovery ratio, R=V /₯  , using c/c =0.1 at the x^ell  is shown
in Table 5.5 along with the recovery ratio  predicted by the immiscible
displacement theory (Eq. 3.119, b=0.1).  Comparisons  with the miscible
theories (Sections 3.3 and 3.4) are also included.  These computations
were based on the matching procedure described in Section 3.5.  Also in-
cluded in Table 5.5 are experimental results reported by Kumer  (1968).
The predicted recovery ratios show trends similar to those observed in
this study and by Kumer, but are somewhat lower.  These results indicate
                                 107

-------
                   TABLE 5.4




OBSERVED VALUES OF -QAt/V  DURING WITHDRAWAL
Probe
Number
1
2
3
5
6
7
8
9
10
Run Number
A-l

0.162
0.166


0.199
0.211
0.211

A-2

0.162
0.164
0.180
0.182
0.196


0.216
A- 6

0.693
0.785
0.760





A- 7
0.376
0.300



0.478



B-2
0.557
0.494
0.554






B-3
0.336
0.283
0.310






                      108

-------
                       TABLE 5.5  A COMPARISON OF THEORETICAL RECOVERY RATIOS WITH EXPERIMENTS
Run Number
Present ,
Experiments
A- 7
B-l
B-2
B-3
B-4
Kumar (1968)
Experiments^
2
3
6
e
r

0.248
0.222
0.277
0.158
0.257

0.334
0.459
0.647
e
w

0.179
0.216
0.260
0.153
0.251

0.296
0.309
0.647
Ap/p

0.021
0.018
0.018
0.018
0.018

0.373
0.223
0.092
Qr
(cc/sec)

58.32
61.20
39.24
120.96
44.64

9.98
3.05
0.8
QW
(cc/sec)

112.80
64.80
44.64
129.36
46.80

12.68
6.73
0.8
Average
K
(cm/sec)

0.49
0.48
0.48
0.48
0.47

0.0081
0.00785
0.0099
Recovery Ratio, R1
Observed

0.790
0.795
0.728
0.820
0.482

0.647
0.578
0.085
Immiscible
Theory
(Eq.3.119)

0.748
0.717
0.631
0.762
0.291

0.545
0.468
0.141
Miscible
Theory

0.700
0.680
0.595
0.720
0.265

0.515
0.440
0.132
1                  —                       2                  —
 Recovery based on c/c  = 0.1 at the well,  Recovery based on c/c
0.03 at the well.

-------
that dispersion actually decreases the predicted recovery ratio and does
not account for £he relatively large observed recovery.  It appears that
the interface curvature may be the important effect in these cases.  If
the behavior shown in Figure 3.2 for the recharge also occurs during
withdrawal, the arrival of the wedge at the well would be delayed.  This
phenomenon during withdrawal could substantially increase the recovery
ratio and account for the decreasing values of At observed in Figure 5.17.
                                 110

-------
                          ACKNOWLEDGMENTS

This investigation was conducted at the Ralph M.  Parsons Laboratory
for Water Resources and Hydrodynamics of the Department of Civil
Engineering at Massachusetts Institute of Technology under a grant
from the Office of Research and Monitoring, Environmental Protection
Agency, Research Grant No. 16060ELJ.

The research was supervised by Dr. Lynn W. Gelhar, Associate Professor
of Civil Engineering, and while Dr. Gelhar was on sabbatical leave by
Drs. Donald R. F. Harleman and Keith D. Stolzenbach.

Mr. Roy Milley assisted in the construction of the laboratory models
and Mr. Edward McCaffrey assisted with electronic instrumentation.
Mrs. Pat Swan assisted in editorial work and typed the manuscript.
The use of the M.I.T. Information Processing Center is acknowledged.

The support and assistance of the Office  of Research and Monitoring,
Environmental Protection Agency through Mr. Jack W. Keeley, Project
Officer, is gratefully acknowledged.
                                  Ill

-------
                               REFERENCES


 1.  Bachmat,  Y.  and J.  Bear,  (1964),  "The General Equations of Hydrodynamic
     Dispersion in Homogeneous  Isotropic Porous Mediums", Journal of Geophys-
     ical Research, Vol.  69, No.  12, pp. 2561-7.

 2.  Bear, J.  and 0. Dagan,  (1964), "Moving  Interface  in Coastal Aquifers",
     Journal of the Hydraulics  Division, Proc. ASCE, Vol. 90, HY 4,
     pp. 193-215.

