A REVIEW OF PHYSICAL AND CHEMICAL PROCESSES PERTAINING TO SOLUTE TRANSPORT Feike J. Leij and J.H. Dane Graduate Student and Professor of Agronomy and Soils Alabama Agricultural Experiment Station Auburn University Auburn University, Alabama Lowell T. Frobish, Director ii CONTENTS page I. INTRODUCTION........................ . II. FORMULATION OF THE TRANSPORT EQUATION............3 Non-reactive Solutes....................3 Reactive Solutes......................6 Solutions of the Transport Equation. .......... 9 Concentrations.................... 10 Boundary and Initial Conditions.............1 Analytical Solutions................. . 13 Non-equilibrium Conditions ...... .......... 18 Physical Non-equilibrium........................ 18 Chemical Non-equilibrium.................19 Apparent Non-equilibrium................. 22 III. DISPERSION AND DIFFUSION... ......................... 26 Molecular Diffusion................... 27 Dispersion........................................ 34 The Concept of Dispersion in Porous Media. ....... 40 Physical Non-equilibrium and Dispersion. ........ 44 Experimental Determination of Dispersion Coefficients . 50 IV. ION EXCHANGE. ........................ 62 Equilibrium Chromatography................ 62 The Combined Effect of Ion Exchange and Dispersion 65 Numerical Solutions of Transport Involving Equilibrium Exchange................. ................. ........ 69 V. TRANSPORT IN STRUCTURED AND LAYERED SOILS....... 78 Mobile and Stagnant Regions. ............... 79 Aggregated Media..................... 82 Fractured Media........................90 Layered Media.........................94 iii VI. PREDICTION OF SOLUTE TRANSPORT UNDER FIELD CONDITIONS Deterministic Models Stochastic Models ......... Mechanistic Models .......... Non-mechanistic Models Concluding Remarks VII. TRANSPORT DURING UNSATURATED FLOW CONDITIONS Steady Flow Transient Flow Experimental Considerations . . VIII. MULTI-PHASE TRANSPORT . .... .. Transport in Vapor and Liquid Phase . . Transport in Immiscible Liquid Phases . IX. EPILOGUE . .. . . .... . NOTATION . . . . . .. . . . . . . . . . BIBLIOGRAPHY .. 0........... S. . . 103 .. . 105 . . 105 .. . . 110 .. . 116 .. . 117 .. . 118 . . . . 120 . . . 122 . .. . . . . 126 . . .. . . . . 126 . ... . . . 129 S .. . . . . 131 . . .. . . 134 . . .. . . 140 FIRST PRINTING, APRIL 1989. Information contained herein is available to all without regard to race, color, sex, or national origin. iv a . 102 1 I. INTRODUCTION Solute transport in soils has received considerable attention during the past decades. Agriculturalists are interested in the behavior and effectiveness of applied fertilizers and pesticides. Leaching of salts (solutes) through the soil profile, which depends on exchange, exclusion, dissolution and precipitation, volatilization, chemical and biological transformations, and other processes, affects both the chemical and physical condition of a soil. Other areas of interest are miscible displacement, such as the mixing of fresh ground water with salt sea water and the secondary recovery of oil. The main reason for the increased interest in the transport of solutes is the concern about contamination of the soil, in particular the movement of contaminants to the ground water. A wide variety of substances is involved, including agricultural chemicals, industrial compounds, radio-active materials, and domestic waste. Solute transport through porous media has been studied by workers from many disciplines: chemical, civil, and petroleum engineering; hydrology, hydrogeology, and geochemistry; and soil science and agronomy. Because of this diversity, research areas include mathematical approaches to solve flow and transport equations, experimental work concerning homogeneous media under laboratory conditions, and field experiments with inherent heterogeneity. Laboratory work has focused on exchange chromatography, the influence 2 Of biological, hydrodynamic, and geochemical processes, and the transport during unsaturated flow conditions. This literature review covers, in very general terms, what has been done to date in the area of solute transport in soils. In particular, it examines both chemical and physical processes and explores transport during saturated as well as. unsaturated flow conditions. II. FORMULATION OF THE TRANSPORT EQUATION Non-reactive Solutes As a starting point we will consider the transport of a chemical species in an isotropic porous medium, which is homogeneous with respect to the relevant transport and flow parameters. The porous medium consists of a rigid solid phase, of which the pores are filled with liquid. It is assumed that the species does not react or interact with the medium and that the solute is completely miscible with the solvent. At the macroscopic level, the equation of conservation of mass, assuming no source/sink terms, leads to: a- -V* JD + = -V*J (2-1) at D V3 - where 0 is the volumetric water content [L3L-3I, C is the concentration of the solute expressed in mass of solute per volume solution phase -3 [ML ], t is time [T], JD is the diffusive-dispersive mass flux -2-1 3 -2-1 [ML T ], J is the volumetric flux of the carrier [LL T ], J is V s the total solute mass flux [ML-2T - 1] and V* denotes the divergence -1 [L I. The autonomous flux J is generally expressed as a Fick-type equation: J = -OD (2-2) D 8x where D is the coefficient of hydrodynamic dispersion [L2T - 1] and x is distance [L]. D represents the random effects due to molecular 3 diffusion and mechanical dispersion, which occur during transport or flow. Many references exist with regard to this topic (15, 53). Although the mechanisms behind these two processes are quite different, they behave similarly and Fick's law for diffusion can be used to describe both effects. Therefore, the coefficient of mechanical 2 -1 dispersion D [L2T-], and the effective coefficient of molecular dis 2-1 diffusion, D [L2T-], are added to give an effective coefficient of e dispersion: D=Ddis + D (2-3) d e D depends, among other things, on water content and pore-water velocity. Because of the intimate relation between microscopic variation of the water velocity and solute spreading, D is also called coefficient of hydrodynamic dispersion (15). One might argue that this is somewhat ambiguous because only Ddi depends on the pore-water velocity. The use of one coefficient is especially convenient when solving the transport equation analytically. Sometimes a third mechanism, which contributes to dispersion, is included; the diffusive transfer of solute between mobile and immobile regions of the liquid phase (135). Substituting Eq.(2-2) into Eq.(2-1) and using the relation between the Darcy flux and the pore-water velocity (Jv=vO), we obtain for one-dimensional transport: aOC_ at X Ov Oacn -- vOC] ( 24 ) -1. where v is the average pore-water velocity [LT ]. Assuming v and 0 and 5 to be uniform with respect to x and assuming steady-state flow, Eq.(2-4) leads to: ac a 2 c ac at - D 2 v (2-5) 8x which is a special form of the Fokker-Planck equation for one variable, namely with v and D independent of x and t (157). It is a linear second-order partial differential equation of parabolic type, which will be referred to as the advection-dispersion equation (ADE); it is also known as convection-dispersion or convection-diffusion equation. It has been widely used in miscible displacement studies (e.g., 121). For multidimensional, anisotropic flow, Eq.(2-S) needs to be rewritten (cf. 165) as: av.C aC _a D aC 1(vc = D 1(2-6)at ax. ij ax. ax. 1 1 where D.. is the dispersion tensor and i,j are indices denoting ij directions. For laboratory conditions the ADE is very useful for modeling solute movement. The equation can be adjusted to include phenomena such as cation-exchange (107), adsorption-desorption of pesticides (198), anion exclusion (29), precipitation-dissolution (159), nutrient uptake by plants (47), microbial induced transformations (41) and radioactive decay (154). For field conditions, Eq.(2-5) is less useful to describe transport, because most soils are macroscopically heterogeneous due to factors such as structural development, heterogeneity of the soil profile, swelling and shrinking, and biological activity. Reactive Solutes Solutes and porous media often interact with each other. Many soils, due to the presence of negatively charged clay minerals, variably charged organic matter, hydrous oxides and mineral edges, exhibit adsorption and exclusion phenomena. These generally cause the solute and the carrier to travel at a different speed. It is assumed that the chemical species is present only in the liquid and the sorbed phase, and that there is only one liquid phase. The process of adsorption/desorption can be included into Eq.(2-1) to yield: a .C + = -V* JD (2-7) where q is the concentration in the sorbed phase expressed in mass of -3 solute per volume of soil [ML -3 ] . Sorbed refers to adsorption and precipitation, although the latter process will be ignored. Actually q can be regarded as a source/sink term. It should be noted that, under unsaturated flow conditions, the vapor phase needs to be considered if volatile solutes are present. For one-dimensional transport, Eq. (2-7) leads to: aa c 1 avec Cat + q xDe x ax(2 If an exchange isotherm is known, which relates the concentration of the species in the adsorbed phase to the concentration of all species in the liquid phase, the dependent variable q can be expressed in terms of C. However, this is only justified in case of a single valued relationship between q and C, Implying that the exchange reaction reaches instantaneous equilibrium and that no hysteresis exists. The importance of hysteresis in the q(C) relationship was illustrated by van Genuchten and Cleary (198). In case of a binary system, S k ) the adsorbed mass of species k per mass of soil [MM-], can be expressed as: S k = f(CT C k ) (2-9) where C T is the total electrolyte concentration and C k is the -3 concentration of species k in solution [ML - 1; both C and Ckare];bt T an k e expressed per volume of solution phase. To express the adsorbed concentration on the basis of volume of soil, the following relationship is used: q = SkPb (2-10) -3 where pb is the soil bulk density [ML . For conditions similar to those stated for the derivation of Eq. (2-5), Eq. (2-8) can be written as: 8C Pb80S 82C - -D -v (2-11) at o at 2ax 8x where C and S refer to the concentrations of solute species k in a binary system. Applying the chain rule and assuming a constant total electrolyte level with respect to time, i.e., Sk=f(Ck) yields: k k' 8S _dS8C (-2 where dS/dC is the slope of the exchange isotherm, which is constant in case of linear adsorption or exchange and which is usually referred to 3 -1 as the distribution coefficient K d [L3M-!. This quantity indicates how the species is distributed between the liquid and adsorbed phase: S = KC (2-13) d Using Eq. (2-12), Eq.(2-11) can be rewritten in terms of one dependent variable: ac a 2 C ac2-14) Ra= D 2 V-x(2 ax where the dimensionless retardation factor R is defined as follows: Pbd R = 1 + dS (2-15) 0wdC Depending on the assumed adsorption mechanism, different transport equations were given by Gupta and Greenkorn (75): 1. Linear adsorption isotherm: q= k C 1 k 2 1ac a C ac 1+ o a-=D 2-V j(2-16) at - x2 a -x 2. Freundlich adsorption isotherm: qk 2 Cn ,-nk 2 2 [1 n-1a ]ac a c ac I1+0(1+bC) 2J at = ax__ -2 (218 In the above three examples k, n, a and b are empirical constants. Dividing Eq. (2-14) by the retardation factor leads again to the ADE, as for the non- reactive case: t D 2 -(2-19)8t 82 Ox ax with D =D/R and v =v/R. In case of non-linear exchange (R=f(C.)), 1 Eq. (2-19) is a non-linear partial differential equation for which analytical solutions are difficult to obtain. A more general equation should include sink/source terms to account for phenomena such as radioactive decay, precipitation, dissolution and chemical reactions. Parker and van Genuchten (130) presented the following, more general expression: P sAPb sPb b as + ac Da2 C C S Ssb(2-20) e at at O - O w w + 0 where j and W are rate constants for first-order decay in liquid and w s -1 -3-1 - solid phase [T - 1, respectively, and 7w [ML-T-] and [T are rate constants for zero-order production in liquid and solid phase, respectively. Solutions of the Transport Equation Before presenting some analytical solutions for Eq.(2-19), it is worthwhile to mention the different types of concentrations that can be considered and to pay attention to the appropriate boundary and initial conditions (131, 200). 10 Concentrat ions The time average of C is defined (60) as: At o 2 '(x yZt ) = At C(x,y,z,t) dt (2-21) At t -- o 2 A spatial or volume average of c, a microscopic variable in this case, at a position (xoYozo) is defined as T cdV C (x yo ,z t) - lim AVet(2-22) v o o o AV->V fdV fAVe where the volume AV and AV represent a "large" chunk of the porous medium and the liquid phase, respectively, and dV and 6V correspond to a microscopic differential volume element and the Representative 3 Elementary Volume (16), respectively. All volumes are expressed in L Furthermore, x, y and z are fixed positions (the center of a 0 0 0 physical or material point). It should be noted that the macroscopic concentration, C, is generally taken as the volume-averaged concentration, C V. At a certain position, the flux-averaged concentration, Cf, is given as the ratio of the transport and flow term: C (t) = Js/Jv (2-23) f s V Cf(t) represents the mass of solute per unit volume of fluid passing through a given cross section during an arbitrary time interval (102). As was stated by Parker and van Genuchten (131), solute flux 11 distributions are in many cases of more interest than pore fluid concentrations. The relationship between flux-averaged concentration, Cf, and volume-averaged concentration, C V , the two types of concentrations pertinent to many displacement experiments, can be derived from Eq.(2-23): aC Cf = C D V (2-24) f V v8x Parker and van Genuchten (131) discussed the importance of distinguishing between both concentration modes in conjunction with the way an experiment is performed and the results are analyzed (cf. 34, 102). Unless stated otherwise, we will assume, as was done previously, that all concentrations are volume averaged, i.e., C=CV V. Boundary and Initial Conditions In order to solve Eq.(2-S), or for that matter (2-19), specific boundary and initial conditions are needed. Following van Genuchten and Alves (197), the initial condition is: C(x,t) = f(x) x>O , t=0 (2-25) where f(x) is an arbitrary function. For the boundary condition at x=O, the first- or concentration- type boundary condition, the Dirichlet problem, is given by: C(x,t) = g(t) x 4 O , t>O (2-26) and the third- or flux-type boundary condition is given by: BC (-D -- + vC) = vg(t) t>O (2-27) xx8O 12 where g(t) is an arbitrary function describing the concentration of the incoming solution. The exit boundary condition can be defined for a semi-infinite or for a finite soil column. For a semi-infinite column the appropriate boundary condition is: ac (x t) = 0 x-) P, t>O (2-28) ax In case of a finite system the following continuity condition at the exit boundary needs to be satisfied: (-D a- + vC) =vC t>O (2-29) xtL ex where C is the exit concentration, which is assumed to be equal to ex CIxtL Hence: ac(x, t) = 0 xtL, t>0 (2-30) ax Analytical solutions of Eq. (2-5) for conditions (2-28) and (2-30) are approximately equal. For the inlet boundary, a third-type condition is generally preferred, because the mathematical solution for this condition leads to conservation of mass. Eq. (2-26) implies equal concentrations in the feed solution and in the porous medium at the boundary. However, in reality it takes some time to attain the same concentration. First, the input solution might not be well mixed resulting in a boundary layer outside the porous medium (24). More important, because at the microscopic level the concentration varies gradually, a transition zone should be taken into account. Obviously a volume averaged concentration of a representative elementary volume (REV) will not immediately be equal to the concentration of the feed solution. In other words, since 13 a (tracer) solution can only be injected at a certain rate, a prescribed concentration will not be established instantaneously. For displacement experiments, involving a step change in concentration (g(tO)=C ), a third-type condition should therefore be 0 used: aC (-D -+ vC) =vC (2-31) 8x o x 0 where C is the concentration of the feed solution. The discontinuity 0 in the concentration across the inlet boundary increases with increasing D/v. Van Genuchten and Parker (200) showed that solutions of the ADE, subject to a first-type boundary condition, lead to flux average concentrations, whereas solutions for a third-type boundary condition lead to volume average concentrations. One is also referred to the work by Parlange and Starr (133, 134), Parlange et al. (132), and Pandey and Gupta (128). Analytical Solutions In case of an infinite system (-oO 0 14 need to be transformed as well. Note that C. is the initial 1 concentration in the soil. The solution of Eq. (2-33) can be found in the areas of heat flow (37) and diffusion (45). Using an alternative transformation: x-vt=(2-35) f4 Dt results in an ordinary differential equation: 2 dC+ 2 dC=0 (2-36) d 2 d= With the transformed initial and boundary concentrations described by: c = c. -o 1 (2-37) C =C 0 -00 the solution of Eq. (2-36) is: C = (C-C.)/(C -C ) = erfc(2-38) 1 0 1 where C is the dimensionless concentration and erfc is the complementary error function. For other boundary conditions, involving semi-infinite or finite media, a coordinate transformation cannot be employed or is inconvenient. Carslaw and Jaeger (37, p. 388) showed how Laplace transforms can be used to solve Eq.(2-5). Van Genuchten and Alves (197) provided a compendium of available analytical solutions to Eq. (2-14), many of them obtained with the help of Laplace transforms. The solutions for four combinations of inlet and exit conditions are listed in table 1. 15 Table 1. Analytical Solutions of Eq.(2-14) for Various Boundary Conditions with Constant R, after van Genuchten and Wierenga (204) I nlet boundary condition C(0. t) = C 1)0,ac- tC 0 ax X 0 = )C 0 Analytical solution crfc 2(L-flt' Ir? KexpAt7I erfc ~ Lu;) t' r%)ep (~ut)'j 1 1 + ? + u't ep( erfc Re t 2 DR11? D J1 2(DRt)'J ti x FIX Wt 1Dt1 213o, s1n (LCxp12 4R L'-RJ Fri 02+(2L +v ;; [cs(lLx) + I;s(azx]ex ty x ' C - I~_T~ -U 02 + 21 D m 21 Om cot (03 +) + - 0 p, 4!) References: A-i: (109), A-2: (113), A-3- (42), A-4: (27) By using the following dimensionless variables T =vt /L Z =x/L P =vL/D C=(C-C. )/(C 0-C.) Eq. (2-14) can be rewritten as 2- R a8- 1aC ac, = P 2 32 (2-39) (2-40) (2-41) (2-42) (2-43) where T is the number of pore volumes, L is the column length [L], Z is the dimensionless distance and P is usually referred to as the Peclet number. Analytical expressions for C =C the dimensionless exit Z=1 ve concentration, are provided in table 2 for different boundary Case Al1 A-2 A-3 A-4 (3ctI3, L 0, 2o1),)4 I \ 16 Table 2. Expressions for the Relative Effluent Concentration, C (T), in e Terms of the Column Peclet Number (P) and Pore Volume (T) for the Four Analytical Solutions Listed in table 1, after van Genuchten and Wierenga (204) Ca e Relative effluent concentration A-I ce(T) erfc )2 (R -T) + exp(P)erfc 2((R T A-2 Ce(T) = erfc [(l (R - T) + ( exp - R T) + P erfe (R T) P P7T 02 T S2/msin(3,) exp 2 - - 4{ PR P A-3 c,(T) = 1 - _ P1 4m cot(3m) + -= 0 02 + P+ - 0"22,sin(I) exp2 P2 2 t in/ 12ep -- H PR PO'm cot(fOm) -t3 + -- 0 A-4 c,(T) = 1 - S_2 4R Pm cot(m) 4 - 0 M=/3 2 - - ? I) 4 conditions. The influence of the different boundary conditions on the concentration profiles is graphically illustrated in figure 1. At small times, the use of the first-type inlet condition results in a considerable error. A difference in the solutions for the semi-infinite and the finite cases occurs once the solute concentration starts to increase at the exit. Just as for the inlet, the concentration at the exit is not likely to be continuous. Unless backmixing occurs from the effluent solution to the soil, the semi-infinite solution is to be preferred. These solutions are useful to predict the solute concentration in homogeneous media and to determine transport parameters in conjunction with an experimentally determined concentration distribution. It should be noted that much more work on the analytical solution of the ADE has been published (e.g., 65, 112, 127). 17 1.0 0.8 z Z H z Z o LUJ z 0 u 0.6 0.4 0.2 REDUCED DISTANCE, Z u z 0 I.