Analyzing Bank Filtration by Deconvoluting Time Series of Electric Conductivity by Olaf A. Cirpka1, Michael N. Fienen3, Markus Hofer2, Eduard Hoehn2, Aronne Tessarini2, Rolf Kipfer2, and Peter K. Kitanidis3 Abstract Knowing the travel-time distributions from infiltrating rivers to pumping wells is important in the management of alluvial aquifers. Commonly, travel-time distributions are determined by releasing a tracer pulse into the river and measuring the breakthrough curve in the wells. As an alternative, one may measure signals of a timevarying natural tracer in the river and in adjacent wells and infer the travel-time distributions by deconvolution. Traditionally this is done by fitting a parametric function such as the solution of the one-dimensional advectiondispersion equation to the data. By choosing a certain parameterization, it is impossible to determine features of the travel-time distribution that do not follow the general shape of the parameterization, i.e., multiple peaks. We present a method to determine travel-time distributions by nonparametric deconvolution of electric-conductivity time series. Smoothness of the inferred transfer function is achieved by a geostatistical approach, in which the transfer function is assumed as a second-order intrinsic random time variable. Nonnegativity is enforced by the method of Lagrange multipliers. We present an approach to directly compute the best nonnegative estimate and to generate sets of plausible solutions. We show how the smoothness of the transfer function can be estimated from the data. The approach is applied to electric-conductivity measurements taken at River Thur, Switzerland, and five wells in the adjacent aquifer, but the method can also be applied to other time-varying natural tracers such as temperature. At our field site, electric-conductivity fluctuations appear to be an excellent natural tracer. Introduction In Switzerland, about 40% of the drinking water is produced by active pumping from alluvial aquifers (SVGW 2002). Most pumping wells have been placed close to the rivers, so that a large fraction of the extracted water is thought to consist of freshly infiltrated river water, characterized by ground water residence times of 1Corresponding author: Swiss Federal Institute of Aquatic Science and Technology (Eawag), Department of Water Resources and Drinking Water, Uberlandstr. 133, 8600 Dübendorf, Switzerland; 41-44-823 5455; fax 141-44-823 521; Olaf. Cirpka@Eawag.CH 2Swiss Federal Institute of Aquatic Science and Technology (Eawag), Überlandstrasse 133, 8600 Dübendorf, Switzerland. 3Department of Civil and Environmental Engineering, Stanford University, Stanford, CA 94305-4020. Received June 2006, accepted November 2006. Copyright ª 2007 The Author(s) Journal compilation ª 2007 National Ground Water Association. doi: 10.1111/j.1745-6584.2006.00293.x 318 a few days. Since many rivers in the area receive treated sewage, there is the potential risk that the infiltrated water contains contaminants and may pollute the wells. In several countries, zones for the protection of wells are defined by means of travel times (e.g., 10 d in Switzerland [BUWAL 2004] and 50 d in Germany [DVGW 1995]). Within these zones, handling of potentially hazardous chemicals is prohibited and restrictions apply to agriculture and land use. Most agencies are concerned when travel times between potentially contaminated rivers and wells fall below the standards set for the delineation of protection zones. The standard technique to determine the travel-time distribution from a river to a well is to add a pulse of an easy-to-measure conservative tracer such as a fluorescent dye into the river and observe the breakthrough curve in the well (e.g., Davis et al. 1980; Lin et al. 2003; Käss 2004). From the breakthrough curve, the fraction of freshly infiltrated water in the extracted water, the time of first arrival, the mean arrival time, and the spread of the Vol. 45, No. 3—GROUND WATER—May–June 2007 (pages 318–328) travel-time distribution can be determined. The results, however, hold only for the hydraulic conditions during the test. To determine the travel-time distribution under different hydrological conditions, a repetition of the tracer test would be necessary. Also, the strong dilution in the river requires adding large amounts of the tracer. An alternative to adding a tracer is the analysis of natural tracers that are already present in the water. In order to determine travel times, natural time-varying signals are needed. Silliman and Booth (1993) used temperature fluctuations in a river and the adjacent aquifer to characterize the hydraulic exchange. Silliman et al. (1995) constructed an analytical solution of temperature in the aquifer for arbitrary temperature signals in the infiltrating river based on the one-dimensional (1D) solution of the advection-dispersion equation for steplike injection. Sheets et al. (2002) analyzed hydraulic heads, temperature, and electric conductivity at a wellfield in Cincinnati, Ohio, by cross-correlation methods. Constantz et al. (2003) compared results of a test with an injected tracer with the analysis of natural temperature fluctuations, using the 1D advective-dispersive model. Although temperature has been established as an easy-to-measure natural tracer for river–ground water interaction (Anderson 2005), it has several disadvantages in comparison to concentrations. Temperature signals are retarded by a factor depending on the porosity and the mineral composition of the medium; they are more smoothed by diffusion-like processes than concentration data; and the main thermal signals, namely, the diurnal and seasonal fluctuations, are not unique to rivers, making it difficult to distinguish river-borne ground water from water of other shallow origin (Hoehn and Cirpka 2006). In the current study, we use electric-conductivity fluctuations as natural tracer. The techniques presented apply also to time series of temperature or other conservative quantities in the water. The basic problem is to characterize the response of the measured signal in the aquifer to changes of the signal observed in the river. In contrast to the pulse-like tracer test, the measured time series in the river exhibits continuous fluctuations, so that the measured signal in the ground water well reflects both the river signal and the transport process between the river and the well. The main