ArticlePDF Available

Dispersion of Groundwater Age in an Alluvial Aquifer System

Wiley
Water Resources Research
Authors:

Abstract and Figures

[1] Interpretation of groundwater ages typically rests on assumptions of minimal mixing of different water ages in the water samples. The effects of three-dimensional, geologic heterogeneity on groundwater mixing and tracer concentrations, however, have not been evaluated. In this study, we use a series of 10 detailed geostatistical realizations along with high-resolution numerical groundwater flow and contaminant transport simulation to model distributions of groundwater ages and chlorofluorocarbon (CFC) ages at wells within a heterogeneous stream-dominated alluvial fan aquifer system. Results show that groundwater reaching a well in the heterogeneous aquifer system typically consists of a wide distribution of groundwater ages (often spanning >50 years), even over short (
Content may be subject to copyright.
Dispersion of groundwater age in an alluvial aquifer system
Gary S. Weissmann,
1
Y. Zhang, Eric M. LaBolle, and Graham E. Fogg
2
Hydrologic Sciences, University of California, Davis, California, USA
Received 28 August 2001; revised 9 April 2002; accepted 25 April 2002; published 18 October 2002.
[1]Interpretation of groundwater ages typically rests on assumptions of minimal mixing
of different water ages in the water samples. The effects of three-dimensional, geologic
heterogeneity on groundwater mixing and tracer concentrations, however, have not been
evaluated. In this study, we use a series of 10 detailed geostatistical realizations along with
high-resolution numerical groundwater flow and contaminant transport simulation to
model distributions of groundwater ages and chlorofluorocarbon (CFC) ages at wells
within a heterogeneous stream-dominated alluvial fan aquifer system. Results show that
groundwater reaching a well in the heterogeneous aquifer system typically consists of a
wide distribution of groundwater ages (often spanning >50 years), even over short (<1.5
m) screened intervals. Additionally, simulated arithmetic mean groundwater ages do not
correspond to mean ages estimated from simulated CFC concentrations. Results
emphasize the potential ambiguity of ‘‘mean’’ groundwater ages estimated from
environmental tracer concentrations in typically heterogeneous geologic systems. The
significant dispersion of groundwater ages also implies that ultimate, maximum effects of
nonpoint source, anthropogenic contamination of groundwater may not be reached until
after many decades or centuries of gradual decline in groundwater quality. INDEX TERMS:
1829 Hydrology: Groundwater hydrology; 1832 Hydrology: Groundwater transport; KEYWORDS:
heterogeneity, groundwater age, environmental tracer, chlorofluorocarbons, transport, modeling
Citation: Weissmann, G. S., Y. Zhang, E. M. LaBolle, and G. E. Fogg, Dispersion of groundwater age in an alluvial aquifer system,
Water Resour. Res.,38(10), 1198, doi:10.1029/2001WR000907, 2002.
1. Introduction
[2] The use of environmental tracers, such as chlorofluor-
ocarbons (CFCs), tritium/helium-3, and carbon-14, for age
dating groundwater has made it possible to estimate mean
ages of groundwater pumped from wells [e.g., Solomon and
Sudicky, 1991; Solomon et al., 1992; Hackley et al., 1992;
Plummer et al., 1993; Busenberg et al., 1993; Cook et al.,
1995; Johnston et al., 1998; Varni and Carrera, 1998;
Plummer and Sprinkle, 2001]. In concept, groundwater
age dates provide important information on sources and
timing of recharge and vulnerability of aquifers to contam-
ination. An age date of 40 years, for example, might suggest
that surface contaminant sources that have only existed for
20 years could not be present in the 40-year-old water.
Accordingly, there is growing interest among regulatory
agencies in using groundwater age dates to map suscepti-
bility of groundwater systems to contamination.
[3] Many papers state or imply that the validity of the
age-dates is suspect if dispersion and/or long screened
intervals create significant mixing of ages [Walker an d
Cook, 1991; Mazor and Nativ, 1992; Goode, 1996; Varni
and Carrera,1998;To m p s o n e t a l . , 1999; Bethke and
Johnson, 2002]. Hinkle and Snyder [1997] noted that
mixing a small portion of younger water that contains
CFC with older water that is CFC-free will result in
reporting of a CFC age date that is between 1944 and the
date of recharge of the modern water in the mixture, even if
the water sample consists of a low percentage of younger
water. Thus groundwater ages determined from CFC data
may not represent the weighted mean age of groundwater
where mixing is significant. Varni and Carrera [1998]
presented similar results indicating this influence of mixing
to be true for any environmental tracer in a heterogeneous
aquifer. Though several writers have indicated that the
influence of heterogeneity and groundwater mixing due to
hydrodynamic dispersion might produce uncertainty in
environmental tracer-based age date results [Busenberg
and Plummer, 1992; Dunkle et al., 1993; Reilly et al.,
1994; Johnston et al., 1998; Varni and Carrera, 1998;
Burow et al., 1999; Tompson et al., 1999], the magnitude
of this effect is largely unknown. It is often assumed,
however, that age dating of waters from short screened
intervals (1 m) will eliminate significant mixing of differ-
ent groundwater ages from different depths.
[4]Bethke and Johnson [2002] used the age-mass
equation of Goode [1996] to analytically model effects
of mass transfer between regional aquifer and aquitard
layers. They showed that for very long times of transport,
a conservative environmental tracer that undergoes advec-
tion, dispersion, and diffusion (Brownian motion) will, in
the limit, become distributed among the aquitard and
1
Now at Department of Geological Sciences, Michigan State University,
East Lansing, Michigan, USA.
2
Also at Department of Geology, University of California, Davis,
California, USA.
Copyright 2002 by the American Geophysical Union.
0043-1397/02/2001WR000907$09.00
16 -1
WATER RESOURCES RESEARCH, VOL. 38, NO. 10, 1198, doi:10.1029/2001WR000907, 2002
aquifer materials in direct proportion to the relative vol-
umes of those materials. Because diffusive mass transfer
into the aquitards must be balanced by diffusive mass
transfer out of the aquitards, mean groundwater ages in the
aquifers can be substantially older than ages that would be
indicated by a simple piston displacement model for the
aquifer alone.
[5] The purpose of our work is to examine the role of
dispersion on groundwater age dates in an alluvial aquifer
complex containing typical heterogeneity in hydraulic con-
ductivity. While Bethke and Johnson [2002] examined
steady state effects of age mixing among aquifer and
aquitard layers, wherein diffusion is the dominant transport
process in the aquitards, we examine the effects of advec-
tion, dispersion, and diffusion in a multiscale model of
heterogeneity containing multiple hydrostratigraphic units,
some of which are dominated by diffusion. Bethke and
Johnson’s [2002] analysis is most applicable for under-
standing potential age mixing of old groundwaters (say,
>10,000 yr) dated by radiometric tracers that have reached a
steady state in terms of mass transfer between tight aqui-
tards and the aquifers. In contrast, the present study applies
more to dating of younger groundwaters with environmental
tracers that have not migrated long enough or far enough to
have reached equilibrium in terms of mass exchange among
the various hydrostratigraphic units.
[6] We also explore whether, as often suggested in the
age-date literature, the age distribution is appropriately
narrow when the screened interval is small. The literature
contains limited quantitative evaluations of these assump-
tions with field data or with models [e.g., Varni and Carrera,
1998; Tompson et al., 1999]. In particular, effects of dis-
persion due to heterogeneity and mixing due to well screen
length on age dates has, to our knowledge, never been
investigated in typically complex alluvial materials.
2. Study Area
[7] We tested the effects of heterogeneity on groundwater
age distributions in the Kings River Alluvial Fan aquifer
system, located southeast of Fresno, California (Figure 1).
The Kings River deposited a stream-dominated alluvial fan
where it exits the Sierra Nevada into the San Joaquin Valley,
thus the aquifer consists of a highly heterogeneous mix of
lithologies ranging from cobble to mud (i.e., silt and clay)
[Weissmann, 1999; Weissmann and Fogg, 1999; Weissmann
et al., 2002]. Five hydrofacies were recognized within this
aquifer system based on core descriptions. These include, in
decreasing order of hydraulic conductivity, gravel, sand,
muddy sand, mud, and paleosol [Burow et al., 1997;
Weissmann, 1999; Weissmann and Fogg, 1999]. Addition-
ally, stratigraphic studies show that the aquifer consists of
five depositional sequences, where each sequence consists
of a heterogeneous composite of the gravel, sand, muddy
sand, and mud hydrofacies, and each sequence is bounded
by laterally continuous unconformities marked by the pale-
osol hydrofacies [Weissmann, 1999; Weissmann and Fogg,
1999; Weissmann et al., 2002].
[8]Burow et al. [1999] reported collection of CFC
samples in 1994 and 1995 from several discrete intervals
within the Kings River alluvial fan. Twenty multilevel
nested sampling wells at six locations (B1, B2, B2.5, B3,
B4, and B5) were used for this sampling. CFC-based age
dates indicate that groundwater ages range between 2 and
>55 years at the time of this 1994-95 sampling (Table 1).
3. CFC Age Dating Method
[9] Age dating of groundwater with CFC is based on
correlation of variable atmospheric tracer concentrations
through time to tracer concentrations in the groundwater
at sampling locations. For instance, CFC-11 (CCl
3
F) and
CFC-12 (CCL
2
F
2
) concentrations in the atmosphere have
steadily increased since the 1940s (Figure 2) [Elkins et al.,
1993, 1999; Plummer et al., 1993]. CFCs are soluble in
water, and the CFC concentration in water (C
CFC
) is related
to the atmospheric CFC partial pressure by Henry’s Law
CCFC ¼KCFC;TPCFC ð1Þ
where K
CFC,T
is the Henry’s Law constant for the CFC
compound at temperature, T, and P
CFC
is the atmospheric
partial pressure of the CFC compound [Warner and Weiss,
1985; Plummer et al., 1993]. Assuming meteoric water
recharged to the groundwater is at equilibrium with the
atmosphere with respect to the CFCs, estimates of CFC
partial pressure in the atmosphere at time of recharge can be
obtained from groundwater sample CFC concentrations by
application of equation (1). A groundwater age is estimated
by relating this estimated past CFC partial pressure to
historic CFC atmospheric concentrations shown in Figure 2
[Plummer et al., 1993; Busenberg et al., 1993].
