Journal of 543 (2016) 17–30

Contents lists available at ScienceDirect

Journal of Hydrology

journal homepage: www.elsevier.com/locate/jhydrol

Hillslope permeability architecture controls on subsurface transit time distribution and flow paths ⇑ A.A. Ameli a,d, , N. Amvrosiadi b, T. Grabs b, H. Laudon c, I.F. Creed a, J.J. McDonnell d,e, K. Bishop b,f a Department of Biology, Western University, London, Ontario, Canada b Department of Earth Sciences, Air Water and Landscape Sciences, Uppsala University, Uppsala, Sweden c Department of Forest Ecology and Management, Swedish University of Agricultural Sciences, Umeå, Sweden d Global Institute for Water Security, School of Environment and Sustainability, University of Saskatchewan, Saskatoon, Saskatchewan, Canada e School of Geosciences, University of Aberdeen, Aberdeen, UK f Department of Aquatic Sciences and Assessment, Swedish University of Agricultural Sciences, Uppsala, Sweden article info summary

Article history: Defining the catchment transit time distribution remains a challenge. Here, we used a new semi- Available online 7 May 2016 analytical physically-based integrated subsurface flow and advective–dispersive particle movement model to assess the subsurface controls on subsurface water flow paths and transit time distributions. Keywords: First, we tested the efficacy of the new model for simulation of the observed dynamics at Time invariant transit time distribution the -studied S-transect hillslope (Västrabäcken sub-catchment, Sweden). This system, like many Flow pathline distribution others, is characterized by exponential decline in saturated hydraulic conductivity and porosity with soil Semi-analytical solution depth. The model performed well relative to a tracer-based estimate of transit time distribution as well as Integrated subsurface flow and transport observed groundwater depth–discharge relationship within 30 m of the . Second, we used the model Svartberget catchment model to assess the effect of changes in the subsurface permeability architecture on flow pathlines Saturated–unsaturated flow and transit time distribution in a set of virtual experiments. Vertical patterns of saturated hydraulic con- ductivity and porosity with soil depth significantly influenced hillslope transit time distribution. Increasing infiltration rates significantly decreased mean groundwater age, but not the distribution of transit times relative to mean groundwater age. The location of hillslope hydrologic boundaries, including the groundwater divide and no-flow boundary underlying the hillslope, changed the transit time distri- bution less markedly. These results can guide future decisions on the degree of complexity that is war- ranted in a physically-based rainfall–runoff model to efficiently and explicitly estimate time invariant subsurface pathlines and transit time distribution. Ó 2016 Elsevier B.V. All rights reserved.