objective is to extract from these data the characteristics exclusively of the transport process. For this purpose, three classical techniques have been applied: cross correlation of the time series, calibration of an advective-dispersive model, and parametric deconvolution. In cross correlation, the correlation coefficient of the two time series is computed after shifting and potentially smoothing one of the two data sets (e.g., Sheets et al. 2002). The time shift with maximum correlation coefficient is considered to represent the characteristic time of transfer. If the input data had been smoothed, the optimum degree of smoothing is an indicator of diffusive processes. The cross-correlation coefficient quantifies to which extent the variability observed at one observation point is linearly related to that of the other observation point (after shifting and smoothing one of the two time series). It is useful also to compute the corresponding coefficients of linear regression, namely, the slope and the intercept. Mathematically, cross-correlating smoothed and shifted time series can be interpreted as applying a parametric transfer function to the time series after subtracting their mean values. The shape of the transfer function is defined by that of the smoothing function, and the integral of the transfer function is given by the slope of linear regression. Identifying the optimal smoothing function is a particular challenge. Analyzing time series of concentration or temperature by fitting advective-dispersive models appears to be a more physical approach than cross-correlating the data. In many applications, 1D transport with uniform and time-invariant coefficients is assumed (see the examples given in the review of Anderson [2005]). The transport model is identical to using the inverse Gaussian distribution as parametric transfer function (Kreft and Zuber 1978). The nonuniformity of real aquifers and the multidimensional nature of solute transport pose major difficulties of this model, so that even if the measurements were taken at points the breakthrough curves of tracer tests with pulse-like injection would differ from the analytical solution. Since wells are typically screened over a range of depths, the breakthrough curves are also affected by sampling an ensemble of streamlines, which may result in long tails or multiple peaks. The problem could be addressed by using multidimensional advectivedispersive models. However, the calibration of such models requires data with high spatial resolution that is missing in most practical applications. For the case of time-invariant transport, the breakthrough curve at an observation well can be computed by convolution of the river signal with an appropriate transfer function (e.g., Jury 1982). The transfer function describes the response to a unit pulse. It may be worth noting that identifying the transfer function for a given pair of input and output signals is mathematically identical to identifying the input signal for a known transfer function and output signal. The latter task has been tackled, for instance, in the reconstruction of contaminant release history (Skaggs and Kabala 1994; Snodgrass and Kitanidis 1997). As stated previously, the solution of the 1D advection-dispersion equation with uniform, constant coefficients for a pulse injection is a particular parametric transfer function, in this case defined by the two parameters velocity and dispersion coefficient. Other parametric models may be more appropriate to capture typical properties of breakthrough curves (Luo et al. 2006). Choosing the most appropriate parametric model is equivalent to identifying the optimal smoothing function in the crosscorrelation method. The major difficulty of parametric deconvolution is that it restricts the inferred transfer function to a predefined shape. As an example, the true transfer function may be multimodal because the well samples water from two distinct layers. Nonetheless, it is possible to unwittingly fit a unimodal parametric transfer function to the data so that major characteristics of the process may be lost (Fienen and Kitanidis 2006). In nonparametric deconvolution, the transfer function is free to adjust to the data O.A. Cirpka et al. GROUND WATER 45, no. 3: 318–328 319 since it does not have to conform to a preselected shape. In discrete form, this implies that the number of estimated parameters is identical to the number of time steps discretizing the transfer function. In order to achieve reasonable estimates, it typically is necessary to enforce a certain smoothness of the transfer function, which can be done by Tikhonov regularization (Skaggs and Kabala 1994) or a geostatistical model (Fienen and Kitanidis 2006). Identifying the optimal degree of smoothness from the data is challenging. Secondary peaks of the transfer function disappear when a tight smoothness criterion is used. Therefore, using a too tight criterion results in loss of information, whereas a too loose criterion may lead to erroneous peaks mapping noise of the two time series onto each other. As a final constraint, one has to ensure that the transfer function is physically reasonable. In particular, it must not contain negative entries. As discussed by Box and Jenkins (1970), the transfer function can also be expressed as the sum of autoregressive (AR) terms, where the output signal at a given time depends linearly on its own previous values, and the so-called moving-average (MA) terms, expressing the dependency on the input time series. This approach, denoted ARMA, has hardly been applied to subsurface transport (Bidwell 1995). If we used the ARMA approach rather than straightforward convolution, we would have to face the same challenges, namely, determining the degree of smoothness (which is related to the number of AR terms considered) and ensuring that the reconstructed transfer function is nonnegative. The latter would require determining the coefficients by constrained optimization (Bidwell 1995). In the present study, we present a nonparametric deconvolution method using a geostatistical smoothness criterion and Lagrange multipliers to enforce nonnegativity. We provide two schemes: in the first, we directly compute the best nonnegative estimate, whereas we generate multiple plausible solutions in the second. The developed methods are applied to time series of electric conductivity measured at a test site at River