[10] Several assumptions in estimating groundwater ages
from CFC data are outlined by Plummer et al. [1993],
including (1) the recharge temperature is known, (2) col-
lection procedures do not allow contact with the atmosphere
or other sources of CFCs, (3) no local sources of CFC
contamination in the groundwater exist outside of the
background atmospheric levels, and (4) groundwater con-
centrations are not significantly affected by hydrodynamic
dispersion, thus piston-type transport of the tracer can be
used to approximate tracer movement in the groundwater
system.
4. Modeling Heterogeneity, Groundwater Flow,
and Environmental Tracer Transport
[11] The modeling approach used to understand the
influence of heterogeneity on groundwater ages and CFC
concentrations at a well was accomplished in three steps: (1)
model the heterogeneity through application of transition
probability geostatistics within a sequence stratigraphic
framework [Weissmann and Fogg, 1999], (2) simulate
steady state groundwater flow using a block-centered finite
difference model, and (3) estimate water travel time from
the water table to wells using backward-in-time random
walk particle tracking [Uffink, 1989; Fogg et al., 1999;
LaBolle et al., 2000]. The following sections outline this
approach.
4.1. Modeling Heterogeneity
[12] To simulate the heterogeneous distribution of hydrof-
acies within the Kings River alluvial fan aquifer system, we
used a transition probability geostatistics approach [Carle
and Fogg, 1996; Carle et al., 1998; Weissmann et al.,
1999]. Weissmann and Fogg [1999] describe application
16 -2 WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE
of this approach to the Kings River Alluvial Fan study area.
The following discussion is a brief summary of the geo-
statistical work.
[13] In the transition probability geostatistical approach,
transition probabilities between categorical data (e.g.,
hydrofacies) are measured in the vertical and horizontal
directions and a Markov chain model is fit to these
measured results. The 3-D Markov chain model is then
used in sequential indicator simulation followed by simu-
lated annealing to produce realizations of the subsurface
heterogeneity. Importantly, ‘‘soft’’ or interpreted geologic
information, such as facies mean lengths, global facies
proportions, and facies juxtaposition relationships, can be
incorporated into the Markov chain model of spatial vari-
ability [Carle et al., 1998]. Because close-spaced data are
plentiful in the vertical direction from core and geophysical
well logs, fitting a model to measured vertical transition
probabilities is straightforward. To estimate the more diffi-
cult horizontal components of the Markov chain model, we
analyzed C-horizon textures from soil surveys in a geologic
Figure 1. A map showing the study area location southeast of Fresno, California. The six USGS well
nest locations are indicated in the study area boundaries (modified from Burow et al. [1999]).
WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE 16 -3
depositional systems context, as described in Weissmann et
al. [1999] and Weissmann and Fogg [1999].
[14] To deal with the nonstationarity created by the multi-
ple stratigraphic sequences and to represent discontinuity
across the unconformity boundaries, we modeled each strati-
graphic sequence separately [Weissmann and Fogg, 1999].
Additionally, the paleosol hydrofacies marks the laterally
continuous unconformity surface and was formed over the
exposed fan surface on the other four hydrofacies; therefore
the paleosols were modeled separately and later overprinted
on the simulated geology of each sequence. Assembly of
conditional simulation results for each sequence together
with the paleosols resulted in a nonstationary, multiscale
model of heterogeneity [Weissmann and Fogg, 1999].
[15] The three-dimensional cell size of 100 m, 200 m, and
0.5 m in the depositional strike, depositional dip, and
vertical directions, respectively, was selected using two
criteria: (1) these lengths are significantly less than the
mean lengths of each facies [Weissmann and Fogg, 1999]
and (2) sufficiently large to avoid computational limitations
during groundwater flow simulation. The overall dimen-
sions of the simulated region are 6300 15,000 100.5 m
in the depositional strike, depositional dip, and vertical
directions, respectively.
[16] A series of 10 equally probable realizations were
developed for this study. Figure 3 shows one realization of
the Kings River Alluvial Fan hydrostratigraphy used in the
groundwater flow and random walk particle tracking sim-
ulations described in the following sections.
4.2. Groundwater Flow Simulation
[17] A three-dimensional steady state modeling approach
was used to simulate groundwater flow conditions in the
aquifer with the finite difference code MODFLOW-96
[McDonald and Harbaugh, 1988; Harbaugh and McDo-
nald, 1996]. The study area was discretized into approx-
imately 950,000 cells within a simulation grid of overall
dimensions of 6300 15,000 100.5 m in the depositional
strike, depositional dip, and vertical directions, respectively,
from the geostatistical realizations. The constant cell dimen-
sions from the geostatistical realizations (100 200 0.5
m) were used for the numerical flow simulation grid.
Table 1. Reported CFC Age Dates From USGS Wells Within the Study Area Compared to Mean Simulated CFC Age Dates for the 10
Realizations From This Study
a
Well and Screen Number Screen Midpoint
Elevation,
m amsl
Simulated Residence
Time, years
Reported CFC Ages,
years
Simulated CFC Ages, years
Mean SD CFC-11 CFC-12 CFC-11 CFC-12
Mean SD Mean SD
B1-1 85.2 22.4 11.3 20 8 20.6 8.7 20.7 9.1
B1-2 58.7 a a 38 41 a a a a
B1-3 28.2 a a 37 48 a a a a
B2-1 84.3 31.6 14.7 16 2 14.0 3.7 13.7 3.8
B2-2 81.8 33.3 16.1 8 4 14.4 3.9 14.1 3.9
B2-3 68.1 57.2 12.5 22 18 26.7 4.5 27.8 4.8
B2.5-1 64.2 81.0 32.2 31 39 33.6 12.0 34.9 12.4
B2.5-2 52.9 144.6 64.8 29 32 49.9 6.0 51.1 4.6
B3-1 85.2 17.7 9.6 6 3 9.7 2.2 9.4 2.3
B3-2 72.1 34.2 11.7 b 16 17.3 3.1 17.5 2.9
B3-3 54.1 87.0 37.7 32 33 32.6 5.5 34.4 6.4
B3-4 46.5 96.9 38.9 29 41 38.1 7.8 40.5 8.9
B3-5 25.8 a a 35 >55 a a a a
B4-1 81.2 18.8 9.2 7 2 10.9 3.3 10.7 3.6
B4-2 70.0 43.6 16.3 b b 18.7 2.2 18.9 2.5
B4-3 48.9 104.4 33.1 b b 36.4 3.9 39.5 4.7
B4-4 25.4 124.7 39.0 b 30 45.3 4.0 49.3 4.0
B5-1 78.2 23.9 18.3 14 6 14.3 7.7 14.5 8.0
B5-2 54.4 52.1 18.1 26 25 25.4 4.9 26.4 5.5
B5-3 20.9 124.5 32.4 b b 49.7 4.8 52.2 3.1
a
Also included are simulated mean ages of groundwater at selected wells. Reported CFC ages are from Burow et al. [1999]. Here a indicates that
significant particles through upgradient boundary (>15%), and therefore groundwater age was indeterminate; b indicates that measured CFC-based age
dates were not reported by Burow et al. [1999].
Figure 2. Atmospheric concentrations of CFC-11 and
CFC-12 since 1940. Concentration values between 1940
and 1977 are from Busenberg et al. [1993]. Concentration
values between 1977 and 1994 were taken from the NOAA/
CMDL Niwot Ridge Station, Colorado, as reported by
Elkins et al. [1999].
16 -4 WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE
[18] Hydraulic conductivity (K) was assigned to each cell
based on the simulated hydrofacies determined by the
geostatistical realization for each scenario. Values used for
K of each hydrofacies (Table 2) were estimated from
pumping test, slug test, laboratory core measurements, and
literature estimates for similar lithologies [Burow et al.,
1999; Weissmann, 1999].
[19] General head boundary conditions [McDonald and
Harbaugh, 1988] were used in the modeling to simulate
inflow and outflow through the lateral and basal boundaries
of the model. Hydraulic head values were defined for these
boundaries using a general gradient of 0.002 parallel to the
stratigraphic dip direction. Minor adjustments to the boun-
dary head values were made along the stratigraphic dip
boundaries to account for local gradient variability reported
in water-elevation maps of the study area [Fresno Irrigation
District, 1993]. No vertical gradient was assigned because
measured vertical gradients in the area are small and
variable in both upward and downward directions [Burow
et al., 1999]. The boundary condition for the top of the
model represents constant recharge to the water table at a
rate of 150 mm/year [Burow et al., 1999].
[20] Groundwater pumpage was not included in the
model for two reasons: (1) the previous characterization
and modeling study of the same area [Burow et al., 1999]
showed no perturbations in hydraulic head suggestive of
significant pumping influence on horizontal or vertical
gradients and (2) wells in the area are not metered. The
main effect of pumping would be to increase downward
groundwater velocities, yet measured heads in the USGS
well nests show no pervasive, downward gradients. Leakage
out of the lower boundary of the model into deeper, finer-
grained alluvial materials may, in fact, be compensating for
lack of specified pumpage in the model. Effects of pumpage
on general characteristics of the local groundwater age
Figure 3. Geostatistical realization 5 from the Kings River Alluvial Fan system.
Table 2. Hydraulic Conductivity Values Assigned to the Five
Hydrofacies
Hydrofacies K, m/s
Gravel 1 10
2
Sand 1 10
3
Muddy sand 1 10
5
Mud 1 10
6
Paleosol 1.3 10
7
Figure 4. Measured hydraulic head [from Burow et al.,
1999] plotted against simulated mean hydraulic head at 20
monitoring well locations for the 10 realizations. Data
points show mean results and error bars show the range of
values from the 10 realizations.
WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE 16 -5
distributions is expected to be minimal, but should be
investigated in future studies.