1. Introduction Vaché and McDonnell, 2006). This combination of information is valuable for constraining where and for how long biogeochemical The transit time distribution (TTD) and water flow pathlines are processes can occur in the riparian zone; for analyzing how the fundamental descriptions of a catchment (McDonnell and Beven, movement of water alters the critical zone itself, and for under- 2014). The TTD (also referred to as the probability density function standing how different sources of water combine to yield the of transit times) describes how hillslopes store, mix and release dynamics of runoff chemistry (Pinay et al., 2015). For example, water and solutes. The flow pathlines define the different knowledge of subsurface TTD and flow pathlines can, in the case sequences of subsurface environments traversed by water entering of mercury contamination, help identify biogeochemical processes the catchment at different points and at different times. The sub- that are more or less likely to play an important role due to the surface TTD and flow pathlines structure are important for both amount of time spent in a particular soil zone relative to when that the quantity and quality of stream flow (Birkel et al., 2011; water reaches the stream (Eklöf et al., 2014). Due to the complexity of unravelling catchment TTD, few runoff generation or water qual- ity models explicitly include such representations. Improved understanding of controls on flow pathlines and TTD ⇑ Corresponding author at: Department of Biology, Biological & Geological is a challenge facing hydrologists. An important step forward is the Sciences Building, Western University, London, Ontario N6A 5B7, Canada. Tel.: +1 6474708182. exploration of how key features of catchment structure, including E-mail address: [email protected] (A.A. Ameli). permeability architecture (defined here as the vertical distribution http://dx.doi.org/10.1016/j.jhydrol.2016.04.071 0022-1694/Ó 2016 Elsevier B.V. All rights reserved. 18 A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30 of hydraulic conductivity and porosity throughout the catchment of flow paths and transit times simulated using these approaches. control volume) and the locations of hydrological boundaries will A systematic assessment of how rapid (exponential) changes in the influence the TTD and flow paths (Basu et al., 2012). Most methods permeability architecture influence TTD and flow pathlines pose currently used to define TTD do not explicitly take into account particular considerations for grid-based numerical approaches subsurface flow physics and flow pathlines, or even basic ground- since many discrete sub-layers are needed to represent the rapid water flow theory controlling subsurface flow movement. Concep- changes in Ks and porosity values. This treatment of vertical tual convolution method (e.g., Hrachowitz et al., 2009, 2010; heterogeneity in permeability architecture can compromise the Kirchner et al., 2000) and storage selection scheme (Botter, 2012; efficiency of grid-based numerical flow and transport schemes Botter et al., 2011; Harman, 2015; Klaus et al., 2015; Rinaldo when systematically testing hypotheses about how interactions et al., 2015, 2011), are based mainly on the assessment of precipi- among different depth functions for Ks and porosity influence hill- tation and stream flow tracer data without explicit characteriza- slope TTD and flow pathlines. tion of the subsurface permeability architecture. Indeed, these Recently, Ameli et al. (2013) and Ameli and Craig (2014) devel- approaches often use the tracer data to compute the best fit oped a new ‘‘grid-free” integrated flow and transport scheme for parameters of an a priori assumed distribution of transit times explicit simulation of 2-D and 3-D time-invariant subsurface flow for a conceptual convolution scheme, or to calibrate storage- pathlines through unconfined , and transit times along related parameters of a priori defined storage functions that are those pathlines. The coupled saturated–unsaturated semi- assumed to control the type of water mixing and release for the analytical solutions satisfy exactly the saturated and unsaturated storage selection scheme. Ali et al. (2014), Basu et al. (2012) and governing equations (including mass balance). The semi- more recently Kirchner (2016) have commented on the limitations analytical solutions also take into account infiltration rate, natural of these implicit schemes for the prediction of TTD. They (and geometry of the unconfined and calculate the a priori others) have noted that these models are difficult to analyze in unknown locations of and seepage faces using a free relation to subsurface physical heterogeneity which can impact boundary condition rather than assuming the water table as a the variability in hillslope flow pathline and hence the correspond- replica of the ground surface (i.e. the topography-driven water ing TTDs (Ali et al., 2014; Birkel et al., 2011; Fiori and Russo, 