Thur, Switzerland. Theory We assume that solutes undergo linear, time-invariant transport from the river to the observation wells. Under these conditions, the superposition principle holds, and the response y(t) of concentration in an observation well to any signal x(t) in the river can be computed by convolution (e.g., Jury 1982): Z T yðtÞ ¼ gðsÞxðt 2 sÞ ds ð1Þ 0 in which g(s) is the transfer function, also denoted as the impulse-response function or Green’s function, and T is the time interval over which it is defined. The transfer function describes the response of the system to a pulselike stimulus, that is, the breakthrough curve that would result from an ideal tracer test in which the signal x(t) in the river is a Dirac pulse. The transfer function can also be interpreted as the probability density function of solute 320 O.A. Cirpka et al. GROUND WATER 45, no. 3: 318–328 travel times. According to Equation 1, the output function y at time t depends on the input x from t 2 T to t. Thus, if the input function x(t) varies continuously in time and measurements are available only for t 0, the output function y(t) is properly defined only for t T. After uniform temporal discretization, we deal with an nx 3 1 vector x and an ny 3 1 vector y of measurements in the river and observation well, respectively, and an ng 3 1 vector g of discrete values of the transfer function. The vector g of transfer function values must be considerably shorter than the vector y of observations, i.e., ng < ny . Now, the convolution integral, Equation 1, becomes a matrix-vector product: y ¼ Xg ð2Þ Xij ¼ txi2j11 ð3Þ with in which the ny 3 ng matrix X, which has Toeplitz structure, is made of the shifted input signal x times the time increment t. The purpose of the present paper is to infer the transfer function g from measured time series x and y, in which both signals continuously vary in time. A least squares solution can be derived but exhibits strong fluctuations because it is sensitive to the noise in the time series data. For regularization, we introduce a measure of smoothness in g(s) as prior knowledge. We do this by assuming that g(s) is a second-order intrinsic random time variable, following a linear semivariogram: gðsÞ ¼ b 1 g9ðsÞ ð4Þ E½g9ðsÞ ¼ 0 "s ð5Þ 1 2 E ðg9ðs1hÞ2g9ðsÞÞ ¼ cg ðhÞ ¼ hjhj 2 ð6Þ in which b is the uniform mean value that is unknown a priori; g9(s) is the deviation from the mean; cg(h) is the semivariogram function, here assumed to follow the linear model; h is the time distance; and h is the slope of the linear variogram cg(h). Now, the conditional probability density function p(g|x,y) of the discretized transfer function g, given the data x and y, can be computed by applying Bayes’ theorem: pðgjx; yÞ ¼ pðyjx; gÞpðgÞ pðyÞ ð7Þ in which p(y|x,g) is the likelihood of the output y given the input x and the transfer function g; p(g) is the prior probability density of g, parameterized by the autocorrelated model described previously; and p(y) is a scalar that does not depend on g. Assuming multi-Gaussian probability density functions (pdf) for the likelihood and the prior terms and expressing g by its mean b and deviation g9, the negative logarithm of p(g|x,y) becomes: Lðg9;bjx; yÞ ¼ ðy 2 Xðg9 1 ubÞÞ ðy 2 Xðg9 1 ubÞÞ r2ep 2 g9T G21 gg g9 1 const ð8Þ in which u is a ng 3 1 vector of unit entries and r2ep is the variance of the deviations between y and the best model outcome. r2ep quantifies the epistemic error, which consists of measurement errors, possible numerical errors, and erroneous conceptual assumptions. We assume that the epistemic error is identical for all elements of y. Ggg is the ng 3 ng matrix of discrete semivariogram values, obtained for all pairs of elements in g; i.e., element (i,j) of Ggg is defined by gg(i, j) ¼ cg(|ti2tj|). The negative posterior log-likelihood L(g9,b|x,y) is the objective function in the direct computation of the most likely value, or best estimate. The solution minimizing L(g9,b|x,y) may lead to negative elements of g ¼ g91ub. Representing the response to a pulse-related tracer test, the transfer function must not be negative because that would imply negative concentrations in the observation well during the tracer test. Thus, we have to enforce nonnegativity: g ¼ g9 1 ub 0 ð9Þ In the application to contaminant-source identification, Snodgrass and Kitanidis (1997) ensured nonnegativity by applying a power-law transformation of the unknown, which includes the log-transformation as a special case. In this approach, it is impossible to get elements of zero in g. Nonnegativity can also be ensured by using a prior distribution p(g) with zero probability density in the negative range (Fienen and Kitanidis 2006). Then, the objective function of Equation 8 is not valid. Fienen and Kitanidis (2006) assumed a reflected Gaussian distribution of the parameters and applied a Markov chain Monte Carlo method with Gibbs sampler, originally derived for source identification (Michalak and Kitanidis 2003), to the deconvolution problem. The latter approach has high computational costs, which makes the method ill-suited for problems in which several hundred parameters need to be identified. In the present study, we use a different approach. We keep Equation 8 and account for the nonnegativity constraints by the method of Lagrange multipliers (e.g., Vogel 2002, chap. 9). Direct Computation of Best Nonnegative Estimate Equation 9 is an inequality constraint. Only when a certain element of g becomes negative in the estimation procedure, the constraint has to be activated. We may express this by: Hðg9 1 ubÞ ¼ 0 ð10Þ in which H is a nt 3 ng selection matrix with nt being the number of active constraints: 1; if gj is affected by the i -th constraint Hij ¼ 0; otherwise ð11Þ Then, the constrained minimization of L(g9,b|x,y) consists of solving the following system of linear equations: 32 3 2 1 T 1 T 21 T g9 X X 2 G X Xu H gg 2 76 7 6 r2ep r ep 76 7 6 76 7 6 1 1 T T 6 T T T 76 b 7 u X Xu u 6 2 u X X t 7 54 5 4 rep r2ep t H ut 0 3 2 1 T X y 7 6 r2ep 7 6 7 6 1 ¼6 ð12Þ T T 7 6 2 u X y7 5 4 rep 0 in which t is the nt 3 1 vector of Lagrange multipliers and ut is a nt 3 1 vector of unit entries. Activation and deactivation of the constraints depend on the behavior of the solution. The following rules apply: gj , 0/add constraint for element j ti 0/keep constraint i ti > 0/remove constraint i Since solute transport cannot be instantaneous, we require