Sunteți pe pagina 1din 15

Ann. nucl. Energy, Vol. 15, No. 4, pp.

175-189, 1988
Printed in Great Britain

0306-4549/88 $3.00+0.00
Pergamon Press plc

1-D MODELLING OF RADIONUCLIDE MIGRATION


THROUGH PERMEABLE A N D FRACTURED ROCK FOR
ARBITRARY LENGTH DECAY CHAINS USING
NUMERICAL INVERSION OF LAPLACE TRANSFORMS
D. P. HODGKINSON a n d P. R. MAULS"
Theoretical Physics Division, AERE, Harwell, Oxfordshire OXl 1 0RA, U.K.

(Received 21 June 1987)


Abstract--This paper presents analytical solutions to Laplace-transformed radionucLide transport equations for arbitrary length decay chains in permeable and fractured rock. These can be combined to give
approximate expressions for transport through heterogeneous media consisting of an arbitrary number of
porous and/or fractured rock units. The time-dependent solutions are obtained by numerical inversion
using Talbot's algorithm. This has been found to be a very efficient, accurate and flexible technique.

NOMENCLATURE

a =
am(s) =
A=
b=
bin(s) =
Bkn(s) =
C~(x, t) =
d~n(x, s) =
CP(x, w, t) =
CP,(x, w,s) =
Cs =
Cn(t) =
C,,~(x, t) =
(',nn(X, S) =

CPm,(X, w, t) =
t2m~(x, w, s) =
d~i =
DL=
Dj =
D =
Di =
Eking(s) =
Ft,(O =
Pl~(s) =
F~, (x, t) =
P~,(x,s) =
gin(s) =
J =
kn =
Kdn =

Length of diffusion path in rock matrix (m)


Functions defined (13) and (48) (mol m-3)
Area of rock column (m 2)
Half-width o f fracture (m)
Functions defined in (16) and (43) (m ~)
Functions defined in (47) (s- ~)
Concentration in flowing pore or fracture water (mol m - 3)
Laplace transform of Cn(x, t) ( m o l ' s m-3)
Concentration in static rock-matrix pore-water (mol m-3)
Laplace transform of CV,(x, w, t)
Solubility limit (mol m-3)
Concentration in vault pore-water (mol m-3)
Concentration of nuclide, n, arising from the source term o f nuclide, m, in flowing pore or fracture water
(mol m - 3)
Laplace transform of Cmn(x, t) (mol" s m - 3)
Concentration of nuclide, n, in static rock-matrix pore-water arising from the source term for nuclide, m
(mol m - 3)
Laplace transform of C~n(x, w, t) (mol- s i n - 3)
Coefficient in (74) (mol s- 1)
Longitudinal dispersion coefficient (m 2 s-~)
Inlet boundary condition parameter (m 2 s- 1)
Generalized outflow rate parameter (m 2 s-~)
Intrinsic diffusion coefficient (m 2 s - i)
Function defined in (54)
Inflow rate (mol s - i)
Laplace transform of Fin(t) (mol)
Generalized outflow rate (tool s - i )
Laplace transform of F~. (x, t) (mol)
Functions defined in (12) and (44) (s- 1)
Total number of rock units
Fraction release rate (s- l)
Mass-based equilibrium distribution coefficient (m 3 kg- 1)

?Central Electricity Generating Board, Generation Development and Construction Division, Sudbury House, 15 Newgate
Street, London, U,K.
175

176

D . P . HODGK1NSON and P. R. MAUL


L = Length of column (m)
Lj = Length of rock u n i t j (m)
M.(t) = Radionuclide inventory in vault (tool)
N = Number of nuclides in chain
Q = Volumetric flowrate from vault (m 3 s- ~)
R. = Retardation factor
s = Laplace transform variable (s- ~)
Sk,..(s, I) = Function defined in (56)
t = Time (s)
/'. = Mean time of F~.(L, t) (s)
T~ = Solubility-limited leach time (s)
Tk,.. (s, I) = Function defined in (55)
Um.(S) = Functions defined in (11) and (45)
v = Velocity of water in flowing pores and fractures (m s-~)
vI = Inlet boundary condition parameter (m s ~)
v = Generalized outflow rate parameter (m s- i)
vm.(s) = Functions defined in (15) and (49)
V = Volume of repository (m 3)
w = Coordinate into rock matrix (m)
W.j = Function defined in (64) (tool)
W~,j = W.j with A set equal to 2 (mol)
x = Coordinate along column (m)
xj = Coordinate along rock u n i t j (m)
y,..(w, s) = Functions defined in (39) and (40)

Greek
~t. =
2. =
A. =
/1.i =
v=
~b =
qSp =
qS,.(s) =
p=
a =
6.z =
z=

Capacity factor of rock matrix


Radioactive decay constant (s t)
Parameters defined in (65) (s- ~)
Coefficients in (72) (s ~)
Parameter in Talbot's algorithm
Flowing porosity of permeable or fractured rock
Porosity of rock matrix
Functions defined in (41) (m ~)
Density of rock (kg m 3)
Parameter in Talbot's algorithm
Variance of F~.(L, t) (s)
Parameter in Talbot's algorithm