- z w z 0 C 1.0 REDUCED DISTANCE, Z FIG. 1. Calculated concentration distribution for R=1 and P-values of 5 and 20, respectively. The curves were obtained with the analytical solutions listed in table 2, after van Genuchten and Alves (197). 18 Non-equilibrium Conditions So far, it has been assumed that instantaneous local equilibrium exists between the solute in the liquid and the adsorbed phase, the distribution of ions in both phases being determined by the exchange isotherm. However, instantaneous equilibrium might not always be achieved. Two different kinds of non-equilibrium exist: physical and chemical. Physical Non-equilibrium Physical non-equilibrium is of particular importance for aggregated media, where flow occurs predominantly between aggregates, and for unsaturated media, where the liquid phase in effect might be discontinuous. During miscible displacement the solute in the feed solution cannot reach all sorption sites immediately and no instantaneous exchange equilibrium can be achieved. This situation is generally classified as physical non-equilibrium. Conceptually, advective flow occurs only in the so called mobile region of the fluid phase. The concentration in the immobile region, both in the liquid and the adsorbed phase, will therefore lag behind the concentration in the mobile region. A large part of the sorption sites are usually only accessible via the immobile or stagnant region of the liquid. Transfer of solute from the mobile to the immobile region, and vice versa, occurs via a diffusion controlled process. For physical non-equilIibr i um, the transport equat ion was reformulated by van Genuchten and Wierenga (202) following Coats and Smith (43): 19 ac ac. as as. a 2 c ac mo im mo im mo mo S+ +p. f - (1-f) = D---- v (2-44) mo at im at b at +Pb t mo 2 mo mo ax where the medium is supposed to be homogeneous with respect to 8 , 8. mo im and v and the flow is steady. The fraction of sorption sites in mo direct contact with the mobile region of the liquid is given by f. Transfer of the species between mobile (mo) and immobile (im) regions of the liquid phase was given by: acm asm im + ( (C -C.(2-45) im at b( at mo im -1 where a is the mass transfer coefficient [T-1 Anion exclusion (29, 103) can be viewed as a situation of physical non-equilibrium, where the exclusion volume roughly corresponds to the immobile region. Eq. (2-44) and (2-45) can be readily adapted (207) to describe the transport of anions which are excluded from certain regions of the liquid phase. A further discussion on the concept of mobile and immobile regions of the liquid phase will be given in the section on mobile and stagnant regions in Chapter V. Chemical Non-equilibrium Chemical non-equilibrium occurs when the adsorption/exchange process requires some time to be completed. No instantaneous equilibrium between the solute concentration in the liquid and the adsorbed phase will occur. To account for chemical non-equilibrium exchange, a kinetic approach can be taken by combining the transport equation with the appropriate rate equation for adsorption of the species by the medium (7, 8, 109). 20 Because of the variety of adsorption sites (clay minerals, organic matter, oxydes) models with two kinds of adsorption sites have been introduced (168). Adsorption occurs almost instantaneously for "type- 1" sites, whereas for "type-2" sites adsorption is time-dependent. Using somewhat arbitrary first-order kinetics, the general sorption rates were given by Selim et al. (168) as: aS 1 _ Stkl C - k2S (2-46) 8t b 21 2 at - k3C - kS 2 (247) t pb3 42 where S and S are the concentrations of solute sorbed to sites 1 and 1 2 2, respectively. Furthermore, k1 and k 3 are forward and k and k are backward reaction rate coefficients [T -1 , respectively. It should be noted that the same assumptions and restrictions apply as were made for the derivation of Eqs.(2-5) and (2-11). At equilibrium, the sorbed concentrations are given by the following linear isotherms: k _ 1 S = -- k2 C = KIC (2-48) S - C == KC (2-49) 2 Pb k 4 2 The sorbed concentration for all sites, at equilibrium, is given by: S = S + S = (K + K2)C = KC (2-50) A fraction F of the total sites belongs to "type-l" (i.e., F=S1/S=K /K). Since "type-i" sites are always at equilibrium, it follows from Eq. (2-48) that: 21 as at = FK a(2-51) For "type-2" sites, the sorption rate may be given by a linear, reversible first-order rate equation (cf. 109). The sorbed concentration for these sites follows from the rate expression given by: at - (K 2 2C - $2 (2-52) where a is a first-order rate coefficient [T-], as before, which is in this case equal to k Eq. (2-51) and (2-52) can be substituted into 4. Eq. (2-11) resulting in the following pair of equations to be solved: Upb ac Pb as 2 a 2 C ac (1+ b) -t+ &at = D -2-v (2-53) e at2 8 ax (2 4(1-F)KC-S 2 (2-54) Various authors (124, 168, 196) were able to describe the breakthrough curves fairly well with a numerical solution of Eq. (2-53) in conjunction with Eq. (2-54). The "one-site" kinetic non-equilibrium model follows from the previous model. In this case, F=O and Eq.(2-53) and (2-54) transform to: ac _PbasD a2 ac (2-55) + - -= D v (-x T eat 2 ax as = (KC - 5) (2-56) at where S still refers to the adsorbed concentration of a particular species. 22 Apparent Non-equilibrium Cameron and Klute (36), in effect, combined the physical and chemical non-equilibrium models. The transition from a chemical reaction at one site (the microscopic viewpoint) to the average observed for a large number of pores (the macroscopic viewpoint) is very complicated because macroscopically uniform soils are generally not microscopically uniform. This involves both the physical and chemical processes pertaining to sorption. In the section on physical non-equilibrium the limited supply of the solute, due to immobile water, was discussed. In fact, a whole range of supply rates exists. The same holds for the actual sorption process; many rate equations might be needed to describe sorption depending on the various soil constituents and the solute. Furthermore, a distinction between physical and chemical non-equilibrium is generally not possible. Therefore, Cameron and Klute (36) proposed a "black box" approach to describe the sorption process. They differentiated between two types of sites, those which appear to react rapidly with the solute and those which appear to react more slowly. Sorption at the latter type is described with a kinetic type of reaction which is used to take into account both chemical and physical non-equilibrium. The sorption sites are divided, as in the section on chemical non-equilibrium, into two fractions. S1 (equilibrium) and S (kinetic). Exchange between ions in the adsorbed and liquid phase occurs via a linear Freundlich (S I ) and , a kinetic (S 2 ) type of process, respect ively: 23 aS e 8C 0 - K -- k - C -kS (2-7 at -1Pbat 3% -4 2b Pb where K 1 (=k /k 2 ), k 3 and k are the equilibrium constant and the adsorption and desorption rate, respectively. S=S +S is the adsorbed 1 2 concentration and C is the liquid concentration, expressed in mass per volume of solution. The transport equation for combined kinetic and equilibrium adsorption is: Pbas 2 - ac 2c ac - (1 +K)- .=D -V-5(2-58) e.at i t 2 ax ax as 2 - k -C-k S ( 9 at 3 4S2 (2-59)? Pb Cameron and Klute (36) expressed these equations in dimensionless form and obtained a solution via Laplace transforms. Application of the model to transport of atrazine, phosphorus, and silver was successful; a purely equilibrium or kinetic model did not fit the data accurately. Nkedi-Kizza et al. (124) fitted two non-equilibrium models, a diffusion controlled and *a first-order reversible kinetic model, through experimental breakthrough curves. How fast equilibrium is attained in the ion exchange process, merely a redistribution of ions rather than a typical chemical reaction, is determined by two mechanisms: the supply of solute through the liquid phase to the liquid/solid interface and the nature of the exchange reaction at that interface. The actual ion exchange reaction is generally not the rate limiting factor in most instances (84). Rather, the diffusion of ions from solution to exchange sites, and vice versa, seems to be the rate limiting step even if no immobile water is present and we deal with 24 chemical non-equilibrium. Presumably, both chemical and physical non-equilibrium models might be used to represent non-equilibrium exchange. In summary, both models can be considered to have two types of adsorption sites. The physical non-equilibrium model has "mobile" sites with instantaneous equilibrium, while for the "immobile" sites exchange is diffusion controlled. This model is described with Eq. (2-44) and (2-45). The chemical non-equilibrium model is described with Eq. (2-53) and (2-54). For "type-2" sites, instantaneous equilibrium is not achieved because of the kinetic nature of the exchange process. When expressed in dimensionless variables it can be shown that the transport models for both models are identical and have equivalent breakthrough curves. Both models can be used when describing ion exchange during transport through aggregated sorbing media. Based on breakthrough curves, obtained via curve fitting, Nkedi-Kizza et al. (124), however, did not want to draw the conclusion that the two models were conceptually similar. Judgment whether local equilibrium exists can be based on exchange data and the physical properties of the medium (cf. 24). Valocchi (191) presented criteria for the validity of the local equilibrium assumption. Presently, much work is being done in the area of transport through structured soils. As far as physical non-equilibrium is concerned, sometimes an effective dispersion coefficient can be used for soils containing aggregates of a particular size and shape. In this way, the sink/source term describing mass 25 transfer from mobile to stagnant regions of the liquid phase can be omitted (cf. 57, 135, 150). On the other hand, - a distinction can be made between the inter-aggregate pore space (macropores), containing the mobile liquid, and the intra-aggregate pore space (micropores), containing the immobile liquid. In the intra-aggregate porespace the predominant transport mechanism consists of diffusion. Generally it is assumed that inside the aggregate instantaneous equilibrium is achieved between the species in the adsorbed and liquid phase. A number of analytical solutions have been published (cf. 153, 186, 201). Some of these solutions will be discussed in Chapter V. III. DISPERSION AND DIFFUSION When a feed solution containing a given solute displaces a resident solution containing another solute or the same solute at a different concentration, a transition zone will develop in which a variation of solute concentration will occur. During displacement experiments, the solute concentration of the effluent is generally monitored to obtain a breakthrough curve. The curve is indicative of the amount of mixing between the two solutions. For a non-reactive solute, the spreading is caused by dispersion and diffusion. In order to describe solute transport, one generally needs to quantify both processes. Solutes can also be useful in hydrological studies, where they are used as tracers to study flow phenomena in porous media. Many references exist about diffusion and dispersion (15, 23, 24, 60, 61, 165). Following Fried and Combarnous (61), two mechanisms of dispersion will be distinguished: 1. Mechanical dispersion. When adopting a microscopic viewpoint, nonuniform velocity profiles exist in a porous medium because of the following reasons: - The velocity of the soil solution is zero at the solid surfaces. Slip flow does not occur because of the relatively high viscosity of water and the small mean free path length (165). 26 27 - The pore dimensions vary, so different maximum velocities occur along the axes of the pores. - Streamlines fluctuate with respect to the mean direction of flow. 2. Physico-chemical dispersion: diffusion. This type of dispersion is due to concentration gradients, actually gradients in chemical potential. The following phenomena take place: - If inside a streamtube a concentration gradient occurs, then diffusion tries to annihilate the gradient. - If concentration gradients exist between two adjacent streamtubes, mass transfer between the streamtubes occurs by diffusion. For both types of dispersion a longitudinal and transverse component can be distinguished. Mechanical dispersion (dispersion) and physico-chemical dispersion (diffusion) are closely related. Diffusion occurs at the molecular level and dispersion at the pore level. Usually it is not possible to distinguish between these levels. Because of their similar nature, diffusion and dispersion are conveniently described together by the coefficient of hydrodynamic dispersion. Molecular Diffusion The process of molecular diffusion is of interest for at least two reasons. First, at low pore-water velocities transport of solutes is dominated by the diffusion process, and second, an analogy exists between molecular diffusion and mechanical dispersion. Because of this analogy, knowledge of the diffusion process is helpful in understanding dispersion. Let us first consider molecular diffusion in a (free) solution, next diffusion in porous media, and conclude with a discussion of mechanical dispersion. 28 The solute molecules in the solution possess random thermal motion which causes an exchange of molecules between adjacent volume elements. If isothermal and isobaric conditions exist, a net. transfer of molecules of some species k occurs when the concentration of k in the adjacent volume elements differs. More particles of k move from elements of higher to elements of lower concentrations than vice versa. Such a net transfer under the influence of a concentration gradient is termed molecular diffusion. The process of molecular diffusion is described by Fick's first law. In the case of one-dimensional diffusion in a free liquid, it can be expressed as: J =-D ac(3-1) D o ax where J is the solute mass flux due to diffusion [ML-2T -] and D is D 0 the coefficient of molecular diffusion [L 2T_]1 Thermodynamically, it is the gradient in the chemical potential that is the driving force for the diffusion process. The chemical potential is defined as: 8G(n.)' P njk(3-2) is -1 where p 1 is the chemical potential of species k [J mol Ink is the number of moles of species k, G is the f ree enthalpy or Gibbs free -1 energy [J mol I and 9'J and P are temperature and pressure, 29 +R9J1in C(3-3) 0 where pk is the chemical potential of species i in a chosen standard state and the mole fraction Ck=Ck/CM the ratio of the molar Mk Mk M concentration of species k to the total molar concentration. Hence, for an ideal thermodynamic solution, the gradient of the chemical potential is proportional to the gradient of the natural logarithm of the concentration: ap a ln C ~ k = M(-k ax]ideal =ax(3-4) However, most solutions do not exhibit ideal mixing behavior. In transport studies, the non-ideality of the solute needs to be addressed for diffusion dominated transport if, for example, the total solute concent rat ion is "high" (sea water) or for the thermodynamic description of ion exchange. For non-ideal systems, Eq. (3-3) is generally rewritten as: =0 + RYJ ln ak(3-5) where ak is the chemical activity of component k, which is related to the concentration by the activity coefficient: a = kCMk(3-6) a k k3-Mk The value of the activity coefficient, 7 k, depends on the nature of the microscopic interactions. Activity coefficients for ions can be estimated with the Debye-Hiickel equation for solutions with ionic strength less than 0.1 N (e.g., 125). The 'Davies' extension of the Debye-H~ckel equation can be used to approximate k for systems 30 containing mainly small ions at ionic strengths less than 0.5 M. For non-ideal solutions, the gradient of the chemical potential is: 3 1[ ] 8 lnak k = k(3-7) a x Jnon-id eal ax The diffusion flux for ideal solutions is given by Fick's first law for a constant total molar concentration CM: N =dC (3-8) k CMD dx where N k is the molar flux density of species k due to diffusion [ML T - 1. The diffusion flux for non-ideal solutions is obtained by inserting the ratio of the non-ideal to the ideal chemical potential gradients in Eq. (3-8): aIn a dC N =-CDk Mk (9 Nk M-CMDo * dx 8 ln CMk The contribution of molecular diffusion to the total molar flux can be shown by substituting Eq.(3-9) into the continuity equation, Eq.(2-1). The result is: CMk _aalIn ak dCMk JvC(Mk3-10) at jx oaIn Ck 1 0x (3- Assuming an ideal system with a liquid velocity of zero and D 0 independent of position and concentration, the well known diffusion equation, Fick's second law, is obtained: ac a 2 c -D (3-11) 3t 0 2 where the subscript notation for Ck has been dropped. 31 It can be shown by differentiation that (45): C(x,t) A xexp (3-12) , -4D t to is a solution of Eq.(3-11), where A is an arbitrary constant. The total amount of substance, m O, diffusing in an infinitely long cylinder with unit cross section, is given by: m= C dx (3-13) -00 With an initial condition C(x,O) = mO6(x), 8(x) being the Dirac delta function (i.e. a spike of solute in an otherwise solute free medium), the concentration distribution becomes (60): m 2 C(x,t) = exp 4D (3-14) 4D t Do Substituting o = 2D t, yields a Gaussian or normal distribution: 0 m 2 C(x,t) = 0 exp[ x (3-15) S2 22 where o- is the standard deviation of the solute concentration distribution. The process of molecular diffusion in porous media is similar to that in a free solution, except that we must now define the mean flow path for diffusion in terms of the structure of the medium. Let us assume that diffusion in other phases than the liquid phase can be ignored. If necessary, the additional contribution to molecular diffusion from exchangeable ions can be accounted for (126). The coefficient of molecular diffusion in soils can be expressed as (24): 32 D = D /A (3-16) e o where A is the tortuosity factor. In simple terms, tortuosity can be viewed as the ratio of the "true" (i.e., circuitous) and straight line flow path distances between two points in a porous medium. Theoretically, its value should only depend on the geometry of the medium. Tortuosity can also be characterized with the so called formation factor, which is frequently used in petroleum engineering. This factor is the ratio of the electrical conductivity of the pure liquid phase to that of the porous medium (cf. 46). It depends on the volume fraction of liquid and the tortuosity of the medium: D /D = /(Mo) (3-17) e o where D is considered an effective coefficient of molecular diffusion e 2 -1 [L T i ] and ? is the formation factor. It should be noted that the validity and usefulness of this concept of tortuosity has been questioned (165). In media with a very low hydraulic conductivity, diffusion is the main mechanism of solute transport. Therefore, the determination of diffusion coefficients for low-permeability media has recently received a considerable amount of interest (e.g., 173). For a concentration profile following a Gaussian distribution, i.e., Fickian diffusion, the coefficient of molecular diffusion of a stagnant liquid in a porous medium can be determined from: -_ 2D (3-18) dt e where D has been substituted for D . This relationship is frequently e o 33 used to obtain the diffusion coefficient from the standard deviation of the concentration distribution as a function of position. In porous media, spreading is caused by various mechanisms. Integration of Eq. (3-18) shows that the variance increases linearly with time for constant diffusion coefficient. Thus, Eq.