that the first element of g, representing g(s ¼ 0), is always zero. That is, the Lagrange multiplier for the first element is always active, regardless of the data. The practical implementation requires an iterative procedure, starting without constraints in the first iteration. Depending on the current estimate, constraints are added, kept, or removed, and the estimation is repeated until the selection of constraints stops changing between two iterations. In our applications, typically, between 5 and 10 iterations were necessary to reach convergence. Generation of Plausible Solutions Since convolution is a linear operation, exact expressions for the uncertainty of the estimated parameter vector g would be possible if the constraints were of the equality type. Because the constraints are inequalities, however, the overall optimization problem is effectively nonlinear. This is expressed in the necessity to identify the set of active constraints by an iterative procedure. In order to evaluate the uncertainty of the estimate, in addition to obtaining the most likely value of g by minimizing the negative log-likelihood L(g9,b|x,y), we also generate multiple plausible solutions, each exhibiting variability consistent with the variogram cg(h) and meeting the measurements y with the required accuracy expressed by the epistemic error rep. We do this by the method of smallest possible modification (e.g., Journel and Huijbregts 1978). In each plausible solution, we express the parameter vector g as the sum of three contributions: g ¼ gu9 1 gc9 1 ub 0 ð13Þ in which gu9 is an unconditional realization of the fluctuations in the transfer function, gc9 is a smooth correction O.A. Cirpka et al. GROUND WATER 45, no. 3: 318–328 321 term, and b is the unknown mean value. Here, gu9 is isomorphic to g9, that is, its expected value is zero, and its generalized covariance function is 2cg(h). The unconditional realization gu9 is generated by multiplying the inverse Cholesky decomposition of 2Ggg with a vector of uncorrelated random numbers drawn from a standard normal distribution (e.g., Rubin 2003, section 3.6.1). For the computation of each plausible solution, we also generate a random deviation ey from the measured data y. The elements of ey have expected values of zero and variances of r2ep . In the conditioning, gu9 remains fixed, whereas g9c and b are estimated by minimizing the following objective function: W¼ ðy1ey 2Xðgu91gc91ubÞÞðy1ey 2Xðgu91g91ubÞÞ c r2ep 2g9cT G21 gg gc9 ð14Þ subject to the nonnegativity constraints. For a given unconditional realization g9; u the optimal set of g9c and b is found by solving the following system of linear equations: 32 3 2 1 T 1 T 21 T g9c X X 2 G X Xu H gg 2 76 7 6 r2ep r ep 76 7 6 76 7 6 1 1 T T 6 T T T 76 b 7 u X Xu u 6 2 u X X t 7 54 5 4 rep r2ep t H ut 0 3 2 1 T X ðy 1 ey 2 Xg9Þ u 7 6 r2ep 7 6 7 6 1 ð15Þ ¼6 7 6 2 uT XT ðy 1 ey 2 Xgu9Þ 7 5 4 rep 2 Hg9u in which the constraints are added, kept, and removed according to the same rules as applied in the direct estimate of the most likely value of g. In our application, we generate 1000 plausible solutions. The resulting conditional statistical distribution of g is analyzed for each element of g, resulting in characteristic percentiles of the distribution P(g(s)), which provides approximate confidence intervals. Estimation of Structural Parameters The optimization problem posed depends on two structural parameters: the variance r2ep quantifying the epistemic error and the slope h of the variogram cg(h) quantifying the smoothness of the transfer function g(s). In principle, these parameters can be derived from the measurements y by maximizing their posterior marginal pdf (e.g., Hoeksema and Kitanidis 1984; Kitanidis 1986). Evaluating the posterior marginal pdf may be complicated in nonlinear estimation problems. Also, the approach involves inversing matrices of the order ny, which is the number of elements in y. In our application, ny is larger than 10,000, which prohibits applying the mentioned approaches. Instead, we follow a simplified approximate procedure. For estimating the epistemic error, assumed identical for all measurements, we enforce that the sum of squared residuals, weighted by their variance r2ep , meets its expected value (see Press et al. 1992, equation 15.1.6): 322 O.A. Cirpka et al. GROUND WATER 45, no. 3: 318–328 r2ep ¼ ðy 2 Xðg9 1 ubÞÞ ðy 2 Xðg9 1 ubÞÞ ny 2 n g 1 nt 2 1 ð16Þ in which ny 2 ng 1 nt 2 1 corresponds to the degrees of freedom in estimating g from y, subject to nt active constraints. This approach is well accepted in optimization of overdetermined problems. Estimating the slope h of the variogram cg(h) in a rigorous way is conceptually and computationally a more demanding task than estimating r2ep . In many practical applications, it may be sufficient to vary h manually and pick a value at which major features of the transfer function, such as larger peaks, are recovered whereas smallscale fluctuations are smoothed. In the present study, we estimate h by the expectationmaximization method (McLachlan and Krishnan 1997). The general idea is to find the slope h that best describes the variability in the set of plausible solutions. For a discretized random variable following a linear semivariogram, it can be shown that the probability density function of the entire vector can be computed by considering the squared difference between two neighboring elements (Michalak and Kitanidis 2003). In our analysis, we consider only the nonzero entries of our estimated vector g. The vector of nonzero elements in the j-th solution is denoted gnz,j. The remaining elements in g are fixed to zero and affect the probability density of gnz, j. The log-probability density ln p(gnz,j|h) of the nonzero elements for given h and Lagrange mulitpliers is: nt; j 2 ng ng 2 1 lnðsÞ lnpðgnz; j jhÞ¼ lnð4phÞ2 2 2 nX t;j 21 1 lnðs0i11; j 2 s0i;j Þ 2 i¼1 ! nX g 21 ðgi11 2gi Þ2 2 1 const 4hðsi11 2si Þ i¼1 1 ð17Þ in which nt,j is the number of Lagrange multipliers in the j-th solution, s is the time-shift increment of the full vector g, and s0i;j is the time shift associated with the i-th zero element in the j-th solution. The constant in Equation 17 does not depend on h. Given a set of solutions, an updated value of h, fitting all plausible solutions best, is computed by minimizing: nr X ð18Þ uðhÞ ¼ 2 ln pðgnz; j jhÞ j¼1 The minimization is done by the Melder-Nead simplex algorithm as