1. INTRODUCTION
T h e r e are m a n y situations in the nuclear industry where it is necessary to consider the consequences o f
radionuclides entering groundwater. The m o s t studied example is the w a t e r - b o r n e m i g r a t i o n of radionuclides
leached f r o m a n u n d e r g r o u n d repository. However, activity could enter the g r o u n d in the event of a n accidental
spill or leak o n a nuclear site. In addition, radioactive material stored in vaults, or active structures such as
s h u t d o w n reactors subject to delayed decommissioning, could come into contact with groundwater.
In all the a b o v e situations, m a t h e m a t i c a l models of radionuclide m i g r a t i o n in g r o u n d w a t e r have a n i m p o r t a n t
role to play in assessing the potential radiological consequences. The degree o f model sophistication needed for
any particular application depends o n the level o f site characterization a n d the availability o f experimental
data, t o g e t h e r with the requirements for accuracy, flexibility, generality a n d c o m p u t a t i o n a l efficiency.
T h r e e - d i m e n s i o n a l finite-element ( A t k i n s o n et al., 1985) a n d finite-difference ( I N T E R A , 1983) codes are
available for p e r f o r m i n g realistic calculations o n well-characterized sites which can be treated as permeable
media. In addition, 3-D codes are being developed to model the t r a n s p o r t o f radionuclides t h r o u g h networks
o f intersecting fractures ( L o n g et al., 1983 ; Schwartz et al., 1983 ; R o b i n s o n , 1984).
Despite the availability o f the a b o v e comprehensive numerical models, there is a c o n t i n u i n g need for simpler
models in order to provide insight a n d u n d e r s t a n d i n g , p e r f o r m sensitivity a n d u n c e r t a i n t y analyses a n d verify
the numerical accuracy o f general-purpose numerical codes. The m o s t a p p r o p r i a t e a p p r o x i m a t i o n s are to
consider a 1-D flow field a n d assume c o n s t a n t physical parameters.
O n e - d i m e n s i o n a l t r a n s p o r t models have played a n i m p o r t a n t role in assessing the consequences o f burying
radioactive waste in permeable ( B u r k h o l d e r et al., 1976; Hill a n d G r i m w o o d , 1978) a n d fractured (Bengtsson

1-D modelling of radionuclide migration

177

et al., 1983; Hadermann and Roesel, 1985) rock masses. In some cases the transport equations have been
solved using numerical techniques (Schwartz, 1977 ; Bo, 1978 ; Rasmuson et al., 1982 ; Hadermann and Roesel,
1983) while in certain special situations analytical solutions have been obtained (Lester et al., 1975 ; Burkholder
and Rosinger, 1980; Harada et al., 1980; Pigford et al., 1980; Rasmuson, 1984). The numerical accuracy of

a number of such codes was tested by comparison with a range of benchmark problems in level 1 of the recent
international I N T R A C O I N project (SKI, 1984). A useful summary of these and other codes is contained in
the compilation by Broyd et al. (1983).
In general, analytical solutions are to be preferred since they are computationally efficient and accurate. For
example, they do not suffer from instabilities, numerical dispersion or have problems coping with a wide range
of time scales. Also, the approximations of using constant values for parameters and linear sorption isotherms
are generally adequate given the typically sparse input data.
However, the range of problems for which the solution can be written down in terms of known analytic
functions is rather limited. Also, the algebra tends to become extremely complicated, and thus prone to errors,
when long decay chains are treated. This is especially true when fractured media models, including rock-matrix
diffusion, are considered. One particular problem with analytical solutions is that oversimplified boundary
conditions are usually used, which are not necessarily compatible with the physical and chemical processes
occurring in the near-field (Bengtsson et al., 1983 ; McGrail et al., 1985 ; Ewart et al., 1985 ; Robinson et al.,
1987).
To overcome these problems, semianalytical solutions have been developed to a subset of the problems
considered in this paper (Hadermann 1980; Gureghian and Jansen, 1985; Sudicky and Frind, 1984). These
generally involve solutions to the equations in terms of auxilliary functions which are evaluated numerically.
An interesting example of this approach is to obtain analytical expressions for a 6-function input (Maul, 1983,
1984) and then to perform a numerical convolution to obtain the result for a realistic input profile (Heinrich,
1984).
However, in the opinion of the present authors, the most appropriate semianalytical solution technique is
to invert numerically the analytical solutions to the Laplace-transformed equations. This has proved to be an
extremely convenient, versatile, efficient and accurate technique for the present type of problem (Barker, 1982 ;
Hodgkinson et al., 1984). Its convenience and versatility arises from the fact that a far greater range of
problems have simple analytical solutions in the Laplace-transform domain than in the time domain. Solutions
for different boundary conditions require straightforward modifications to the code rather than requiring the
derivation and programming of a different solution.
Essentially, any linear process can be included in the model and any boundary condition which can be
represented as a sum of time-dependent analytic functions can be used. Moreover, the reduction in complexity
in Laplace-transform space means that arbitrary length decay chains can be considered. The efficiency and
accuracy of the method derive from the power of the numerical inversion algorithm due to Talbot (1979). In
fact, Talbot's algorithm can yield answers as accurately and efficiently as obtained by evaluating some analytical
solutions involving special functions. The present models are suitable for incorporation in total systems analysis
codes such as SYVAC (Dormuth and Sherman, 1981 ; Thompson et al., 1984) or LISA (Saltelli et al., 1985).
An additional advantage of employing Laplace transformed solutions is that the net transport through the
system, the mean transport time and its variance can be calculated straightforwardly. This enables an estimate
to be made of the times when there will be significant output from the geosphere before the solution is inverted
for any particular time. This can be used to improve the efficiency of the numerical inversion procedure.
In this paper, the above approach is used to derive solutions to the radionuclide transport equations.
The phenomena included for permeable rock are 1-D advection, linear equilibrium sorption, hydrodynamic
dispersion and radioactive decay. In addition, matrix diffusion (Gluekauf, 1955 ; Neretnieks, 1980) is included
for fractured rock masses. In contrast to previous approaches, with the exception of the solution of Heinrich
(1984) for permeable rock, the results are presented for arbitrary length decay chains. Permeable and fractured
rock solutions are presented in Sections 2 and 3 respectively. These results are combined in Section 4 to give
an approximate solution for transport through a series of different rock types. Possible source terms are
discussed in Section 5, and the methods used to invert the Laplace-transformed solution and verify the
numerical accuracy of the results are decribed in Section 6. Section 7 identifies some potential extensions to
the present models, while some conclusions of this work are drawn together in Section 8. Details of computer
programs and applications to a variety of radionuclide transport problems, will be given in forthcoming papers
and reports.