 3.  De Josselin de Jong, G.,  (1960),  "Singularity Distributions for the
     Analysis of Multiple-Fluid Flow Through Porous Media", Journal of
     Geophysical Research, Vol. 65, No. 11,  pp. 3739-58.

 4.  Esmail, 0. J. and 0. K. Kimbler,  (1967),  "Investigation of the Tech-
     nical Feasibility of Storing Fresh Water  in  Saline Aquifers", Water
     Resources Research,  Vol.  3,  No. 3, pp.  683-95.

 5.  Gardner, G.  H., J.  Downie, and H. A. Kendall,  (1962),  "Gravity Segre-
     gation of Miscible Fluids  in Linear Models", Society of Petroleum
     Engineers Journal,  Trans.  AIME, Vol. 225, pp.  95-104.

 6.  Gelhar, L. W.  and M. A.  Collins,  (1971),  "A General Analysis of Longi-
     tudinal Dispersion in Nonuniform  Flow", Water  Resources Research,
     Vol. 7, No.  6, pp.  1511-21.

 7.  Green, D. W. and R.  Cox,  (1968),  "Storage of Fresh Water  in Underground
     Reservoirs Containing Saline Water, Phase II", Kansas  Water Resources
     Institute, Contribution No.  36.

 8.  Guymon, G. L., V. H. Scott and A. Hermann,  (1970),  "A  General  Solution
     of the Two Dimensional  Diffusion-Convection  Equation by  the Finite
     Element Method", Water  Resources  Research, Vol.  6, No. 6,  pp.  1611-17.

 9.  Hamrick, J.  M., (1970), "Density  Induced  Mixing  in Linear Flow Through
     Porous Media", M.I.T. Department  of Civil Eng.,  MS Thesis.

10.  Harleman, D. R. F., P.  F.  Melhorn, and  R.  R. Rumer,  (1963),  "Dispersion
     Permeability Correlation  in  Porous Media", Journal of  the Hydraulics
     Division. Proc. ASCE, Vol.  89, HY 2, pp.  67-85.

11.  Harleman, D. R. F.  and  R.  R. Rumer,  (1963),  "Longitudinal and  Lateral
     Dispersion in an Isotropic Porous Medium", Journal  of  Fluid Mechanics,
     Vol. 16, Part III.

12.  Henry, H. R., (1960), "Salt  Intrusion  Into Coastal Aquifers",  Inter-
     national Association of Sci. Hyd., Pub. No.  52,  pp.  478-87.

                                   113

-------
 13.   Keulegan,  G.  H.,  (1954),  "Ninth Progress Report on Model Laws for
      Density Currents;  An Example  of Density Current Flow in Permeable
      Media", National  Bureau of  Standards Report No. 3411, Washington, B.C.

 14.   Kohout, F.  A.,  (1970),  "Reorientation of Our Saline Water Resources
      Thinking",  Water  Resources  Research, Vol. 6, No. 5, pp. 1442-48.

 15.   Kumer,  Anil,  (1968),  "Dispersion and Gravity Segregation of Miscible
      Fluids  in  Porous Media  for  Stratified Radial Flow Systems", MS Thesis,
      Louisiana  State University, Baton Rouge.

 16.   Kumer,  A.  and 0. K. Kimbler,  (1970), "Effect of Dispersion, Gravity
      Segregation,  and Formation  Stratification on the Recovery of Freshwater
      Stored  in  Saline Aquifers", Water Resources Research, Vol. 6, No. 6,
      pp.  1689-1700.

 17.   Li,  W.  H.  and G. T. Yeh,  (1968), "Dispersion at the Interface of Miscible
      Liquids in  a  Soil", Water Resources Research, Vol. 4, No. 2, pp. 369-77.

 18.   List, E. J. and N. H. Brooks,  (1967), "Lateral Dispersion in Saturated
      Porous  Media", Journal  of Geophysical Research, Vol. 72, No. 10,
      pp.  2531-41.

 19.   Mercado, A.,  (1967),  "The Spreading Pattern of Injected Water in a
      Permeability  Stratified Aquifer", Symposium of Haifa, International
      Assoc.  of Sci. and Hydrology, Pub. No. 72.

 20.   Moulder, E. A., (1970), "Freshwater Bubbles:  A Possibility for Using
      Saline  Aquifers to Store Water", Water Resources Research, Vol. 6,
      No.  5,  pp.  1528-31.