implemented in Matlab (Lagarias et al. 1998). After updating h, a new set of solutions has to be generated. With the new set, h is updated again. The procedure is repeated until the relative change of h between two sets of solutions is less than 1%. Application to Experimental Data from a Site at River Thur Description of the Study Site We analyzed data collected from the capture zone of a pumping well producing drinking water for about 30,000 people in the city of Frauenfeld, Canton Thurgau, Switzerland. The site is located in the floodplain of the central Thur valley, which is about 30 km long and 2 km wide. During the Pleistocene, glaciers cut into the underlying tertiary bedrock. The retreating glacier left a lake behind, filling the valley with fine lacustrine sediments, acting now as aquitard. These clays are overlain by an approximately 10-m-thick layer of bedded glaciofluvial gravel and sand, forming a productive aquifer. From pumping tests, the mean hydraulic conductivity of this material is known to range between 1 3 1023 m/s and 5 3 1023 m/s. The top layer of approximately 2 m thickness consists of alluvial loam. At our study site, the aquifer may be considered as semiconfined. The upper catchment of River Thur is prealpine with Mount Säntis (2502 m above sea level) being the highest point. In the central Thur valley (altitude ~ 400 m above sea level), the river was channelized in the 1890s. At our site, River Thur flows from east to west along the northern edge of the valley. The main channel is about 40 to 45 m wide. There is no overbank on the right-hand side, whereas the left overbank is about 130 m wide, defined by a levee, behind which a side channel was installed to capture discharge from tributaries and drain agricultural land. The pumping well is located outside the overbank, with a distance to the levee of approximately 80 m. It is a horizontal well with nine arms, typically producing approximately 9000 m3/d. In the fall of 2003, four monitoring wells (PVC pipes with 4.5 inch diameter) were instrumented in the capture zone of the pumping well. Figure 1 gives a schematic overview of the field site. Monitoring wells 1 and 2 are located directly at the upper edge of the main channel, with well 1 being screened over 2 m at the bottom of the aquifer, and well 2 at the uppermost 2 m of the ground water (depth to water table, 2 to 4 m). The fully screened monitoring well 3 is placed in the middle of the overbank, on the presumed streamline leading from monitoring wells 1 and 2 to the pumping well. Monitoring well 4 is also fully screened; this well was installed outside of the overbank, close to a local upwelling spring feeding a small riverine, denoted Giessen. At the study site, River Thur infiltrates year-round. Ground water is assumed to be stratified, with a local flow component on top that is caused by the infiltration of river water and a regional flow component at lower sections that follow the main direction of the valley. The overall direction of flow is from northeast to southwest. The hydrological regime of River Thur has alpine characteristic: precipitation in the upper catchment causes rapid fluctuations of the water level along the entire river course. These fluctuations are not dampened by natural or anthropogenic reservoirs. Base flow is highest during snowmelt in the spring, but flow peaks can occur at any time of the year in response to rainfall in the upper catchment. Associated with changes of discharge are fluctuations of electric conductivity during high river flow. Rain water contains lower concentrations of total dissolved solids than soil water and ground water or effluents of waste water treatment plants. As a consequence, electric conductivity in River Thur decreases during peak flows (Figure 2). The conductivity signal is propagated into the aquifer and can be used as a natural tracer. The monitoring wells, the pumping wells, and a station in River Thur have been equipped with data loggers, recording hydraulic head, temperature, and Figure 1. Overview over the test site. Northing and easting are in meters according to the Swiss terrestrial reference system. Gray lines: contours of hydraulic head (0.1 m/line) resulting from two-dimensional simulation of ground water flow. O.A. Cirpka et al. GROUND WATER 45, no. 3: 318–328 323 A: Electric Conductivity 700 σ [µS/cm] 600 500 400 300 200 01−01−04 01−07−04 01−01−05 01−07−05 B: Hydraulic Head h [m a.s.l.] 396 395 electric-conductivity signal measured in the river (gray lines). All wells react to fluctuations of electric conductivity in the river. The fastest and strongest response is observed in monitoring well 2, which is close to the river and screened in the uppermost part of the aquifer. The electric-conductivity signal in the other wells is more dampened. It is also apparent that the electric conductivities measured in the wells differ systematically from those in the river in the springtimes of 2004 and 2005. We attribute this to snowmelt water in the river, which is not in chemical equilibrium with the sediment material. Upon infiltration, this water partially dissolves minerals, leading to increased concentrations of total dissolved solids and thus to increased values of electric conductivity; that is, electric conductivity is not always a conservative tracer. 394 393 392 01−01−04 01−07−04 01−01−05 01−07−05 Figure 2. Electric conductivity and hydraulic heads in River Thur at our experimental site during the study period. electric conductivity at time intervals of 1 h. The time series begin on November 20, 2003, and end at different times, depending on the well. All stations cover at least the time period until April 21, 2005. Figure 3 shows the time series of electric conductivity in the wells. For comparison, the plots are underlain by the Preparation and Analysis of the Data As can be seen from Figure 3, the discrepancies in electric conductivity between ground water and the river appear only in an annually repeating period in the spring. High-frequency components are less affected by the fact that electric conductivity is a nonconservative tracer. In order to eliminate the effects of springtime discrepancies, we remove the seasonal trend from all time series. For this purpose, we fit a uniform mean as well as sine and cosine functions with frequencies 1/year, 2/year, 3/year, and 4/year to the data by standard least square fitting. These trend signals