178

D . P . HODGKINSONand P. R. MAUL
2. PERMEABLE M E D I U M

The migration of radionuclides through a 1-D permeable medium has been considered by a large number
of authors, as discussed in the last section. The justification for adding to the literature on this topic is that
the present approach is very general and encompasses most previous results, with the notable exception of
those treating nonlinear phenomena such as concentration-dependent sorption.
2.1. Conceptual model
The geosphere transport path between the source and biosphere is assumed to be a 1-D column of watersaturated permeable rock (Bear, 1972) with a cross-sectional area, A. The water flows through an interconnected
porosity (qg) with a constant mean velocity, v. In addition to being advected with the flow, dissolved radionuclides are subject to hydrodynamic dispersion and diffusion, with a longitudinal dispersion coefficient, D L.
Transverse dispersion is neglected in the present 1-D approach, although its effects could be accounted for by
an overall dilution factor, if relevant data are available.
Also, linear equilibrium sorption onto the rock is taken into account. This results in a constant retardation
factor, R,, within each rock unit for each member of a decay chain, with n = 1 to N. Of course, in practice
the retardation factor will not be a constant but will be dependent upon a large number of factors such as the
aqueous concentrations of the sorbing element and competing ions, the chemical composition of the groundwater and the physical and chemical properties of the adjacent rock. Since all these factors are likely to vary
in an indeterminate way along the flow path, a reasonable approach is to assume piecewise constant retardation
factors and perform appropriate sensitivity and uncertainty analyses. Finally, radioactive decay and in-growth,
with decay constants 2,, are included in the analysis.
A flexible inlet boundary condition is used, which can account for a wide range of time-dependent inputs,
specified in terms of a linear combination of the pore-water concentration and its first derivative. The outlet
boundary condition is assumed to be zero concentration at an infinite distance from the source. This has been
chosen in the interests of simplicity, in the knowledge that for advectively dominated transport the results are
insensitive to the choice of the outlet boundary condition. Prior to the release of nuclides into the geosphere,
all concentrations are taken to be zero.
2.2. Mathematical model
The basic radionuclide transport equations for the pore-water concentrations, C, (x, t), are of the form (e.g.
SKI, 1984) :
R.

aC.(x,t)
Ot
-

aC.(x,t)
D Lc32C"(X't)
v ~
+
Ox 2

R.2.C.(x,t)+R._12._,C._~(x,t),

(1)

for n = 1 to N, where 2o = 0. A complete list of symbols and units is given at the beginning of the paper.
The initial conditions are :

Co(x, 0) = 0,

(2)

while the boundary conditions are :

[71Cn(0, t) --D ~c~C.(x, t)


3x -

x=

F~(t)
0 = ~bA-

(3)

and
lira C, (x, t) = 0.

(4)

The time-dependent functions, F~,(t), in (3) are the total flowrates of each nuclide into the column. The
constants vI and D t will usually be taken equal to v and D L, in which case the advective and dispersive flux is
specified. However, alternatives such as taking D ~ = 0 to represent a specified concentration boundary
condition, are also possible. A further possibility would be to take v~ = v and D I equal to a typical diffusion
constant, since this is likely to be the major Fickian process close to a repository, while in the geosphere it is
usually appropriate to use a dispersion coefficient, D L, which reflects larger scale inhomogeneities.

1-D

modelling of radionuclide migration

179

The quantities of interest are the generalized radionuclide outflow rates :

~ ( x , t) = 4,A [ v c . ( ~ , t ) - D aC.(x, t)],


j

(5)

where v and D o are constants which will usually be set equal to v and D L. However, if D o = 0 and v = 1/dpA
then the concentration can be calculated. The above expression is evaluated at the length, L, of the column.

2.3. Laplace-transformed solution


The above equations are most easily solved by applying a Laplace transformation (Watson, 1981) which
changes the time derivative in (1) to a multiplication by the Laplace transform variable, s. The equations (1
4) then become :

R.(s+ 2.)C.(x,s) = - v

a~.(x,s) ~-DLa2C"(x's) + R . _ , 2 .
Ox
~x 2

, a~.(x,s) x=

, _

v C,(O,s)-O

~xx

P.(s)

0 = q~-

,~. ,(x,s),

(6)

(7)

and

lirn C,(x, s) = 0,

(8)

where C(x, s) is the Laplace-transformed pore-water concentration.


These equations can be solved by taking account of all the parent nuclide source terms together, or by
considering them one at a time. Both approaches have advantages when coding the equations into a computer
program. The solution for all~parent source terms considered simultaneously is :

C.(x,s) = E Umn(S)am(S)exp [-bm(s)x],

(9)

m= 1,n

where
u..(s) = 1,

(lO)

and
u,,,(s) =

l-I

k. . . .

R,2k

(11)

, gk+,(s)-m.(s)] '

f o r m < n.
The functions 9k(S) in the denominator of (11) are:

9k(s) = Rk(s+2k).

(12)

The coefficients am(s) in (9) are :

am(s)

[V1

Pl(s)

q-l~bm(s)] i=~l,mVim(S)( ~ a

'

(13)

where

Vmm(S) = 1

(14)

and

V,m(S) =

l--I
Rk2k
k=i,m- J [gk (S) --O.,(S)] '

for i < m. Finally, the functions b,.(s) are given by :

(15)

180

D.P. HODGKINSONand P. R. MAUL


t~

bin(s) = 2 ~ c { - 1 + [1 +4DLg,,(s)/v 2] ,/2},

(16)

which in the zero dispersion limit becomes :


lim bin(s) = gm(S)/V.

(17)

DL~ 0

The alternative form of the solution, considering one parent source term at a time, is now presented. The
concentration of nuclide, n, arising from the source term for nuclide, m ( < n), can be written as :

( l-I
) ~
exp[--bk(s)x]
Cmn(X,S) FI(s)
A k=m,n--I~kRk k=m,. [vI+D'bk(s)] 1-[ [gi(s)-g~(s)]

(18)

i=m,n

while for m = n,

C,,(x, s) - P~(s) exp [-b,(s)x]


~bA [vI + D'b.(s)] "

(19)

The total solution is obtained by summing over the parent sources, i.e.

C(x,s)=

Y~ ~,..(x,s).

(20)

m = I,n

The calculated quantities are the Laplace-transformed generalized outflow rates :

(x, s) = q~A[v C. (x, s) - D O

OC.(x,

Ox s) j,

(21)

at a distance, L, from the source.

3. FRACTURED M E D I U M

The physical and chemical processes which occur during the passage of radionuclides through systems of
interconnected fractures are rather poorly understood (Neretnieks, 1985). Nevertheless, it is important to
make some estimates of their behaviour, since fractured rocks are being considered as potential repository
sites in many countries (CEC, 1984). Thus an idealized approach, based on current understanding and the
availability of data, is taken in this paper.