 21.   Perrine, R. L. and G. M. Gay,  (1964), "Radial Motion of a Surface
      Between Fluids Within a Porous Media", Journal of Canadian Petroleum
      Technology, Vol. 3, No. 1, pp. 33-38.

 22.   Reddel, D. L. and D. K. Sunada, (1970), "Numerical Simulation of
      Dispersion in Ground Water Aquifers", Colorado State Univ., Hydrology
      Paper No. 41.

 23.   Rose, D. A. and J. B. Passioura, (1971), "Gravity Segregation During
     Miscible Displacement Experiments", Soil and Science, Vol. 3, No. 4,
     pp.  258-65.

 24.  Rumer, R. R. and D. R.  F.  Harleman, (1963), "intruded Salt-Wedge in
      Coastal Aquifers", Journal of the Hydraulics Division,  Proc. ASCE,
     Vol. 89, HY 6, pp. 193-220.

25.   Schiedegger, A.  E., (1961), "General Theory of Dispersion in Porous
     Media", Journal of Geophysical Research, Vol. 66,  No.  10, pp. 3273-8.

26.  Shamir, V.  Y. and G. Dagan, (1971), "Motion of the Seawater Interface
     in Coastal Aquifers; A Numerical Solution", Water  Resources Research,
     Vol. 7, No. 3, pp. 644-57.
                                    114

-------
27.  Shamir, V.  Y.  and D.  R.  F.  Harleman,  (1967),  "Numerical  Solutions for
     Dispersion in  Porous  Mediums",  Water  Resources  Research,  Vol.  3,  No.  2,
     pp. 557-81.

28.  Theis, C. V.,  (1967), "Aquifers and Models",  Proc.  National Symposium
     on Ground Water Hydrology,  American Water Resources Assoc., pp.  138-48.
                                    115

-------
                         DEFINITION  OF SYMBOLS
           The  primary  symbols  are defined below.  A  few  additional
     special symbols are defined where used.

Symbol              Definition                                 Dimension
A                   Q/27rnH                                      (L2/T)
A                   Q    /2-rrnH                                   (L2/T)
 r,w                 r,w                                       v
b                   n/H  at the well
c                   Concentration of material
c                   Concentration of material in liquid  2
c                   Depth average of concentration
C                   Matching constant between cycles  (n=l,2,.--)
d,-n                 Mean grain diameter                         (L)
D                   KApH/np                                     (L2/T)
                                                                  2
D                   Longitudinal grain dispersion coefficient   (L /T)
                                                                  2
D9                  Lateral grain dispersion coefficient       (L /T)
                                                                  2
Dm                  Molecular  diffusion coefficient             (L /T)
                                                                  2
E                   Density induced dispersion coefficient      (L /T)
                                                                  2
EI                  A  constant dispersion coefficient           (L /T)
—                                                                 2
E                   Average density dispersion coefficient      (L /T)
                                                                  2
E                   Total (gross) dispersion coefficient       (L /T)
f                   Relative interface height  (n/H)             (L/L)
F(z),F(2T)          Dependent  similarity functions  (n/H)
                                                                    2
g                   Gravitational constant                      (L/T )
H                   Aquifer height                              (L)
j(x,t)              Constant of integration
                                                                     2
j                   Mass flux  of material relative  to  mean      (M/TL  )
                     Linear flow
_                                                                    2
J                   Mass flux  of material relative  to  linear   (M/TL  )
                     mean flow, for immiscible displacement
                                                                  2
k                   Intrinsic  permeability                      (L )
K                   Hydraulic  conductivity                      (L/T)
L                   Interface  length for linear  flow           (L)
n                   Porosity
                                                                    2
p                   Pressure                                    (M/T L)
                                  117