are subtracted from the original data. The remaining time series still contain characteristic high-frequency signals that can be related to each other. B: Monitoring Well 2 700 600 600 σ[µS/cm] σ[µS/cm] A: Monitoring Well 1 700 500 400 500 400 300 300 200 01−01−04 01−07−04 01−01−05 01−07−05 200 01−01−04 01−07−04 01−01−05 01−07−05 D: Monitoring Well 4 700 600 600 σ[µS/cm] σ[µS/cm] C: Monitoring Well 3 700 500 400 500 400 300 300 200 01−01−04 01−07−04 01−01−05 01−07−05 200 01−01−04 01−07−04 01−01−05 01−07−05 E: Pumping Well 700 σ[µS/cm] 600 500 400 300 200 01−01−04 01−07−04 01−01−05 01−07−05 Figure 3. Time series of electric conductivity in the monitoring wells and the pumping well. Gray lines: river data; black lines: well data. 324 O.A. Cirpka et al. GROUND WATER 45, no. 3: 318–328 In the procedure of estimating the transfer function, we generate sets of 1000 solutions. In order to estimate the slope h of the semivariogram cg(h), we have to analyze the set of solutions, update h, and generate a new set of plausible solutions. Typically, this requires a few iterations. To speed up the process, we start with sets of 100 solutions until the estimates of h change only by a few percent from one iteration to the next and continue with sets of 1000 solutions. The program is implemented in Matlab. Results of Deconvolution Figure 4 shows the estimates of the transfer functions g(s) for all monitoring wells and the pumping well. The solid lines indicate mean values ÆgðsÞjx; yæ of 1000 plausible solutions. The crosses denote the directly computed best nonnegative estimate g^ðsÞ. The gray area marks the range bounded by the 16 and 84 percentiles of the statistical distributions given by the set of plausible solutions. That is, 16% of all plausible solutions fall below the gray area and another 16% above. We choose these limits because they mark deviations by 61 standard deviation in the case of a Gaussian distribution. The dotted lines indicate the minimum and maximum values obtained in all solutions. The most striking result of Figure 4 is the good agreement between the mean ÆgðsÞjx; yæ of the plausible solutions and the best estimate g^ðsÞ: We attribute this to the fact that the estimation problem is almost linear. Nonlinearity is exclusively introduced by the nonnegativity of the transfer function. For times at which the best estimate g^ðsÞ is clearly larger than zero, only a few plausible solutions give values of zero. The main differences between the best estimate g^ðsÞ and the mean of the solutions ÆgðsÞjx; yæ can be found in regions where the best estimate requires an active constraint. Here, the mean ÆgðsÞjx; yæ of the plausible solutions typically is nonzero. These periods include early times of the transfer functions for monitoring wells 3 and 4 and the pumping well. The mentioned differences may be significant for water resources management because they concern the earliest breakthrough of solutes in the observation point. According to the scheme to directly compute the best estimate, there is a definite time before which no breakthrough is possible. Using linearized uncertainty propagation, the associated uncertainty would be zero, since at these times Lagrange multipliers are active. The plausible solutions, by contrast, reveal a finite probability of earlier breakthrough. The latter might be important for probabilistic risk assessment because the concentration of a degrading contaminant decreases with travel time. That is, a finite probability of early breakthrough for a pulse of a conservative compound is associated with an increased probability of contaminant breakthrough. Table 1 contains the identified structural parameters, and Table 2 lists characteristic values for the directly computed best estimates g^ðsÞ of the transfer functions. Monitoring well 2 reacts the fastest to conductivity B: Monitoring Well 2 1 0.8 g [1/d] g [1/d] A: Monitoring Well 1 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0 0.6 0.4 0.2 0 0 2 4 6 8 10 12 14 0 1 2 τ [d] 4 5 6 7 τ [d] C: Monitoring Well 3 D: Monitoring Well 4 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0 0.05 0.04 g [1/d] g [1/d] 3 0.03 0.02 0.01 0 0 5 10 15 20 τ [d] 0 10 20 30 40 50 60 τ [d] E: Pumping Well 0.03 g [1/d] mean of plausible solutions 0.02 16% − 84% probability min, max 0.01 best nonnegative estimate 0 0 5 10 15 20 25 30 35 40 τ [d] Figure 4. Estimates of the transfer functions g(s) for all monitoring wells and the pumping well. Solid line: mean of 1000 conditional realizations; gray area: range between 16 and 84 percentiles of the conditional statistical distributions; dotted lines: minimum and maximum values obtained in all realizations; crosses: best nonnegative estimate. O.A. Cirpka et al. GROUND WATER 45, no. 3: 318–328 325 Table 1 Structural Parameters Estimated for the Data Sets Epistemic Error rep (lS/cm) h of Transfer Function (d23) 14.5 20.1 20.3 21.4 19.5 3.24 3 1024 4.76 3 1022 1.03 3 1024 4.89 3 1026 1.33 3 1026 Monitoring well 1 Monitoring well 2 Monitoring well 3 Monitoring well 4 Pumping well changes in the river. The peak occurs already after 10 h and the center of the breakthrough curve after 1.5 d. Monitoring well 2 is also the well with the highest zeroth moment m0 ðg^Þ of the transfer function. Ninety-four percent of the fluctuations in River Thur is recovered in this monitoring well. This is a direct indication that almost all water passing the monitoring well is freshly infiltrated river water. Monitoring well 1 has the same distance to the river as monitoring well 2 but is screened at greater depth. Well 1 reacts considerably slower than well 2 because the path length from the river to the well is longer. We assume that the fluctuating water level of the river determines to what extent the water passing monitoring well 1 at a given time is freshly infiltrated river water. For both monitoring wells 1 and 2, it is not possible to determine a time of first breakthrough; that is, the response of the wells to electric-conductivity fluctuations starts immediately. The transfer functions determined for monitoring wells 3 and 4, as well as for the pumping well, are delayed. According to the directly computed best estimate g^ðsÞ, the times of first and maximum breakthrough increase with increasing distance to the river (Table 2). Considering the plausible solutions, the probability of earlier breakthrough is fairly high, even though only small values of the