2008; table assumption). Without the need for implementation of verti- Hrachowitz et al., 2009, 2010). cal discretization (i.e. sub-layers), these grid-free approaches more Useful simulations of mean transit time (MTT) and TTD through recently have been extended to account explicitly and exactly for an integrated subsurface flow and transport simulation have been various rates of exponential decline in saturated hydraulic conduc- done using numerical approaches (e.g., Ali et al., 2014; Basu et al., tivity with soil depth (Ameli et al., 2016); a characteristic feature of 2012; Cardenas and Jiang, 2010; Fiori and Russo, 2008; Kollet and many till-mantled ‘‘critical zone” environments. This steady-state Maxwell, 2008; Molénat et al., 2013; Sayama and McDonnell, scheme provides a continuous map of head and velocity in the 2009). Numerical approaches were also used to directly model entire hillslope without the need for interpolation, efficiently gen- groundwater age through solving novel groundwater age govern- erating subsurface flow pathlines toward a stream and transit ing equations (Goode, 1996; Woolfenden and Ginn, 2009). And, times along these pathlines. Thus this approach allows for an expli- more recent work has explored the effect of climate on TTD of cit and systematic exploration of the effects of subsurface vertical groundwater more broadly (Maxwell et al., 2015). These heterogeneity, hydrological boundaries and infiltration rate on physically-based models are able to take into account the basic ‘‘time invariant” TTD and flow pathlines in a 2-D hillslope. groundwater flow and transport theories. Numerical approaches Here we provide the first field test of this semi-analytical model were also used to assess the impact of subsurface vertical and lat- against the observed hillslope flow and transport dynamics in an eral heterogeneity on TTD. Kollet and Maxwell (2008) coupled the extensively studied till-mantled hillslope in the Västrabäcken ParFlow numerical steady-state model with a Lagrangian particle sub-catchment of the Krycklan Basin (Laudon et al., 2013). We then tracking approach to assess the impact of changes in macro- use this physically based model as the test bed for virtual experi- dispersion values on TTD. Cardenas and Jiang (2010) linked a ments that systematically explore the impacts of changing physical topography-driven steady-state flow model within the numerical features and permeability architecture of the hillslope and infiltra- COMSOL platform with an Advection–Dispersion–Diffusion trans- tion rate on the transit time distribution and flow pathlines. Specif- port equation to assess the impact of the rate of exponential ically we: decline in saturated hydraulic conductivity (Ks) with depth on TTD. Fiori and Russo (2008) and Fiori et al. (2009) also developed (1) Assess the performance of the semi-analytical approach pre- 3-D numerical flow and transport models to assess the impact of sented by Ameli et al. (2016) in simulating internal hydro- statistically-driven heterogeneity in saturated hydraulic conduc- metric observations along the hillslope and isotopic tivity on the shape of TTD. observations at the catchment. These numerical modeling approaches, however, considered (2) Use virtual experiments to explore the influence of changing relatively smooth changes in subsurface material properties both subsurface conditions (saturated hydraulic conductivity and laterally and vertically. For example, Fiori and Russo (2008) clearly porosity change with soil depth, mechanical dispersion, explained that the level of subsurface heterogeneity that they con- location of no-flow boundary underlying the hillslope, hill- sidered in their numerical experiment was ‘‘moderate” compared slope length and infiltration rate) on subsurface TTD, mean to the much stronger heterogeneity that can be found in many groundwater age and water flow pathlines. catchments. Cardenas and Jiang (2010) also considered gradual changes in Ks with depth; the largest rate of exponential decline 2. Material and methods in Ks with soil depth they considered was a = 0.01. But much more rapid changes in Ks with depth are typical in forested catchments 2.1. Hillslope description (e.g., Harr, 1977; McGuire and McDonnell, 2010), especially in gla- cial tills soils where exponential change values of a up to 4 are The S-transect hillslope is located on the 12 ha Västrabäcken more typical (e.g., Grip, 2015; Lundin, 1982; Nyberg, 1995; sub-catchment (denoted C2 in the Krycklan Basin, Fig. 1a–c), which Seibert et al., 2011). Ali et al. (2014) noted the lack of characteriza- is part of the Svartberget catchment (Laudon et al., 2013). The sub- tion of large scale spatial heterogeneity as one of the major limita- catchment topography is characterized by gentle slopes, and the tions of numerical methods which may impact the verisimilitude subsurface consists of well-developed podzols, overlying glacial A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30 19