3.1. Conceptual model


Rather than consider a network of intersecting fractures with distributions of lengths, orientations and
positions (Long et al., 1983 ; Schwartz et al., 1983 ; Robinson, 1984 ; Bourke et al., 1985b), the present approach
assumes a set of identical parallel flow paths through the geosphere (Lever et al., 1983). The substructure of
fractures, such as the existence of relatively narrow channels within the fracture plane (Neretnieks, 1985;
Bourke et al., 1985a), is thought to have an important effect on radionuclide transport (Herbert et al., 1985).
However, such effects are very poorly characterized at the present time, and thus in this paper the flow is
assumed to take place between two smooth parallel surfaces separated by a distance, 2b. As a further
approximation, the water velocity, v, is assumed to be uniform across the fracture, with the effects of
nonuniformities being incorporated into the overall hydrodynamic dispersion term (Taylor, 1954). Despite
these drastic simplifications, it should be possible to choose the flow parameters such that pessimistic estimates
are obtained (Bengtsson et al., 1983 ; Hadermann and Roesel, 1985).
In addition to advection through the fractures, the following physical and chemical phenomena are included
in the model. First, linear equilibrium sorption is assumed to take place onto the fracture surfaces and infill
material. This is represented by a constant retardation factor, R,, for each nuclide in a chain. This should be
determined from measurements of surface distribution coefficients (Lever et al., 1983) and estimates of the
true surface area per unit volume of fracture, rather than that implied by the parallel plate aperture, 2b. It is
straightforward to include first-order kinetic sorption in the present model (Hodgkinson et al., 1984) but this

I-D modelling of radionuclide migration

18I

has not been done since such processes are likely to be relatively rapid compared to the long migration times
through the geosphere.
Secondly, hydrodynamic dispersion is included using the diffusion-like approximation with a constant
longitudinal dispersion coefficient, D L. A potentially important retardation phenomenon is the diffusion of
nuclides from water flowing through fractures into stagnant water in the micropores of the rock matrix (Sudicky
and Frind, 1984; Neretnieks, 1980; Lever et al., 1983; Rae and Lever, 1980; Bibby, 1981). Diffusion
perpendicular to the fracture plane is included in the present model, together with linear equilibrium bulk
sorption onto the rock matrix. These are characterized by an intrinsic diffusion coeffcient, D~,, and a capacity
factor, co,, which reflects the physical and chemical storage in the rock matrix (Lever et al., 1983). Diffusion is
assumed to occur up to a distance, a, from the fracture surface. This distance can be regarded as being due to
a symmetry boundary condition between parallel fractures separated by 2a, or else the length-scale for limited
matrix diffusion (Hadermann and Roesel, 1985).
Finally, radionuclide decay and in-growth, with decay constants 2,, are included for chains with an arbitrary
number, N, of members.

3.2. Mathematical model


With the above assumptions, the basic radionuclide transport equations for the concentrations in the
fracture, Ca(x, t) and in the pore water of the rock matrix, C~(x, w, t), are

R.

OC.(x,t)
Ot
-

OC.(x,t)
c32C.(x,t)
D~ c3CP(x,w,t)
-R.2.C.(x,t)+R._,2._,C._,(x,t)
~x
+ DL
Ox 2
{- b
ffww
w:0
(22)

and

c~.

OCP.(x, w, t)
~t

. O2C~(x, w, t)
D'.
-c~.2.CP.(x,w.t)+c~.
~w 2

,2. ,C e. ,(x,w,t),

(23)

where w is the coordinate into the rock marix perpendicular to the fracture and 20 = 0. Also :

~. = dpp+ PKan,

(24)

is the rock capacity factor which is the sum of a physical contribution due to retention in the rock matrix
porosity (q~p) and a chemical term due to linear equilibrium bulk sorption, where Ks. is the mass-based
distribution coefficient and p is the density.
The initial and boundary conditions are :
C . ( x , o) = o,

(25)

C~(x, w, 0) = 0,

(26)

v'C.(O, t ) - D ' OC.(x, t) x=o


~x

F'.(t)

~bA'

(27)

lirn C.(x, t) = 0,

(28)

CP.(x,O, t) = C.(x, t)

(29)

and

OCP (x, w, t)
~ww

~=o = 0.

(30)

The time-dependent functions, F'.(t), in (27) are the total flowrates of each radionuclide into the column,
where qb is the fracture porosity and A the cross-sectional area. As with the permeable rock case, the constants
vI and D l will usually be set equal to v and D ' , but may be set to other values if required.
The quantities of interest are the generalized outflow rates of radionuclides from the rock column,
ANE

13:3-C

182

D . P . HODGKINSONand P. R. MAUL

where v and D will usually be set equal to v and D L. However, the outflow concentration can be calculated
by setting D = 0 and v = 1/4A.

3.3. Laplace-transformed solution


In the absence of matrix diffusion (i.e, when D~, = 0) the transport equations for porous and fractured rock
are identical, and thus the solution is the same as given in Section 2.3. When matrix diffusion is included, the
Laplace-transformed equations are :

R.(s+2.)C2.(x,s)=-v

c~C'~(x,s)
8~+D

I"

c~22.(x,s)
D~. O(?.(x,w,s)w=0
~x 2 -~- b
Ow
+R._I2n_IC.

I(x,s)

(32)

and

~.(s+ 2.)(2~(x,w,s) = Di.

c~w2

+ n_l~,n_l~n I(X,W,S),

(33)

with the boundary conditions :


-

v C.(O,s)--O

i ~Cn(X, S) x=
P(S)
~x
o - 4)A '

(34)

lira (7.(x, s) = 0,

(35)

C~(x, 0, s) = Cn(x, s)

(36)

,~C~(x, w, s)
8w
w=~ = 0.