(3-17) can be used to evaluate whether these mechanisms exhibit Fickian behavior. An approach which is not restricted to profiles following a Gaussian distribution is the method of moments. The moments of a concentration distribution, resulting from a pulse type of solute input, can be used to determine the mean and variance from the observed or calculated distribution (9, 10). No assumption about the distribution of the solute concentration needs to be made in order to determine these moments. The p-th moment of a concentration distribution with respect to x, m , is defined as: m = x C(x,t) dx (3-19) p I --00 This operator m can be applied to each term of the advection-dispersion equation. Aris (10) showed how these moments can be determined without evaluating the integrals, which is convenient if the distribution of C(x,t) cannot be described by a mathematical relationship. Some properties which can be determined with these 2 2 moments are the mean, . or i , and the variance, o or , of the x x concentration distribution, where = ml/mo (3-20) 2 2 a = (m /m ) - 4 (3-21) 34 Aris' moment method is also an Important tool In obtaining solutions for the diffusion/dispersion equation in stratified media (80, 81, 115). Dispersion The description of dispersion as a diffusion type process has been shown plausible in the classical paper by Taylor (189). Taylor's analysis, concerning flow in a circular tube, is often discussed in the literature (e.g., 60, 61, 73,- 121). Because dispersion in a free solution provides a qualitative explanation of dispersion in a porous medium, a brief treatment of Taylor's paper will be presented. Taylor- (189) considered laminar flow in a circular tube with radius a, having the following parabolic velocity profile: u(r) =u r ](3-22) a-1 wh Iere u0 is the maximum velocity at the axis [LT ] and r is the radial distance from the axis [L]. It can be seen that the mean velocity over a cross section of the tube, u, is equal to u /2. m 0 First, transport of a solute by advection only will be considered. When a solute is introduced into the tube, the concentration profile develops similarly to the velocity profile, since no transport in the radial direction occurs. Taylor (189) discussed three case for whic the initil1coditins-ad2th1mea 35 V Co Al CO 0 a L*- u --t A3 rA A2 [ ott >i t+x FIG. 2. Distribution of mean concentration in three cases in absence of molecular diffusion, after Taylor (189). In the first case, Al, the solute is originally only present in a short segment of the tube, with width X and concentration C . The 0 solute will be distorted into a parabola with its shape depending on u(r). The mean concentration over a cross-section is now given by: C =CX/(ut) OO m 0 0 0(3-24) C = 0 x>u t t>O mo In the second case, A2, a solute with concentration C enters the tube. 0 This case can be solved by assuming that the constant initial concentration for xO (3-25) m 0 010 C = 0 x>u t m 0 The third case, A3, deals with a solute which is initially confined to a greater distance X than for the first case (189). In all three cases, the concentration profiles are determined merely by advection, i.e., by the shape of the velocity profile. However, at low velocities the concentration is determined by molecular diffusion as well. Subsequently, the second step is to include molecular diffusion in the transport equation. Taylor assumed that diffusion was significant for radial transport and negligible for longitudinal transport. Therefore, the following respective restrictions were imposed: 2a 2 2D At-2a or At D -D(3-26-a) 00 2D At-L or At2 At (3-47) dis where At is the mean residence time of a solute particle in a tube. The relative spread of velocities, (a-,)2, is given by 2/ 2 , which is a constant for a particular medium with average velocity and variations about that velocity of Av =Iv,I. This part of the 46 dispersion in porous media was first developed by Taylor (189). By reversing the direction of flow, this type of dispersion can be annihilated (cf. 167). Obviously, dispersion in a porous medium is much more complex because pore channels are interconnected and not straight. This situation can be modeled by using so called mixing compartments, figure 3, where complete mixing occurs at points x=O, x=L, x=2L, etc. The effective dispersion coefficient has a value of zero at x=O and reaches a maximum at x=L. With the mean residence time of a solute particle in a tube between two mixing points being At=L/, an average dispersion coefficient can be used: 2 D = (o-')L/2 (3-48) dis Figure 3 illustrates the effect of mixing, at regular intervals L, on the coefficient of mechanical dispersion according to Eq. (3-48). As can be seen, 'unlimited' spreading occurs only over a finite distance, L, between mixing points. An effective, average dispersion coefficient for such a distance is used for computational purposes. If transport in separate channels is the main mechanism behind dispersion, this dispersion will be limited by mixing at intervals with length L. Then the mechanical dispersion coefficient is presumably of the form: D =L (3-49) whreLcan be approximated from Eq. (3-48), i.e. , from (a ')2L/2. whr dis Bolt (24) suggested that a value equal to the grain diameter be used. 47 -- 0 0 L 2L 3L x [L] FIG. 3. The value for D determined by mixing and the averaged value dis for D according to Eq. (3-48) (dashed line). dis Usually, however, L has to be quantified by experimental procedures. dis Theoretically, if the flow were to be reversed along the same steamlines as before the reversal, only the contribution of molecular diffusion to mixing would not vanish. In reality, a larger part of this mixing is not reversible. In addition to mixing, advective spreading is limited by transverse diffusion. This mechanism becomes important if flow occurs in a single tube or if no mixing occurs because of the absence of inter-capillary pore space, as inside the aggregate. It is still assumed that the porous medium consists of simple tubes for which Taylor's (189) results can be used. An expression for Ddi s was obtained by combining Eq. (3-47) and an appropriate expression for the diffusion 48 time needed to traverse the width R over which the velocity profile is spread. Bolt proposed: D (') 2 2v>2gR2/D (3-50) dis e where g is a geometry factor. This expression is similar to Taylor's result for a circular tube (Eq.(3-41)). The dispersion length becomes: L ( ') 2gR /D (3-51) dis e In porous media, mixing and transverse diffusion (i.e., in a single tube) occur simultaneously. The question then arises as to which of the two prevails. This can be estimated by evaluating Eq. (3-48) and (3-50). Since both processes counteract spreading, it can be assumed that the process which results in the smallest value for Ddis dominates. Physical non-equilibrium occurs when flow in the intra-capillary pore space is negligible. The stagnant phase of the liquid in the intra-capillary pores does not participate in any (turbulent) mixing. Rather, mixing between mobile and immobile phases of the liquid occurs via transverse diffusion only. A description of dispersion in the stagnant phase can be obtained by adapting Eq.(3-50) to account for the limited accessibility of the liquid in the immobile phase. This was done using a functional relationship between mobile and stagnant liquid concentrations. This relationship can be presented, assuming a linear increase of the surface concentration of the aggregate with time, C =kt, as (45, 135): mo 49 C - = kg(R )2/D (3-52) mo im a e where Cmo is the concentration in the mobile phase [ML 3, is themo ' im average concentration in the immobile or stagnant phase [ML-3], k is a ratecoeficent[ M-3T - 1 rate coefficient [ML T ] and R is the radius of the aggregate [L]. a The geometry factor of the aggregate, g, varies between 1/8 for infinitely extended cylinders to 1/15 for a sphere. By using the solute flux over the entire liquid phase, Passioura (1971) obtained the following dispersion coefficient, as caused by the stagnant phase effect: 2 2 Ddis = go. R2 2/OD (3-53) dis im a e with accompanying dispersion length: L = ge. R 2 /OD (3-54) r Im a e The effects of longitudinal diffusion, advective dispersion (microscopic variation in the velocity profile in combination with mixing and/or transverse diffusion) and the presence of a stagnant phase can presumably be added since they act more or less independently. It should be noted, however, that this last assumption deserves more investigation (123). In the case of autonomous effects, the diffusion/dispersion flux can be written as: BC D = -JVLD x (3-) D diff dis r e V dis av e 50 It appears that, in many instances, mixing is the major counteracting factor of advective dispersion. Ldis can therefore be assumed to be constant ( 2R ) L is only important at small fluxes, a dif while L (dispersion length due to stagnant phase) becomes significant r for larger values of JV and increasing aggregate size R .a An increasing aggregate size also causes the breakthrough curve to lose its sigmoidal shape (22). Because of the relative importance of the larger (inter-capillary) pores for transport, the tracer will appear sooner in the effluent with larger aggregate sizes. It should be noted that Ldiff is inversely proportional to JV (Eq.(3-46)), and assuming that mixing prevails (Eq.(3-48)), Ldis is approximately equal to the aggregate diameter, i.e., independent of the flow, and Lr is proportional to JV (Eq. (3-54)). Despite the fact that a number of simplifications and assumptions were made, the correspondence between the experimental and theoretical behavior of L D was fairly good, figure 9.8 (24). In terms of the Pe-number, defined as Pe=vL/D e (L=2R), diffusion is the dominant term for Pe<1 and the stagnant phase a effect becomes dominant for Pe>120. Experimental Determination of Dispersion Coefficients To describe the dispersion process and to solve the ADE, a numerical value for the longitudinal, and sometimes for the transverse, dispersion coefficient needs to be determined. For one-dimensional flow in a homogeneous medium with two-dimensional transport, the ADE can be written as: 51 ac ac ac ac (357) at - DL 2 x DT .a52 ax 2 - ay2 D L and D T can be obtained by relating experimental results of the concentration profile with explicit expressions for C. For this purpose, displacement experiments are carried out, either in the laboratory or in the field. For the determination of D these experiments often involve a laboratory column filled with a packed soil. The solute concentration of the resident solution in the column is known, and a feed solution, with generally the same concentration of a different solute, is leached through the column. In most cases the effluent concentration is determined as a function of time, although the concentration in the soil solution, at different locations along the direction of flow, can be determined as well. A variety of methods, used to obtain dispersion coefficients from these experimental dat., will be presented here. A review of some of these methods can also be found in van Genuchten and Wierenga (203). Because macroscopic variations in water content and pore water velocity result in additional spreading, the column should be packed in such a way that the soil is homogeneous with respect to the advection term. Although this situation will never be reached totally, a constant 'effective' pore water velocity can be used for many practical purposes. Deviations from this velocity during the experiment obviously lead to incorrect results. Mixing due to viscosity and density differences between resident and feed solution are not to be included in the coefficient of 52 hydrodynamic dispersion. Therefore, many studies involve the so called tracer case (e.g., 20). The low concentration of a non-reactive solute (i.e., a tracer) does not affect the density and the viscosity of the solvent. Displacing and displaced fluids have the same density and viscosity. Differences in viscosity and density lead to instabilities (fingering). Biggar and Nielsen (21) showed that unstable flow -3 -3 dominates for a density difference of 3.4x10 g cm and a difference in viscosity of 0.003 cP, obscuring the effect of molecular diffusion. Rose and Passioura (161) demonstrated that, during horizontal displacement experiments, small differences in density of the displaced and displacing fluids lead to quite different dispersion coefficients. The displacing liquids consisted of water, 0.1, 0.2, and 0.4 M NaCl solutions with densities of 0.99707, 1.00116, 1.00523, and 1.0133 g -3 cm , respectively. The amount of gravity segregation could be characterized with a gravity segregation factor 0: = gkhAp/(rqvL) (3-58) -2 where g is the acceleration due to gravity [LT ], k is the permeability [L2], h the height of the porous medium in the transverse -3 direction of flow, [L], Ap is the difference in liquid density [ML ], ^t is the mean of the viscosities of the resident and feed solution [ML T-1, v is the mean pore-water velocity [LT - ] and L is the length of the porous medium in the direction of flow, [L]. Since k decreases rapidly during desaturation, gravity segregation is not much of a factor during unsaturated flow. An increase in D by a factor 2 was -4 reported by these authors for their experiments, with Ap=8.2x10 g 53 -3 cm compared to Ap=O. Gravity segregation might also be important when contaminated salt water intrudes horizontally into a fresh water aquifer. Because of the stratified flow and the subsequent mixing by diffusion at the interface, contamination will extend further into the fresh water than expected on the basis of the conventional hydrodynamic dispersion process. Further discussions on viscosity and density effects can be found in Krupp and Elrick (105), Bachmat (12), and Scheidegger (165). Due to the temperature dependence of these parameters, displacement experiments need to be carried out at constant temperature. One should also be aware of apparatus-induced dispersion (87), which is especially important for short-column experiments and for unsaturated media. These authors studied a conceptual porous medium consisting of two different layers: the soil and the apparatus. Apparatus-induced dispersion was assessed by carrying out displacement experiments in the absence of the porous medium. The dispersion coefficient of the porous medium was obtained by solving the advection-dispersion equation for a two-layer system. The value of this dispersion coefficient was as much as 40% less than obtained with the one layer equation, where mixing was assumed to occur in the medium only. The so called dead-volume, inside the apparatus but outside the porous medium, should therefore always be kept to a minimum. Rather than relying on effluent concentrations, in situ determinations of the solute concentration can also be made. Harleman and Rumer (82) measured electrical resistance with a conductivity probe 54 to determine NaCl-concentrations. Gupta et al. (76) used an Ag-AgCl electrode to measure Cl concentrations in an unsaturated glass bead medium. Electrical potential, in contrast with electrical resistance, is not greatly affected by the water content of the medium. Kirda et al. (95) used labeled Cl in the feed solution, of which the concentration was measured in situ with a Geiger-Muller tube. Grismer et al. (70) used a dual-source gamma-attenuation system to determine salt and water content. In an error analysis, thse authors showed that the accuracy of solute concentrations determined at low water contents was limited. Grismer (69) used the same technique to determine SrCl 2 and Nal concentrations in a displacement study during horizontal transient unsaturated flow. Initially, no salt was present and the feed solutions had a molality of 0.205 and 0.1 m, respectively. No mention was made of density or viscosity effects. Following Fried and Combarnous (61), some methods to determine DL and D will be discussed. The solute (tracer) concentration can be determined as a function of position at a certain time or for various times at a certain position, as with breakthrough experiments. For both cases, analytical solutions exist. For a uniform medium and steady flow (v and 8 are constant), the 1-D ADE was given as: c D 2C -vc (3-59) at L x2 8x subject to the following initial and boundary conditions C(x,t) = C x=0 t>O 0 C(x,t) = 0 x- t>0 (3-60) C(x,t) = 0 x>0 t=0 55 Utilizing the definition of the complementary error function, the solution of Eq. (3-59) may be written as (i.e., an approximation of the solution by Lapidus and Amundson (109)): 1 x- vtl C/C =2erfc [ (3-61) /4D t L For a given time, the solution follows a normal distribution 1-N[(x-p)/r] with mean displacement =vt and standard deviation - = -2D t. N[ J is the probability density function for a normal L distribution with values N[-1]=O.16 and N[1]=0.84. The width of the transition zone, 2o-, can be determined by plotting C/Co versus x, with: 2o = x0 -x0 = 8Dt (3-62) 0.16 084 L where x and x denote positions for which C/C is equal to 0.16 0.16 0.84 0 and 0.84, respectively. The value of the longitudinal dispersion coefficient follows from: D=(x -x 2/8t (3-63) L 0.84 0.16 If the concentration is observed at a certain position as a function of time, the following expression for D can be used: L x-vt x-vt DL 0 i[/ .16 _ t0.84] (-4 t t 0.16 0.84 After the discussion of the determination of the longitudinal dispersion coefficient, D, Fried and Combarnous (61) continue with the determination of the transverse dispersion coefficient, D T . It should be noted that relatively little in-depth work has been done in the area 56 of transverse dispersion, because of the increased experimental difficulties and the much smaller effect of transverse dispersion in comparison with longitudinal dispersion (e.g., 82). For the determination of D in a semi-infinite 2-D system T consisting of a uniform soil, it is assumed that 1-D steady state flow conditions exist and that steady solute transport is established (C=C(x,y)), in which case the longitudinal dispersive flux is usually negligible. Under these circumstances the transport equation becomes: 2 v D C (3-65) 8ax T 2 8y where y is the distance in the direction transverse of the flow [L]. For the following boundary conditions: a= 01x>O , y4?0 C(O,y) = C x=O , OO t=O i C = C x=O t a-0 (3-74) 0 C = C. x-W t>0 1 where D (8) is an alternative expression for the dispersion s coefficient. It is assumed that the medium is semi-infinite; the initial and exit boundary conditions transform to the same condition. For a homogeneous soil with horizontal, non-hysteretic flow we have: Jv = -D ( ) e (3-75) w ax therefore Eq. (2-132) was rewritten as: 0 ac _ a [D () ac + D () ae ac (3-76) t a Ds) a J w axax where D is the soil water diffusivity [L2T-1. Using the Boltzmann w transformation (w =x/V/t), Eq. (3-76) becomes d dC] dc D (0) + -- 0 (3-77) d- s d- + 2 dw while the boundary and initial conditions, Eq. (3-73), become C=C. -o 1 ( 3-78 ) The function g is defined as: dO 8 g = Ow + 2D (o) -e = Iw O(-9 w d-w f dcC-9 60 The dispersion coefficient can then be obtained from C D (e) = d g(3-80) s 2 - dC (-0 C. 1 in a way as described by Bruce and Klute (35) for D . D Ce) may be w s determined by carrying out experiments using segmented soil columns. For each segment, average values for e, C, and g can be determined as a function of w. Laryea et al. (110) determined dispersion coefficients for both cations and anions during horizontal infiltration. Elrick et al. (59) extended the above analysis to vertical flow. A power series solution similar to that of Philip (139) was used to express the solutions of flow and transport equations. The dispersion process depends on viscosity, density, velocity, water content, molecular diffusion, and permeability. In order to assess the influence of velocity and particle size, experimental results are commonly analyzed by plotting DL/Do versus the Peclet number (Pe=vd/D) on a log-log scale graph. In this case, d is a 0 characteristic pore or particle size dimension. Bear (16) distinguished the following dispersion regimes (61) based on values of the Peclet number: - Molecular diffusion is dominant for Pe<0.4. - Molecular diffusion and mechanical dispersion are of the same order for O.41.5x10 . Several authors have provided graphical relationships between velocity or the Peclet number and the dispersion coefficient (e.g., 22, 24, 61, 138, 160). Finally, the so-called numerical dispersion should be mentioned. When the transport equation is solved by numerical methods, using experimental values of the dispersion coefficient, the solute front exhibits additional spreading. Especially at high values of the Peclet number, 'smearing' of the numerical solution occurs around the front. This is an artifact due to the numerical procedure (26). A number of approaches can be taken to reduce this phenomenon (cf. 1, 29, 39, 195). IV. ION EXCHANGE Equilibrium transport of linearly exchanging solutes can be described in a straightforward way with Eq.