( − ) ( , )= 0

ɳ( ) ( , )= 0

Fig. 1. Layout of the S-transect. (a) Study location, (b) plan view of 12 ha Västrabäcken sub-catchment and (c) plan view of the S-transect along with the location of discharge measurement site and groundwater measurement . Groundwater wells, referred to as S4, S12, S22 and S140, located at 4, 12, 22 and 140 m (sub-catchment divide) from the stream. (d) 2-D cross section of the S-transect used here with a length of L. The topographic surface Zt ðxÞ was generated from the original 5 m LIDAR DEM. Saturated hydraulic conductivity (Ks) and porosity (hs) decay exponentially with depth. a and Þ are the parameters of the exponential relationship between saturated hydraulic 1 conductivity and porosity with depth, respectively. Ks0 [L T ] and hs0 are the saturated hydraulic conductivity and porosity along the topographic surface (zt ðxÞ). The topographic surface (zt ðxÞ) is subject to a Dirichlet condition along the surface water course (red circle) with a constant head of 0.45 m, and specified infiltration function

(RðxÞ) along the remaining part. The bottom boundary zbðxÞ with a slope of 0.5% and divide (right side of the hillslope at X = 140 m) are assumed impermeable. The a priori unknown water table zWT ðxÞ is also calculated as a boundary with zero pressure head (green line). The blue lines show the alternative locations of the no-flow boundary underlying the hillslope used in the Section 3.7. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

till. This hillslope has been the subject of a large number of previ- retention data which estimated the Gardner sorptive number as ous hydrological (Laudon et al., 2004; Peralta-Tapia et al., 2014; b = 1 1/m and an air entry pressure of ue = 0.05 m. The water con-

Stähli et al., 2001) and biogeochemical (Bishop et al., 1995; Cory tent at zero matric tension was used as the porosity (hs) at each et al., 2007; Klaminder et al., 2006; Leith et al., 2014; Peralta- depth. The best exponential fit function to porosity–depth mea-

0:26ðzZt Þ Tapia et al., 2016; Seibert et al., 2009) studies. The sub- surements of the mineral soil was hsðx; zÞ¼0:49e , where catchment vegetation is dominated by Norway Spruce and Scots (z zt) refers to the soil depth. The Ks-depth relationship was also Pine with an undercover of bilberry. Average daily precipitation measured in the sub-catchment using the permeameter method and actual evapotranspiration during the study period (Bishop, 1991). The best fit exponential function to the observed

2:46ðzZt Þ (13.10.2013–22.09.2014) were 1.79 and 0.96 mm/day, respec- Ks-depth data was ksðx; zÞ¼86e m/d. tively, where daily actual evapotranspiration was calculated using Peralta-Tapia et al. (2016) analyzed a 10-year time series iso- the HBV model (Seibert, 2000). Precipitation during the study per- topic data (d18O and d2H) for stream and precipitation waters iod was 653 mm, slightly higher than the long-term (1990–2012) within the Svartberget catchment to estimate TTD using a convolu- mean annual precipitation (630 mm). Temperature during the tion approach; the stream data was collected from the same study period was 4.4 °C, higher than the long-term (1990–2012) stream as S-transect discharged into. A Gamma distribution with mean annual temperature (2.2 °C). Therefore, the study period a shape parameter of 0.59 was the best fit to their observed iso- was slightly wetter and warmer than the average conditions of topic data (black line in Fig. 2b). In addition, in an analysis of soil the 23 previous years. water d18O during snowmelt/ flood on the S-transect, Daily stream discharge was taken from the continuous mea- Laudon et al. (2004) showed that the infiltrating rainfall/snowmelt surements at a V-notch weir located 0.5 km downstream of the does not penetrate deeper than 90 cm at S22 before and after S-transect (Fig. 1a). During the study period, the 10th percentile, spring flood. These findings were also used here to assess the effi- median, average, and 90th percentile daily discharge were 0.12, cacy of the subsurface flow and transport model in simulating the 0.50, 0.80 and 1.80 mm/d, respectively, which were consistent TTD and flow pathlines. with long term daily stream discharge. Average daily groundwater A 5-m LiDAR Digital elevation model (DEM) was used to depth measurements were collected with pressure transducers at derive sub-catchment divides and to create a 2-D representation four groundwater wells located along the hillslope transect (as of the surface of the S-transect along the topographic fall line defined by the topographic fall line) at 4 m, 12 m, 22 m, and (Fig. 1d).The S-transect is characterized by a planar geometry 140 m (sub-catchment divide) from the stream. Sites below these with an almost uniform width and slope. The ratio of hillslope wells were referred to as S4, S12, S22, and S140. Statistical t- width variation with respect to the hillslope length was almost tests showed that groundwater depth was significantly related to 0.03 and the slope of the hillslope in the direction perpendicular stream discharge (p < 0.001) (Fig. 2, black lines represent the best to the slope varied in a range of less than 2%. The terrain analysis fit to the observed groundwater depth–discharge relationship). of the DEM also suggested that the flow direction in the transect These stream-discharge relationships were used to parameterize is North–South. Thus, we believe that the 2-D representation the subsurface flow model in this paper. Soil samples were col- should be an appropriate emulation of the natural subsurface lected from different depths at S4, S12, S22, and the water content behavior within the S-transect. Groundwater well locations was measured at different matric tensions in order to calculate soil (Fig. 1c) were then projected onto this 2-D representation; obser- moisture retention curves (Nyberg et al., 2001). The unsaturated vations showed there was no net recharge lower than 30 cm Gardner model (Gardner, 1958), used here to characterize the below the streambed and so we placed the no-flow boundary behavior, was fitted to the observed soil moisture at this depth (Fig. 1d). 20 A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30