(37)

and

As in the permeable medium case, the equations can be solved with all the parent nuclide source terms
considered at once, or one at a time. Taking all the source terms together, the solution to the pore-water
concentration equation (33) with boundary conditions (36) and (37) can be written as :

~.(x,w,s)= ~, (2~(x,s)y,..(w,s),
m=

(38)

I,n

where for m = n,

y,,(w, s) =

cosh [4),(s) (a-- w)l


cosh [4~,(s)al

(39)

and for m < n,

ym.(w.s)

ek2k ~
=

ykk(W, s)
1-I

(40)
'

l=m,~t
I#-k

with
q~m(s) = [(s+ 2,.)~/D~,] ,/2.

(41)

In the fracture the general solution is :

C'.(x, s) = Z
m=

with

u,..(s)a,,,(s) exp [-b.,(s)x],


I,n

(42)

1-D modelling of radionuclide migration

b,,(s)

/)

= ~

{ - 1 + [1 +4DLg,.(S)/V 2] 1/2},

183

(43)

just as for the permeable medium case. However, the definitions of y,. (s), u,., (s) and am(s) are more complicated
than before due to the presence of matrix diffusion. In the fractured case, y,,(s) is given by:
y,.(s) = R,,(s+2,,) + ~ - ~b,.(s) tanh
The functions

u,.,(s)

[4Jm(S)a].

(44)

are defined by the recursion relation


[ffn(S)--gm(S)]Umn(S) = Rn

l~n--lUm,n--I(S)~-

2
Okn(S)Umk(S)'
k=m,n- 1

(45)

with

u,~(s)

= 1.

(46)

where
Di

~l~l

B,(s) = - ~ -

~b,tanh

,=~_, D~+~/ ,=,,. I-I

[dpt(s)a]

[~(s)-~i(s)]

(47)

j=k,n
j#l

Finally,

a,.(s)

is written in the same form as for permeable media, namely

am(S)
where the functions

Vim(S) are

1
/~(s)
[v' + D'b.,(s)] i=~I,m ve,.(s) qS~A-'

(48)

defined by the recursion relation for m > i,


vim(s)=-

uj,.(s)vq(s),

(49)

j~i,m-- I

with
Vmm(S) = 1.

(50)

The alternative solution, taking one source term at a time is as follows. In the pores,

CX(x,w,s)=

~ C.,k(X,S)y~.(W,S),

(51)

k=m,n

while in the fractures :

C...(x,s)

P~ (s) ~

Ek,,~(s) exp [-- bk (s)x]

(52)

4~A k~.. [v' +D'bk(s)] l--I [g,(s)-gk(s)]"


i=m,n
i~k

The multiplicative functions Ek,..(S) are much more complicated than in the permeable medium case, but can
be written in closed form as follows :

Ek,,(s)

= 1

(53)

and

E,,..(s) = ~ Tk,..(s, I)Sk,..(s, I).

(54)

The summation in (54) is over all possible choices of the nuclides ( m + 1)... ( n - 1), excluding the nuclide k.
If the chosen nuclides are put in the set L and the remaining nuclides m . . . n are put in the set I', the terms
Tk,..(s, I) and Sk,..(s, I) can be written as :

184

D.P. HODGKINSONand P. R. MAUL

Tkmn(S,I) =

ely 1-I [gt(s) --gk(S)]

i,j~l"

(55)

lel

(i+ l)...(j-- I ) e l

and

Sk,,,(s,t)=

1-[ (B,,,+, +2,R3.

(56)

i,i+ 1~I"

In the expression for Tk,,,,(s, I), if there is no choice of nuclides satisfying i,je I' and (i+ 1)... ( j - 1) e/, then
the term is taken to be unity. A similar consideration applies to Sk,,,(s, I).
The evaluation of the expression for Ekm, is best understood by means of an example. Taking a four nuclide
decay chain with k = 1, m = 1 and n = 4 one obtains :

B~4[gz(s)--g,(s)] [g3(s)--g,(s)l+Bl3[g2(s)--g,(s)] x (B34+23R3)


+Bz4[g3(s)--gl(s)] (B~2+~RI)+(Btz+,~R~) (B23+22R~) (B34+23R3). (57)
The above equation is written in a form so that the contributions from the terms T~ ~4(s,/) and S~4(s, I) can
E l 14(S) =

be seen separately.
The total concentration is again obtained by summing over the parent source terms :

C,(x,s) = ~

Cm,(x,s).

(58)

m-- l,n

The calculated quantities are the Laplace-transformed generalized outflow rates,

l~,(x,s) -= c~A[vC,(x,s)

D O~"(x'
~xx s)-1"

(59)

4. HETEROGENEOUS MEDIA

The exact solutions presented in Sections 2 and 3 assume that the water-borne radionuclides pass through
a single type of rock, with constant physical properties, between the repository and biosphere. For some
geological scenarios this is not an adequate approximation, for example due to the presence of sedimentary
layers with different characteristics or to heterogeneous physical and chemical properties within a single rock
type.
A better approximation in such situations is to consider a path consisting of a succession of rock types with
piecewise constant parameters. In particular, these rock units could be either permeable or fractured.
Thus, we consider a succession of rock units with lengths, Lj, and coordinates, xj, (0 ~< xj ~< Lj) for j = 1
to J. These are ordered such that x ~ = 0 is the inlet from the repository and x~ = Lj is the outlet to the biosphere.
An important question is what boundary conditions are to be used at the interfaces between two rock types.
Since we are dealing with second-order differential equations, two boundary conditions are required. The most
rigorous approach would be to enforce the continuity of the concentrations and the flowrates of each nuclide
through each interface. However, in common with previous approaches of this type (Hadermann, 1980;
Gureghian and Jansen, 1985) an approximate procedure is used here. This is to take one of the boundary
conditions to be zero concentration at infinity. This should be reasonably accurate at high Peclet numbers,
where the transport is mainly by advection and the downstream boundary condition has little effect.
There remains the question of what to take for the second boundary condition. Hadermann (1980) chose
to enforce continuity of the concentrations. However, a more general approach is adopted here. It is assumed
that the generalized outflow rate of each nuclide across an interface is conserved. Thus :
F~,(xj = Lj, t) = F~,(x j+ ~ --=0, t),