(2-19). However, for transport of non-linearly exchanging solutes, the situation is somewhat more complicated. First, the influence of exchange on the solute front without dispersion/diffusion is reviewed. Second, the effect of dispersion is included. Several approximate analytical techniques can be used to study transport under these conditions, which illustrate the effect of non-linear exchange. In a majority of the studies, however, the transport problem is solved numerically by using experimentally determined exchange isotherms. Some of these studies will be discussed. Equilibrium Chromatography Following Reiniger and Bolt (156) and Bolt (24), the qualitative influence of the exchange isotherm on the concentration front will be studied for a relatively simple case. It is assumed that only two different cation species are present during miscible displacement under steady flow conditions. This provides a fairly representative picture for many soil-water systems. The adsorbed concentration q of a solute species depends on the liquid concentration of that species, C, and the total concentration of all solute species, C It is convenient to -3 express q in mol m , i.e., moles of charge per volume of porous c 62 63 medium, and C in mol m , i.e., moles of charge per volume of solution. The total concentration is constant during ion exchange. When diffusion and dispersion are neglected, a step change in concentration occurs at the "concentration front" located around Jvt/0, the V penetration depth. If C is assumed to be constant across this T concentration front, q depends only on C. For a particular species Eq. (2-7) then becomes: .(q' + ) acJ ac at - v -a1 where q' (=dq/dC) is the differential capacity of the exchanger for the exchanging species, JV is the Darcy flux which is assumed to be positive, and x is the (positive) distance from the inlet. In order to solve Eq.(4-1), and to justify neglecting the dispersive flux, aC/ax needs to be finite. With the chain rule for partial derivatives, ac ax at (-) (c) (-x)c= -1, Eq.(4-1) can be rewritten as (ax _ JV(4-2) at-C -q' +-0 If the condition that ax be finite is violated, i.e. if the Ox concentration profile exhibits jumps, the conservation of mass (Eq. (4-1)) needs to be expressed in an alternative way: (Aq + 8AC) dx = AC dV (4-3) where V is the input volume per unit area of the column. The rate of propagation, given by Eq. (4-2), then needs to be rewritten as: dx JV(4) ( d -) a = a q ( 4 -4 AC 64 ac The position of a particular solute concentration (for finite) follows from integration of Eq. (4-2). For a soil with q and 0 homogeneous with respect to position and dV=Jvdt, the position for a particular concentration is given by: tj V-V (C) Xc= J qi+ dt = o___ xC q+ t q')+0 ( 4-5 ) 0 where the feed solution, with concentration C, enters a column having an initial concentration C ; V (C) is an inverse feed function, which 0 is the volume of solution applied to the column at the moment that the concentration at x=O reaches C (for step type displacement V =0). The 0 average depth of the concentration fronts in the adsorbed and liquid phase is given by: C C .0 0 ={XC (q'+ 0) dC / (q'+ 0) dC (4-6) C. C. 1 1 The rate of propagation of the solute in the liquid phase follows from Eq.(4-2). Equilibrium chromatography enables us to determine the propagation in the adsorbed phase as well by using the exchange isotherm. To study the shape of the solute front, Eq. (4-5) is differentiated (V =0): 0 ax _ Vq'' (4.7) OC V (q'+ 0) At this point, a distinction needs to be made among favorable, unfavorable, and linear exchange. If the incoming cation has a convex 65 2 2 isotherm, q' '=d q/dC2<0, we deal with so-called favorable exchange, while the exchange is said to be unfavorable if q''>O and linear if q' '=0. In case of favorable exchange there is a dilemma. According to Eq.(4-1), the slope of the solute front is negative (aC/axO. The latter implies that the solute travels faster at higher concentrations than at lower concentrations. This is physically not possible for a step front and, as already mentioned, the rate of propagation should then be determined based on Eq.(4-4). For cases other than a step change in the concentration of the eluent, a "self sharpening" effect will occur until a step front has been established. For linear exchange (q''=0), we can see from Eq. (4-7) that ax/aC=O for any applied volume. The initial profile will therefore not be altered during passage through the porous medium. If the initial profile contains jumps, Eq. (4-3) needs to be used, otherwise Eq. (4-1) suffices. For unfavorable exchange we can conclude from Eq. (4-7) that for a particular value of C, the front flattens with increasing V. Unfavorable exchange induces solute spreading (i.e., decreases the concentration gradient). The Combined Effect of Ion Exchange and Dispersion Bolit (24) examined the combined effect of exchange and diffusion/dispersion using analytical techniques. Having expressed all (physical) dispersion effects in one dispersion length, L , he 66 attempted to include exchange as well. The ADE in terms of the dispersion length is obtained with Eq. (2-7) and (3-55): raC aC a__(L 1C (q'+e) -c +J a-c-( x -) 1=0 (4-8) V[8x -X D axJ Although the 'effective' value of L D varies through the soil column, Frissel et al. (62) showed that a column averaged value might be used. For a constant LD Eq.(4-8) might be written as Eq.(2-14). If, in addition, dq/dC=constant (linear exchange), the same analytical solutions as for Eq. (2-19) can be used. In case of favorable exchange, the solute front in the adsorbed phase is generally located ahead of the solute front in the liquid phase. In terms of the effective retarded velocity, v , the "high" concentrations will travel at a velocity greater than v and the "low" concentrations will travel at a velocity less than v . In contrast, diffusion/dispersion causes just the opposite effect. Eventually, a steady front will develop with respect to the moving coordinate x-v t and all concentrations will travel at a velocity equal to v . For a constant L Eq. (4-8) can be rewritten to obtain a general expression for the propragation of the solute front: ( - qO _1-LD a(ac/ax) (4-9) Once steady state has been established, we obtain: * x J v v = (8 - Aq (4-10) AC Based on these two expressions, Bolt (24) presented a method to analytically determine adsorbed and liquid concentrations. 67 In case of unfavorable exchange, no steady front will develop with respect to the moving coordinate, because dispersion/diffusion and exchange both cause the "low" concentrations to travel faster than the "high" concentrations. The ADE must now be solved numerically to get an accurate picture of the solute front in the adsorbed and liquid phase (Cf. 106, 107). For estimation purposes, one can add the two effects by first solving the front position as determined by Eq.(4-5) (exchange only) and then adding the effect of diffusion/dispersion. The latter effect causes the solute front to spread around x =v t. With Eq.(2-38), p we can estimate this spreading according to: x-x = 2 D t inverfc(2C) (4-11) p where C is the dimensionless concentration defined in Eq.(2-38). This term can be treated as a perturbation term to be added to the front position according to Eq.(4-3). In this way, a conservative estimate was obtained for the maximum amount of spreading. Finally, the occurrence of non-equilibrium during ion exchange should be mentioned. In aggregated soils, the deviation from local equilibrium is most likely to be caused by physical non-equilibrium, i.e. the limited accessibility of exchange or adsorption sites. Bolt (24) assumed that all adsorption takes place in the stagnant region of the liquid. Examples of adsorption in both mobile and stagnant region were treated by van Genuchten and Wierenga (202), Tang et al. (186) and Selim et al. (170). Transport from the mobile to the stagnant region occurs via a diffusion process (Eq. (2-44)), which was described by Bolt (24) with: 68 a(q.m +m .mC =im imCim k 0 (C -C.) (4-12) at a mo mo im where qim is the concentration in the adsorbed phase (immobile sites only), k is a rate constant for diffusion inside the aggregate, and a instantaneous chemical equilibrium is assumed inside the aggregate. A distribution ratio between mobile and immobile phase, defined as: Kmi (Aq + 0. 0o(4-13) D A AC im mo can be used to rewrite Eq. (4-12) as ac. k im - a (C -C.) (4-14) at mi mo im Without longitudinal diffusion/dispersion the transport equation becomes (Eq. (4-1)): aC. aC aCSmi im mo mo 0 K + m-.0 =+ o 0 (4-15) moKD at mo at vax Equations. (4-14) and (4-15) can be solved by introducing a position dependent time and by using a transformation in order to obtain scaled variables. The solution is expressed in the so called Goldstein J-function (cf. 56, 196). Instead of using the above approach, the stagnant phase effect can be described by an equivalent length parameter L . For spherical r aggregates with radius Ra the expression of Crank (45) is adapted for a situation where exchange occurs: (q.m + 0. C. im i m- iO. D (C -C. ) (4-16) at R 2 im mo I 69 Combination of Eqs.(4-12) and (2-16) results in the following expression for k? a 0. 15D im (4-17) k- 2 0to R mo a With the use of Eq. (3-46), L can now be expressed in terms of k a Two ' r a" different expressions were given by Bolt (24), depending on whether adsorption/exchange takes place. The advantage of this approach is, as was mentioned in the section on physical non-equilibrium and dispersion, that all effects are accounted for by one effective value of the dispersion length. Consequently, analytical solutions can be used to solve the advection-dispersion equation. Numerical Solutions of Transport Involving Equilibrium Exchange The two preceding sections, discussed some of the fundamentals of transport of exchanging solutes with and without dispersion. This was done in a way so that explicit expressions for the position of the solute front were obtained, and that (approximate) analytical solutions were available. However, because of the non-linearity of the (measured) exchange isotherms, numerical techniques are usually needed to solve the transport equation. For that reason, this section focuses on the (numerical) simulation of exchanging solutes. One of the earlier works on transport of reactive solutes was reported by Kay and Elrick (94). These authors determined adsorption isotherms of lindane for various soils and soil fractions. Linear adsorption was observed; the l indane was particularly strong adsorbed by organic matter. A chromatographic model developed by Hashimoto et 70 al. (83) was used to predict the movement of lindane through the soil. A reasonably good agreement between predicted and experimental breakthrough curves was obtained. At higher pore water velocities, the prediction was somewhat poorer, perhaps because of physical non-equilibrium. Displacement of the lindane (elution curve) was described poorly, possibly because of inadequate knowledge of the desorpt ion curve. Lai and Jurinak (106, 107) solved the transport equation for non-linear adsorption with an explicit finite difference method. These authors considered homovalent exchange in a binary system for one-dimensional steady flow. The dimensionless variables Xk=Ck/C T and Y =S /S were introduced for the solute concentration in the liquid and adsorbed phase, respectively. The following equation was obtained, using a similar relation as Eq. (2-12): 2 aXk a2Xk aXk at k D(X2k kV(Xk ax (4-18) ax where the dispersion coefficient and velocity were adapted according to: D(Xk) = D (4-19) S1+ (PbST/CT) J (Xk) v(X k ) = v (4-20) S1+ (PbST/OCT) f' (Xk) with f' (Xk)=dYk/dXk . The denominator of the latter two terms can be recognized as the retardation factor; its value depends on X k because of non-linear exchange. Following Helfferich (84), various adsorption functions Yk=f(Xk) were used. For a number of isotherms (linear, 71 concave and convex), numerical solutions were obtained for Eq. (4-18) subject to specified boundary and initial conditions. Column studies were carried out involving Ca and Mg, which exhibit a slightly non-linear exchange isotherm. The comparison between experimental and numerical results was good. Rubin and James (163) presented a comprehensive analysis of multi-component equilibrium exchange during one-dimensional steady flow in layered and homogeneous profiles with a variabl e C T The equations to describe exchange and transport were solved numerically with the Galerkin method. These authors showed some interesting features, such as multiple fronts and plateau zones during multi-species transport with a varying C T. It was also demonstrated how ion exchange and hydrodynamic dispersion influence solute transport.. Along the same lines, Valocchi et al. (192) presented an analytical framework for transport of various species of ion exchanging solutes, based upon the theory of chromatography. In contrast with some previous work in this' area, the effect of hydrodynamic dispersion was included and the total electrolyte level was not assumed to be constant. The development of the transport equations is quite similar to that presented by Rubin and James (163). In case of n exchanging ions, the governing set of equations is: ac k as k a 2 a 0 atk OD k kJk=1) 2). . n(4-21) nt bR 2 Va 72 C T is the variable total solute concentration in.the liquid phase and S is the constant, total solute concentration in the adsorbed phase or T the cation exchange capacity. In case an adsorbed species k, with valence uk is exchanged for species j, with valence uj, the exchange coefficient can be defined as: Kjk = (Y/X.)k (Xk/Yk)k j (4-24) where Y and X again denote dimensionless concentrations. It should be noted that a more accurate solution might be obtained by using chemical activities. Furthermore, the assumption that Kjk is constant is strictly not correct. To solve the transport equations, we need to reduce the number of dependent variables, by expressing S k in terms of C k . This can be done by using the n-1 independent equilibrium expressions, and by using Eq. (4-23). The multi-component exchange isotherm for a species k is given as: Sk = Fk(C 1 C 2 ,. C ) k=1,2,....,n (4-25) Eq. (4-25) can be substituted into Eq. (4-21), which leads to a set of n transport equations of the following form: 8 Ck n ac. a2Ck 8 Ck(4- it bEjkD 2tJDx(x4-26) k=1 a 8 x 2 aFk ask where fJk = aC. -C." This system has been solved for binary and 3 3 ternary exchange using the Galerkin finite element method (e.g., 163). 73 As was mentioned earlier, the value of Kjk is important as it affects the shape of the front. For a concave isotherm, Kjk< 1 the Xk-profile travels at a speed proportional to t instead of t , 1/2 as for Fickian dispersion. For a convex isotherm, Kjk>1, the Xk-profile becomes steeper when the front travels through the medium. Although the situation with a varying CT is somewhat more complex, one can assume that each particular concentration travels in a constant CT environment. A qualitatively good comparison of theoretically predicted concentrations, using multi-component exchange, with experimental results, involving the monitoring of Ca and Mg in a ground water aquifer, was reported by Valocchi et al. (192). Finally, these authors reported that dispersion induced exchange, a phenomenon reported by Lake and Helfferich (108), which occurs as a result of dispersive mixing across the solute front in case C T varies and which leads to changes in the adsorbed phase, was of minor importance. Selim et al. (169) examined the concentration of 2,4-D during infiltration and redistribution. This was one of the first studies concerning reactive solute movement during transient, unsaturated flow. Numerical solutions were obtained with an explicit-implicit finite difference technique. The dispersion coefficient was obtained as a function of pore-water velocity and the retardation factor was described similarly as by Lai and Jurinak (106, 107). The concentration profiles obtained from a field experiment, involving the application of 2 cm of an aqueous solution containing 50 ig/ml of 2,4-D, were in good agreement with the simulated profiles. 74 Although the retardation factor, R, is not constant for non-linear exchange, one may want to use an effective, constant value for R which can be used over the whole concentration range. Valocchi (190) introduced an effective distribution coefficient, Kd, as the ratio of the step change in aqueous phase concentration and the step change in sorbed phase concentration. The effective velocity of the solute front can be given by (Eq. (4-2)): v =v/R = dx /dt (4-27) p which provides a way to determine a (constant) value for R. The use of a constant value for R is plausible for a number of cases. For "trace" quantities of the species (i.e., C.1.1 s, D was independent of velocity under the conditions stated by Smiles and Philip (178). Groenevelt et al. (72) studied dispersion in a dispersed clay paste, using a slit model to represent the flow between clay platelets. The dispersion coefficient was found to be virtually velocity independent and could be equated to the coefficient of molecular diffusion. It should be noted that in their case the Peclet number (Pe=wv/D where w is half the width of the slit), was also very small (<1) because of the absence of large pores. Findings by Watson and Jones (212), who conducted similar experiments as Smiles et al. (179), seem to indicate that hydrodynamic dispersion is velocity dependent. In the literature, a linear dependence of D on v has often been reported for (steady) flow at higher values of the Pe-number. Yule and Gardner (219), for example, reported a longitudinal dispersion coefficient for a Plainfield sand which was linearly related to v. At large Pe-numbers (e.g. Pe>35) it is plausible to assume a linear relationship between v and D (cf. 15, 82, 206). Aggregated Media The study of transport in aggregated media has received considerable in recent years. Agronomists seek to optimally use irrigation water and applied chemicals, which might be easily lost in aggregated soils because of "bypass" flow. Chemical properties, such as pe, p11, and pO 2 , might be quite different for inter- than for 83 intra-aggregate pores, leading to more complex sink/source terms. The performance of ion exchange columns is the topic of many studies in the chemical engineering literature (e.g., 155). Transport in these columns is usually described in terms of a transport equation containing a sink/source term which describes the exchange inside the solid particles of the column. This section will examine how the transport equation is formulated for aggregated media. Passioura (135) distinguished between micropores, inside the soil aggregates where solute movement occurs only via diffusion, and in macropores, located between the aggregates where transport occurs via viscous (advective) flow. The latter type of transport can be completely dominant, and considerable work has been devoted towards a better understanding of flow in macropores (e.g., 18). The liquid present in the micropores is supposed to be immobile or stagnant, whereas the liquid in the macropores is said to be mobile. Passioura (135) used the following equation to describe transport in an aggregated medium (cf. Eq. (2-44)): o. aC.. a 2 C ac im Im + mo =D mo mo (5-2) o at at mo 2 mo ax mo ax where D is the dispersion coefficient of the mobile liquid phase and mo where the immobile phase, inside the aggregate, is treated as a distributed sink. In general, C and C. are not known and one has to mo 1im resort to a simpler model (Eq. (2-5)) and use an effective value for D. An expression for D should combine the advect ive dispersion in the mobile region with the stagnant phase effect in the immobile region. 84 Therefore, Passioura (135) adapted the expression of Aris (9) for a dispersion coefficient in a tube, to obtain the following expression for "overall" dispersion in an aggregated porous medium: D =D 0 /e + gv 2 a2D ( mo mo T ma/D .(5-3) where a is the radius of the aggregate [L], g is a coefficient characterizing shape, 0 T is the total (liquid) volume fraction available for the solute, 0 is the volume fraction of the mobile mo liquid phase, and Dei is the coefficient of molecular diffusion in elm the immobile liquid phase [L2T -I ]. Following Taylor (189), one can introduce a moving coordinate =x-vt. For large t, steady-state transport will occur so that C -C. = mo im constant. As advective dispersion can be neglected for a plane moving at velocity v (g=constant), the solute flux across such a plane is: 0. 22 aC im v a mo dif 0 T 15D 8(5-4ax where Passioura (135) erroneously omitted the negative sign, and the expression of Crank (45) was used for the concentration inside the spherical aggregates. Because the steady-state assumption implies that aC /ax=aC/ax, the following expression for the effective dispersion mo coefficient inside the aggregate was suggested by Passioura (135): 0. 2 2 D. - m va (5-5) im T 15D eim The effective dispersion coefficient for the liquid in the macropores, which characterizes advective dispersion, was described with (cf. Eq. (3-44)): 85 D =D + v a (5-6) mo emo mo where D is the coefficient of molecular diffusion in the mobile emo 2 -1 phase [LT ] and v the velocity in the mobile phase (=vOT / m). This mo T mo leads to the following overall dispersion coefficient: D = Dmo mo/T + Dim (-7) Passioura and Rose (136) performed experiments to evaluate D according to Eq.