Fig. 2. Comparison between observed and simulated hydrological and transport processes. (a) Observed (circles) and simulated (triangles) relationship between groundwater depth and specific discharge at sections S4, S12, S22 and S140. The black lines represent the best fit to the observed groundwater depth–discharge relationship based on data collected from October 2013 to October 2014. Red and blue triangles represent the simulated relationship between groundwater depth and discharge in the calibration (red) and validation (blue). (b) Observed and simulated (weighted ensemble) dimensionless TTD. The former was obtained by imposing a convolution approach on the 10-year time series of isotopic data (d18O and d2H) for stream water and precipitation within the Svartberget catchment of which the S-transect and the Västrabäcken sub-catchment are a part; the stream data was collected from the same stream as S-transect discharged into. The simulated weighted ensemble TTD was calculated by assembling simulated transit times of water particles discharged into the stream in response to various flow rates. The transit times were weighted based on the frequency of occurrence of the corresponding stream discharge rates in annual discharge frequency distribution. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

2.2. Modeling method zones, respectively, can be coupled with a uniform RWPT scheme (e.g., Salamon et al., 2006) to generate flow pathlines from the The 2D schematic of S-transect hillslope with an exponential topographic surface to the surface water course and calculate tran- decay in saturated hydraulic conductivity and porosity with depth sit time along the pathlines. To do that, first, continuous maps of ð ; Þ is shown in Fig. 1d. The semi-analytical series solution method of Darcy fluxes throughout the entire saturated zone (qsx x z & ð ; Þ Ameli et al. (2016) was used to calculate the continuous fields of qsz x z ) and Darcy–Buckingham fluxes in the unsaturated zone ð ; Þ ð ; Þ saturated and unsaturated hydraulic head and velocity in the (qux x z & quz x z ) are required. These fluxes can be calculated as: entire hillslope (Appendix A). This solution was coupled with a að z Þ d/ ðx; zÞ að z Þ d/ ðx; zÞ Random Walk Particle Tracking (RWPT) transport method to gen- q ðx; zÞ¼e z t s & q ðx; zÞ¼e z t s ð1aÞ sx dx sz dz erate the flow pathlines and transit time along these pathlines.

að z Þ d/ ðx; zÞ ð ; Þ¼ z t u & ð ; Þ qux x z e quz x z 2.2.1. Uniform Random Walk Particle Tracking (RWPT) method dx að z Þ d/ ðx; zÞ The calculated discharge potential function (Eq. (A.1)) and ¼ z t ½ u þ b/ ð ; Þ ð Þ e u x z 1b Kirchhoff potential (Eq. (A.2)) in the saturated and unsaturated dz A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30 21