(60)

where F~,(x, t) is defined by (5) and (31) for permeable and fractured media. Ifv = v and D = 0 in the above
expressions, and the volumetric flowrates of water (vdpA) in the two media are set equal, then Hadermann's
condition is achieved. However, in most applications it would seem appropriate to choose v = v and D = D L
in each medium, in which case there will be no loss of mass from the system. Furthermore, we do not necessarily
assume equal volumetric water flowrates in each unit. Thus the velocity, porosity and area of the column in

1-D

modelling of radionuclide migration

185

each section is allowed to vary in an unconstrained manner. One reason for allowing this flexibility is that
when a radionuclide plume moves from one sedimentary layer to another, it can become diluted by an
additional flow of uncontaminated water.
It follows from the foregoing discussion that in the second rock unit, the only modifications that are required
to the solutions presented in Sections 2 and 3, are that the Laplace transform of the generalized nuclide inflow
rates, P~.(s), are replaced by the Laplace transform of the nuclide outflow rates from the first rock unit,
F'~.(L~, s). In turn, the output functions from the second unit are used as inputs to the third unit, if required.
In this way, the Laplace transforms of the generalized radionuclide outflow rates for any number of rock
sections can be built up. When this has been done, the complete Laplace-transformed expression is numerically
inverted to the time domain, as described in Section 6.

5. SOURCE TERMS

In the expressions given in earlier sections, the inlet boundary conditions have just been specified by arbitrary
time-dependent functions, F~(t). The method of evaluation described in the next section allows any function
to be used provided that its Laplace transform, r~(s), can be written in terms of analytic functions and has
suitable convergence properties as [sl --' ~ . This substantially increases the power and flexibility of the present
technique. Some examples of possible source terms are discussed below.

5.1. Simple leaching model


A number of authors (Burkholder et al., 1976; Hill and Grimwood, 1978; Lester et al., 1975; Pigford et
al., 1980), have used a simple leaching model for which :

F'.(t) = k . g . ( t ) ,

(61)

where M.(t) is the time-dependent inventory of each nuclide, which satisfies:

r~(t)_2.M.(t)+A._,M._l(t)"

dM.(t)
dt

(62)

Thus for each nuclide the time-dependent inventory of each nuclide is leached by a constant fraction per unit
time, k.. The Laplace-transformed inflow rates for this model are simply :

P.(s) = k. ~" W.fl(s+Aj),

(63)

j ~ l,n

where

W.j = ~. ~

,.=lj

M,.(0)

1~ ( A t - Aj)

(64)

l=m,n
I#j

and
Aj = kj + 2j.

(65)

5.2. Decaying band-release model


A related source-term model which has found wide application, for example in the INTRACOIN project
(SKI, 1984) assumes that the outflow to the geosphere for all nuclides occurs over a finite time, T. During this
time the outflow rate is taken to be the time-dependent radionuclide inventory, arising from radioactive decay
but anaffected by transport to the geosphere, divided by T. In this case :

r~.(s) = ~

~
j == 1 ,n

W'.j
(s+~j)

[1-exp[-(s+2j)T]},

where W~,jis obtained from W.j by replacing terms in A by 2.

(66)

186

D . P . HODGKINSONand P. R. MAUL

5.3. Advection-diffusion models


One of the strengths of the present approach is that it can accommodate a wide variety of source term
models based on advection, dispersion, equilibrium sorption, etc. In fact any transport process can be included
as long as it is linear.
F o r example, the first of the rock units treated in the previous sections could be considered to be a model
of transport through a repository. In this case, the concentration at the far edge of the vault would need to be
specified, perhaps by an estimate of the solubility limit. However, it should be remembered that the downstream
boundary conditions used in the present work are not suitable for diffusively dominated transport.
A number of authors (Chambr6 et al., 1982 ; Sharland and Tasker, 1984) have used advection~liffusion
models, with various geometries, to define the source of nuclides escaping from a repository. These and many
others could be used in conjunction with the present geosphere model.

5.4. Solubility-limited models


One problem with the source term models considered above is that they take no explicit account of elemental
solubility limits, except in the special case where there is only a single important isotope of an element.
In this special case, an extension of the leaching model considered in Section 5.1 (Robinson et al., 1987)
would be to assume that initially the inflow rate to the geosphere is solubility limited, with :

F ~,(t) = QC s,

(67)

where Q is the volumetric flowrate through the repository. However, eventually when most of the nuclide has
been removed from the repository the concentration in the vault pore-water will be limited to

C, (t) = M, (t)/c~, V,

(68)

where a t and V are the capacity factor and volume of the repository. In this case,

r~ (t) = QC, (t).

(69)

Using (62) it can be shown that F~ (t) is given by (67) for 0 ~< t ~< T1, where :

[I+2,M,(O)/QCS]

TI = ~ In [_- 1 ~ - ~

V~

-J'

(70)

T0I,

(71)

and that for t > T1 :

r~ (t) = QC s exp [ - A , ( t where

1 =

~" 1

"~ Q/a, V.

(72)

The Laplace-transform of this solubility-limited inflow rate for a single member chain is :

F',(s)

QC]s 1

A,

exp
(--STl)-j.]
~

(73)

However, in general it is not possible to derive general analytical solutions for actinide source terms, where
solubility limits are important. This is because solubility limits apply to elements, but various isotopes of the
same element occur in different decay chains. Thus the source term equations for all chains must be solved
simultaneously. At times when there are insufficient amounts of the element in the repository to saturate the
water to its limit, the models of the type discussed in Sections 5.1 and 5.2 can be used. However, at other times
the elemental solubility limit becomes the determining factor. This switching between mechanisms means that
the equations become nonlinear and in general have to be solved numerically. A model of this type has recently
been proposed for cemented intermediate-level waste (Robinson et al., 1987).
Even though the solution is generally only defined by a series of numerical values, it should usually be
possible to fit these to a sum of time-dependent functions which can be Laplace-transformed. A reasonable
approximation for many situations would be to write :

1-D modelling of radionuclide migration

187

FI,(t) = ~ d , i exp (-/~,i t),

(74)

whose Laplace-transform is :
/~.(s) = ~

d.~

(75)

(s+#.i)'

where d.~ and #nl are constants which are fitted to the numerical results. This is of the same general form as
(63).