[21] Figure 4 shows the simulated steady state hydraulic
head at the midpoint of the 20 well screens plotted against
measured head for these locations from Burow et al. [1999].
The average root-mean square error in head at the 20
measuring points for the 10 realizations is 2.41 meters, with
r
2
value of near 1. Although modeled head values do not
exactly match the measured values, simulated head values
are close and the flow simulations preserved both the
variability in vertical gradients and the overall observed
horizontal gradients within the study area.
[22] The measured heads indicate no pervasive trend in
vertical hydraulic gradients [Burow et al., 1999], but rather
shifting gradients from upward to downward, apparently
due to flow around lenticular heterogeneities in three-
dimensions. The simulated heads exhibit similar behavior,
except at location B4, where the model tends to have
downward gradients. In addition to flow simulations
through the heterogeneous realizations, we simulated flow
through a homogeneous K field for comparison purposes.
Anisotropic values for K of 3.6 10
4
m/s, 4.3 10
4
m/
s, and 1.7 10
5
m/s for the depositional strike, deposi-
tional dip, and vertical directions, respectively, were esti-
mated from effective K properties of the geostatistical
realizations. Effective K values in all three directions were
determined through simulation of flow in the strike, dip, and
vertical directions, as described by Weissmann [1999].
Similar boundary conditions to those previously described
were used for these ‘‘homogeneous’’ simulations.
4.3. Particle Tracking Simulation Approach
[23] Transport simulations employed a random-walk par-
ticle tracking method (RWPM), described by LaBolle et al.
[1998, 2000]. Application of the RWPM to simulate the
forward-time advection-dispersion equation (ADE) models
future particle distributions, with density representing solute
concentration given knowledge of past distributions. When
applied to simulate the backward-time ADE, the RWPM
equation models past particle locations, with density repre-
senting probability given that future positions are known
[Uffink, 1989; Fogg et al., 1999; Neupauer and Wilson,
1999]. This procedure accurately models transport in a
heterogeneous porous media, where sharp contrasts exist
between lithologies. Details of the RWPM are described by
Uffink [1989], LaBolle et al. [1996, 1998, 2000], and
LaBolle and Fogg [2001], and are summarized in this
section.
[24] In this approach, we assume that transport is
described by an ADE of the form:
@nc
@t¼@vic
@xi
þ@
@xi
nDij
@c
@xj
 ð2aÞ
Dij ¼aTv
jj
þD0
d

dij þaLaT
ðÞvivjv
jj=ð2bÞ
where cis the concentration, vis pore water velocity, D
ij
is
a dispersion tensor, nis effective porosity, D
d
0
is the
molecular diffusivity in a porous media, d
ij
is the Dirac
delta function, and a
L
and a
T
are longitudinal and
transverse dispersivities, respectively. The dependent vari-
able of (2a) can be replaced with the conditional probability
p(x,t|s,y), where c(x,t)=p(x,t) = p(x,t|s,y)p(s,y) with initial
condition [Arnold, 1974]
lim
t!spx;tjy;sðÞ¼dxyðÞ ð3Þ
where xis the final location of the particle at time tand yis
the initial location of the particle at time s. The form of the
backward-time ADE for constant nis given by the adjoint
of (2a) [Arnold, 1974]
@p
@s¼vi
@p
@yi
@
@yi
Dij
@p
@yj
 ð4aÞ
lim
s!tpx;tjy;sðÞ¼dxyðÞ ð4bÞ
The diffusion operator is said to be self-adjoint because it
takes the same form in both the forward and backward
processes. The advection term is not self-adjoint unless the
velocity is divergence free, in which case (4a) is equivalent
to
@p
@s¼@vip
@yi
@
@yi
Dij
@p
@yj
 ð5Þ
In reference to the forward equation, pin the backward
equation (5) is evolving in sfrom some final time to some
initial time. If, for convenience, one reverses time in (5) such
that positive values of strepresent evolution backward in
time, the backward ADE (5) simply takes the form of the
forward ADE (2a) with the velocity reversed. Both
equations are based on conditional probabilities. The
backward equation gives information on locations where
contaminants may have originated in the past [Uffink, 1989;
Fogg et al., 1999]. This backward-time ADE was used to
compute the ages of groundwater entering a monitoring well
from travel time between the well screen and water table.
[25] Parameter values required for the RWPM include
velocities calculated from the flow simulations, molecular
diffusivities, effective porosity, and longitudinal and trans-
verse dispersivities. The molecular diffusivity (D
d
0
)was
estimated from
D0
d¼Dd
t2ð6Þ
where D
d
is the diffusion coefficient in a liquid [Grathwohl,
1998] and tis tortuosity defined as
t¼Le
Lð7Þ
where L
e
is the flow path length and L is the straight-line
distance [Carman, 1956; Grathwohl, 1998; Fetter, 1999].
Using D
d
=210
9
m
2
/s, an average of diffusion
coefficients for common ions cited by Domenico and
Schwartz [1998], and a tortuosity of 1.7 [Carman, 1956], D
d
0
was estimated to be approximately 6.9 10
10
m
2
/s. An
effective porosity value of 0.33 was applied for these
simulations.
[26]LaBolle [1999] and LaBolle and Fogg [2001]
showed that simulation results are insensitive to the value
16 -6 WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE
of local-scale a
L
. This is true because the spreading due to
longitudinal dispersion is insignificant compared to that
caused by the hydrofacies-scale heterogeneity that is cap-
tured in the geostatistical simulations. Furthermore, the
particular random walk particle algorithm that accurately
balances mass across sharp interfaces runs faster when
dispersivity is isotropic. Dispersivity was therefore treated
as isotropic, with a
L
=a
T
= 0.04 m, a value similar in
magnitude to reported transverse dispersivities at field sites
similar to the scale of our cells [e.g., Freyberg,1986;
Sudicky, 1986; Garabedian et al., 1991; Farrell et al.,
1994; Mallants et al., 2000]. Because some uncertainty in
this value exists, we conducted sensitivity runs with aset to
0.1 and 0.01 m. These simulations showed minimal effects
on the results reported here (Figure 5).
4.4. Numerical Simulation of Groundwater Ages
[27] To assess the age distribution of water entering a well
screen, we simulated backward-in-time release of 5000
particles from a discrete point located at the midpoint of
17 monitoring well screens at various depths in the study
area (Table 1). Thus these simulations reflect groundwater
age distributions at a discrete point in the aquifer. Particles
in these simulations represent parcels of water of various
ages. Using the steady state velocity fields along with the
backward-time particle tracking approach previously
described, the travel times for particles between these
monitoring well screen midpoints and the water table were
simulated for each of the 10 heterogeneous realizations and
the homogeneous system. Because the backward-moving
particles reach the water table at various times, the results of
this procedure produces probability distribution of ground-
water residence time at a well (Figure 6). Average residence
time of groundwater reaching each well screen was calcu-
lated as the arithmetic mean of particle travel times
(Table 1). Assuming that CFC atmospheric concentrations
are the same as vadose zone air concentrations, and that
water recharged at the water table is in equilibrium with this
Figure 5. Sensitivity results of simulated CFC-based age estimates for 10 realizations using a
T
= 0.01,
0.04, and 0.1 m for (a) CFC-11 ages at well B1-1, (b) CFC-12 ages at well B1-1, (c) CFC-11 ages at well
B4-2, and (d) CFC-12 ages at well B4-2.
WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE 16 -7
concentration, this residence time should be equivalent to
the CFC-based groundwater age.
[28] In several groundwater residence time simulations, a
significant number of particles left the simulation through
the upgradient boundary, never reaching the water table.
Therefore we could not directly determine the residence
time of those particles. Instead, we estimated their residence
times by adding the estimated residence time based on a
relationship between particle age and depth in the system. A
depth versus residence time relationship was determined by
averaging simulated groundwater residence time at several
depths along the downgradient boundary (Figure 7). This
depth versus residence time relationship carries much
uncertainty for several reasons, including (1) upgradient
geology may be significantly different since this will
include the apex region of the alluvial fan and thus be
dominated by a greater percentage of coarse-grained mate-
rial; and (2) heterogeneity within a system, as shown by this
research, causes significant variance around the depth
versus residence time relationship at any depth. However,
even with these potential sources of error for these particles,
the overall influence of these particles is minimal for most
wells and simulations (Figure 6) since these particles
represent water with residence times >55 years and would
thus contain no CFC. This assumption allows a means of
estimating the full distribution of groundwater residence
Figure 6. Simulated groundwater residence time distributions from the midpoint of selected well
screens for each of the 10 realizations: (a) B1-1, (b) B4-1, (c) B4-2, (d) B5-1, and (e) B5-2. Dashed
portions of the residence time distribution curves indicate that >10% of the particles required estimation
of groundwater residence times using the depth versus residence time relationship of Figure 7. The thick
dashed line shows simulated residence times assuming a homogeneous aquifer system. See color version
of this figure at back of this issue.
16 -8 WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE
time while maintaining a simulation space small enough to
conduct the numerical experiments within a reasonable
amount of time.
[29] Simulated CFC age dates were obtained by finding
in Figure 2 the reported atmospheric CFC concentration
values (C
CFC,p
) corresponding to residence time of each
particle. A total CFC concentration was estimated at each
well using
CCFC ¼P
Np
CCFC;p
Np
ð8Þ
where N
p
is the number of particles. The resulting
‘‘simulated CFC-age’’ of groundwater was estimated by
determining the year in Figure 2 that corresponds to this
calculated CFC concentration at the well. In this applica-
tion, we assume that the temperature at recharge depth
remains constant over the length of time simulated in order
to honor Henry’s Law.
5. Results
5.1. General Observations
[30] Simulations of groundwater residence time distribu-
tions show that groundwater tapped by each well consists of
a wide mix of water with various residence times, even
though the particles were released from points in space rather
than from the entire 1.5-m-long well screens (Table 1 and
Figure 6). A high degree of residence time distribution
variability between the 10 different realizations was
observed (Figure 6) and is also reflected in standard devia-
tion of residence times computed for the various scenarios
(Table 1). This variability reflects the uncertainty inherent to
our conditional simulations of heterogeneity in the system.