Substitution of Eqs. (A.1) and (A.2) into Eq. (1) yields:

XN ÀÁÂÃÀÁ ÂÃÀÁ XN ÀÁÂÃÀÁ ÂÃÀÁ að Þ p p p að Þ p p ð ; Þ¼ z zt n n ðc Þ þ n ðc Þ & ð ; Þ¼ z zt c n ðc Þ þ c n ðc Þ qsx x z e L An sin L x exp nz Bn sin L x exp nz qsz x z e nAn cos L x exp nz nBn cos L x exp nz n¼1 n¼1 ð2aÞ

XM ÂÃÀÁ ÂÃÀÁ XM ÀÁÂÃÀÁ ÂÃÀÁ að Þ p p p að Þ p p p ð ; Þ¼ z zt m m ð Þ £m þ m ð Þ £m & ð ; Þ¼ z zt m m ð Þ þ m ð Þ qux x z e Cm sin x exp £mz mp Dm sin x exp £mz mp quz x z e Cm cos x exp £mz Dm cos x exp £mz L L L L L L L L m¼1 m¼1 ð2bÞ

Continuous fields of pore water velocity are then calculated as: old waters (both tails of the TTD) relative to mean transit time

ð ; Þ ð ; Þ (Godsey et al., 2010). To schematize the degree of variability in tran- qsx x z qsz x z  V sxðx; zÞ¼ & V szðx; zÞ¼ ð3aÞ hsðx;zÞ hsðx;zÞ s sit times, the dimensionless transit time distribution (q s s0) with  0 ð ; Þ ð ; Þ s qux x z quz x z V uxðx; zÞ¼ & V uzðx; zÞ¼ ð3bÞ respect to scaled transit time s (as has done in Fiori and Russo huðx;zÞ huðx;zÞ 0 (2008)) is also shown in this paper. where the saturated moisture content (hs) is equal to the porosity h ð ; Þ¼ and is obtained as a function of soil depth s x z 2.2.2. Virtual experiment Þð ð ÞÞ h ð ; Þ z Zt x ðh Þ s0 x z e . The unsaturated moisture content u is also One possible way forward for synthetic work aimed at under- u obtained based on both the suction pressure head ( ) and soil standing the controls on hillslope flow and transport processes, Þð Þ ðbðuueÞÞ ðh ð ; ; uÞ¼h ð ; Þ z Zt depth at each location u x z s0 x z e e ). Using together with the potential interaction among controls, is virtual the calculated continuous fields of V x and V z (Eq. (3)) in the entire experimentation. Weiler and McDonnell (2004) defined this as hillslope, the uniform random walk step of a water particle is given experiments driven by a collective field intelligence and performed by: using robust modeling approaches. Once our semi-analytical series pffiffiffiffiffiffiffiffiffiffiffiffiffiffi V pffiffiffiffiffiffiffiffiffiffiffiffiffiffi V solution model was parameterized and assessed based on the field xk ¼ xk1 þ V Dt þ 2D DtX x 2D DtX z ð4aÞ p p x L L jVj T T jVj data, the impact of different controls was assessed by changing the value of the control (all else being equal) without any further cal- pffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffi ibration. The interaction among controls was also assessed using a k k1 V z V x z ¼ z þ V zDt þ 2DLDtXL 2DT DtXT ð4bÞ similar approach. p p jVj jVj qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi This virtual experimentation approach improves our under- standing of how different subsurface structures and infiltration where jVj¼ V 2 þ V 2 & D ¼ a jVj & D ¼ a jVj x z L L T T rates influence time-invariant TTD. While the list of possible vari- k k xp and zp are the particle position at the kth time step, DL and DT ables is long, some of the more important factors are: (1) the rate of 2 1 [L T ] are the longitudinal and transverse mechanical dispersion exponential decline in K with soil depth (a); (2) the rate of expo- a a s coefficients, respectively, and L and T [L] refer to longitudinal and nential decline in porosity with soil depth (Þ); (3) mechanical dis- transverse dispersivity of the porous medium. XL and XT denote persion (aL); (4) infiltration rate (R); (5) hillslope length, L (i.e., the random numbers drawn from a normal distribution with zero distance along the hillslope between the stream and the local sub- mean and unit variance for each particle and each time step (Dt). catchment water divide); and (6) the location of the no-flow