6. NUMERICAL INVERSION AND VERIFICATION

The analytical solutions to the Laplace-transformed equations, presented in Sections 2-4, have been numerically inverted to the time domain using a method proposed by Talbot (1979). This procedure has previously
been found to be highly accurate and efficient for a number of related problems (Barker, 1982; Hodgkinson
and Lever, 1983; Maul, 1986). Talbot's algorithm involves a numerical integration along a contour in the
complex s-plane, which is specified by the dimensionless parameters 3, v and a. For a typical problem the
choice r = 6, v = 1 and a = 0 gives ten-figure accuracy using 32 integration points with exponentially increasing
precision above this.
For some problems, notably those with Peclet numbers exceeding a hundred, the above parameters are not
suitable. Nevertheless, it has generally been found possible to invert such cases by choosing alternative values
for z and/or v. It is often better to invert a summation of exponential terms individually rather than together.
These questions will be discussed in more detail in a later paper.
Terms of the form exp ( - ? s ) , which arise in the zero dispersion limit [see (12) and (17)] and in some source
terms [equation (73)], give rise to exponentially growing terms as Re s ~ - oo which cannot be inverted using
Talbot's algorithm. However, they are easily removed by appropriate use of the relation :
l[exp ( - ? s ) f ( s ) ] = f ( t - ? ) H ( t - ? ) ,

(76)

for 7 > 0, where ~ - 1 symbolises the inverse Laplace transform, H ( t - ? ) is the Heaviside unit step function
and f ( t - 7 ) is a general function.
The numerical accuracy of the method has been verified by comparison with some analytical solutions
(Burkholder and Rosinger, 1980; Pigford et al., 1980) and with some of the benchmark problems used in the
I N T R A C O I N study (SKI, 1984). Furthermore, two independent programs, based on the formulae presented
here, have been compared for a range of further problems.
Before inverting the Laplace transform for any particular time, it is useful to have an estimate of the times
for which the generalized output rates can be expected to be significant. This can be obtained in a straightforward way by evaluating the first three moments as follows,
oF~,(L, t) dt = ff~,(L, 0),

(77)

a ~ (L, s)
f. -

~s

.,= 0

(78)

and
,~

~ 2 ~Os
( L2 , s) s= o -t2..

(79)

Thus the time integrals of the generalized outflow rates, their mean times (T.) and variances (6.2) are known.
Analytical expressions can be obtained for the derivatives in (78) and (79), but in practice it is found more
efficient to evaluate these numerically.

188

D.P. HODGKINSONand P. R. MAUL


7. POSSIBLE EXTENSIONS

The models considered in this paper include first-order approximations to the major migration phenomena
expected to occur in permeable and fractured rock. However, the models of the various migration mechanisms
could be refined and extended in a number of ways. The most important requirement is that they do not make
the equations nonlinear. Some possible extensions are discussed below.

7.1. Kinetic sorption


The present models assume that sorption reactions reach equilibrium instantaneously. For safety assessment
studies this is likely to be a good approximation since the characteristic reaction times are likely to be
considerably less than geosphere migration times. However, in some special circumstances it may be necessary
to include kinetic terms. A particular example is the analysis of field experiments (Hodgkinson and Lever,
1983). The permeable medium model discussed in Section 2 can be extended to include first-order kinetics for
bulk sorption. F o r the fractured case, the kinetics of sorption onto fracture surfaces, and kinetic bulk sorption
in conjunction with matrix diffusion,can similarly be incorporated.

7.2. Matrix diffusion


In the model considered in Section 3, it was assumed that diffusion occurred sideways from a series of
identical parallel fractures. However, other geometrical configurations could be considered. For instance,
diffusion into spherical blocks (Bengtsson et al., 1983) and diffusion from a cylindrical hole (Hadermann and
Roesel, 1985) could be incorporated into the present approach in a straightforward manner. A general
formalism for such models has been developed by Barker (1985).
A further possible refinement would be to include a surface layer close to the fracture having different
characteristics from the bulk rock matrix. Also, the effect of dead-end pores could be included (Hemingway
et al., 1983).
7.3. Boundary conditions
The boundary condition at the end of the column, or at the interface between two rock units could be made
more general. However, when advection is more important than dispersion, the results are rather insensitive
to the outlet boundary condition, while the algebraic expressions become much more complicated.

8. CONCLUSIONS

The technique of numerically inverting analytical solutions to Laplace-transported radionuclide transport


equations has been found to be very efficient, accurate and flexible. Moreover, in the Laplace-transform
domain, the analytical solutions have a much simpler algebraic form than in the time domain. This has enabled
us to write down closed expressions for the transport of nuclides in arbitrary length decay chains through both
permeable and fractured media. Moreover, these solutions can be combined to given an approximate treatment
of transport through heterogeneous media consisting of an arbitrary number of rock units with piecewise
constant parameters. The compact and general nature of the solutions simplifies the computer programming
and thereby lessens the likelihood of errors. This is to be contrasted with the extremely complicated analytical
expressions in the time domain for three member chains in a permeable medium with a straightforward inlet
boundary condition (Burkholder and Rosinger, 1980). The inlet boundary conditions used in the present work
can be very general functions of time, or solutions to differential equations involving advection, dispersion
and sorption. Furthermore, the model presented here can be extended in a straightforward manner to other
simple geometries, or to include other linear processes. The numerical accuracy of the method has been checked
by comparison with analytical solutions and benchmark problems which formed part of the I N T R A C O I N
project.

Acknowledgements--We would like to thank D. A. Lever for useful discussions. This work is published with the permission
of the Central Electricity Generating Board.