Even with this uncertainty, most curves associated with a
well retain similar shape and breadth of age distribution.
[31] With the exception of Well B1-1 and Well B5-1, the
broad age distributions and their asymmetry indicate that
non-Fickian dispersion caused by facies-scale heterogeneity
is significant. In the case of Wells B1-1, most simulated age
results agree fairly closely with the normal age distribution
produced under homogeneous conditions (Figure 6a). In the
Well B1-1 simulations that display this normal distribution,
the particles traveled primarily through one hydrofacies type
with minor excursions into other hydrofacies. Simulations
that produced a broader residence time distribution (e.g.,
realizations 4, 6, and 10) were the result of particles
traveling through several hydrofacies between the well
screen and water table. Similar responses were observed
in simulations conducted from Well B5-1, where all but four
of the simulations showed Fickian response related to
transport through primarily one hydrofacies. Therefore the
residence time distribution shape appears to be highly
dependant on the degree of heterogeneity encountered by
the tracer.
[32] Simulated CFC-based groundwater ages favorably
match measured CFC-based groundwater ages reported by
Burow et al. [1999] (Figure 8 and Table 1), with some
anomalies that are explainable. Simulated CFC-based
groundwater ages are generally slightly older than reported
CFC ages, especially for ages from CFC-12. This may stem
from the lack of regional pumping in our simulations that
would potentially draw a greater proportion of younger
water to the deeper screened intervals in the model.
5.2. Anomalies
[33] Only simulation results for wells B2-2, B2.5-2, B3-4,
and B4-4 show noticeably greater CFC ages than those
reported by Burow et al. [1999]. Sample measurement error,
possibly caused by CFC contamination of the sample (thus
producing an artificially young age result since concentra-
tions would be increased), may explain some of these
discrepancies. For example, the measured groundwater
age from CFC-12 for Well B4-4 is 29.5 yrs, which is
inconsistent with the lack of detectable tritium [Burow et
al., 1999] and the depth of this well. Furthermore, 1,2-
Dibromo-3-Chloropropane (DBCP) and nitrate concentra-
tions, contaminants determined to be associated with
younger groundwater [Burow et al., 1999], along with
specific conductance in groundwater sampled from this well
suggest that the water from this well is older than 29.5 yrs
[Burow et al., 1999]. Finally, each of the 10 simulated ages
from the 10 geostatistical realizations for Well B4-4 is older
than measured value (Table 1). The discrepancy between
measured and modeled age may also be related to a problem
with the geostatistical representation of heterogeneity at that
location since simulated vertical gradients at location B4
were consistently downward in contrast to the measured
gradients at that location. Burow et al. [1999], however,
noted that some of the samples from the study site were
probably contaminated with CFC-11 or CFC-12.
[34] Discrepancies between our simulated groundwater
ages and observed CFC ages were also noted in the CFC-11
age dates for well B3-4. The simulated CFC-11 ground-
water age of 38.1 years is greater than the observed 29 years
for CFC-11 in this well. However, the simulated CFC-12
Figure 7. Estimated residence time versus depth relation-
ship for the study area produced by estimating groundwater
residence times for cells along the downgradient boundary
of the study area.
WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE 16 -9
age of 40.5 favorably matches the observed 41-year CFC-12
age date. Additionally, the age trend between Wells B3-1,
B3-2, B3-3, B3-4, and B3-5 show the expected tendency of
increasing age with depth at the B3 location, and the
measured B3-4 age date appears to be too young to fit this
trend. Burow et al. [1999] indicate that CFC-11 is more
susceptible to microbial transformation, however this was
not thought to be a dominant factor affecting CFC-11
concentrations due to low organic content and well-oxy-
genated groundwater. Additionally, CFC-12 is less likely
than CFC-11 to be contaminated by sampling [Bartolino,
1997]. Burow et al. [1999] assigned a recharge date to this
older value from CFC-12 data. Our results suggest this older
age date is more appropriate.
[35] Finally, the measured CFC-11 age of water in Well
B2-2 (8 years) is less than that observed in Well B2-1 (16
years), indicating that the deeper aquifer holds younger
water at the B2 location. This same phenomenon was
reported for measured ages at Well B2.5-1 and Well B2.5-
2. None of the simulations within the 10 realizations
reflected this reversed groundwater age, thus simulated
CFC-11 and CFC-12 ages are greater than the measured
values. These age discrepancies may be either due to
contamination of the CFC samples, thus producing a
younger age date, or presence of short-circuit flow paths
close to wells B2 and B2.5 in the field but not in the model.
5.3. Simulated CFC Age and the Mean Age
[36] In order to assess whether CFC ages reflect the mean
groundwater age of a sample in a heterogeneous aquifer, we
compared the simulated CFC age results to the simulated
overall mean groundwater residence time at a well
(Figure 9). Simulated CFC-based age dates are consistently
lower than the actual mean residence time of the simulated
samples. This is because the CFC age dating cannot differ-
entiate ages of water older than 55 years. Therefore it
appears that CFC-based ages do not typically reflect the
mean age of the groundwater in a typically heterogeneous
alluvial system. This result supports similar observations by
Varni and Carrera [1998], Tompson et al. [1999], and
Bethke and Johnson [2002].
6. Discussion and Conclusions
6.1. Residence Time Dispersion and Mean Age
[37] The simulated residence time distributions (Figure 6
and Table 1) exhibit substantial variance and positive skew-
ness (tailing), with some water ‘‘samples’’ containing water
ages ranging from 5 10 yrs to >100 yrs. The resulting CFC
age dates represent biased sampling of these distributions,
thereby underestimating actual mean age. The bias increases
with mean age (Figure 9), owing to the growth in residence
time variance with mean age.
[38] The substantial variance of residence times is caused
primarily by spatially variable groundwater velocities that
are inherent to the three-dimensional, heterogeneous allu-
vial system (Figure 3). Mixing of residence times along the
sampling interval is not a significant factor here because the
well screens are short (1.5 m). The tailing behavior is
caused by presence of very low-K units (paleosols and
muds) in which relatively slow advection and diffusion
dominate transport [LaBolle and Fogg, 2001].
[39] The results clearly illustrate how representation of
groundwater age by a single date can be misleading. Only
when the flow paths are relatively short, as in the case of
wells B1-1 and B5-1, does the groundwater residence
time distribution approximate a normal distribution with
small variance for some realizations of the heterogeneity
Figure 8. Mean simulated CFC-based groundwater ages plotted against measured CFC-based
groundwater ages reported by Burow et al. [1999] for (a) CFC-11 and (b) CFC-12. Ranges around the
mean show the range of simulated CFC-based ages within the 10 realizations.
16 -10 WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE
(Figure 6), and, thus only in these cases will simulated CFC
ages reasonably correspond to the mean groundwater ages.
[40] Even if the age date technique is capable of unbiased
sampling, interpretation of the field measured ages can be
confounded by the broad range of ages within the sample.
For example, the notion that an age date of, say, 40 years is
always indicative of less vulnerability to recent, anthropo-
genic contamination than an age of 20 years is flawed. If the
source of contamination began 30 years ago, depending on
the groundwater residence time distributions, the 40-year
water could contain as much or more water that is younger
than 30 years as the 20-year water. Furthermore, if the age
of a water is estimated to be very old (e.g., thousands of
years), the tailing phenomenon could have the effect of
including very young water in the same sample. Such
extreme tailing of residence times is consistent with Bethke
and Johnson’s [2002] analysis, which indicates that diffu-
sion induced tailing can cause mixing of young and very old
waters locally within an aquifer. Even without diffusion-
induced tailing, it is expected that regional groundwater
flow systems would commonly include both old water that
has traveled great distances and younger water that enters
the regional aquifer system via leakage or preferential,
vertical flow through discontinuities in confining beds.
Groundwater pumpage would accelerate the downward
flow of relatively young water and enhance the mixing of
disparate groundwater ages. Future research should be
directed at assessing potential age dispersion and age
sampling bias for a variety of environmental tracers and
aquifer systems.
[41] The hypothesis that CFC-based age dates represent
an average age appears to be with little merit for all but the
most simple flow systems involving relatively young
(<55 years) groundwater. Several writers have indicated
that this may be the case [e.g., Walker and Cook, 1991;
Mazor and Nativ, 1992; Goode, 1996; Varni and Carrera,
1998; Fogg et al., 1999; Tompson et al., 1999; Bethke and
Johnson, 2002]. The present study models the full residence
time distributions, thereby quantifying and elucidating the
potential errors and ambiguities for a highly characterized,
heterogeneous system.
[42] The bias problem in age dating techniques that
preferentially detect the relatively young waters (e.g., CFC
and tritium-based methods) could be evaluated more
directly with an age dating technique that works well for
groundwater ages in the intermediate range (e.g.,<1000 yr).
While techniques exist for dating much older waters (e.g.,
14
C,
36
Cl), no techniques are available for dating the
intermediate range. Clearly, intermediate-age dating techni-
ques are needed.
6.2. Implications for Groundwater Quality
Sustainability
[43]Fogg et al. [1999] simulated dispersion of ground-
water residence times in a regional-scale (45 km) analysis
of a portion of the Salinas Valley. Similarly, their results
indicated significant dispersion of groundwater residence
time within simulated water ‘‘samples,’’ with residence
times in individual ‘‘samples’’ often ranging from 10 to
>500 yrs. Their model indicated such mixing over most of
the Salinas Valley study area, which receives nonpoint
source contaminant (nitrate) loading primarily from fertil-
izer sources. Fogg et al. [1999] pointed out a weighty
implication of such extreme mixing - that wells exhibiting
contamination today in alluvial aquifers similar to the
Salinas Valley (e.g., San Joaquin Valley, basin and range
aquifers, Gulf Coast aquifers, etc.) commonly derive only a
fraction of their production from the waters that are young
Figure 9. Simulated mean groundwater residence times plotted against mean simulated CFC-based
groundwater ages from the 10 realizations for (a) CFC-11 and (b) CFC-12. The error bars show the
ranges of simulated groundwater residence times and CFC ages for each of the 10 realizations.
WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE 16 -11
enough to be contaminated. Thus, if the sources of recalci-
trant contaminants (nitrates and salinity in the Salinas Valley
case) have not diminished appreciably since their introduc-
tion starting in the 1940s and 1950s, one can expect steadily
declining water quality many decades into the future, even
if the sources are reduced or eliminated today. In other
words, if the groundwater contamination that we see today
in the Salinas Valley is due only to contamination of that
fraction of the water pumped by wells that has been exposed
to anthropogenic contaminants, this fraction will increase
for many decades or centuries into the future, leading to
gradually rising concentrations at the wellhead. Because of
the slow response time of basin-scale groundwater quality
in deep, alluvial systems like the Salinas and San Joaquin
Valleys, this degradation would be virtually irreversible on
the scale of decades to centuries.
[44] Such a scenario has played out with respect to
pesticide contamination in the Kings River fan aquifer
[Burow et al., 1999], where use of the chemical DBCP
circa 1950 1972 has left a legacy of contamination in wells
up through the present. DBCP concentrations in this system
appear to have been declining in recent years owing to
chemical transformation and/or recycling of groundwater
through pumpage and reapplication for irrigation. The
average groundwater residence times in this comparatively
thin, coarse-grained system are young relative to many other
portions of the San Joaquin Valley and many other alluvial
basins. Thus the contamination and subsequent recovery of
groundwater quality in the Kings River fan aquifer system
within 30 to >50 yrs can be considered relatively fast. It is
likely that sources of contamination in finer-grained aquifer
systems (e.g., west side of the San Joaquin Valley) where
the time constants for transport are much greater have not
yet had full impact on deeper aquifers.
[45] The present study was conducted in part to inves-
tigate the age dispersion processes in greater detail, with
much more detailed hydrostratigraphic characterization and
a more carefully calibrated model, including some valida-
tion with measured CFC-based ages. Much like the results
of Fogg et al. [1999], we find significant dispersion of ages
within individual water ‘‘samples,’’ albeit for less regional
length and timescales. Thus the work of this study reinfor-
ces the conclusions of Fogg et al. [1999] regarding ground-
water sustainability in the presence of nonpoint source,
recalcitrant contaminants.
[46] Such trends in groundwater quality can only be
detected through very long-term monitoring, which is
typically rare. Although groundwater hydrologists and
managers intuitively know that groundwater quality
changes can be very slow, there is a strong tendency to
stop monitoring water quality when certain chemical con-
stituents show little variation for several months or years.
Although data on long-term water quality trends are rare,
some of the available data show an alarming consistency
with the above hypothesis regarding long-term degradation
of groundwater quality. For example, Dubrovsky et al.
[1998, p. 17] show average groundwater nitrate concen-
trations rising steadily from the 1950s through the 1980s
and 1990s in the eastern San Joaquin Valley. Such data
together with our improved understanding of groundwater
residence time dispersion suggest that in many basins we
may presently be viewing only the ‘‘early breakthrough’’ of
this centuries long ‘‘tracer test.’’ Continued long-term water
quality monitoring and process investigation are therefore
paramount.
[47]Acknowledgments. This research was supported by the Coop-
erative State Research Service, U.S. Department of Agriculture, under
agreement 93-38420-8792, the National Science Foundation (EAR-98-
70342), University of California Water Resources Center (W-901), Occi-
dental Chemical Company, the U.S. EPA (R819658) Center for Ecological
Health Research at U.C. Davis, NIEHS Superfund Grand (ES-04699),
University of California Toxic Substances Teaching and Research Program,
and Lawrence Livermore National Laboratory. Although the information in
this document has been funded in part by the United States Environmental
Protection Agency, it may not necessarily reflect the views of the Agency
and no official endorsement should be inferred. We benefited greatly from
discussions with Karen Burow, Steve Carle, and Thomas Harter and from
comments on the manuscript from Karen Burow. Comments by two
anonymous reviewers and the associate editor also led to significant
improvements to the manuscript.
References
Arnold, L., Stochastic Differential Equations: Theory and Applications,
228 pp., John Wiley, New York, 1974.
Bartolino, J. R., Chlorofluorocarbon and tritium age determination of
ground-water recharge in the Ryan Flat subbasin, Trans-Pecos, Texas,
U.S. Geol. Surv. Water Res. Invest.,96-4245, 29 pp., 1997.
Bethke, C. M., and T. M. Johnson, Paradox of groundwater age, Geology,
30(2), 107 – 110, 2002.
Burow, K. R., G. S. Weissmann, R. D. Miller, and G. Placzek, Hydrogeo-
logic facies characterization of an alluvial fan near Fresno, California,
using geophysical techniques, U.S. Geol. Surv. Open File Rep.,97– 46,
15 pp., 1997.
Burow, K. R., S. Y. Panshin, N. M. Dubrovsky, D. VanBrocklin, and G. E.
Fogg, Evaluation of processes affecting 1,2-dibromo-3-chloropropane
(DBCP) concentrations in ground water in the eastern San Joaquin
Valley, California: Analysis of chemical data and ground-water flow
and transport simulations, U.S. Geol. Surv. Water Resour. Invest.,
99 – 4059, 1999.
Busenberg, E., and L. N. Plummer, Use of chlorofluorocarbons (CCl
3
F and
CCl
2
F
2
) as hydrologic tracers and age-dating tools: The alluvium and
terrace system of central Oklahoma, Water Resour. Res.,28, 2257 – 2283,
1992.
Busenberg, E., E. P. Weeks, L. N. Plummer, and R. C. Bartholomay, Age
dating ground water by use of chlorofluorocarbons (CCl
3
F and CCl
2
F
2
),
and distribution of chlorofluorocarbons in the unsaturated zone, Snake
River Plain Aquifer, Idaho National Engineering Laboratory, Idaho, U.S.
Geol. Surv. Water Resour. Invest. Rep.,93– 4054, 47 pp., 1993.
Carle, S. F., and G. E. Fogg, Transition probability-based indicator geos-
tatistics, Math. Geol.,28, 453 – 476, 1996.
Carle, S. F., E. M. LaBolle, G. S. Weissmann, D. Van Brocklin, and G. E.
Fogg, Conditional simulation of hydrofacies architecture: A transition
probability/Markov approach, in Hydrogeologic Models of Sedimentary
Aquifers, Concepts in Hydrogeol. and Environ. Geol., vol. 1, edited by
G. S. Fraser and J. M. Davis, Spec. Publ. SEPM Soc. Sediment. Geol.,
147 – 170, 1998.
Carman, P. C., Flow of Gases Through Porous Media, 182 pp., Academic,
San Diego, Calif., 1956.
Cook, P. G., D. K. Solomon, L. N. Plummer, E. Busenberg, and S. L.
Schiff, Chlorofluorocarbons as tracers of groundwater transport pro-
cesses in a shallow, silty sand aquifer, Water Resour. Res.,31, 425–
434, 1995.
Domenico, P. A., and F. W. Schwartz, Physical and Chemical Hydrogeol-
ogy, 2nd ed., John Wiley, New York, 1998.
Dubrovsky, N. M., C. R. Kratzer, L. R. Brown, J. M. Gronberg, and K. R.
Burow, Water quality in the San Joaquin-Tulare Basins, California,
1992 – 95, U.S. Geol. Surv. Circ.,1159, 38 pp., 1998.
Dunkle, S. A., L. N. Plummer, E. Busenberg, P. J. Phillips, J. M. Denver, P.
A. Hamilton, R. L. Michel, and T. B. Coplen, Chlorofluorocarbons
(CCl
3
F and CCl
2
F
2
) as dating tools and hydrologic tracers in shallow
groundwater of the Delmarva Peninsula, Atlantic Coastal Plain, United
States, Water Resour. Res.,29, 3837 – 3860, 1993.
Elkins, J. W., T. M. Thompson, T. H. Swanson, J. H. Butler, B. D. Hall, S.
O. Cummings, D. A. Fisher, and A. G. Raffo, Decrease in the growth
16 -12 WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE
rates of atmospheric chlorofluorocarbons 11 and 12, Nature,364, 780 –
783, 1993.
Elkins, J. W., T. M. Thompson, J. H. Butler, and R. C. Myers, Atmospheric
Measurements of CFC-11 and CFC-12 from the NOAA/CMDL Flask
Program, Natl. Oceanic and Atmos. Admin, Silver Spring, Md., 1999.
(Available at ftp://ftp.cmdl.noaa.gov/noah/cfcs/)
Farrell, D. A., A. D. Woodbury, and E. A. Sudicky, The 1978 Borden tracer
experiment: Analysis of the spatial moments, Water Resour. Res.,30,
3213 – 3223, 1994.
Fetter, C. W., Applied Hydrogeology, 4th ed., 598 pp., Macmillan, New
York, 1999.
Fogg, G. E., E. M. LaBolle, and G. S. Weissmann, Groundwater vulner-
ability assessment: hydrogeologic perspective and example from Salinas
Valley, California, in Assessment of Non-Point Source Pollution in the
Vadose Zone,Geophys. Monogr. Ser., vol. 108, edited by D. L. Corwin,
K. Loague, and T. R. Ellsworth, pp. 45– 61, AGU, Washington, D. C.,
1999.
Fresno Irrigation District, Ground-water Elevation Map, January, 1993,
scale 1 inch = 2.5 miles, Fresno, Calif., 1993.
Freyburg, D. L., A natural gradient experiment on solute transport in a sand
aquifer, 2, Spatial moments and the advection and dispersion of nonreac-
tive tracers, Water Resour. Res.,22(13), 2031 – 2046, 1986.
Garabedian, S. P., D. R. LeBlanc, L. W. Gelhar, and M. A. Celia, Large-
scale natural gradient tracer test in sand and gravel. Cape Cod, Massa-
chusetts, 2, Analysis of spatial moments for a nonreactive tracer, Water
Resour. Res.,27, 911– 924, 1991.