(S-7). Water retention data were used to estimate 8 im, which is approximately the volumetric water content at a suction of 75 cm of water. Pores which drained at suctions less than 75 cm were considered to be macropores. D was obtained via the technique described -6 -6 by Rose and Passioura (160). Values for Dei . between 4x10 and 6x10 elm 2 -1 cm s were used, depending on the porous material. It is of interest to examine the Brenner number (B) as a function of va. For low values of va, the main contribution to B will be from advective dispersion, whereas at higher values of va the stagnant phase effect becomes more important. Therefore, we will have somewhere an absolute maximum for B. 1/2 This maximum will occur when va=(1D D 0 /8. ) . Displacement is elm emo mo im likely to be most efficient, i.e., closest to piston flow, for this value of va. The study of transport during flow of other fluids than water might provide independent ways to characterize soil structure. Millington and Shearer (118) discussed the effect of aggregation in porous media on gas diffusion. These authors distinguished between a solid phase and an intra- and inter-aggregate porespace filled with 86 waiter or gas. Various expressions for D eIDo were derived, where De and D0 are diffusivities of the fluid in presence and absence of the porous medium, respectively. Millington and Shearer (118) illustrated the contrasting behavior of diffusion in the gas phase and ion diffusion in the liquid phase. The ratio De /Do was found to be higher for non-aggregated than for aggregated media as far as diffusion in the liquid phase is concerned, while the opposite seemed to hold for gas diffusion, except during very dry conditions. Scotter (166), studying the preferential solute movement through larger soil voids (cf. 174), considered two pore geometries: cylindrical channels and planar voids. Theoretical breakthrough curves for a non-reactive solute showed that a substantial amount of the solute appeared in the effluent before one pore volume had leached through, especially for large channel diameters. Van Genuchten et al. (201) used the development by Scotter (166) to solve transport through cylindrical macropores analytically. The cylindrical pore 'is supposed to contain the mobile region of the liquid phase, while the surrounding porous medium contains small pores in which the immobile region resides. The adsorption sites were divided into a fraction in close contact with the mobile liquid phase, and a fraction in contact with the immobile region. A separate retardation factor was used for' both fractions. A set of equations similar to Eq. (2-44) and (2-45) was 87 furthermore, dispersion in the macropore can usually be Ignored without loss of accuracy. The approach taken by van Genuchten et al. (201) seems to be generally accepted to describe solute transport in aggregated porous media. The overall transport equation contains both the concentration in the mobile and immobile phase. The geometry of the aggregates influences the diffusive transport between the two regions as well as the average solute concentration inside the aggregates. Valocchi (191) studied the validity of the assumption of local physical and chemical equilibrium by comparing solutions of the non-equilibrium and equilibrium type transport equations for aggregated media. Two cases of physical non-equilibrium were considered, namely diffusion in spherical aggregates and first order transport between mobile and immobile regions of the liquid phase. The general equation for reactive solute transport, assuming linear adsorption in both mobile and immobile region, can be written as: ac ac. a2c aceR mo iR m mo mo 0 R mot+0 mR =0im =0 D -0 v (5-8) momoat imim at mo 2 mo 8x 8x For spherical aggregates the physical non-equilibrium diffusion model is: a Ci -- (x,r,t) r 2 dr (5-9) J0 Ca(x,r,t) r= C (x,t) (5-10) AC AC Rm a =D 1 (2 a) (5-11) im~t eim 2r Ar r 88 where a is the aggregate radius and C is the local concentration a inside the aggregate. In case the transfer of solute between mobile and immobile zones can be described as a first order process we get: ac. 0. R(C - CRm)_(5-12) im im at mo im The transport problem is thus posed by Eq. (5-8) in combination with Eq. (5-9) to (5-19) or in combination with Eq. (5-12). One case of chemical non-equilibrium, involving first order kinetics, was studied by Valocchi (191). The ADE was described as 0 ac as a 2 c ac (-30 - + P b at -OD 2 ev a (5-13) ax 2 where. S is governed by the first-order linear kinetic expression as - k 1 C - k 2 S (5-14) with k and k as the forward and reverse rate coefficient, 1 2 respectively. This is referred to as the linear chemical non-equilibrium model. These models are well documented (150, 151). They were written in dimensionless form and solved in the Laplace domain by Valocchi (191). In order to quantify deviations from local equilibrium, a time moment analysis was carried out. Aris (10) showed how absolute time moments, m which are defined analagous to moments for position dependent concentrations (Eq. (3-18)), can be determined with the solution in the Lapl ace domain: m= (-1)nl d- c(x s)] (5-15) p s L ds p 89 where X is the dimensionless position and C(X,s) is the Laplace transform of C(X,t) with s as the transformation variable. The advantage of this approach is that the solution in the Laplace domain is fairly easy to obtain. Based on these moments, Valocchi (191) obtained expressions for the first normalized moment and the second and third central moment for the equilibrium and the three non-equilibrium models for a Dirac-type input. In this way, mean breakthrough time, spreading, and tailing could be characterized. It was found that sorption rate limitations (non-equilibrium) caused enhanced spreading and tailing. Non-equilibrium, however, does not influence the mean breakthrough time. At large values for the dimensionless rate parameter, F=k2L/v, where L is the column length, 2 the non-equilibrium and equilibrium results are similar. By comparing the equilibrium solution with the more realistic non-equilibrium solution, the error, which is made by using the local equilibrium solution, can be established. This error increases linearly with incresing Pe/F and decreases with increasing retardation factor. Following Passioura (135), Rao et al. (151), and de Smedt and Wierenga (57), Valocchi (191) obtained effective dispersion coefficients for the non-equilibrium situations. Before discussing transport in fractured media, which is very similar to transport in aggregated media, a few other studies should be mentioned. Rasmuson and Neret ni eks (153) obtained an anal yt ical solution for the ADE in a porous medium consisting of spherical particles. Rasmuson (152) extended this analytical solution to a case 90 where two-dimensional dispersion occurred. Other solutions involving spherical particles were given by Skopp and Warrick (175) and Tang et al. (186), while Sudicky and Frind (184) provided solutions for porous media with rectangular voids. One of the first attempts to combine analytical and experimental work was reported by Rao et al. (150), who studied transport in a medium consisting of spherical aggregates. Fractured Media Because of interest in the disposal of hazardous wastes in fractured bedrock, geologists and engineers have lately been studying transport in fractured media. Fractures are important for the advective transport of contaminants because of their relatively high water conductance (bypass). Formulation of the transport problem is very similar to that for aggregated media, i.e. , the two components are advective transport in the fissure (the mobile region) and diffusion into the rock (the immobile region). For a number of cases, analytical solutions are available. In other instances, numerical methods need to be used. The finite element method is particularly convenient to obtain, numerical solutions for transport in fractured media (86). Tang et al. (186) presented an analytical -solution for contaminant transport in a single fracture. They considered a radio-active contaminant with a finite migration distance because of decay. Diffusion of the contaminant into the porous rock matrix slows 91 the fracture, (3) molecular diffusion within the fracture, in the direction of the fracture axis, (4) molecular diffusion from the fracture into the porous matrix, (5) adsorption onto the face of the matrix, (6) adsorption within the matrix, and (7) radio-active decay. For solute transport in the fracture, the following equation was used: :C 82C OD . C. C mo D mo + eim Im v mo -t R - C + - (5-16) at R 2 w mo bR 8x R 8z mo az mo x=w mo -1 where 2w is the fracture width [L], pw is a decay constant [T 1, R =1 w mo + - K is a face retardation factor, and K is the distribution 0 dmo dmo coefficient in the fracture. For transport inside the matrix the equation was: ac. D . d2C. im _ elm im C (6-17) at R. 2 w im Im ax where D. is the effective diffusion coefficient in the matrix, R. = 1 elm im + K is the matrix retardation factor, and K is the 0 dim dim distribution coefficient in the porous matrix. In both cases the adsorption was assumed to follow a linear adsorption isotherm. Equations (5-16) and (5-17) were solved analytically by Tang et al. (186) for specified initial and boundary conditions with the help of Laplace transforms. Some numerical examples were given as well. It was found that the effect of dispersion cannot be ignored. If D=0 2 -1 m s , the solution lags substantially behind the general solution, which used an expression for D as Eq. (3-36). This is in agreement with the findings of Rasmuson and Neretnieks (154) who also investigated the migration of decaying radionuclides in fissured rocks. They found 92 that advective dispersion in the fissures results in larger travel distances for the radionuclides, therefore decreasing diffusion into the matrix and subsequent decay inside the rock matrix. Agreement between the analytical solution by Tang et al. (186) and a finite element method solution by Grisak and Pickens (68) was generally good, except for the transient case. The results demonstrated that matrix diffusion can prevent severe contamination of underlying aquifers with the (decaying) radioactive substances. Neretnieks (119) treated dispersion in fissured rock, idealizing the medium by assuming flow through parallel channels of different size. The fissure width, w, is given by a distribution function f(w). A pulse with concentration C 0 , introduced at the inlet, travels a distance x in time t. in fissures with width w.. Therefore, w can be 11 expressed in terms of t. If the residence time for a fissure is less than a certain time t, it delivers tracer. At the outlet, mixing among solutions exiting from all fissures occurs, resulting in the following dimensionless exit concentration: T f(w)Jv(W) dw C(t)/Co = wt - J(t)/J v (5-18) 0 00 V V T f(w)Jv(W) dw V 0 The flow of water containing tracer, Jv(t), from fissures with widths V w(t) = C f(z,t,C) dC (6-3) JO 2C (z,t)= (C-) f(z,t,C) dC (6-4) C J 0 where is the average concentration over the entire field in a given x,y-plane. The randomness of the flow term at the input boundary (rainfall, irrigation) was taken into account by introducing a random leaching rate, which is linked to the random behavior of the hydraulic conductivity to give a probability distribution, P(v), for the pore water velocity. Because dispersion/diffusion effects are neglected, the concentration profile obeys a step function (C(z,t)=O or 1 for z>vt and z= 1 - P(z/t) = 1 - P(v) (6-5) where P(v) is expressed in terms of variables x.. The field scale dispersion, caused by random changes in 0 and v across the field, exceeds the mechanical dispersion at the pore scale. If desirable, the 108 pore scale dispersion could be taken into account by introducing an 'effective' dispersion coefficient to make up for vertical heterogeneity. Dagan and Bresler (49) argued that the large transition zone, (0< <1), which is commonly encountered under field conditions, can now be described correctly. When using the advection-dispersion equation an unrealistically high value for the dispersion coefficient would be needed. Bresler and Dagan (30) compared the solute spreading effect,. due to heterogeneity of soil properties, with Taylor's (189) theory of dispersion during flow in a capillary tube. The analogy is that (transverse) diffusion initially cannot annihilate the longitudinal concentration gradient caused by the advective term. However, for larger times, diffusion becomes the dominant process. For conditions normally encountered during infiltration, they concluded that only the advection term is needed to describe the transport process, but that for very large travel distances, e.g. 100 m, dispersion should be included. The model derived by Dagan and Bresler (49) was used by Bresler and Dagan (30) to predict solute concentrations for various rates of infiltration using data of Warrick et al. (210). For a travel distance of 1 m, an effective dispersivity between 16.1 and 334.2 cm was found. If one had used the ADE, assuming Fickian dispersion, the transition zone (the width of the solute front)would grow proportional with t / . However, Bresler and Dagan (30) found this zone to grow linearly with t. It should be noted that the work by these authors did not include an independent field study to validate the model by comparing measured and calculated solute distributions. 109 Bresler and Dagan (31) added pore scale dispersion, as measured in- the laboratory, or the much larger average field value for dispersion, which reflects mainly heterogeneity of the profile, to the previously mentioned mechanisms of spreading (horizontal heterogeneity of' soil properties and a variable infiltration rate). They concluded that the use of the conventional ADE, with constant coefficients, poorly predicted in a heterogeneous field. To model transport, one should focus on the variability of the hydraulic conductivity, the average rate of recharge, and the value for the (average) field dispersion. The influence of laboratory-measured pore scale dispersion is negligible, and the variability of recharge on the soil surface, provided that it is of modest value, had little influence on the predicted solute profiles. In a subsequent series of papers, a more general approach was given for the stochastic modeling of unsteady flow and transport. Dagan and Bresler (50) derived a model for infiltration and redistribution in a heterogeneous field, which was applied to two spatially variable soils (32). Bresler and Dagan (33) derived a model to predict solute transport using the results of their work on water flow modeling. Expected solute concentrations were calculated by using either an accurate numerical scheme or a simplified model, or by representing the variable field by an equivalent uniform field. The simplified model 110 hydrodynamic dispersion process, do not seem to have an appreciable improvement upon prediction of statistical moments in spatially variable fields. Using an equivalent uniform medium resulted in a much larger error. In the simplified model, the solute concentration was predicted with statistical moments rather than deterministic values. The advantage of this procedure is that, although it might not accurately predict concentrations at specific positions, the averaging procedure resulted in reasonably accurate concentrations for an entire field. This is often of more interest than specific values at given locations. Non-mechanist ic Models A recent approach to model solute transport in the field ignores the actual mechanism of flow and transport and merely focuses on the solute concentration as a result of these unknown processes. In this "black box" approach, the transport and flow processes are represented by a transfer function, which characterizes systems whose internal mechanisms are unknown or unknowable (90). This approach has been used for some time by hydrologists and engineers (51). Raats (147) applied the concept of transfer functions to describe advective solute transport during steady flow in soils. He described the movement of a certain water parcel in the course of time by focusing on the travel time, usually of primary interest in contamination studies. The position, x, that a parcel, , occupies at time t can be given by: x = x (,t) (6-6) 111 with the velocity for X: v = (6-7) Denoting s as the arc-length along the streamline [L], Raats (147) characterized v by its magnitude, v, and unit tangent vector, T: V =a (6-8-a) 4 = vi/v (6-8-b) The travel time between two points on the streamline was given by: t-t = i ds (6-9) s 0 where s and s are an arbitrary and a reference position, respectively. 0 For steady flow without sinks and sources, Raats (147) obtained the following alternative expression for the travel time, making use of Eq. (6-8-b): s s t-t -1 expT ds ds (6-10) 0 (Ov) [ j wher (O) 0is the flux at s 0 Raats (147) noted that the divergence of the unit tangent vector' field, V-T,. is a characteristic of the flow pattern and a measure of the divergence or convergence of infinitesimal stream tubes. For such a stream tube, Raats (147) obtained: A(s) = Aexp {V- ds (6-11) 00 derived: 112 1 s t-t (ev)A0 Ads (6-12) 0 The integral on the right-hand side corresponds to the total amount of water in the stream tube between s and s, whereas (Sv) A is the 0 00 steady input flux. Eq. (6-12) gives the ratio of these quantities, being the travel time between s and s . Having established a link between the 0 exact analytical expression for the travel time in terms of r, and an expression for approximate graphical analyes, viz. Eq. (6-12), Raats (147) proceeded by considering more specific flow problems. After evaluating the travel time for one parcel, the movement of collections of particles forming a surface was studied. This is useful to examine the effect of solute application as a pulse or step front. In order to predict whether the solute passes a certain point, e.g., the outlet of a soil column, the time needed to pass the column (transit time) and the time the solute resides in the column (residence time) need to be known. Therefore, Raats (147) used expressions for the transit time, T, for each streamline and the residence time for a parcel in the system. In order to study the effect of a solute application, one can follow a collection of particles which enter the system simultaneously, forming an isochrone. Obviously, the frequency distribution of transit times in the medium, consisting of various tubes, is of importance. It is also important to determine the contribution of each tube to the total transport. This can be done by monitoring the output concentration as a function of time. The cumulative transit time distribution function, q, 113 is defined as the fraction of the streamtubes with transit times smaller than T. At any time, a fraction q of the output will be "younger" than z and a fraction (l-q) will be "older" than T. The transit time density distribution function is equal to dT/dq, i.e., the frequency distribution of the volume of the streamtubes. It takes into account the rate of solute movement in a particular tube and the contribution of that tube to the total solute transport. Finally, the transfer function is the product of the input and transit time density distribution function. Raats (148) applied the theory to a number of flow patterns. Jury (90) determined the solute concentration at various depths in the soil by means of an output function, which was the result of a solute flux applied at the soil surface, i.e., an input function. These two functions were related by using a transfer function, which could be determined by using measurements of the solute concentration at a particular depth. The concentration at other depths could be predicted with the transfer function. A brief derivation of the transfer function follows. Instead of using time as the independent variable, Jury (90) used the cumulative amount of surface-applied water, I, as the independent random variable for solute transport at a particular location (214). The probability that a tracer injected at the surface will reach a depth L after an amount of water I has been applied, is L fL()') dI' (6-13) 0 114 I' is a dummy integration variable and fL(I) is the probability density function, denoting that the injected tracer will arrive at depth L upon application of an amount of water between I and I+dI, i.e., a distribution of travel times. The inlet concentration is given by C. =C =8(I), a narrow pulse injected at I=O at the soil surface. The in o average concentration, at z=L, for arbitrary variations in C. and in spatially uniform water input is given by: CO C(I) = C (I-I ) f(I') d' (6-14) where f (I') is the probability of the solute reaching z=L between L 'time' I' and I'+dI', i.e., the cumulative infiltration, and C. (I-I') in is the concentration discharging at depth L if for a 'breakthrough time' I'. Note that this concentration remains equal to the inlet concentration, because front spreading is ignored. The rate of water input is not uniformly distributed over the field, therefore a second probability density function, g(i), is needed to describe the variabilty of input rate for various locations. The relation between time and cumulative infiltration follows from i=dI/dt. The probability that a unit area of soil will receive a water input flux between i and i+di is ig(i). The probability that the solute reaches a depth z=L between t and t+dt depends on the soil properties (via fL (I)) and on the input rate (via g(i)). This probability is given by the travel time density function, hLCt), which is the joint probability function: 115 hL t) = Jig(i)fL(it) di (6-15) %0 The average concentration at a given depth L, for a spatially uniform input concentration C. t), follows from: in Cout Ct) = {C in (t-t' )hLt') dt' (6-16) where t' is a dummy integration variable. Jury (90) assumed that the distribution of physical processes contributing to the density function, fL(I), between z=0 and z=L is equal for all depths. In that case, the probability that a tracer will reach a given depth z, after a cumulative infiltration 1=I, is equal to the probability of reaching a reference depth z=L after a cumulative infiltration of I=I1L/z: IL/z Pz M = PL(IL/z) = { fLI') dl' (6-17) 0 and by combining Eq.