-------
 Symbol              Definition                                 Dimension
                                                                  o
 Q                   Recharge (+) or withdrawal (-) rate in     (L /T)
                      radial flow
                                                                  3
 Q ,0                Recharge rate, withdrawal rate magnitude   (L /T)
 r                   Horizontal radial coordinate               (L)
 r.                  Radius to interface                        (L)
 r^                  Radius to convective front                 (L)
 r                   Radius to probe                            (L)
 r                   Well radius                                (L)
  w
 R                   Recovery ratio (=₯.„/₯ )
                                       K  O
 s                   Piezometric difference between two points
 S                   Interface slope for linear flow (=L/H)
 S,-n                  Slo£e_ of concentration breakthrough curve  (T~l)
  5                   at c/c =0.5
                            o
 t                   Time                                       (T)
 t                   Time at the start of a process             (T)
 trQ                  Time from the start of a process till      (T)
                      c/c =0.5 at radius r
                         o
 u                   Horizontal seepage velocity in linear flow (L/T)
 u1 0                 Horizontal seepage velocity in liquid 1,2  (L/T)
  1 sz
 u.                   Velocity of the interface                  (L/T)
 u                   Seepage velocity relative to mean flow     (L/T)
 U                   Mean linear convective seepage velocity    (L/T)
 v                   Vertical seepage velocity in linear flow   (L/T)
 v                   Horizontal radial seepage velocity         (L/T)
 v                   Vertical seepage velocity in radial flow   (L/T)
  y                                               2                i
 V                   Volumetric coordinate (=7rnHr )             (LJ)
 V.                   Volumetric distance to convective front    (L3)
 V                   =  V - VA                                   (L3)
 V.                   Volumetric distance to the interface from  (L )
  1                   V origin
 ₯£                   =  V - Vi                                   (L3)
V                    Net injected Volume at the start of a      (L )
  0                   process
                                                                  o
V                    Volumetric distance to a probe in the      (L )
 P                   model
V                    Volume of liquid recovered                 (L )
 R
                                  118

-------
Symbol              Definition                                 Dimension
x                   Horizontal coordinate in linear flow       (L)
x.                  Linear distance to interface               (L)
x-L                  = x - x±                                   (L)
3c                   = x - Ut                                   (L)
y                   Vertical coordinate                        (L)
z,z'                Independent similarity variables
a                   Longitudinal dispersivity                  (L)
<*„                  Lateral dispersivity                       (L)
3                   Inverse slope of the interface = L/H in    (L2)
                     linear flow, = AV/H in radial flow
                                                                  2
3                   Inverse slope at the start of a process in (L )
                     radial flow
6                   Width of the dispersed zone about the      (L)
                     interface
A                   Volumetric width of the dispersed zone     (L )
A                   Volumetric width of the dispersed zone at  (L3)
                     the start of a process
At                  = 1/S5Q                                    (T)
AV                  Volumetric length of the interface         (L3)
Ap                  Density difference between liquids 1       (M/L  )
                     and 2
AT                  = KApAt/npH
t,                   Distance to interface from top of aquifer  (L)
n                   Distance to interface from bottom of       (L)
                     aquifer
                          1/7                                     1/2
                    = B/V. '  in radial flow                    (L
                         '*
y                   Dynamic viscosity                           (M/LT)
v                   Kinematic viscosity  (=p/p)                  (L2/T)
£                   Transformed  time coordinate
p                   Density                                     (M/L3)
t>l}2                Density of liquid  1,2                       (M/L3)
p                   Depth  average  density                       (M/L3)
T                   Dimensionless  time (=KApt/npH)
e                   Parameter characterizing radial flow (=/D/A)
e                   =/D/A
 r,w                      r,w
<£  „                Piezometric  Head in liquid 1,2             (L)
 •'•> ^
                                   119

-------
                            APPENDICES
                                                           Page No.
A.   Range of Validity of Approximation in Eq.  2.55           120


B.   A Note on Boundary Conditions                            123
C.   An Estimate of Electronic Dispersion for the Averaging
     Probes                                                   126
                                121

-------
                             Appendix A
          Range of Validity of Approximation in Eg. (2.55)

 In Section 2.3 it was assumed that the interface would remain linear
 (Eq. 2.38), and that the net effect of grain dispersion would be to
 decrease the rate of interfacial tilting.  Thereafter, the primary
 approximation used to develop the coupled solution involved the appli-
 cation of 3u/3y (Eq. 2.53) at the mean interfacial position in order
 to determine 3u./9y (Eq. 2.56).  The approximation
                            _3u
                            3y
                               x=x.
                      3u.
                     	i
                      3y
                                (A.I)
 was introduced in Eq.  (2.55).   The following analysis indicates when
 this approximation is  valid.