Goode, D. J., Direct simulation of groundwater age, Water Resour. Res.,32,
289 – 296, 1996.
Grathwohl, P., Diffusion in Natural Porous Media: Contaminant Transport,
Sorption/Desorption and Dissolution Kinetics, 207 pp., Kluwer Acad.,
Norwell, Mass., 1998.
Harbaugh, A. W., and M. G. McDonald, User’s documentation for MOD-
FLOW-96, an update to the U.S. Geological Survey Modular Finite_Dif-
ference Ground-Water Flow Model, U.S. Geol. Surv. Open File Rep.,
96 – 485, 1996.
Hackley, K. C., C. L. Liu, and D. D. Coleman, C-14 dating of groundwater
containing microbial CH
4
,Radiocarbon,34(3), 686 – 695, 1992.
Hinkle, S. R., and D. T. Snyder, Comparison of Chlorofluorocarbon age
dating with particle tracking results of a regional ground-water flow
model of the Portland Basin, Oregon and Washington, U.S. Geol. Surv.
Water Supply Pap.,2483, 1997.
Johnston, C. T., P. G. Cook, S. K. Frape, L. N. Plummer, E. Busenberg, and
R. J. Blackport, Ground water age and nitrate distribution within a glacial
aquifer beneath a thick unsaturated zone, Ground Water,36, 171– 180,
1998.
LaBolle, E. M., Theory and simulation of diffusion processes in porous
media, Ph.D. dissertation, 202 pp., Univ. of Calif., Davis, 1999.
LaBolle, E. M., and G. E. Fogg, Role of molecular diffusion in contaminant
migration and recovery in an alluvial aquifer system, Transp. Porous
Media,42, 155 – 179, 2001.
LaBolle, E. M., G. E. Fogg, and A. F. B. Tompson, Random-walk simula-
tion of transport in heterogeneous porous media: Local mass-conserva-
tion problem and implementation methods, Water Resour. Res.,32, 583 –
593, 1996.
LaBolle, E. M., J. Quastel, and G. E. Fogg, Diffusion theory for transport in
porous media: Transition probability densities of diffusion processes
corresponding to advection-dispersion equations, Water Resour. Res.,
34, 1685 – 1693, 1998.
LaBolle, E. M., J. Quastel, G. E. Fogg, and J. Gravner, Diffusion processes
in composite porous media and their numerical integration by random
walks: Generalized stochastic differential equations with discontinuous
coefficients, Water Reour. Res.,36(3), 651 – 662, 2000.
Mallants, E., A. Espino, M. Van Hoorick, J. Feyen, N. Vandenberghe, and
W. Loy, Dispersivity estimates from a tracer experiment in a sandy
aquifer, Ground Water,38(2), 304 – 310, 2000.
Mazor, E., and R. Nativ, Hydraulic calculation of groundwater flow velo-
city and age examination of the basic premises, J. Hydrol.,138,211
222, 1992.
McDonald, M. G., and A. W. Harbaugh, A modular three-dimensional
finite difference ground-water flow model, U.S. Geol. Surv. Tech. Water
Resour. Invest.,book 6(chap. A1), 1988.
Neupauer, R. M., and J. L. Wilson, Adjoint method for obtaining back-
ward-in-time location and travel time probabilities of a conservative
groundwater contaminant, Water Resour. Res.,35(11), 3389 – 3398,
1999.
Plummer, L. N., and C. L. Sprinkle, Radiocarbon dating of dissolved in-
organic carbon in groundwater from confined parts of the upper Floridan
aquifer, Florida, USA, Hydrogeol. J.,9(2), 127 – 150, 2001.
Plummer, L. N., R. L. Michel, E. M. Thurman, and P. D. Glynn, Environ-
mental tracers for age dating young ground water, in Regional Ground-
Water Quality, edited by W. M. Alley, pp. 255 –294, Van Nostrand Re-
inhold, New York, 1993.
Reilly, T. E., L. N. Plummer, P. J. Phillips, and E. Busenberg, The use of
simulation and multiple environmental tracers to quantify groundwater
flow in a shallow aquifer, Water Resour. Res.,30, 421 – 433, 1994.
Solomon, D. K., and E. A. Sudicky, Tritium and helium-3 isotope ratios for
direct estimation of spatial variations in groundwater recharge, Water
Resour. Res.,27, 2309 – 2319, 1991.
Solomon, D. K., R. J. Poreda, S. L. Schiff, and J. A. Cherry, Tritium and
helium-3 as groundwater age tracers in the Borden Aquifer, Water Re-
sour. Res.,28, 741 – 755, 1992.
Sudicky, E. A., A natural gradient experiment on solute transport in a sand
aquifer: Spatial variability of hydraulic conductivity and its role in the
dispersion process, Water Resour. Res.,22(13), 2069 – 2082, 1986.
Tompson, A. F. B., S. F. Carle, N. D. Rosenberg, and R. M. Maxwell,
Analysis of groundwater migration from artificial recharge in a large
urban aquifer: A simulation perspective, Wa ter R e s o ur. Re s . ,35(10),
2981 – 2998, 1999.
Uffink, G. J. M., Application of Kolmogorov’s backward equation in ran-
dom walk simulations of groundwater contaminant transport, in Con-
taminant Transport in Groundwater: Proceedings of the International
Symposium on Contaminant Transport in Groundwater, edited by H.
E. Kobus and W. Kinzelbach, pp. 283298, A. A. Balkema, Brookfield,
Vt., 1989.
Varni, M., and J. Carrera, Simulation of groundwater age distributions,
Water Resour. Res.,34, 3271 – 3281, 1998.
Walker, G. R., and P. G. Cook, The importance of considering diffusion
when using carbon-14 to estimate groundwater recharge to an unconfined
aquifer, J. Hydrol.,128, 41 – 48, 1991.
Warner, M. J., and R. F. Weiss, Solubilities of chlorofluorocarbons 11 and
12 in water and seawater, Deep Sea Res.,32, 1485 – 1497, 1985.
Weissmann, G. S., Toward new models of subsurface heterogeneity: An
alluvial fan sequence stratigraphic framework with transition probability
geostatistics, Ph.D. dissertation, Univ. of Calif., Davis, 1999.
Weissmann, G. S., and G. E. Fogg, Multi-scale alluvial fan heterogeneity
modeled with transition probability geostatistics in a sequence strati-
graphic framework, J. Hydrol.,226, 48 – 65, 1999.
Weissmann, G. S., S. F. Carle, and G. E. Fogg, Three-dimensional hydro-
facies modeling based on soil surveys and transition probability geosta-
tistics, Water Resour. Res.,35, 1761 – 1770, 1999.
Weissmann, G. S., J. F. Mount, and G. E. Fogg, Glacially-driven cycles in
accumulation and sequence stratigraphy of a stream-dominated alluvial
fan, San Joaquin Valley, California, J. Sediment. Res.,72(2), 240 – 251,
2002.
G. Fogg, E. M. LaBolle, and Y. Zhang, Hydrologic Sciences, University
of California, One Shields Avenue, Davis, CA 95616, USA.
G. Weissmann, Department of Geological Sciences, Michigan State
University, 206 Natural Science Building, East Lansing, MI 48824 – 1115,
USA. (weissman@msu.edu)
WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE 16 -13
Figure 6. Simulated groundwater residence time distributions from the midpoint of selected well
screens for each of the 10 realizations: (a) B1-1, (b) B4-1, (c) B4-2, (d) B5-1, and (e) B5-2. Dashed
portions of the residence time distribution curves indicate that >10% of the particles required estimation
of groundwater residence times using the depth versus residence time relationship of Figure 7. The thick
dashed line shows simulated residence times assuming a homogeneous aquifer system.
WEISSMANN ET AL.: DISPERSION OF GROUNDWATER AGE
16 -8
... For instance, a common application of BTTP is to recover contamination history and identify responsible parties, where the BTTP's peak captures the most likely release time of contaminants from the source (Skaggs and Kabala, 1994;Woodbury and Ulrych, 1996;Woodbury 35 et al., 1998;Sun et al., 2006aSun et al., , 2006bJha and Datta, 2015;Yeh et al., 2015;Jamshidi et al., 2020;Chen et al., 2023). BTTP can also be used to date groundwater since BTTP characterizes the age distribution of groundwater due to borehole mixture and/or hydrodynamic dispersion in regional-scale aquifers (Weissmann et al., 2002;Cornaton and Perrochet, 2006;LaBolle et al., 2006;Konikow, 2007a, 2007b;Janssen et al., 2008;McMahon et al., 2008;Maxwell et al., 2016;Ponprasit et al., 2022;Mao et al., 2023). In addition, BTTP provides a more comprehensive method to assess aquifer vulnerability than 40 classical statistics-based approaches through the generation of three-dimensional (3-d), transient vulnerability maps for groundwater to non-point source contamination (Fogg et al., 1999;Zhang et al., 2018). ...
... represents the groundwater age distribution and provides core information for groundwater sustainability assessment (Fogg et al., 1999;Weissmann et al., 2002;Fogg and LaBolle, 2006). ...
... Figures 6a~6d show the calculated BTTP for the USGS wells sampled by Burow et al. (1999) (listed in Figure 4). Both the adjoint S-FDE (11a) and the adjoint ADE (15a) were first calibrated to fit the measured CFC-11 age of Well B41 (the modeling of CFC ages followed the methodology proposed by Weissmann et al. (2002)). Preliminary tests showed that the simulated CCF-11 age is insensitive to the two truncation parameters, since the subordination truncation parameter (or the temporal truncation 410 parameter ) mainly affects the very early time, i.e., < 1 day (or very late time, i.e., > 50 yrs) in the BTTP. ...