I-D modelling of radionuclide migration

189

REFERENCES

Atkinson R., Herbert A. W., Jackson C. P. and Robinson P. C. (1985) Report AERE-R.11365.
Barker, J. A. (1982) Adv. War. Resource. 5, 98.
Barker J. A. (1985) J. Hydrol. 77, 263.
Bear J. (1972) Dynamics o f Fluids in Porous Media. Elsevier, Amsterdam.
Bengtsson A., Magnusson M., Neretnieks I. and Rasmuson A. (1983) SKBF Technical Report KBS TR83-48. Swedish
Nuclear Fuel Supply Co./Division KBS.
Bibby R. (1981) Wat. Resource Res. 17, 1075.
Bo P. (1978) COLUMN: Numerical Solution o f Migration Equations Involving Various Physiochemical Processes. Chemistry
Department, Riso National Laboratory.
Bourke P. J., Derlich S., Goblet P. and le Manse D. (1985a) Proc. SecondEuro. Community Conf. Radioactive Waste Mgmnt
Disposal, Luxembourg.
Bourke P. J., Durrance E. M., Heath M. J. and Hodgkinson D. P. (1985b) Report AERE-R.11414.
Broyd T. W., Dean R. B., Hobbs G. D., Knowles N. C., Putney J. M. and Wringley J. (1983) Nuclear Science and Technology
Report EUR 8669, Commission of the European Communities.
Burkholder H. C. and Rosinger E. L. J. (1980) NucL Technol. 49, 150.
Burkholder H. C., Cloninger M. O., Baker D. A. and Janson G. (1976) Nud. Technol. 31,202.
Chambr6 P. L., Pigford T. H., Fujita A., Kanki T., Kobayashi A., Lung H., Ting D., Sato Y. and Zavoshy S. J. (1982)
Report LBL-13842, Vol. II, Lawrence Berkeley Laboratory.
Commission of the European Communities (CEC) and the OECD Nuclear Energy Agency (NEA) (1984) Geological Disposal
o f Radioactive Waste : An Overview o f the Current Status of Understanding and Development (OECD).
Dormuth K. W. and Sherman G. R. (1981) Report AECL-6814, Atomic Energy of Canada Limited.
Ewart F. T., Sharland S. M. and Tasker P. W. (1985) Proc. Ninth Syrup. Scientific Basis NucL Waste Mgmnt, Stockholm.
Glueckauf E. (1955) Trans Faraday Soc. 51, 1541.
Gureghian A. B. and Jansen G. (t985) Wat. Resource Res. 21, 733.
Hadermann J. (1980) Nucl. TechnoL 47, 312.
Hadermann J. and Roesel F. (1983) EIR Report 506, Swiss Federal Institute for Reactor Research.
Hadermann J. and Roesel F. (1985) EIR Report 551, Swiss Federal Institute for Reactor Research.
Harada M., Chambr6 P. L., Foglia M., Higashi K., Iwamoto F., Leung D., Pigford T. H. and Ting D. (1980) Report LBL10500, Lawrence Berkeley Laboratory.
Heinrich W. F. (1984) Technical Record TR-286, Atomic Energy of Canada Limited.
Hemingway S. J., Bradbury M. H. and Lever D. A. (1983) Report AERE-R.10591.
Herbert A. W., Hodgl(inson D. P., Jackson C. P., Lever D. A., Rae J., Robinson P. C., Sharland S. M. and Tasker P. W.
(1985) Report TP.1131, U.K. Atomic Energy Authority.
Hill M. D. and Grimwood P. D. (1978) Report NRPB-R.69, National Radiological Protection Board.
Hodgkinson D. P. and Lever D. A. (1983) Radioact. Waste Mgmnt Nucl. Fuel Cycle 4, 129.
Hodgkinson D. P., Lever D. A. and England T. M. (1984) Ann. nucl. Energy 11, I 11.
INTERA Environmental Consultants Inc. (1983) Report ONWI-457.
Lester D. H., Jansen G. and Burkholder H. C. (1975) AIChE Symp. Ser. 71,202.
Lever D. A., Bradbury M. H. and Hemingway S. J. (1983) Prog. Nucl. Energy 12, 85.
Long J. C. S., Remer J., Wilson C. and Witherspoon P. A. (1983) Wat. Resource Res. 18, 645.
Maul P. R. (1983) Report TPRD/B/0189/N82, Central Electricity Generating Board.
Maul P. R. (1984) Report TPRD/B/0555/N84, Central Electricity Generating Board.
Maul P. R. (1986) Rad. Prot. Dosim. 14, 207.
McGrail B. P., Chick L. A. and McVey G. L. (1985) Nucl. Technol. 69, 114.
Neretnieks I. (1980) J. Geophys Res. 85, 4379.
Neretnieks I. (1985) Paper presented at the 1AH 17th Int. Cong. Hydrol. Rocks Low Permeability, Tucson, Arizona.
Pigford T. H., Chambr6 P. L., Albert M., Foglia M., Harada M., Iwamoto F., Kanki T., Leung D., Masuda S., Muraoka
S. and Ting D. (1980) Report LBL-11616, Lawrence Berkeley Laboratory.
Rae J. and Lever D. A. (1980) Report TP.853, U.K. Atomic Energy Authority.
Rasmuson A. (1984) Wat. Resource Res. 20, 1435.
Rasmuson A., Narasimhan T. N. and Neretnieks 1. (1982) War. Resource Res. 18, 1479.
Robinson P. C. (1984) Report TP.1072, U.K. Atomic Energy Authority.
Robinson P. C., Hodgkinson D. P., Tasker P. W., Lever D. A., Windsor M. E., Grime P. W. and Herbert A. W. (1987)
Report AERE-R.11854.
Saltelli A., Bertozzi G. and Stanners D. A. (1985) LISA ." A Code for Safety Assessment in Nuclear Waste Di~sposals."
Program Description and user Guide (Commission of the European Communities Report).
Schwartz F. W. (1977) J. HydroL 32, 257.
Schwartz F. W., Smith L. and Crow S. S. (1983) Wat. Resource Res. 19, 1253.
Sharland S. M. and Tasker P. W. (1984) Report AERE-R.11240.
SKI (1984) Report SKI 85 : 3, Swedish Nuclear Power Inspectorate.
Sudicky E. A. and Frind E. O. (1984) War. Resource Res. 20, 1021.
Talbot A. (1979) J. Inst. Math. Applic. 23, 97.
Taylor G. I. (1954) Proc. R. Soc. A225, 473.
Thompson B. G. J., Hall P. and Duncan A. G. (1984) BNES ConJi, London.
Watson E. J. (1981) Laplace Transforms. VNR New Mathematics Library.
ANE 1 3 : 3 - D

S-ar putea să vă placă și