 If  the interface x.(y,t)  is of a linear form as given in Eq.  (2.38)
 with 3 unspecified,  the analysis of Section 2.2 applies and the density
 gradient  can be obtained from Eqs.  (2.21)  and (2.48) as given in
 Eq.  (2.58).   Thus Eq.  (2.53)  can be written
  _
  3y
                        K
                        n p

                                                  (A.2)
 In  this equation  x.  =  x.(y,t)  is  the only function of y,  and the equation
 has  the integrated  form
            u = —
K Ap 1
n p  23
                          erf
                              x-Ut
- erf
x-x.
	1

/4T
(A.3)
where j(x,t)  is  a  constant  of  integration.

From the continuity  equation  (Eq.  2.51)

                                f y       a.
                          v =  -
                            j
                         •r— dy
                         3x  J
                                  y=-H/2
which satisfies the condition v =  0  at  y  =  -H/2.   The  requirement  that
v = 0 at y = H/2 implies  that
                         3
                         3x
                              .H/2
                              -H/2
                   u dy
        = 0
                                122

-------
Expressing the velocity in terms of the convective coordinate  system,
                            u = u - U
this flux condition becomes
                             l-H/2
                            '-H/2
                                  u dy = 0
This condition is used  to evaluate  j(x,t), with  the  result
           = K Ap  1

         U   n p   28
H
                           +H/2    x-x.
                          -H/2
      i .           X-X.


erf 	 dy - erf 	
                                           (A.4)
Along the interface  (x=x.),  the_interfacial velocity  relative to the

convective front  is  defined  at u.,
                        u. = u. -  U  H  u
                         i     i
                                       x=x.
                                           (A.5)
which is  combined with  Eq.  (A.4)  to  yield
_
u . —
1
K Ap
n p
1
2BH
r+H/2
J-H/2
s±
erf —
A
:r dy
£
                                              —ix=x.
                                                                  (A.6)
Equation  (A.6)  is  evaluated  using
                                        -z
                  erf  z  dz  =  z  erf  z  +
                      const
and can be expressed as
            p   (-282H)
                                                                  (A. 7)
where
                              -  B(y-H/2)
                              -  B(y+H/2)
                                                                  (A. 8)
                                123

-------
 Differentiating this equation with respect to y yields


                   ^i   K Ap -1
                                                                  (A.9)
 The approximation in Eq. (2.55)  involves replacing quantity
                         jhi
                         3y
                            x=x.
                               i
     = K A£
       n p
                                 (A.10)
 from Eqs.  (2.54)  and (2.60)  by the expression in Eq.  (A.9).  This will
 be  valid when
                                                                  (A.11)
 in which  case,  using  erfy .  = — y.,
                      3y
n p
/47T£
       8U
       3y
                                                                  (A.12)
                                           x=x.
                                              i
The maximum absolute values  of y   and p_  occur at y = ±H/2 and thus the
restrictions on y.. ,y»  are  seen to  be  equivalent to the requirement that
                            /4T
                                      «  1
Since the approximate width of  the  dispersed  zone  about the interface,
5, is given by  ATT £  (Eq.  2.47)  the restriction implied by the approxi-
mation in Eq. (2.55) is L  « 6.   Therefore, the result obtained in Sec-
tion 2.3 is expected to be valid when  the  dispersed zone is wide in
comparison to the wedge length.
                                124

-------
                            Appendix B

                   A Note on Boundary Conditions

For a confined aquifer of constant thickness,  the boundary condition
to be satisfied by the solution of the convective dispersion equation
at the upper and lower boundaries is


                         — = 0,   y = ± H/2                     (B.I)

The approximate solution to the convective dispersion equation for the
appropriate boundary conditions considered in  Sections 2.2 and 2.3 was
as given by Eqs. (2.44) or (2.58).  For this solution

                  „             R        (x-x )

                                              ->                (B'2)
where x-x. = x-gy.  At the lower boundary, y = -H/2, and Eq.  (B.2)
becomes


The consequence of not satisfying the boundary conditions is the crea-
tion of an artificial mass flux through the aquifer in the vertical
direction.  At y = -H/2 this flux is proportional to Eq. (B.3).