Preprint
Full-text available
Backward probabilities such as backward travel time probability density function for pollutants in natural aquifers/rivers had been used by hydrologists for decades in water-quality related applications. Reliable calculation of backward probabilities, however, has been challenged by non-Fickian pollutant transport dynamics and variability in the resolution of velocity at study sites. To address these two issues, we built an adjoint model by deriving a backward-in-time fractional-derivative transport equation subordinated to regional flow, developed a Lagrangian solver, and applied the model/solver to backtrack pollutant transport in various flow systems. The adjoint model applies subordination to a reversed regional flow field, converts forward-in-time boundaries to either absorbing or reflective boundaries, and reverses the tempered stable density to define backward mechanical dispersion. The corresponding Lagrangian solver is computationally efficient in projecting backward super-diffusive mechanical dispersion along streamlines. Field applications demonstrate that the adjoint subordination model can successfully recover release history, dated groundwater age, and spatial location(s) of pollutant source(s) for flow systems with either upscaled constant velocity, non-uniform divergent flow field, or fine-resolution velocities in a non-stationary, regional-scale aquifer, where non-Fickian transport significantly affects pollutant dynamics and backward probability characteristics. Caution is needed when identifying the phase-sensitive (aqueous versus absorbed) pollutant source in natural media. Possible extensions of the adjoint subordination model are also discussed and tested for quantifying backward probabilities of pollutants in more complex media, such as discrete fracture networks.
... In this context, it is worth mentioning that, although molecular diffusion is a process that is sometimes accounted for in numerical experiments (Cremer et al., 2016), its effect on transport is often disregarded. Nevertheless, some theoret ical studies have highlighted diffusive transport as a potentially important process controlling factor of solute behavior under both unsaturated and saturated conditions (Cirpka & Kitanidis, 2002;Nissan & Berkowitz, 2019;Weissmann et al., 2002). ...
... This observation has been made at different spatial scales and in both saturated and unsaturated conditions. For instance, studies by Nissan and Berkowitz (2019) at the (saturated) pore scale, Cirpka and Kitanidis (2002) at the (unsaturated) site scale and Weissmann et al. (2002) in a regional aquifer show that high Pe values (i.e., a predominance of advection over diffusion) leads to more anomalous behavior compared to low Pe values. Inversely, transport at low Pe (i.e., diffusion-dominant) is characterized by shorter residence times in stagnant zones, which reduces the anomalous behavior of transport. ...
... Increasing σ χ renders more complex the impact of diffusion on transport: Early arrival times are mostly unchanged but the plume spreading is decreased by adding diffusion, decreasing the very pronounced tailing (i.e., elongated late arrivals) generated by the heterogeneity in the advective flux. This phenomena has been previously observed by few studies under various conditions (Cirpka & Kitanidis, 2002;Nissan & Berkowitz, 2019;Weissmann et al., 2002) and is explained by the capacity of diffusive motion to move mass away from quasi-stagnant zones, reducing in this way the potential for very late arrivals (i.e., tailing). Interestingly, the effect of small scale dispersivity, being scaled by the local advective flux, can be virtually canceled in quasi-stagnant zones, unlike velocity-independent . ...
Article
Full-text available
Physical properties of soils are ubiquitously heterogeneous. This spatial variability has a profound, yet still partially understood, impact on conservative transport. Moreover, molecular diffusion is often a disregarded process that can have an important counter‐intuitive effect on transport: diffusion can prevent non‐Fickian tailing by mobilizing mass otherwise trapped in low velocity zones. Here, we focus on macroscopically homogeneous soils presenting small scale heterogeneity, as described by the Miller‐Miller theory. We then analyze the dynamic control of soil heterogeneity, advection and diffusion on conservative transport. We focus especially on the importance of diffusion and of its tortuosity‐dependent spatial variability on the overall transport. Our results confirm previous finding that relatively high Péclet number systems are highly sensitive to the degree of heterogeneity, which promotes non‐Fickian transport. Also, diffusion appears to have a profound impact on transport, depending on both the degree of heterogeneity and the Péclet number. For a high Péclet number and a very heterogeneous system, diffusion leads to the counter‐intuitive decrease of non‐Fickian spreading described previously. This is not observed for a low Péclet number due to the impact of the spatial variability in the diffusion coefficient, which appears to be a significant controlling factor of transport by promoting or preventing the accumulation of mass in low velocity zones. Globally, this work (a) highlights the complex, synergistic effect of soil heterogeneity, advective fluxes and diffusion on transport and (b), alerts on potential upscaling challenges when the spatial variability of such key processes cannot be properly described.
... Enfin, à l'échelle même des alluvions, les propriétés hydrodynamiques peuvent s'avérer très hétérogènes dans toutes les directions de l'espace (Weissmann et al., 2002 ;Yin et al., 2023) et avoir un impact direct sur les propriétés de transport du milieu (Painter et al., 2001 ;Sarris et al., 2018). Différentes méthodes peuvent être utilisées afin de mettre en évidence ces hétérogénéités à une échelle inférieure au mètre et ainsi, mieux comprendre comment ces dernières influent sur le comportement de la nappe alluviale. ...
... Ces couches sont appelées hydrofaciès (Boumaiza, 2008) Du fait de cette hétérogénéité, les propriétés hydrodynamiques (telles que la conductivité hydraulique K) peuvent varier de plusieurs ordres de grandeur sur quelques mètres seulement (Boggs et al., 1992 ;Koltermann and Gorelick, 1996 ;Eggleston and Rojstaczer, 1998). Chaque hydrofaciès peut donc être caractérisé par une conductivité hydraulique K propre (Weissmann et al., 2002). Les hydrofaciès fortement connectés à gros grains (galets, graviers) et caractérisés par une continuité spatiale (Yin et al., 2023), permettent la formation de voies d'écoulement ...
Thesis
Full-text available
This thesis aims to study the vulnerability of alluvial hydrosystems (a river and its alluvial aquifer) in the context of climate and socio-economic change. In most cases, alluvial aquifers are shallow and productive water resources, which makes them easily exploitable. These same characteristics make them sensitive to drought and surface pollution. In this context, the case of the Allier hydrosystem (Auvergne, France) is used as an example, it is a major regional water resource for domestic water supply, irrigation, and industries. The Naussac dam (a 185 million cubic meter reservoir) has been used since 1983 to ensure a minimum river flow in summer. The approach adopted in this thesis is a multi-scale study using modeling tools to answer the questions addressed by stakeholders. Thus, a rainfall-runoff model (GARDENIA, BRGM) is used at the watershed scale, MODFLOW (Groundwater Vistas 8) is used at the alluvial aquifer scale (a well field), and Distributed Temperature Sensing (DTS) with fiber optic is used at the alluvium scale (at the Port-Douvot experimental site, Besançon, France). This thesis presents all the field investigations (geophysics, geology, hydrodynamics) and the main modeling results in three dedicated chapters. This work was funded by the Agence de l'Eau Loire-Bretagne and I-SITE CAP20-25, and is part of the AUVERWATCH Observation System (SNO H+, SNO RENOIR, IR OZCAR, eLTER, GNIP AIEA).
... These methods assume that the primary mechanism for tracer transport is through groundwater advection. However, tracers are affected by the same processes that affect solutes in aquifers (Underwood et al., 2018;Weissmann et al., 2002). This process results in the dispersion of tracers along flow paths rather than the movement of discrete parcels of uniform age (Ekwurzel et al., 1994;Underwood et al., 2018;Weissmann et al., 2002). ...
... However, tracers are affected by the same processes that affect solutes in aquifers (Underwood et al., 2018;Weissmann et al., 2002). This process results in the dispersion of tracers along flow paths rather than the movement of discrete parcels of uniform age (Ekwurzel et al., 1994;Underwood et al., 2018;Weissmann et al., 2002). The tracer concentration in the aquifer may be affected by advection, dispersion, and exchanges between the mobile and immobile parts of an aquifer, and recharge estimates should consider transport processes (Underwood et al., 2018). ...
Article
Study region: The study site, the Mount Bruce Flats, is located in the Pilbara, a semiarid region of northern Western Australia. Study focus: Semiarid climates are characterised by an extreme water deficit, with evaporation exceeding precipitation several-fold. Groundwater recharge is episodical and occurs only after infrequent but very large volume precipitation, while all other precipitation events are quickly lost to evaporation. We investigated the contribution of an internally drained basin to localised groundwater recharge by combining the results of calculations from various tracer methods. We developed a new mobile-immobile recharge transport model coupled with a probability model based on Markov Chain Monte Carlo simulations that combines several age and hydrochemical tracers to estimate localised infiltration. New hydrological insights for the region: The actual recharge varied greatly across the basin, being 4-5 times higher in the central part compared to the margins. The highest probability for recharge was calculated for 2.5 mm×y − 1 and 12.6 mm×y − 1 , with a median value of 6.8 mm×y − 1 while the median mobile effective porosity was low at ~0.5 %. The flood plains overlying freshwater aquifers may contribute several times more to localised recharge than the upper parts of the catchments, which are characterised by fast runoff and a lower water retention time. Therefore, these endorheic basins could be a valuable source of water for groundwater recharge or constitute an additional hydrological challenge for mining.
... As in previous work, we neglect the effects of subgrid-scale mechanical dispersion, assuming that the detailed discretization of the velocity field captures the bulk of dispersive transport (Henri & Harter, 2019;Labolle & Fogg, 2001;Weissmann et al., 2002). However, these previous studies used a finer vertical discretization scheme (0.4-0.5-m cell thicknesses) than the present study (1.0-4.6-m ...
Article
Full-text available
In water‐stressed regions of the world, the inundation of working landscapes to replenish aquifers—known as flood‐managed aquifer recharge (flood‐MAR)—has become a valuable tool for sustainable groundwater management. Due to their diverse land use histories, however, many potential recharge sites host nonpoint source contaminants (such as salts, pesticides, and fertilizers) within the vadose zone that may flush to groundwater during recharge operations. To identify the controls on contaminant migration, we perform stochastic simulations of flood‐MAR through a heterogeneous alluvial aquifer and apply transient particle tracking to evaluate conservative and reactive contaminant transport over 80 years of recharge operations. With semi‐annual recharge events, the water table begins to rise 0.13–1.84 years after the first inundation event while solutes take much longer (11 to >$>$80 years) to transit the 45‐m thick unsaturated zone. We derive a parametric expression for the ratio of celerity (or rate of pressure transmission) to velocity of the flood‐MAR wetting front and show that this simplified expression agrees with values calculated from heterogeneous model simulations. Slow solute velocities (0.25–1.75 m year⁻¹) allow for significant contaminant removal through denitrification, but the contaminant plume experiences minimal dispersion or dilution over this time, reaching the water table as a sharp front. Our results suggest that minimizing groundwater velocity and maximizing groundwater celerity during flood‐MAR should optimize increases in water supply while limiting water quality degradation.