transfer function are estimated at these times. The transfer function for monitoring well 4 extends over the longest time, exhibiting a distinct tail and a small secondary peak at a travel time of about 50 d (Figure 4). As noted previously, monitoring well 4 is located in the direct vicinity of a small spring, which indicates a local anomaly of the hydraulic properties. The computed zeroth moments of the transfer functions are 60%, 68%, and 26% for monitoring wells 3 and 4 and the pumping well, respectively. For perfectly conservative transport, the zeroth moments may be interpreted as recovery rate. Values considerably smaller than 100% could indicate mixing with water of other origin. However, the missing 40% and 32% in monitoring wells 3 and 4, respectively, may also partially be in the truncated tails of the transfer functions. The possibility of the latter becomes clear if one considers the tails of the plausible solutions rather than that of the best estimate g^ðsÞ. Finally, one may speculate to what extent the incomplete recovery of electric-conductivity fluctuations is attributed to hydrogeochemical reactions of the infiltrating river water with the mineral surfaces of the aquifer. For comparison, Table 2 lists also the time shift tcorr with highest correlation coefficient when the same data sets are analyzed by cross correlation. In these evaluations, the river signals without seasonal trend are smoothed by a rectangular filter function and cross-correlated to the well signals without seasonal trend. The optimal filter width is individually identified for each well (data not shown). The mean travel times tcorr estimated by crosscorrelation fall between the peak time tmod ðg^Þ and the mean travel time tc ðg^Þ according to nonparametric deconvolution. The differences in the mean travel times obtained by cross correlation and nonparametric deconvolution indicate that the rectangular filter chosen for cross correlation is not optimal. The uncertainty of all transfer functions, expressed by the error bounds (lines expressing 16% and 84% probability of exceedance), is fairly small. This is caused by the small values of h, obtained by the expectation-maximization method (Table 1). Small values of h imply smooth estimates of the transfer function. However, the low uncertainty in g(s) comes with epistemic errors rep in the range of 20 lS/cm (Table 1). As explained in the Theory section, the latter includes the model error. That is, the uncertainty in g(s) is conditional on the validity of the transfer function concept, including the assumption of time-invariant transport. In order to accept the concept, a fairly large discrepancy between simulated and measured time curves of electric conductivity must be accepted. Discussion and Conclusions We have presented a method to infer nonparametric transfer functions from measured time series, guaranteeing Table 2 Characteristics of the Directly Determined Best Nonnegative Estimates of Transfer Functions g^ðsÞ Monitoring well 1 Monitoring well 2 Monitoring well 3 Monitoring well 4 Pumping well T (d) ng m0 ð^ gÞ 14 7 20 60 40 336 168 480 1440 960 0.51 0.94 0.60 0.68 0.26 t0 ð^ gÞ tc ð^ gÞ tmod ð^ gÞ rð^ gÞ instant. instant. 1d6h 4 d 22 h 7 d 17 h 3 d 18 h 1 d 11 h 6d2h 21 d 3 h 18 d 2 h 1 d 18 h 8h 4d3h 10 d 7 h 15 d 12 h 3 d 16 h 1 d 13 h 3 d 22 h 15 d 7 h 10 d 0 h tcorr 3d7h 1d2h 5 d 14 h 14 d 2 h 17 d 10 h T, truncation time; ng, number of estimated values; m0 ð^ gÞ, zeroth moment (recovery rate); t0 ð^ gÞ, time of first breakthrough; tc ð^ gÞ, center of gravity; tmodaf ; ð^ gÞ, peak time; rð^ gÞ, standard deviation of determined travel-time distribution (measure of spread); tcorr, optimum time shift of cross-correlation method; instant., instantaneous. 326 O.A. Cirpka et al. GROUND WATER 45, no. 3: 318–328 smoothness and nonnegativity by introducing secondorder intrinsic autocorrelated behavior of the transfer function as prior knowledge and applying Lagrange multipliers. We have applied the method to time series of electric conductivity in a river (input signal) and in wells in the corresponding alluvial aquifer (output signals). In this context, the inferred transfer functions may be interpreted as the response to a tracer test with pulse-like injection or as the probability density function of solute travel time. The physical interpretation as breakthrough curve and the statistical interpretation as probability density function exemplify the necessity of nonnegativity because neither negative concentrations nor negative probability densities are permitted. In contrast to cross-correlation methods, deconvolution yields a full distribution of travel times rather than a single optimal time-shift value. Such distributions may be used to predict the breakthrough of contaminant in case of an accidental spill into the river. They may also be used to quantify the risk that a degrading river-borne compound reaches a production well. Choosing a nonparametric transfer function gives the opportunity to identify special features of the travel-time distribution like the secondary peak of the transfer function for monitoring well 4 or the plateau-like shape identified for the transfer function of monitoring well 1. Such features would be lost in parametric deconvolution. Strictly speaking, the convolution principle holds only for time-invariant processes. As seen in Figure 2, the hydraulic head in River Thur is fairly dynamic. The heads in the wells react very fast to head changes in River Thur. Nonetheless, the hydraulic gradient and therefore the velocity of ground water are not constant. Even so, we could obtain good results with a model that neglects such fluctuations. This is so because the time scale of velocity fluctuations is smaller than the typical travel times to the wells. That is, fluctuations of velocity are already partially averaged out while the signal is propagated from the river to the wells. In the current study, we have applied the deconvolution to the full data sets of electric conductivity measured in River Thur and five wells. If different types of behavior at different time periods can be observed, one may subdivide the time series into several fractions and perform the analysis for each of them. A prerequisite for the latter approach is that the considered fractions are significantly longer than the duration of the inferred transfer function. At our field site, electric conductivity has been