The possible local effects of not satisfying this boundary condition
can be evaluated by comparing the magnitude of the ficticious vertical
flux to the horizontal dispersive mass flux at a given point.  The
ratio of these two fluxes is, from Eqs. (2.58) and (B.3),



                       D2 If /D1 If = 6VD1                       

Thus, it is seen that, even though the ratio D/D.. is generally small
   "-*- or less), the implied vertical flux of mass at the boundary will
eventually become equal to or larger than the horizontal flux because
3 increases with time as indicated in Eqs.  (2.63) and  (2.68).  The im-
plication is that the effects of the no-flux condition at the boundaries
may become important when 3 > D-,/!^-

Another method of estimating the possible effects of the boundary con-
ditions is to compare the horizontal and vertical components of  the
total mass transport through the aquifer.   Hamrick  (1970) computed the
artificial vertical flux and compared it with the total horizontal flux
                                 125

-------
 evaluated at its maximum, x = 0.  The total vertical dispersive mass
 flux is given by
                                       dx =
                                              (B.5)
 and the maximum horizontal masj^ flux relative to the mean motion is,
 assuming p to be constant, at x = 0,
 r+H/2

 '-H/2
                           8x
          fH/2
_  dy + p      (cu)
x=0      '-H/2
                                                                  (B.6)
o 1
                             /7rL,    3 E  A/L
                             — )  +-_*(-
where  the  integral
                                 r
                               erf
                                               TJ =
                                                   H
                                                   Lw/rT
                                             (B.7)
may be  evaluated numerically.   However, for small •  .  ,  the error function
can be  approximated  by  the first term of its series expansion and $ is
given approximately  as
                                                                  (B.8)
At its maximum, within  the  aquifer
                                    26
Under these conditions  the  error  function in Eq.  (B.6)  is  approximated
by erf(/TrL/2<5 ^ L/6.  Thus,  Eq.  (B.8)  is  reasonably accurate for the
range of 6/L for which  the  interaction solution  is  applicable (6/L»l).
As a consequence J   can be  approximated  by
                  X

                      T  -   PCoDlrL    1  E  ,1.2.
                      Jx ~     3   16 +  2  D(T)  J
                                             (B.9)
The ratio of the absolute values of the  total mass  flux  is  given  by
                                  126

-------
                  J
                  JL
                  J
                   x
2 6
                                        (B.10)
Again, it is seen that the implied ratio of vertical to horizontal mass
transport will eventually become large,  in this case because both 6/L
and 3 increase with time.  Thus, from consideration of both local fluxes
and total mass transport, there is the implication that the boundary
conditions may, for sufficiently large times, influence interfacial
tilting and overall mixing for the coupled analysis (Section 2.3).
                                  127

-------
                             Appendix C

    An Estimate of Electronic Dispersion for the Averaging Probes

 The following analysis is based on the solution for the concentration in
 a uniform linear flow with no density difference between the two fluids,
     ,     1,1-,X-Ut
   c/c  = -=- + -=• erf (	
      o   2   z      /rtr—
                                                                   (C-l)
 in which x is the longitudinal coordinate, u is the seepage velocity  and
 D  is the longitudinal dispersion coefficient.  It will be assumed  that
 a probe at x = x1 measures the average concentration in a region x.>a
 around the probe because of the generated electronic field.  The indi-
 cated average concentration, c, from the probe readout becomes
                            =x +a
- =     (l/2a)(l+erf
                                        '4D t
                                               dx
                                            (C-2)
 and after integration
                        4a
                               x+a-ut
                         ,+a-ut
                                       erf
                                 x..— a-ut     x..-a-ut
                                 —	 erf —	
              /4D t
                                               4D t
                                            2               2
                                  -(x +a-ut)      -(x -a-ut)
                                      4D-t
                                 VTT
                                  4D,t
                             e      1
                                                                   (C-3)
From Eq.  (C-3)  we  find that
                        „ c
               (At)
                   -1
                         r-
                = -  - erf
                              X;L=ut
                                            (C-4)
and as a-K), Eq.  (C-4)  reduces  to
                   (At)
     9 —
-1     Co
      3t
                                              u
                                              (C-5)
                                 128

-------
xjhich corresponds to a point measurement with no electronic dispersion.
The effects of electronic dispersion can be assessed by considering the
ratio of the indicated and actual rates of change of concentration (Eqs.
(C-3) and (C-4))
                      At/At = (6/2a)erf

where 6 is the dispersed zone thickness defined by
                       (C-6)
                          i£
                          3x
,-1/2
                             x=ut
Figure C-l shows this ratio as a function of a/5.   It can be seen that
when a/6 < 0.2, the indicated and average rates of change differ by
less than 5%.  The possible error in At is of importance because the
values of At from experiments are used to determine dispersion
coefficients.