... The analytical uncertainty for the CFC's concentrations was estimated to be around 3% for recent water (<20 years) and 5% for old water (between 55 and 30 years). However, it is important to note that the groundwater age obtained by CFC's-SF 6 analyses may differ from the true mean groundwater age in the case of mixing processes or matrix diffusion (Katz, 2004;Weissmann et al., 2002). Therefore, the term "apparent age" is used throughout the paper. ...
Article
Full-text available
Backward probabilities, such as the backward travel time probability density function for pollutants in natural aquifers/rivers, have been used by hydrologists for decades in water quality applications. Calculating these backward probabilities, however, is challenging due to non-Fickian pollutant transport dynamics and velocity resolution variability at study sites. To address these issues, we built an adjoint model by deriving a backward-in-time fractional-derivative transport equation subordinated to regional flow, developed a Lagrangian solver, and applied the model/solver to trace pollutant transport in diverse flow systems. The adjoint model subordinates to a reversed regional flow field, transforms forward-in-time boundaries into either absorbing or reflective boundaries, and reverses the tempered stable density to define backward mechanical dispersion. The corresponding Lagrangian solver efficiently projects backward super-diffusive mechanical dispersion along streamlines. Field applications demonstrate the adjoint subordination model's success with respect to recovering release history, groundwater age, and pollutant source locations for various flow systems. These include systems with upscaled constant velocity, nonuniform divergent flow fields, or fine-resolution velocities in a nonstationary, regional-scale aquifer, where non-Fickian transport significantly affects pollutant dynamics and backward probabilities. Caution is needed when identifying the phase-sensitive (aqueous vs. absorbed) pollutant source in natural media. The study also explores possible extensions of the adjoint subordination model for quantifying backward probabilities of pollutants in more complex media, such as discrete fracture networks.
Article
Full-text available
Groundwater age or residence time is important for identifying flow and contaminant pathways through groundwater systems. Typically, groundwater age and age distributions are inferred via lumped parameter models based on measured age tracer concentrations. However, due to cost and time constraints, age tracers are usually only sampled at a small percentage of the wells in a catchment. This paper describes and compares two methods to increase the number of groundwater age data points and assist with validating age distributions inferred from lumped parameter models. Two machine learning techniques with different strengths were applied to develop two independent metamodels that each aim to establish relationships between the hydrochemical parameters and the modelled groundwater age distributions in one test catchment. Ensemble medians from the best model realisations per age distribution percentile were used for comparison with the results from traditional lumped parameter models based on age tracers. Results show that both metamodelling techniques predict age distributions from hydrochemistry with good correspondence to traditional lumped parameter model (LPM)-derived age distributions. Therefore, these techniques can be used to assist with the interpretation of lumped parameter models where age tracers have been sampled, and they can also be applied to predict groundwater age distributions for wells in a similar hydrogeological regime that have hydrochemistry data available but no age tracer data.
Article
The occurrence of various N-related human activities increases the difficulty in distinguishing the major sources of NO3- contamination in groundwater, especially in areas with mixed land uses. In addition, the estimation of the timing and pathways of NO3- is necessary to better understand the processes of NO3- contamination in the subsurface aquifer system. This study applied environmental tracers, such as stable isotopes and age tracers (δ15N and δ18O of NO3-, δ11B, chlorofluorocarbons, and 3H), to elucidate the sources, timing, and pathways of NO3- contamination in the groundwaters of the Hanrim area, which has suffered from illegal disposal of livestock wastes since the 1980s, and also characterizes them based on mixed N-contaminant sources such as chemical fertilizers and sewage. The combined use of δ15N and δ11B overcame the limitation of using only NO3- isotopes for the identification of overlapping sources of N and successfully identified the major source of N as livestock wastes. The lumped parameter model (LPM) estimated the binary mixing of the young (age: 23-40 years, NO3-N: 2.55-15.10 mg/L) and old (age: >60 years, NO3-N: <3 mg/L) groundwaters, and explained their age mixing behaviors. The young groundwater was highly affected by livestock-derived N loading during 1987-1998, which coincides with the period of improper dumping of livestock wastes. Furthermore, the young groundwater with elevated NO3-N followed the historical NO3-N curves with younger ages (6 and 16 years) than those derived from the LPM, suggesting the possibility of faster inflows of livestock wastes through the permeable volcanic structures. This study demonstrated that a comprehensive understanding of NO3- contamination processes can be achieved using environmental tracer methods, which enables the efficient management of groundwater resources in areas with multiple N sources.
Article
3H and 35S activity in precipitation was analyzed to evaluate short and medium groundwater age. Precipitation sample was collected from Feb-2020 to May-2021 at rainy day. Tritium activity in precipitation ranged from 0.5 to 11.2 TU and 35S activity ranged from 10.1 to 97.7 mBq/L. 3H and 35S activity correlated and their contents are high in spring while Jun ~ Aug precipitation sample was low due to marine air mass introduction. The 3H and 35S activity of some karst groundwaters were analyzed.
Article
Diffusion in Natural Porous Media: Contaminant Transport, Sorption/Desorption and Dissolution Kinetics introduces the general principles of diffusion in the subsurface environment and discusses the implications for the fate and transport of contaminants in soils and groundwater. Emphasis is placed on sorption/desorption and the dissolution kinetics of organic contaminants, both of which are limited by the slow speed of molecular diffusion. Diffusion in Natural Porous Media: Contaminant Transport, Sorption/Desorption and Dissolution Kinetics compiles methods for calculating the diffusion coefficients of organic compounds (in aqueous solution or vapor phase) in natural porous media. The author uses analytical solutions of Fick's 2nd law and some simple numerical models to model diffusive transport under various initial and boundary conditions. A number of these models may be solved using spreadsheets. The book examines sorption/desorption rates of organic compounds in various soils and aquifer materials, and also examines the dissolution kinetics of nonaqueous phase liquids in aquifers, in both the trapped residual phase and in pools. Diffusion in Natural Porous Media: Contaminant Transport, Sorption/Desorption and Dissolution Kinetics concludes with a discussion of the impact of slow diffusion processes on soil and groundwater decontamination and the implications of these processes for groundwater risk assessment.
Article
Chlorofluorocarbon (CFC)-age dating was used to determine the presence or absence of modern water (water containing CFCs) in samples from 54 water wells in the Portland Basin. The accuracy of the CFC-dating method was confirmed by tritium analyses of water from six wells. CFC-dating results from 51 of the 54 wells were compared with minimum travel times of the water estimated using particle-tracking-simulation techniques. Particle tracking was accomplished using a program that combines the results of a regional, ground-water flow model of the Portland Basin with an estimated porosity field, in order to determine ground-water flow paths and to calculate particle travel times. The particle tracker can be used to delineate areas that, on the basis of short modeled travel times and known or perceived upgradient-contaminant loadings, may be affected by contaminants. A sensitivity analysis of various particle densities used in the simulations indicated that about 500 particles per model cell provides an adequate description of the minimum time of travel for water particles tracked from the 51 wells. The CFC results and the particle-tracker results were comparable at 39 (76 percent) of the 51 wells, if water that has entered the ground-water system since 1944 (the limit of detection using CFCs) is considered modern. Furthermore, although the particle-tracker-model results and the CFC-model results were not comparable at 12 sites, the particle-tracker model appeared to err on the conservative side when compared with the CFC model. If the CFC-model results are correct, the particle-tracker results are conservative at 9 of the 12 sites. At those nine sites, the particle-tracker model indicated the presence of modern water, but the CFC-model ages were old (pre-1944). At the other three sites, the particle-tracker model failed to indicate the apparent presence of modern water. On the basis of this comparison, it appears that particle-tracking techniques can be used to identify parts of the Portland Basin likely to yield modern water to wells, and that CFC-age dating can provide a useful check on the reliability of the particle-tracker results. Results of both particle-tracking simulations and CFC-age dating indicate that modern ground water flows to depths of hundreds of feet in the Portland Basin. Although wells with deep open intervals generally yield old water more often than wells with shallow open intervals, many wells with deep open intervals do produce modern water. Because many wells with deep open intervals in the Portland Basin produce modern water, aquifer depth alone is not a reliable indicator of the vulnerability of ground water to contamination. Other factors, such as contaminant loadings, and contaminant transport and fate, need to be considered when determining ground-water vulnerability.
Article
From theory on Brownian motion it is known that random walks are described by a forward and a backward equation. The forward equation is commonly used to predict the future state of a system. Applications of the backward equation are less common. It may be used to examine the history of a system. In terms of contaminant transport: the backward equation describes the probability distribution of a contaminant's initial position, given its presence at a certain time and place. This information is useful in designing preventive measures, such as protection zones around groundwater pumping stations. -Author
Article
The DIC was precipitated as barium carbonate and analyzed for both δ13C and 14C. Due to the formation of microbial CH4, some of the DIC had unusually heavy δ13C values of -3 to -1‰. the standard groundwater 14C-age correction models were developed to account for carbonate dissolution as the primary carbon input beneath the soil zone. If the heavy δ13C values observed in this study are used in readily available groundwater dating models without accounting for the effects of microbial methane formation, many of the resultant calculated ages are negative (future ages) or mathematically unsolvable. Isotopic and analytical results show a positive correlation (r2 = 0.90) between the δ13C of the DIC and the concentration of methane in the groundwater. With this correlation, we were able to correct the δ13C values of the DIC which were altered due to microbial CH4 formation. This adjustment of δ13C values, along with estimations of dead carbon input from the redox processes, allowed us to calculate 14C ages using standard groundwater age correction models. -from Authors