proven a valuable and easy-to-measure natural tracer. Presumably because of hydrogeochemical reactions of the infiltrating river water with the mineral matrix, electric conductivity is not fully conservative, but the discrepancies between the river and ground water data sets could be eliminated by removing the seasonal trend. The method of time series analysis presented here does also apply to temperature data, such as those measured at our site (data not shown). Electric conductivity, however, has the advantage that it propagates faster than temperature and is less smoothed. For applications at other sites, we recommend measuring both temperature and electric conductivity at time intervals of 1 h or smaller. If the river does not exhibit significant fluctuations in electric conductivity, or the electric-conductivity signals in the wells are too strongly disturbed by reactive processes, the temperature series may be used. In other cases, we prefer analyzing the electric-conductivity data. Further improvement in natural-tracer analysis could be achieved by considering time series of truly conservative quantities such as sodium and chloride ions, which, unfortunately, are more difficult to measure continuously by probes. Acknowledgments This research was made possible by a grant from Canton Thurgau, Switzerland. We are indebted to Romeo Favero, Andreas Scholtis, and Marco Baumann of Amt für Umwelt, Canton Thurgau, for installing, equipping, and maintaining the observation wells. We thank Mark Bakker and a second reviewer for their constructive remarks in helping improve the paper. References Anderson, M.P. 2005. Heat as a ground water tracer. Ground Water 43, no. 6: 951–968. Bidwell, V.J. 1995. Constrained identification of the aggregated mixing zone model of solute transport in soil. Environment International 21, no. 5: 731–736. Box, G.E.P., and G.M. Jenkins. 1970. Time Series Analysis, Forecasting and Control. San Francisco, California: Holden-Day. BUWAL. 2004. Wegleitung Grundwasserschutz. Bern, Switzerland: Vollzug Umwelt, Bundesamt für Umwelt, Wald und Landschaft. Constantz, J., M.H. Cox, and G.W. Su. 2003. Comparison of heat and bromide as ground water tracers near streams. Ground Water 41, no. 5: 647–656. Davis, S., G. Thompson, H. Bentley, and G. Stiles. 1980. Ground-water tracers—A short review. Ground Water 18, no. 1: 14–23. DVGW. 1995. Richtlinie für Trinkwasserschutzgebiete, 1. Teil: Schutzgebiete für Grundwasser, Richtlinie W 101. Bonn, Germany: Deutscher Verein des Gas- und Wasserfachs. Fienen, M.N., and P.K. Kitanidis. 2006. A Bayesian geostatistical transfer function approach to tracer test analysis. Water Resources Research 42, no. 7: W07426. Hoehn, E., and O.A. Cirpka. 2006. Assessing hyporheic zone dynamics in two alluvial flood plains of the Southern Alps using water temperature and tracers. Hydrology and Earth System Sciences 10, no. 4: 553–563. Hoeksema, R.J., and P.K. Kitanidis. 1984. An application of the geostatistical approach to the inverse problem in twodimensional groundwater modeling. Water Resources Research 20, no. 7: 1003–1020. Journel, A.J., and C.J. Huijbregts. 1978. Mining Geostatistics. London, UK: Academic Press. Jury, W.A. 1982. Simulation of solute transport using a transfer function model. Water Resources Research 18, no. 2: 363– 368. Käss, W. 2004. Geohydrologische Markierungstechnik [Textbook of Geohydrological Marking and Tracing Techniques]. In Lehrbuch der Hydrogeologie, ed. G. Matthess, vol. 9, 2nd ed. Berlin, Germany: Bornträger. Kitanidis, P.K. 1986. Parameter uncertainty in estimation of spatial functions: Bayesian analysis. Water Resources Research 22, no. 4: 499–507. Kreft, A., and A. Zuber. 1978. On the physical meaning of the dispersion equation and its solutions for different initial and O.A. Cirpka et al. GROUND WATER 45, no. 3: 318–328 327 boundary conditions. Chemical Engineering Science 33, no. 11: 1471–1480. Lagarias, J.C., J.A. Reeds, M.H. Wright, and P.E. Wright. 1998. Convergence properties of the Nelder-Mead simplex method in low dimensions. SIAM Journal on Optimization 9, no. 1: 112–147. Lin, A.Y.C., J.F. Debroux, J.A. Cunningham, and M. Reinhard. 2003. Comparison of rhodamine WT and bromide in the determination of hydraulic characteristics of constructed wetlands. Ecological Engineering 20, no. 1: 75–88. Luo, J., O.A. Cirpka, M.N. Fienen, W. Wu, T.L. Mehlhorn, J. Carley, P.M. Jardine, C.S. Criddle, and P.K. Kitanidis. 2006. A parametric transfer function methodology for analyzing reactive transport in nonuniform flow. Journal of Contaminant Hydrology 83, no. 1–2: 27–41. McLachlan, G.J., and T. Krishnan. 1997. The EM Algorithm and Extensions. New York: John Wiley. Michalak, A.M., and P.K. Kitanidis. 2003. A method for enforcing parameter nonnegativity in Bayesian inverse problems with an application to contaminant source identification. Water Resources Research 39, no. 2: 1033. 10.1029/2005 WR 004576. Press, W.H., S.A. Teukolsky, W.T. Vetterling, and B.P. Flannery. 1992. Numerical Recipes in FORTRAN 77: The Art of Scientific Computation, 2nd ed. New York: Cambridge University Press. 328 O.A. Cirpka et al. GROUND WATER 45, no. 3: 318–328 Rubin, Y. 2003. Applied Stochastic Hydrogeology. Oxford, UK: Oxford University Press. Sheets, R.A., R.A. Darnera, and B.L. Whitteberry. 2002. Lag times of bank filtration at a well field, Cincinnati, Ohio, USA. Journal of Hydrology 266, no. 3–4: 162–174. Silliman, S.E., and D.F. Booth. 1993. Analysis of time-series measurements of sediment temperature for identification of gaining vs losing portions of Juday-Creek, Indiana. Journal of Hydrology 146, no. 1–4: 131–148. Silliman, S.E., J. Ramirez, and R.L. McCabe. 1995. Quantifying downflow through creek sediments using temperature timeseries—One-dimensional solution incorporating measured surface-temperature. Journal of Hydrology 167, no. 1–4: 99–119. Skaggs, T.H., and Z.J. Kabala. 1994. Recovering the release history of a groundwater contaminant. Water Resources Research 30, no. 1: 71–79. Snodgrass, M.F., and P.K. Kitanidis. 1997. A geostatistical approach to contaminant source identification. Water Resources Research 33, no. 4: 537–546. SVGW. 2002. Infoblatt TWI 12. Zürich, Switzerland: Schweizerischer Verein des Gas- und Wasserfachs. Vogel, C.R. 2002. Computational Methods for Inverse Problems. Philadelphia, Pennsylvania: Frontiers in Applied Mathematics, Society for Industrial and Applied Mathematics SIAM.