(6-15) and (6-16), the concentration for a spatially variable water application is given by: 00 00 C(znt) { C.(t-t) {L ig(i)fL(it') di dt' (6-18) 0 n 0 for a water application I=it. To obtain f a calibration at one depth is necessary. For this purpose, Jury et al. (93) obtained soil solution samples by vacuum extraction at a depth of 30 cm at 14 locations. Usually a lognormal distribution seemed to fit fL(I) , but other functional re lat ionships could be used as well. Once fL( I) was obtained, the performance of the model was tested by comparing predicted values of C (z I) with measured values at other depths than out for which the model was calibrated. Good agreement was found between 116 results obtained by using the transfer function and experimental values. It is not clear how well the model works for stratified soils. The model also used for unsaturated conditions (92). Concluding Remarks A review of some modeling approaches was presented by Addiscott and Wagenet (4), who concluded that among the many models available, few have been exhaustively tested under field conditions. Tests by others than the original authors, under different circumstances, have been rare. Most analytical and numerical models are deterministic and mechanistic by nature. Because of the spatial variability of soil properties in the field, stochastic models seem to be more attractive. One such stochastic model, involving transport of a reactive solute, was presented by van der Zee and van Riemsdijk (193). For fundamental laboratory studies, the mechanistic deterministic model is the most appropriate. Combination of the deterministic ADE with a stochastic flow model could be a useful approach in future field studies. Of course, many more efforts have been made to model water flow for field conditions. Some involved scaling (180), others incorporated geostatistics or the theory of stochastic processes (99, 100, 114, 208, 216, 217, 218). Charbeneau (38) applied the kinematic theory, i.e., a functional relationship between flux and water content. A similar theory was developed for solute transport. The kinematic waves can be distinguished in self sharpening, during infilitrat ion, and self spreading, during drainage. 118 Although much of the material already discussed includes transport during unsaturated flow, some specific work appearin in the soil science literature is discussed in this section. In particular, some of the laboratory experiments are reviewed. A distinction is made between steady state and transient flow conditions. A few approaches to control the flow and/or water content for steady state conditions will be mentioned as well. Steady Flow One of the first results on miscible displacement during unsaturated flow was reported by Nielsen and Biggar (120). They observed that the magnitude of water not readily. displaced, the hold back, increased with desaturation. The inclusion of dispersion phenomena in the transport equation seems more appropriate for unsaturated than for saturated flow conditions. Biggar and Nielsen (20) emphasized the role of, diffusion, pointing out that, although mechanical dispersion will decrease at lower velocities, the process of molecular diffusion tends to enhance spreading. Krupp and Elrick (104) explained tailing during unsaturated steady flow with the 'stagnant liquid concept'. According to these authors, the dispersion coefficient is not related to 0 in a simple manner. Variation in the sequences of fully filled and part ially filled pores and transport in the surface films contribute to spreading. 119 the dispersion coefficient accurately in this case. Also, at very low water contents, the variance in pore-water velocity is likely to vanish since there is no pore sequence- for liquid flow. Thus, the amount of spreading is reduced (104). Yule and Gardner (219) studied both longitudinal and transverse dispersion under unsaturated conditions. Transverse dispersion, i.e., diffusion, gains importance during unsaturated flow due to the lower pore water velocity. The value for the transverse dispersion coefficient was found to be nearly independent of pore water velocity. Van Genuchten and Wierenga (202) noted that more tailing is to be expected in unsaturated sorbing media. Not only will the relative amount of stagnant water increase, the fraction of sorption sites located in the stagnant region will increase as well. Among others, they used the concept of a mobile and 'immobile region of the liquid phase to formulate the transport equations for unsaturated soils. De Smedt and Wierenga (57) observed early breakthrough and prolonged tailing during unsaturated condit ions. Dispersion coefficients obtained from saturated displacement experiments, with the same pore-water velocity, could be used to fit the results from unsaturated experiments if the transport equation accounted for mobile and immobile water. When this distinction was not made, dispersion coefficients many times larger than for saturated conditions were needed to fit the 120 qualitatively explained the asymmetry of the experimental concentration profile. By varying the velocity and water content independently, using different gravitational heads, Gupta et al. (78) found that D increased with v and g enerally decreased with increasing 8. However, no functional relationship was derived. Awad (11) examined dispersion phenomena i n a medium fine sand. He found that the hydrodynamic dispersion coefficient appeared to increase with decreasing water content, but found no basis for the presence of a mobile and stagnant region of the liqui d phase. A theoretical investigation of the relationship between the hydrodynamic dispersion coefficient and water content was pursued,. but not established. Transient Flow This section briefly reviews transport during infiltration, the development of leaching strategies and the exploration of the dependence of D on v. Wierenga (214) showed that transient flow could be represented by an effective steady state flow to describe solute transport. This simplifies considerably the task of obtaining a correct flow term to predict solute movement considerably (38). Kirda et al. (95, 96) determined the displacement of chloride during infiltration with chloride-free water in soil columns, both experimentally and theoretically. For a. given amount of cumulative infiltration, the Cl salt was leached more efficiently (i.e., with less 121 the salt than larger, less frequent water applications. Warrick et al. (209) observed that the water content at the soil surface determines the propagation of the solute present in the irrigation water, while the influence of the initial water content in the remainder of the soil profile was shown to be small. For equal quantities of water infiltrated, the depth of the maximum solute concentration increases with decreasing surface water content. Smiles et al. (179) investigated the absorption of a KCl solution by initially rather dry soils. They observed that for a low Peclet-number, the dispersion coefficient, D, was virtually s velocity-independent and that piston displacement of the solute occurred. Smiles and Philip (178), using low initial salt and water contents, noticed that the C1 concentration profile was somewhat asymmetrical due to changes in the advection-term across the front. However, the prediction of the solute front was relatively insensitive to changes in D . Quantitatively, diffusion and dispersion are s identical. Smiles et al. (177) transformed the transport equation to describe (transient) absorption by using a constant surface flux v . 0 2 The solution, expressed in terms of position, v0x, and time, v t, was unique, suggesting that D is velOcity-independent. Smiles and Gardiner 5 (176) also observed close to piston displacement of the solute in a clay soil. Thin films of water on the clay surface were not accessible to the invading solute. The Cl front was located ahead of the piston front, which was attributed to anion exclusion. Bond (25) analytically solved transport of a solute, applied as a pulse, during unsteady flow. 122 He used a velocity dependent D. Experimentally and analytically determined curves matched fairly well for 3 H 0 but not for Cl. No 2 explanation could be offered for the poor prediction of the Cl profile. The initial volumetric water content was 0.176. Finally, Grismer (69) studied the absorption of water containing Nal or SrCl2 by a silt loam. The dispersion coefficients, determined according to Smiles et al. (179), depended on 0 in a way similar to the diffusion coefficient. C omplete displacement of the initial water was observed, although for the initial water content of 0.124, a higher flux was required than for an initial water content of 0.035 to achieve complete displacement. Apparently, solute transport during steady flow differs from transport during transient flow, at least for infiltration processes. In the latter case, the dispersion coefficient is reported to be virtually independent of the velocity and a complete displacement of the resident solution occurs (piston displacement). Since the solute acts as a tracer of relative water. movement, it is assumed that the same holds for the flow process in general. Experimental Considerations A. major task when obtaining- breakthrough curves for steady, unsaturated flow conditions is the independent control of water content and pore water velocity. Because both factors influence dispersion, it 123 The driving force for water flow is a hydraulic head gradient, VH, with H given by: H = h + z (7-1) where h is the pressure head [L] and z the gravitational head [L]. The osmotic head is ignored in this expression. The flow of water is given by Darcy's law: JV = -K(O) VH (7-2) where K(e) is the hydraulic conductivity [LT I. It follows from Eq.(7-1) and (7-2) that the water flow can be manipulated directly by adjusting h or z, or indirectly via the water content which influences K and h. Nielsen and Biggar (120) used hanging water columns and a negative air pressure at the outlet to obtain a constant average water content and flowrate. The glass bead medium required a change in head of only 3 cm of water to obtain identical flowrates at full and approximately half of the saturated water content. Fritted glass plates, which have a negligible exchange capacity, a large capillary conductivity, and a narrow pore size distribution, were used to maintain unsaturated conditions. Elrick et al. (58) and Krupp and Elrick (104) regulated the water content by placing a perforated sample holder in a pressure chamber, while the desired flowrate was obtained with a constant volume pump and a head tube. Both ends of the sample were in close contact with a cellulose acetate membrane filter, supported by a porous screen and a 124 stainless steel end cap containing three plastic nipples. The sample and pressure chamber were placed on a balance to allow continuous monitoring of the sample weight and hence water content. A driving head, equal to the sample length, was established by increasing the air pressure in the chamber(unit gradient). The gradient in pressure head was equal to zero; a constant water content was obtained throughout the sample. This meant that only one flowrate was possible for each water content. Yule and Gardner (219) stated that variations in the hydraulic gradient can only be achieved by variations in water content. This is not quite true, as will be discussed shortly, but it accurately points out the problem. Their experimental setup consisted of a rectangular plexiglasss soil container with an inflow and outflow control section. Each section consisted of 11 one-bar porous ceramic tubes. At the inlet, the inflow in each tube was controlled by a Mariotte bottle and at the outlet, the outflow from each tube was collected in a plastic vial maintained under suction controlled with a bubbling tower. A unit gradient (gravitational head) was established to obtain a uniform moisture content throughout the column. Gupta et al. (77) used tilted soil columns to study dispersion phenomena in an unsaturated glass bead medium. Solutions with different KCl concentrations were applied with a pump. The water flow could be 125 Awad (11) used a two phase steady-state flow system to obtain independent control of the volumetric water content and pore water velocity. A horizontal system was used. Awad (11) proposed to use changes in the pneumatic head to vary the flowrate. The pressure head is given by: h =h + h (7-3) m a where h is the matric pressure head, which governs the volumetric m water content, and h is the pneumatic head which does not influence a the water content (101). Hence, by manipulating ha) the flow can be regulated (with a pump) without affecting the water content. A flow cell was constructed with fritted glass plates. The pneumatic head was regulated by maintaining a gradient in air pressure in the medium between the plates. A constant flowrate of water was established with a pump. Water content and bulk density were measured with a gamma attenuation system. The hydraulic head was obtained by using tensiometers connected to pressure transducers. The method could not be used at very low water contents, due to the occurence of a high resistance boundary layer between plates and medium. VIII. MULTI-PHASE TRANSPORT A logical extension of the previously discussed unsaturated transport is the study of transport of immiscible fluids. Multiphase transport involves vapor and liquid phase or various liquid phases. There is a need to study these processes in order to predict movement of (volatile) substances such as pesticides, oil products, and other (organic) compounds. First, pesticide movement in liquid and vapor phase will be discussed based on the work by Jury et al. (91), followed by a brief mention of research in the area of immiscible liquid transport. Transport in Vapor and Liquid Phase Jury et al. (91) determined the loss of pesticide via volatilization and diffusion in the vapor phase and via advection and diffusion in the liquid phase. Their investigation involved the movement of triallate in soil samples placed in a volatilization chamber at relative humidities of 50 and 100%. The vaporized triallate was collected on polyurethane plugs. The total concentration of the solute in the soil was given by: C t = PbS +%0 Ce +O09C (8-1) g g where C is the total concentration expresses as mass of solute per t -3 volume of soil [ML ], S is the adsorbed concentration, expressed as 126 127 mass of solute per mass of soil [MM_], C and C9 are liquid and gas concentration, respectively, expressed as mass of solute per volume of fluid [ML -], and 0 and 0 are the liquid' and gas content, 3 -3 respectively, expressed as volume of fluid per volume of soil [L L The solute flux was expressed as ac 9 ac, JS -Dg ax -etax JVct(82 where Js is the solute flux, D9 is the gas diffusion coefficient [L 2T ],p D is the liquid diffusion coefficient [L 2T_ l1and J is the, t V 3 -2 -1 volumetric water flux [L L T I. It should be noted that advective transport in the vapor phase was considered to be negligible. Under some circumstances, -such as non-isothermal flow, this might not be justified. The continuity equation for one-dimensional transport in the absence of a sink/source term is: at+ a S=0 (8-3) at ax Eq. (8-1) and (8-2) can be substituted in Eq. (8-3) to describe mass transport. In order to solve the resulting equation, it should contain only one dependent variable,, viz, a concentration in a particular phase. The following relationships are useful for this purpose: 1) Gas and liquid concentration are related by Henry's law 128 2) The adsorption isotherm is linear, which seems reasonable for the- trace amounts of solute encountered: q = aCt + b (8-5) Some additional assumptions were made by Jury et al. (91) to rewrite Eq. (8-3) in terms of C : g ac a 2 cac S g D _ g+ V g (8-6) at e 2 e ax ax where c = pbaKH+0K +0, De=Dg+KHD and ve=JvK For appropriate b H t H g e g H e V H boundary and initial conditions, Eq. (8-6) can be solved analytically for C . Of course, due to hysteresis and non-equilibrium, Eq. (8-4) and g (8-5) might not be representative. This would subsequently influence the accuracy of the solutions of Eq. (8-6). Jury et al. (91) considered two cases, viz. transport by diffusion only (Jv=O) and transport by advection-diffusion (Jv O), which obviously led to different solutions for C . An expression for g the solute flux at the surface, J (O,t), was found by using Eq. (8-2), 5 in conjunction with the solution for C and Eq.(2-4). The cumulative g loss was then obtained via integration of the flux over time. The experimentally measured cumulative loss was used to determine the effective diffusion coefficient, D . Knowledge of this D enables one e e to make theoretical predictions of the mass flow. It was shown that advection (evaporation) caused the vapor loss to increase slightly. Loss of' the pesticide was mainly due to depletion from the upper soil layer. Because of the increase in the concentration gradient, diffusion becomes more important close to the soil surface. 