This analysis can be adapted to the observations in radial flow if we
determine 6 during recharge from Eq. (5.3)
                                              -1/2
and estimate that 2a is equal to the spacing of the probe electrodes
(a=3/8 in.).  Using these relationships and the estimated dispersivity
from Section 5.4  (a =0.15 cm), the radius rA corresponding to a given
value of a/6 can be determined as shown in Figure C-l.  Probe locations
are also shown.

From Figure C-l it can be seen that during recharge the distortion due
to electronic dispersion will be significant for only the first few
probes near the wall.  For tests with a density difference or measure-
ments during withdrawal, this distortion will be unimportant because the
thickness of the dispersed zone will be larger under these conditions.
                                 129

-------
       1.0
       0.8 -
       0.6 -
At/At
                 Equation  C-6
       0.4
       0.2
c
o
•H
•u  o\
ra  =>fc
o
o o
                          0)

                          o
                                  =«=
                                    =«=
                                   J.
                                                _L
                      0.1
          0.2
0.5
0.6
                      0.3         0.4


                           a/6



Figure C-l   Estimate of Electronic Distortion During Recharge
                                                                                               0.7

-------
  SELECTED WATER
  RESOURCES ABSTRACTS
  INPUT TRANSACTION FORM
                     /. Report No.
              2.
              DENSITY INDUCED MIXING IN CONFINED AQUIFERS
  7.  Author(s)   Gelharj  L} wilson, J.L., Miller, J.S., and
               Hamrick, J.M.
  9. Organization
              Massachusetts Institute of Technology, Cambridge,
              Department of Civil Engineering
3.  Accession No.

w

5.  Report Date
6.
8.  Performing Organization
   Report No.
10.  Project No.
   EPA.  WQO.  16060ELJ
  12.  Sponsoring Organization

  15.  Supplementary Notes
                                                                  11. Contract/Grant No.
                                                                   13.  Type of Report and
                                                                      Period Covered
                      Final Report, EPA, WQO, Project  #16060ELJ, 1972.
                      137 p, 36 fig, 7 tab, 28 ref,  3  append
  16. Abstract .
   Analytical  techniques are developed to describe the mixing between two fluids of  dif-
   ferent density in a confined aquifer, in which one fluid  is introduced to the aquifer
   by well recharge.   The immiscible displacement process  in both linear and radial  flows
   is analyzed and the effects of longitudinal and lateral dispersion are included using
   a boundary  layer approximation.  The theoretical results  demonstrate the effect of  hy-
   drodynamic  dispersion in retarding gravity segregation  due to density differences.  The
   theoretical results are compared with observations of aquifer mixing in linear and  ra-
   dial flow laboratory models.  During recharge excellent agreement between the theore-
   tical predictions and experimental results is found and the predicted retarding effect
   of longitudinal dispersion are verified.  During withdrawal some systematic  difference
   between the theory and observation are noted.  Theoretical predictions of recovery  ef-
   ficiency during a recharge-storage-withdrawal sequence  show trends similar to those ob
   served but  are typically 5 to 10% lower than those observed.  Direct theoretical  pre-
   dictions of recovery efficiency during single or multiple sequences of recharge-
   storage-withdrawal are developed for an immiscible system, and similar developments
   outlined for miscible displacement.  The results are  suggested for use in developing
   preliminary estimates of operating conditions in the  field.  This report was submitted
   in fulfillment of Project Number 16060ELJ under the sponsorship of the Water Quality
   Office, Environmental Protection Agency.  (Gelhar, MIT)	
  17a. Descriptors
         Groundwater Recharge, Water Storage,  Dispersion,   Density Currents,
        Aquifer,  Water Quality, Recharge Wells, Underground Storage
                                                                                 *Confined
  17b. Identifiers
  17C.COWRR Field & Group  Q4B  05E  02F
  18. Availability
19. Security Class.
   (Report)
                          20. Security Class.
21. No. of
   Pages

22. Price
                                                       Send To:
                             WATER RESOURCES SCIENTIFIC INFORMATION CENTER
                             U.S. DEPARTMENT OF THE INTERIOR
                             WASHINGTON. D. C. 20240
  Abstractor  Lynn W.  Gelhar
             ^Institution  Massachusetts Institute of Technology
WRSIC 102 (REV. JUNE 1971)
                                                                                  6PO 9 J 3.2«f

-------