129 Transport in Immiscible Liquid Phases Recently, an increasing amount of research has been initiated in the area of non-aqueous phase liquids, NAPL (2, 129). The situation is slightly different for this type of flow than for flow in systems containing a single liquid phase. Advective transport plays a role in all phases, and knowledge of the flow properties for each phase is needed. Multiple phase flow is, among others, discussed by Scheidegger (165). Once the flow problem has been solved, an attempt can be made to describe transport. The distribution coefficients of the solute between the various phases, as well as in some cases the exchange and adsorption parameters, need to be known. In some instances, the immiscible fluids may consist of multiple components with distinct coefficients describing the distribution between the various fluids (44). Baehr and Corapcioglu (13) studied transport of the various components of gasoline for which they predicted quite different travel times. Numerical solutions for multiphase flow and transport can be found in Huyakorn and Pinder (86). Abriola and Pinder (2, 3) treated the simultaneous transport of organic compounds in three phases: the non-aqueous, aqueous, and vapor phases. These authors derived a system of three non-linear partial differential equations with five independent variables (i.e., two capillary pressures and three mass fractions). The capillary pressures occur as a result of interfacial tension. The mass balance equation for a particular species within a certain phase contains a storage term, an advective and autonomous flux, a sink or source term, and exchange 130 resulting from a phase change or interphase diffusion. The mass balance needs to be derived for all species in each phase. Often a number of terms can be neglected, like the advective transport in the gasphase as mentioned before. These authors did not consider the (ad)sorption of the solute to the solid phase. An implicit numerical model was used to solve the resulting system of non-linear equations with Newton-Raphson iteration. The study was restricted to one-dimensional tranport. The primary mechanism by which the volatile component was transported through the medium was via gas diffusion. Since it was assumed that local equilibrium existed between gas and liquid phases at all times, the concentration in the liquid phases changed accordingly. IX. EPILOGUE It is hoped that this review demonstrated that the problem of solute transport is a very broad one with many interesting aspects. Combined with the urgency of ground water pollution, it is therefore not surprising that a great deal of work has been done in this area. Obviously, it was impossible to cover all aspects of transport modeling and to discuss fundamental processes in great detail in a review like this. Instead, it was restricted to a discussion of some basic processes which affect transport. These processes are of a chemical, physical, and biological nature, which explains why often an interdisciplinary approach is needed (Chapter I). It should be noted that, because of thisinterdisciplinary nature, the terminology and the focus of the research vary widely. Some useful references were provided in the discussions on the basic transport equation (Chapter II and III). The reader is encouraged to consult the more advanced and comprehensive discussions in the literature. The principles of transport, i.e. using the principle of conservation of mass, can be applied to transport under "complicating" conditions: with non-linear exchange (Chapter IV), in structured and layered media (Chapter V), during unsaturated flow (Chapter VII), and for multi-phase systems (Chapter VIII). Although the fundamental processes are essentially the same, these (and other) conditions offer challenges for additional research. As our understanding of transport 131 132 progresses, new questions will undoubtedly arise. Because this report's objective was to formulate transport problems, or at least refer to it, relatively little attention was paid to the mathematical, in particular the numerical, solution of transport problems Many good references exist on the mathematical tools which are available to solve transport problems. The ultimate goal of much of the reasearch is geared towards a better understanding of solute transport in natural porous media. Some of the basic research might be helpful to solve "real problems," some might not. Chapter VI pointed out that many of the traditional deterministic models are of limited value. It is clear that modeling should be accompanied by measurements, which makes such studies very challenging for researchers because of the amount of effort involved (time, money, expertise). Furthermore, some of the research findings are not applicable under different conditions. To put research efforts in the area of solute transport in perspective, it is noted that many social and economical aspects are involved as well. Some pollutants can safely be dispersed in large water bodies, whereas others need to be contained. For some materials (e.g., radioactive waste), one might argue that nonproduction and nonuse is the only sensible approach. The short term economic benefits of, for instance, agricultural chemicals, do not warrant their use in view of possible adverse effects in the long term. Transport modelers can provide part of the information needed to make a rational decision. Finally, in light of all the difficulties associated with transport 133 modeling, it seems only logical to improve management practices in dealing with da ngerous chemicals and to educate potential developers, producers, and users. This requires input from researchers in the area of solute transport as well. NOTATION This section contains the description and units of most of the symbols used in this review. Because the notation in the literature is not uniform, occasionally some unusual symbols were used or symbols were used twice. Symbol Description Units a (aggregate) radius L ak chemical activity of species k B Brenner number (column Peclet number) C concentration of solute in liquid phase ML 3 C dimensionless concentration C(x,s) concentration in Laplace domain ML -3 Ca local concentration inside aggregate ML-3 C dimensionless exit concentration e C exit concentration ML -3 ex Cf flux-averaged concentration ML -3 f-3 C concentration of solute in gas phase ML -3 C. initial concentration ML -3 C. inlet concentration ML in -3 C concentration of solute in liquid phase (=C) ML C mean concentration ML m CimCmo C for immobile and mobile region ML -3 mo3 C molar concentration of species ML C mole fraction of species k (C k/C M M Mk M-3 C concentration of feed solution (eluent) ML - C outlet concentration ML 3 out C t total mass of solute per bulk volume ML 134 135 Symbol Description Units -3 C t time average of C ML C T total concentration in liquid phase ML -3 C V volume averaged concentration ML d depth or characteristic particle size L D,D coefficient of (hydrodynamic) dispersion L2T - 1 D effective (retarded) value for D (=D/R) LT 1 D coefficient of mechanical dispersion LT - dis D (effective) coefficient of molecular diffusion LT - D D for immobile region LT eim e D D for mobile region LT emo e 2 -1 D gas diffusion coefficient LT g 2-1 De liquid diffusion coefficient LT 2 -1 D L coefficient of longitudinal dispersion L2T D. ,D D for immobile and mobile region L2T imt mo 2-1 D coefficient of molecular diffusion in the L2T free liquid D (O) dispersion coefficient (=OD(e)) LT - D T coefficient of transverse dispersion LT - D (0) soil water diffusivity LT - f probability density function f (I) probability density function for movement in a soil as f fraction of sorption sites in direct contact with the mobile region f' slope of exchange isotherm (dS/dC or dY/dX) F fraction of adsorption sites belonging to "type-i" (=S 1 /S) F dimensionless rate parameter (k2L/v) g acceleration due to gravity LT g geometry factor g(i) probability density function for infiltration G Gibbs free energy J mol- 136 Symbol Description Units h(O) soil water retention function h pressure head or height L h pneumatic head L hL(t) travel time density function h matric head L m H hydraulic head L 0~- 1 i infiltration rate LT I cumulative infiltration L JD diffusive/dispersive mass flux ML T D-2 J. ,J cumulative diffusive flux in immobile and ML mobile region J total solute mass flux ML-2T - 1 S-1 JV volumetric water flux (Darcy flux) LT -1 k rate coefficient T k permeability L 2 -1 k rate constant in the aggregate T a K(8) hydraulic conductivity LT K equilibrium constant K d distribution coefficient LM Kdi m K d for immobile region LM K K for mobile region LM .dmo d K distribution ratio of ions between mobile and D immobile region KH partition coefficient in Henry's law K exchange coefficient jk L column or mixing length L L D effective diffusion/dispersion length L Ldi f diffusion length L L d dispersion length L L dispersion length due to stagnant phase effect L r-2 m 0 total mass of solute per area of column ML m p-th moment of temporal or spatial distribution 137 Symbolj Description Units m p n N[ ] N. 1 P P P or Pe q q' qim r R R R. ,R im, mo R a s S S. ,S im mo ST t 17 T u u o v v V V v mo v m w ML-2T-1 _1 _) ML T6 depth-averaged moment number of moles normal distribution function diffusive molar flux probability pressure Peclet number (vd/D, vL/D) concentration of solute in adsorbed phase differential capacity of exchanger (=dq/dC) q for immobile sites radial distance gas constant retardation factor R for immobile and mobile region aggregate radius position along a streamline concentration of solute in adsorbed phase on a mass basis S for immobile and mobile region total solute concentration in adsorbed phase time temperature number of pore volumes (vt/L) velocity in a free liquid maximum value for u pore-water velocity effective (retarded) velocity (v/R) volume volume per unit cross sectional area pore-water velocity in mobile region (veT /e ) mean width L-3 ML 3 -3 L J mol K L L MM - 3 MM - 3 MM 3 mm-3 T K LT-1 LT 1 -1 LT -1 LT -1 LT L n aL3L -2 excha-1 regiLT I , I 138 Symbol I Description Units x Cartesian coordinate in direction of flow L X width L xC position at which a particular concentration L resides X k dimensionless concentration in liquid phase for species k (=Ck/C T x equivalent depth of penetration of a solute L front y space coordinate L Yk dimensionless concentration in adsorbed phase for species k (=Sk/S z space coordinate L z gravitational head L Z dimensionless distance (r/a, x/L) a mass transfer coefficient T a. random hydraulic parameter 1 aL longitudinal dispersivity L a T transversal dispersivity L 3 gravity segregation factor activity coefficient w -s rate constant for production in liquid and MLT- 1 W adsorbed phase T 6 Dirac delta function Sviscosity ML- 1T - 1 0 volumetric water content or volume fraction LL - 3 3 -3 0 im 0 moe for immobile and mobile region LL - 3 e ,9 volume fraction in liquid and gas phase L3L - 3 g gS-3 0 total volume fraction LL A tortuosity factor -1 pk chemical potential of species k J mol o -1 pk k in a chosen standard state J mol mean of spatial (i ) or temporal (It) L or T distribution 139 Symbol Description Units Awl s rate constant for first order decay in liquid T-1 and adsorbed phase Vk valence of species k Cartesian coordinate in a moving system L P density ML-3 Pb dry bulk density ML-3 T standard deviation of spatial ((Y ) or temporal L or T (0-t) distribution x 0' relative spread in velocities T (transit) time T -T unit vector field (v/v) formation factor probability function Boltzmann variable (x/ff) LT I BIBLIOGRAPHY (1) ABRIOLA, L.M. 1987. Modeling Contaminant Transport in the Subsurface: An Interdisciplinary Challenge. Rev. of Geophys. 25:125-134. (2) and G.F. PINDER. 1985. A Multiphase Approach to the Modeling of Porous Media Contamination by Organic Compounds. 1. Equation Development. Water Resour. Res. 21:11-18. (3) . 1985. A Multiphase Approach to the Modeling of Porous Media Contamination by Organic Compounds. 2. Numerical Simulation. Water Resour. Res. 21:19-26. (4) ADDISCOTT, T.M. and R.J. WAGENET. 1985. Concepts of Solute Leaching in Soils: A Review of Modelling Approaches. J. of Soil Sci. 36:411-424. (5) AMOOZEGAR-FARD, A., D.R. NIELSEN, and A.W. WARRICK. 1982. Soil Solute Concentration Distributions for Spatially Varying Pore Water Velocities and Apparent Diffusion Coefficients. Soil Sci. Soc. Amer. J. 46:3-9. (6) , A.W. WARRICK, and W.H. FULLER. 1983. A Simplified Model for Solute Movement through Soils. Soil Sci. Soc. Amer. J. 47:1047-1049. (7) AMUNDSON, N.R. 1948. A Note on the Mathematics of Adsorption in Beds. J. Phys. Chem. 52:1153-1157. (8) . 1950. Mathematics of Adsorption in Beds. II. J. Phys. Chem. 54:812-820. (9) ARIS, R. 1956. On the Dispersion of a Solute in a Fluid Flowing through a Tube. Proc. R. Soc. London Ser. A. 235:67-77. (10) . 1958. On the Dispersion of Linear Kinematic Waves. Proc. R. Soc. London Ser. A. 245:268-277. (11) AWAD, M.M. 1986. Effect of Solution Content and Seepage Velocity on Hydrodynamic Dispersion. Ph.D. Thesis. Colorado State Univ. Fort Collins. (12) BACHMAT, Y. 1969. Hydrodynamic Dispersion in a Saturated Homogeneous Porous Medium at Low Peclet Numbers and Nonhomogeneous Solution. Water Resour. Res. 5:139-143. 140 141 (13) BAEHR, A.L. and M.Y. CORAPCIOGLU. 1987. A Compositional Multiphase Model for Groundwater Contaminants by Petroleum Products. 2. Numerical solution. Water Resour. Res. 23:201-213. (14) BEAR, J. 1969. Hydrodynamic Dispersion in Flow Through Porous Media. In Flow through Porous Media. J. deWiest (Ed.) Academic Press. New York. (15) . 1975. Dynamics of Fluids in Porous Media. Elsevier. New York. (16) . 1979. Hydraulics of Groundwater. McGraw-Hill, New York. (17) and Y. BACHMAT. 1967. A Generalized Theory on Hydrodynamic Dispersion in Porous Media. Proc. Int. Assoc. Sci. Hydrol. Symp. Haifa. Publ. 72. (18) BEVEN, K. and P. GERMANN. 1982. Macropores and Water Flow in Soils. Water Resour. Res. 18:1311-1325. (19) BHATTACHARYA, R.N. and V.K. GUPTA. 1983. A Theoretical Explanation of Solute Dispersion in Saturated Porous Media at the Darcy Scale. Water Resour. Res. 19:938-944. (20) BIGGAR, J.W. and D.R. NIELSEN. 1962. Miscible Displacement. II. Behavior of Tracers. Soil Sci. Soc. Amer. Proc. 26:125-128. (21) . 1964. Chloride-36 Diffusion During Stable and Unstable Flow Through Glass Beads. Soil Sci. Soc. Amer. Proc. 28:591-594. (22) . 1976. Spatial Variability of the Leaching Characteristics of a Field Soil. Water Resour. Res. 12:78-84. (23) BOAST, C.W. 1973. Modeling the Movement of Chemicals in Soils by Water. Soil Sci. 115:224-230. (24) BOLT, G.H. 1982. Movement of Solutes in Soils: Principles of Adsorption/Exchange Chromatography. In Soil Chemistry. B. Physico-Chemical Models. G.H. Bolt (Ed.). Elsevier. Amsterdam. (25) BOND, W.J. 1987. Solute Transport During Unsteady Unsaturated Soil Water Flow: The Pulse Input. Aust. J. Soil Res. 25: 233-241. (26) BREDEHOEFT, J.D. 1971. Numerical Dispersion Comments. Water Resour. Res. 7:755-756. (27) BRENNER, H. 1962. The Diffusion Model of Longitudinal Mixing in Beds of Finite Length. Numerical Values. Chem. Eng. Sci. 17: 229-243. 142 (28) BRESLER, E. 1973. Simultaneous Transport of Solutes and Water Under Transient Unsaturated Flow Conditions. Water Resour. Res. 9:975-986. (29) . 1973. Anion Exclusion and Coupling Effects in Nonsteady Transport through Unsaturated Soil. I. Theory. Soil Sci. Soc. Amer. Proc. 37:663-669. (30) and G. DAGAN. 1979. Solute Dispersion in Unsaturated Heterogeneous Soil at Field Scale. II. Application. Soil Sci. Soc. Amer. J. 43: 467-472. (31) . 1981. Convective and Pore Scale Dispersive Solute Transport in Unsaturated Heterogeneous Fields. Water Resour. Res. 17: 1683-1693. (32) . 1983. Unsaturated Flow in Spatially Variable Fields. 2. Application of Water Flow Models to Various Fields. Water Resour. Res. 19:421-428. (33) . 1983. Unsaturated Flow in Spatially Variable Fields. 3. Solute Transport Models and Their Applications to Two Fields. Water Resour. Res. 19:429-435. (34) BRIGHAM, W.E. 1974. Mixing Equations in Short Laboratory Cores. Soc. Pet. Eng. J. 14:91-99. (35) BRUCE, R.R. and A. KLUTE. 1956. The Measurement of Soil Moisture Diffusivity. Soil Sci. Soc. Amer. Proc. 20:458-462. (36) CAMERON, D.R. and A. KLUTE. 1977. Convective-Dispersive Solute Transport with a Combined Equilibrium and Kinetic Adsorption Model. Water Resour. Res. 13:183-188. (37) CARSLAW, H.S. and J.C. JAEGER. 1959. Conduction of Heat in Solids. Clarendon Press, Oxford. (38) CHARBENEAU, R.J. 1984. Kinematic Models for Soil Moisture and Solute Transport. Water Resour. Res. 20:699-706. (39) CHAUDHARI, N.M. 1971. An Improved Numerical Technique for Solving Multidimensional Miscible Displacement Equations. Soc. of Petrol. Eng. J. 11:277-284. (40) CHILDS, S.W. and R.J. HANKS. 1975. Model of Salinity Effects on Crop Growth. Soil Sci. Soc. Amer. Proc. 39:617-622. (41) CHO, C.M. 1971. Convective Tranport of Ammonium with Nitrification in Soil. Can J. Soil Sci. 51:339-350. 143 (42) CLEARY, R.W. and D.D. ADRIAN. 1973. Analytical Solution of the Convective-Dispersive Equation for Cation Adsorption in Soils. Soil Sci. Soc. Amer. Proc. 37:197-199. (43) COATS, K.H. and B.D. SMITH. 1964. Dead End Pore Volume and Dispersion in Porous Media. Soc. Petrol. Eng. J. 4:73-84. (44) CORAPCIOGLU, M.Y and A.L. BAEHR. 1987. A Compositional Multiphase Model for Groundwater Contaminants by Petroleum Products. 1. Theoretical Considerations. Water Resour. Res. 23:191-200. (45) CRANK, J. 1975. The Mathemat'ics of Diffusion. Clarendon Press, Oxford. (46) CREMERS, A. and H. LAUDELOUT. 1965. On the 'Isoconductivity Value of Clay Gels. Soil Sci. 100: 298-299. (47) CUSHMAN, J.H. 1982. Nutrient Transport Inside and Outside the Root Rhizosphere. Theory. Soil Sci. Soc. Amer. J. 46:704-709. (48) DAGAN, G. 1984. Solute Transport in Heterogeneous Porous Formations. J. Fluid Mech. 145:151-177. (49) and E. BRESLER. 1979. Solute Dispersion in Unsaturated Heterogeneous Soil at Field Scale. I. Theory. Soil Sci. Soc. Amer. J. 43:461-467. (50) . 1983. Unsaturated Flow in Spatially Variable Fields. 1. Derivation of Models of Infiltration and Redistribution. Water Resour. Res. 19:413-420. (51) DANCKWERTS, P.V. 1953. Continuous Flow Systems. Chem. Eng. Sci. 2:1-13. (52) DANE, J.H. and A. KLUTE. 1977. Salt Effects on the Hydraulic Properties of a Swelling Soil. Soil Sci. Soc. Amer. J. 41:1043-1049. (53) DAVIDSON, J.M., P.S.C. RAO, and P. NKEDI-KIZZA. 1983. Physical Processes Influencing Water and Solute Transport in Soils. In Chemical Mobility and Reactivity in Soil Systems. Soil Sci. Soc. Amer., Spec. Publ. 11. Madison, Wis. (54) DE JOSSELIN DE JONG, G. 1958. Longitudinal and Transverse Diffusion in Granular Deposits. Trans. Amer. Geophys. Union. 39:67-74. (55) DE SMEDT, F. and P.J. WIERENGA. 1979. Mass Transfer in Porous Media with Immobile Water. J. of Hydr. 41:59-67. 144 (56) . 1979. A Generalized Solution for Solute Flow in Soils with Mobile and Immobile Water. Water Resour. Res. 15: 1137-1141. (57) . 1984. Solute Transfer through Columns of Glass Beads. Water Resour. Res. 20: 225-232. (58) ELRICK, D.E., K.T. ERH, and H.K. KRUPP. 1966. Applications of Miscible Displacement Techniques to Soils. Water Resour. Res. 2:717-727. (59) , K.B. LARYEA, and P.H. GROENEVELT. 1979. Hydrodynamic Dispersion During Infiltration of Water into Soil. Soil Sci. Soc. Amer. J. 43:856-865. (60) FISCHER, H.B., E. LIST, R.C.Y. KOH, J. IMBERGER, and N.H. BROOKS. 1979. Mixing in Inland and Coastal Waters. Academic Press. (61) FRIED, J.J. and M.A. COMBARNOUS. 1971. Dispersion in Porous Media. In Adv. Hydrosci. 9:169-282. Academic, New York. (62) FRISSEL, M.J., P. POELSTRA, and P. REINIGER. 1970. Chromatographic Transport Through soils. III. A Simulation Model for the Evaluation of the Apparent Diffusion Coefficient in Undisturbed Soils with Titrated Water. Plant and Soil 33:161-176. (63) CAPON, E.N. 1933. On the Theory of Exchange Adsorption in Soil. J. Gen. Chem. U.S.S.R. 3:144-152. (64) GELHAR, L.W., A.L. GUTJAHR, and R.L. NAFF. 1979. Stochastic Analysis of Macrodispersion in a Stratified Aquifer. Water Resour. Res. 15:1387-1396. (65) GERSHON, N.D. and A. NIR. 1969. Effects of Boundary Conditions of Models on Tracer Distributions in Flow through Porous Mediums. Water Resour. Res. 5:830-839. (66) GIDDINGS, J.C. 1965. Dynamics of Chromatography. Part I. Principles and Theory. Marcel Dekker. New York. (67) GILLHAM, R.W., E.A. SUDICKY, J.A. CHERRY, and E.O. FRIND. 1984. An Advection-Diffusion Concept for Solute Transport in Heterogeneous Unconsolidated Geological Deposits. Water Resour. Res. 20:369-378. (68) GRISAK, G.E. and J.F. PICKENS. 1980. Solute Transport through Fractured Media. 1. The Effect of Matrix Diffusion. Water Resour. Res. 16:719-730. (69) GRISMER, M.E. 1986. Nondestructive Observations of Solution Displacement in Soils. Soil Sci. 141:185-189. 145 (70) , D.B. MCWHORTER, and A. KLUTE. 1986. Monitoring Water and Salt Movement in Soils at Low Solution Contents. Soil Sci. 141:163-171. (71) GROENEVELT, P.H. and G.H. BOLT. 1969. Non-Equilibrium Thermodynamics of the Soil-Water System. J. Hydrol. 7:358-388. (72) , D.E. ELRICK, and K.B. LARYEA. 1980. Coupling Phenomena in Saturated Homo-Ionic Montmorillonite. IV. The Dispersion Coefficient. Soil Sci. Soc. Amer. J. 44:1168-1173. (73) GUPTA, V.K. and R.N. BHATTACHARYA. 1983. A New Derivation in the Taylor-Aris Theory of Solute Dispersion in a Capillary. Water Resour. Res. 19:945-951. (74) . 1986. Solute Dispersion in Multidimensional Periodic Saturated Porous Media. Water Resour. Res. 22:156-164. (75) GUPTA, S.P. and R.A. GREENKORN. 1974, Determination of Dispersion and Nonlinear Adsorption Parameters for Flow in Porous Media. Water Resour. Res. 10:839-846. (76) GUPTA, R.K., A. KLUTE, and R.J. MILLINGTON. 1972. An Electrode System for Transient Dispersion Studies in Unsaturated Glass Bead Mediums. J. Indian Soc. Soil Sci. 20: 297-304. (77) , R.J. MILLINGTON, and A. KLUTE. 1973. Hydrodynamic Dispersion in Unsaturated Porous Media. I. Concentration Distribution during Dispersion. J. Indian Soc. Soil Sci. 21:1-7. (78) . 1973. Hydrodynamic Dispersion in Unsaturated Porous Media. II. The Stagnant Zone Concept and the Dispersion Co-efficient. J. Indian Soc. Soil Sci. 21:121-125. (79) COVEN, O., R.W. FALTA, F.J. MOLZ, and J.G. MELVILLE. 1985. Analysis and Interpretation of Single Well Tracer Tests in Stratified Aquifers. Water Resour. Res. 21:676-684. (80) and F.J. Molz. 1986. Deterministic and Stochastic Analyses of Dispersion in an Unbounded Stratified Porous Medium. Water Resour. Res. 22:1566-1574. (81) , and J.G. MELVILLE. 1984. An Analysis of Dispersion in a Stratified Aquifer. Water Resour. Res. 20: 1337-1354. (82) HARLEMAN, D.R.F. and R.R. RUMER. 1963. Longitudinal and Lateral Dispersion in an Isotropic Porous Medium. J. Fluid Mech. 16: 385-394. 146 (83) HASHIMITO, I., DESHPANDE, K.B., and H.C. THOMAS. 1964. Peclet Numbers and Retardation Factors for Ion Exchange Columns. Ind. Eng. Chem. Fund. 3:213-218. (84) HELFFERICH, F. 1962. Ion Exchange. McGraw-Hill. New York. (85) HOPMANS, J.W. and J.H. DANE. 1985. Effect of Temperature- dependent Hydraulic Properties on Soil Water Movement. Soil Sci. Soc. Amer. J. 49:51-58. (86) HUYAKORN, P.S. and G.F. PINDER. 1983. Computational Methods In Subsurface Flow. Academic Press. New York. (87) JAMES, R.V. and J. RUBIN. 1972. Accounting for Apparatus-induced Dispersion in Analyses of Miscible Displacement Experiments. Water Resour. Res. 8:717-721. (88) JENNINGS, A.A., D.J. KIRKNER, and T.L. THEIS. 1982. Multicomponent Equilibrium Chemistry in Groundwater Quality Models. Water Resour. Res. 18:1089-1096. (89) JOURNEL, A.G. and CH.J. HUIJBREGTS. 1978. Mining Geostatistics. Academic Press. London. (90) JURY, W.A. 1982. Simulation of Solute Transport Using a Transfer Function Model. Water Resour. Res. 18:363-368. (91) , R. GROVER, W.F. SPENCER, and W.J. FARMER. 1980. Modeling Vapor Losses of Soil-incorporated Triallate. Soil Sci. Soc. Amer. J. 44:445-450. (92) and G. SPOSITO. 1985. Field Calibration and Validation of Solute Transport Models for the Unsaturated Zone. Soil Sci. Soc. Amer. J. 49:1331-1341. (93) , L.H. Stolzy, and P. Shouse. 1982. A Field Test of the Transfer Function Model for Predicting Solute Transport. Water Resour. Res. 18:369-375. (94) KAY, B.D. and D.E. ELRICK. 1967. Adsorption and Movement of Lindane in Soils. Soil Sci. 104:314-322. (95) KIRDA, C., D.R. NIELSEN, and J.W. BIGGAR. 1973. Simultaneous Transport of Chloride and Water During Infiltration. Soil Sci. Soc. Amer. Proc. 37:339-34S. (96) . 1974. The Combined Effects of Infiltration and Redistribution of Leaching. Soil Sci. 117: 323-329. 147 (97) KIRKNER, D.J., A.A. JENNINGS, and T.L. THEIS. 1985. Multlsolute Mass Transport with Chemical Interaction Kinetics. J. Hydrol. 76:107-117. (98) KLINKENBERG, L.J. 1957. Pore Size Distribution of Porous Media and Displacement Experiments with Miscible Liquids. J. Petrol. Techn. 210:63-66. (99) KNIGHTON, R.E. and R.J. WAGENET. 1987. Simulation of Solute Transport Using a Continuous Time Markow Process. 1. Theory and Steady State Application. Water Resour.Res. 23:1911-1916, (100) . Simulation of Solute Transport Using a Continuous Time Markow Process. 2. Application to Transient Field Conditions. Water Resour. Res. 23:1917-1925. (101) KOOREVAAR, P., G. MENELIK, and C. DIRKSEN. 1983. Elements of Soil Physics. Elsevier. Amsterdam. (102) KREFT, A. and A. ZUBER. 1978. On the Physical Meaning of the Dispersion Equations and its Solutions for Different Initial and Boundary Conditions. Chem. Eng. Sci. 33:1471-1480. (103) KRUPP, H.K., J.W. BIGGAR, and D.R. NIELSEN. 1972. Relative Flow Rates of Salt and Water in Soil. Soil Sci. Soc. Amer. Proc. 36:412-417. (104) and D.E. ELRICK. 1968. Miscible Displacement in an Unsaturated Glass Bead Medium. Water Resour. Res. 4:809-815. (105) . 1969. Density Effects in Miscible Displacement Experiments. Soil Sci. 107:372-380. (106) LAI, S.H. and J.J. JURINAK. 1971. Numerical Approximation of Cation Exchange in Miscible Displacement through Soil Columns. Soil Sci. Soc. Amer. Proc. 35:894-899. (107) . 1972. Cation Adsorption in One-dimensional Flow through Soils: A Numerical Solution. Water Resour. Res. 8:99-107. (108) LAKE, L.W. and F. HELFFERICH. 1978. Cation Exchange in Chemical Flooding. 2. The Effects of Dispersion, Cation Exchange, and Polymer/Surfactant Adsorption on Chemical Flood Environment. Soc. Pet. Eng. J.:435-444. (109) LAPIDUS, L. and N.R. AMUNDSON. 1952. Mathematics of Adsorption in beds. VI. The Effect of Longitudinal Diffusion in Ion Exchange and Chromatographic Columns. J. Phys. Chem. 56:984-988. 148 (110) LARYEA, K.B., D.E. ELRICK, and M.J.L. ROBIN. 1982. Hydrodynamic Dispersion Involving Cationic Adsorption during Unsaturated Transient Water Flow in Soil. Soil Sci. Soc. Amer. Proc. 46:667-671. (111) LEVENSPIEL, 0. and W.K. SMITH. 1957. Notes on the Diffusion-type Model for the Longitudinal Mixing of Fluids in Flow. Chem. Eng. Sci. 6:227-233. (112) LINDSTROM, F.T. 1976. Pulse Dispersion of Trace Chemical Concentrations in a Saturated Sorbing Porous Medium. Water Resour. Res. 12:229-238. (113) , R. HAQUE, V.H. FREED, and L. BOERSMA. 1967. Theory on the Movement of Some Herbicides in Soils. Environ. Sci. Techn. 1:561-565. (114) LUXMOORE, R.J. and M.L. SHARMA. 1980. Runoff Response to Soil Heterogeneity: Experimental and Simulation Comparisons for Two Contrasting Watersheds. Water Resour. Res. 16:675-684. (115) MARLE, C., P. SIMANDOUX, J. PACSIRSKY, and C. GAULIER. 1967. Etude du D6placement de Fluides Miscibles en Milieu Poreux Stratifie. Rev. Inst. Fr. Petrol. 22: 272-294. (116) MATHERON, G. and G. DE MARSILY. 1980. Is Transport in Porous Media Always Diffusive? A Counterexample. Water Resour. Res. 16:901-917. (117) MERCADO, A. 1967. The Spreading Pattern of Injected Water in a Permeability Stratified Aquifer. Int. Ass. Sci. Hydrol. Symp. Haifa. Publ. 72. (118) MILLINGTON, R.J. and R.C. SHEARER. 1971. Diffusion in Aggregated Porous Media. Soil Sci. 111: 372-378. (119) NERETNIEKS, I. 1983. A Note on Fracture Flow Dispersion Mechanisms in the Ground. Water Resour. Res. 19:364-370. (120) NIELSEN, D.R. and J.W. BIGGAR. 1961. Miscible Displacement in Soils. I. Experimental Information. Soil Sci. Soc. Amer. Proc. 25:1-5. (121) . 1962. Miscible Displacement. III. Theoretical Considerations. Soil Sci. Soc. Amer. Proc. 26: 216-221. (122) - - , R.D. JACKSON, J.W. CARY, and D.D. EVANS (Eds.). 1972. Soil Water. Amer. Soc. of Agron. Madison, Wis. 149 (123) , M.Th. VAN GENUCHTEN, and J.W. BIGGAR. 1986. Water Flow and Solute Transport Processes in the Unsaturated Zone. Water Resour. Res. 22:89S-108S. (124) NKEDI-KIZZA, P., J.W. BIGGAR, M.Th. VAN GENUCHTEN, P.J. WIERENGA, H.M. SELIM, J.M. DAVIDSON, and D.R. NIELSEN. 1983. Modeling Tritium and Chloride 36 Transport through an Aggregated Oxisol. Water Resour. Res. 19:691-700. (125) NOVOZAMSKY, I., J. BEEK, and G.H. BOLT. 1978. Chemical Equilibria. In Soil Chemistry. A. Basic Elements. G.H. Bolt and M.G.M. Bruggenwert (Eds.). Elsevier. Amsterdam. (126) NYE, P.H. 1979. Diffusion of Ions and Uncharged Solutes in Soils and Soil Clays. Adv. Agron. 31: 225-272. (127) OGATA, A. and R.B. BANKS. 1961. Solution of the Differential Equation of Longitudinal Dispersion in Porous Media. U.S. Geol. Surv. Prof. Paper 411-A:1-7. (128) PANDEY, R.N. and S.K. GUPTA. 1980. Comment on 'Dispersion in Soil Columns: Effect of Boundary Conditions and Irreversible Reactions'. Soil Sci. Soc. Amer. J. 44: 661. (129) PARKER, J.C., R.J. LENHARD, and T. KUPPUSAMY. 1987. A Parametric Model for Constitutive Properties Governing Multiphase Flow in Porous Media. Water. Resour. Res. 23:618-624. (130) and M.TH. VAN GENUCHTEN. 1984. Determining Transport Parameters from Laboratory and Field Tracer Experiments. Virg. Agric. Exp. Stat. Bull. 84-3. (131) . 1984. Flux-averaged Concentrations in Continuum Approaches to Solute Transport. Water Resour. Res. 20:866-872. (132) PARLANGE, J.-Y., D.A. BARRY, and J.L. STARR. 1985. Comments on "Boundary Conditions for Displacement Experiments through Short Laboratory Columns". Soil Sci. Soc. Amer. J. 49:1325. (133) and J.L. STARR. 1975. Linear Dispersion in Finite Columns. Soil Sci. Soc. Amer. Proc. 39:817-819. (134) . 1978. Dispersion in Soil Columns: Effect of the Boundary Conditions and Irreversible Reactions. Soil Sci. Soc. Amer. J. 42:15-18. (135) PASSIOURA, J.B. 1971. Hydrodynamic Dispersion in Aggregated Media. 1.Theory. Soil Sci. 111:339-344. 150 (136) and D.A. ROSE. 1971. Hydrodynamic Dispersion in Aggregated media. 2. Effects of Velocity and Aggregated Size. Soil Sci. 111:345-351. (137) PERSAUD, N., J.V. GIRALDEZ, and A.C. CHANG. 1985. Monte-Carlo Simulation of Noninteracting Solute Transport in a Spatially Heterogeneous Soil. Soil Sci. Soc. Amer. J. 49:562-568. (138) PFANNKUCH, H.O. 1963. Contribution a 1'Etude des D6placements de Fluides dans un Milieu poreux. Rev. Inst. Fr. Petr. 18:215-270. (139) PHILIP, J.R. 1957. Numerical Solution of Equations of the Diffusion Type with Diffusivity Concentration-dependent. II. Aust. J. Phys. 10: 29-42. (140) . 1968. The Theory of Absorption in Aggregated Media. Aust. J. Soil Res. 6:1-19. (141) . 1968. Diffusion, Dead-end-pores and Linearized Absorption in Aggregated Media. Aust. J. Soil Res. 6:21-30. (142) PICKENS, J.F. and G.E. GRISAK. 1981. Scale-dependent Dispersion in a Stratified Granular Aquifer. Water Resour. Res. 17:1191-1211. (143) . 1981. Modeling of Scale-dependent Dispersion in Hydrogeologic Systems. Water Resour. Res. 17:1701-1711. (144) PINDER, G.F. and W.G. GRAY. 1977. Finite Elements in Surface and Subsurface Hydrology. Academic Press. New York. (145) PLUMB, O.A. and S. WHITAKER. 1988. Dispersion in Heterogeneous Porous Media. 1. Local Volume Averaging and Large-scale Averaging. Water Resour. Res. 24:913-924. (146) . 1988. Dispersion in Heterogeneous Porous Media. 2. Predictions for Stratified and Two-dimensional Spatially Periodic Systems. Water Resour. Res. 24:927-938. (147) RAATS, P.A.C. 1978. Convective Transport of Solutes by Steady Flows. 1.General Theory. Agr. Water Manage. 1:201-218. (148) . 1978. Convective Transport of Solutes by Steady Flows. 2. Specific Flow Problems. Agr. Water Manage. 1:219-232. (149) RAO, P.S.C., J.M. DAVIDSON, R.E. JESSUP, and H.M. SELIM. 1979. Evaluation of Conceptual Models for Describing Non-equilibrium Adsorption-Desorption of Pesticides during Steady Flow in Soils. Soil Sci. Soc. Amer. J. 43:22-28. 151 (150) , R.E. JESSUP, D.E. ROLSTON, J.M. DAVIDSON, and D.P. KILCREASE. 1980. Experimental and Mathematical Description of Non-adsorbed Solute Transfer by Diffusion in Spherical Aggregates. Soil Sci. Soc. Amer. J.44:684-688. (151) , D.E. ROLSTON, R.E. JESSUP, and J.M. DAVIDSON. 1980. Solute Transport in Aggregated Porous Media. Theoretical and Experimental Evaluation. Soil Sci. Soc. Amer. J. 44:1139-1146. (152) RASMUSON, A. 1981. Diffusion and Sorption in Particles and Two-dimensional Dispersion in a Porous Medium. Water Resour. Res. 17:321-328. (153) and I. NERETNIEKS. 1980. Exact Solution of a Model for Diffusion in Particles and Longitudinal Dispersion in Packed Beds. AIChE Journal 26:686-690. (154) . 1981. Migration of Radionuclides in Fissured Rock: The Influence of Micropore Diffusion and Longitudinal Dispersion. J. of Geophys. Res. 86:3749-3758. (155) RAZAVI, M.-S., B.J. MCCOY, and R.G. CARBONELL. 1978. Moment Theory of Breakthrough Curves for Fixed-bed Adsorbers and Reactors. Chem. Eng. J. 16:211-222. (156) REINIGER, P. and G.H. BOLT. 1972. Theory of Chromatography and its Application to Cation Exchange in Soils. Neth. J. Agric. Sci. 20:301-313. (157) RISKEN, H. 1984. The Fokker-Planck Equation. Methods of Solution and Application. Springer-Verlag. New York. (158) ROBBINS, C.W., J.J. JURINAK, and R.J. WAGENET. 1980. Calculating Cation Exchange in a Salt Transport Model. Soil Sci. Soc. Amer. J. 44: 1195-1199. (159) , R.J. WAGENET, and J.J. JURINAK. 1980. A Combined Salt Transport-chemical Equilibrium Model for Calcareous and Gypsiferous Soils. Soil Sci. Soc. Amer. J. 44:1191-1194. (160) ROSE, D.A. and J.B. PASSIOURA. 1971. The Analysis of Experiments on Hydrodynamic Dispersion. Soil Sci. 111:252-257. (161) . 1971. Gravity Segregation during Miscible Displacement. Soil Sci. 111:258-265. (162) RUBIN, J. 1983. Transport of Reacting Solutes in Porous Media: Relation between Mathematical Nature of Problem Formulation and Chemical Nature of Reactions. Water Resour. Res. 19:1231-1252. 152 (163) and R.V. JAMES. 1973. Dispersion-affected Transport of Reacting Solutes in Saturated Porous Media: Galerkin Method Applied to Equilibrium-controlled Exchange in Unidirectional Steady Water Flow. Water Resour. Res. 9:1332-1356. (164) SAFFMAN, P.G. 1959. A Theory of Dispersion in a Porous Medium. J. Fluid Mech. 6:321-349. (165) SCHEIDEGGER, A.E. 1974. The Physics of Flow Through Porous Media. University of Toronto Press. (166) SCOTTER, D.R. 1978. Preferential Solute Movement through Larger Soil Voids. I. Some Computations Using Simple Theory. Aust. J. Soil Res. 16:257-267. (167) and P.A.C. RAATS. 1968. Dispersion in Porous Mediums Due to Oscillating Flow. Water Resour. Res. 6:1201-1206. (168) SELIM, H.M., J.M. DAVIDSON, and R.S. MANSELL. 1976. Evaluation of a Two-site Adsorption-Desorption Model for Describing Solute Transport in Soil. Proc. Summer Conf., Washington D.C. pp 444-448. (169) , R.S. MANSELL, and A. ELZEFTWAY. 1976 . Distribution of 2,4-D and Water in Soil during Infiltration and Redistribution. Soil Sci. 121:176-183. (170) , R. SCHULIN, and H. FLUHLER. 1987. Transport and Ion Exchange in an Aggregated Soil. Soil Sci. Soc. Amer. J. 51:876-884. (171) SHAMIR, U.Y. and D.R.F. HARLEMAN. 1967. Dispersion in Layered Porous Media. J. Hydraul. Div. ASCE. 93(HY5):237-260. (172) SIMMONS, C.S. 1982. A Stochastic-Convective Transport Representation of Dispersion in One-dimensional Porous Media Systems. Water Resour. Res. 18:1193-1214. (173) SKAGIUS, K. and I. NERETNIEKS. 1986. Porosities and Diffusivities of Some Nonsorbing Species in Crystalline Rocks. Water Resour. Res. 22: 389-398. (174) SKOPP, J., W.R. GARDNER, and E.J. Tyler. 1981. Solute Movement in Structured Soils: Two-region Model with Small Interaction. Soil Sci. Soc. Amer. J. 45:837-842. (175) SKOPP, J. and A.W. WARRICK. 1974. A Two-phase Model for the Miscible Displacement of Reactive Solutes in Soil. Soil Sci. Soc. Amer. J. 138:545-550. 153 (176) SMILES, D.E. and B.N. GARDINER. 1982. Hydrodynamic Dispersion during Unsteady, Unsaturated Water Flow in a Clay Soil. Soil Sci. Soc. Amer. J. 46: 9-14. (177) , K.M. PERROUX, S.J. ZEGELIN, and P.A.C. RAATS. 1981. Hydrodynamic Dispersion During Constant Rate Absorption of Water by Soil. Soil Sci. Soc. Amer. J. 42:229-234. (178) and J.R. PHILIP. 1978. Solute Transport during Absorption of Water by Soil: Laboratory Studies and Their Practical Implications. Soil Sci. Soc. Amer. J. 42:537-544. (179) , J.H. KNIGHT, and D.E. ELRICK. 1978. Hydrodynamic Dispersion during Absorption of Water by Soil. Soil Sci. Soc. Amer. J. 42:229-234. (180) SPOSITO, G. and W.A. JURY. 1985. Inspectional Analysis in the Theory of Water Flow through Unsaturated Soil. Soil Sci. Soc. Amer. J. 49:791-798. (181) , and V.K. GUPTA. 1986. Fundamental Problems in the Stochastic Convection-Dispersion Model of Solute Transport in Aquifers and Field Soils. Water Resour. Res. 22:77-88. (182) STARR, R.C., R.W. GILLHAM, and E.A. SUDICKY. 1985. Experimental Investigation of Solute Transport in Stratified Porous Media. 2. The Reactive Case. Water Resour. Res. 21:1043-1050. (183) SUDICKY, E.A. 1986. A Natural Gradient Experiment on Solute Transport in a Sandy Aquifer: Spatial Variability of Hydraulic Conductivity and its Role in the Dispersion Process. Water Resour. Res. 22:2069-2082. (184) and E.O. FRIND. 1982. Contaminant Transport in Fractured Porous Media: Analytical Solutions for a System of Parallel Fractures. Water Resour. Res. 18:1634-1642. (185) , R.W. GILLHAM, and E.O. FRIND. 1985. Experimental Investigation of Solute Transport in Stratified Porous Media. 1. The Nonreactive Case. Water Resour. Res. 21:1035-1041. (186) TANG, D.H., E.O. FRIND, and E.A. SUDICKY. 1981. Contaminant Transport in Fractured Porous Media: An Analytical Solution for a Single Fracture. Water Resour. Res. 17:SSS-S64. (187) , F.W. SCHWARTZ, and L. SMITH. 1982. Stochastic Modeling of Mass Transport in a Random Velocity Field. Water Resour. Res. 18:231-244. 154 (188) TANJI, K.K. 1970. A Computer Analysis on the Leaching of Boron from Stratified Soil Columns. Soil Sci. 110:44-51. (189) TAYLOR, G.I. 1953. Dispersion of Soluble Matter in Solvent Flowing through a Tube. Proc. R. Soc. London Ser. A. 219:186-203. (190) VALOCCHI, A.J. 1984. Describing the Transport of Ion-exchanging Contaminants Using an Effective Kd Approach. Water Resour. Res. 20: 499-503. (191) . 1985. Validity of the Local Equilibrium Assumption for Modeling Sorbing Solute Transport through Homogeneous Soils. Water Resour. Res. 21:808-820. (192) , R.L. STREET, and P.V. ROBERTS. 1981. Transport of Ion-exchanging Solutes in Groundwater: Chromatographic Theory and Field Simulation. Water Resour. Res. 17:1517-1527. (193) VAN DER ZEE, S.E.A.T.M. and W.H. VAN RIEMSDIJK. 1987. Transport of Reactive Solute in Spatially Variable Soil Systems. Water Resour. Res. 23:2059-2069. (194) VAN EYKEREN, J.C.H. and J.P.G. LOCH. 1984. Transport of' Cationic Solutes in Sorbing Porous Media. Water Resour. Res. 20:714-718. (195) VAN GENUCHTEN, M.Th. 1978. Mass Transport in Saturated- Unsaturated Media. I. One-dimensional Solutions. Res. Rep. 78-WR-11. Princeton Univ. (196) . 1981. Non-equilibrium Transport Parameters from Miscible Displacement Experiments. Research Report No.119, U.S. Salinity Laboratory, Riverside, Calif. (197) and W.J. ALVES. 1982. Analytical Solutions of the One-dimensional Convective-Dispersive Solute Transport Equation. USDA Tech. Bull. No. 1661. (198) and R.W. CLEARY. 1982. Movement of Solutes in Soil: Computer-Simulated and Laboratory Results. In Soil Chemistry. B. Physico-chemical models. G.H. Bolt (Ed.). Elsevier. Amsterdam. (199) and W.A. JURY. 1987. Progress in Unsaturated Flow and Transient Modeling. Rev. of Geophys. 25:101-105. Amer. Geophys. Union. (200) and J.C. PARKER. 1984. Boundary Conditions for Displacement Experiments through Short Laboratory Columns. Soil Sci. Soc. Amer. J. 48:703-708. 1ss (201) , D.H. TANG, and R. GUENNELON. 1984. Some Exact Solutions for Solute Transport through Soils Containing Large Cylindrical Macropores. Water Resour. Res. 20:335-346. (202) and P.J. WIERENGA. 1976. Mass Transfer Studies in Sorbing Porous Media. I. Analytical Solutions. Soil Sci. Soc. Amer. J. 40:473-480. (203) . 1977. Mass Transfer Studies in Sorbing Porous Media. II. Experimental Evaluation with Tritium ( H 0). Soil Sci. Soc. Amer. J. 41:272-278. 2 (204) . 1986. Solute Dispersion Coefficients and Retardation Factors. In Methods of Soil Analysis. I. Physical and Mineralogical Methods. Soil Sci. Soc. Madison, Wis. (205) , and G.A. O'CONNOR. 1977. Mass Transfer Studies in Sorbing Porous Media. III. Experimental Evaluation with 2,4,5-T. Soil Sci. Soc. Amer. J. 41: 278-284. (206) VILLERMAUX, J. and W.P.M. VAN SWAAY. 1969. Modele Representatif de la Distribution des Temps de Sejour dans un Reacteur Semi-infini a Dispersion Axiale avec Zones Stagnantes. Application a 1'Ecoulement Ruisselant dans des Colonnes d'Anneaux Raschig. Chem. Eng. Sci. 24:1097-1111 (207) WAGENET, R.J. 1983. Principles of Salt Movement in Soils. In Chemical Mobility and Reactivity in Soil Systems. Soil Sci. Soc. Amer. Publ. 11. Madison, Wis. (208) WARRICK, A.W. and A. AMOOZEGAR-FARD. 1979. Infiltration and Drainage Calculations Using Spatially Scaled Hydraulic Properties. Water Resour. Res. 15:1116-1120. (209) , J.W. BIGGAR, and D.R. NIELSEN. 1971. Simultaneous Solute and Water Transfer for an Unsaturated Soil. Water Resour. Res. 7:1216-1225. (210) , G.J. MULLEN, and D.R. NIELSEN. 1977. Scaling Field-measured Soil Hydraulic Properties Using a Similar Media Concept. Water Resour. Res. 13:355-362. (211) and D.R. NIELSEN. 1980. Spatial Variability of Soil Physical Properties in the Field. In D. Hillel. Applications of Soil Physics. Academic Press. New York. (212) WATSON, K.K. and M.J. JONES. 1982. Hydrodynamic Dispersion during Absorption in a Fine Sand. 2. The Constant Flux Case. Water Resour. Res. 18:1435-1443. 156 (213) WHEATCRAFT, S.W. and S.W. TYLER. 1988. An Explanation of Scale Dependent Dispersivity Using Concepts of Fractal Geometry. Water Resour. Res. 24:566-578 (214) WIERENGA, P.J. 1977. Solute Distribution Profiles Computed with Steady-state and Transient Water Movement. Soil Sci. Soc. Amer. J. 41:1045-1051. (215) WILLIAMS, J.K. and R.A. DAWE. 1986. Fractals-An Overview of Potential Applications to Transport in Porous Media. Transport Porous Media 1:201-210. (216) YEH, T.C., L.W. GELHAR, and A.L. GUTJAHR. 1985. Stochastic Analysis of Unsaturated Flow in Heterogeneous Soils. 1. Statistically Isotropic Media. Water Resour. Res. 21:447-456. (217) . 1985. Stochastic Analysis of Unsaturated Flow in Heterogeneous Soils. 2. Statistically Anisotropic Media with Variable a. Water Resour. Res. 21:457-464. (218) . 1985. Stochastic Analysis of Unsaturated Flow in Heterogeneous Soils. 3. Observations and Applications. Water Resour. Res. 21:465-471. (219) YULE, D.F. and W.R. GARDNER. 1978. Longitudinal and Transverse Dispersion Coefficients in Unsaturated Plainfield Sand. Water Resour. Res. 14:582-S88.