All particles were initially released at evenly-spaced locations boundary underlying the hillslope where Za refers to the minimum along the topographic surface. The optimum number of initially vertical distance between stream bed and the no-flow boundary. released particles along the topographic surface was first obtained The modeling approach we used here can efficiently take into by performing a sensitivity analysis which assesses the impact of account systematic vertical changes in Ks and porosity with depth the number of particles on MTT and TTD. The pathlines generated as well as these other aforementioned factors. As will be shown and the residence times along these pathlines were then used to later, the impact of changes in Ks with soil depth (factor 1) on sub- calculate the transit times of water particles discharged into the surface flow characteristics is significant. Thus, the effects of fac- watercourse. We then fitted various distributions including tors 2–6 were assessed for two end members of exponential power-law, Weibull and Gamma distributions to the simulated decline in saturated hydraulic conductivity with soil depth (i.e. transit times to characterize the transit time probability density extreme vertical heterogeneity in ks which is defined in this paper function. As expected, the Gamma distribution was the best fit to by a = 3 and a homogenous ks with a = 0) to incorporate the inter- the simulated transit times for all examples we solved in this action among these factors. Table 1 reports the list of controlling paper. The expression for the Gamma distribution probability den- factors that were varied in the virtual experiments and their tested sity function is expressed as a function of the transit time (s) as: levels. The selected values of controlling factors used in the virtual as a experiments were from the range observed in till environments ðs Þ s 0 as qðsÞ¼ e 0 ð5Þ with the exception of zero values for a, Þ and aL which represent sCðaÞ a pure homogenous media as well as Þ = 0.75, which represents where a is the Gamma distribution shape parameter and s0 is mean extreme heterogeneity in porosity. transit time. The shape parameter (a) in the gamma distribution describes how much weight is found in the tails of the distribution, 3. Results versus near the center, and is a measure of the degree of variability of subsurface transit times (Kirchner, 2016). The ratio of the stan- The efficacy the semi-analytical series solution method of Ameli dard deviation to the mean of transit times equals the square root et al. (2016) in simulating the groundwater depth–discharge rela- of 1/a. Thus, as the Gamma shape parameter decreases, the variabil- tionship at S4, S12, S22, and S140 as well as TTD of water particles ity in transit times increases with higher proportions of young and discharged into the stream is assessed. The model is then used to 22 A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30

Table 1

List of controlling factors (i.e. permeability, hydrological boundary and flow rate parameters) that were varied in the virtual experiments and their tested levels. Za is the minimum vertical distance calculated between stream bed and no-flow boundary at the bottom of the hillslope (Section 3.7). The original and alternative bottom boundaries used in this section are shown in Fig. 1d.

Parameter a h Þ a a Factor ks0 [m/ [1/m] s0 [] [] R [mm/d] L [cm] T [cm] L [m] za [m] d] a (Section 3.2) 100 0; 1; 2; 3 0.49 0.26 1.8 1 0.01 140 0.30 Þ (Section 3.3) 100 0; 3 0.49 0; 0.25; 0.5; 0.75 1.8 1 0.01 140 0.30 R (Section 3.4) 100 0; 3 0.49 0.26 0.12; 0.5; 0.8; 1.8 1 0.01 140 0.30

aL (Section 3.5) 100 0; 3 0.49 0.26 1.8 0; 0.1; 1; 10 0; 0.001; 0.01; 0.1 140 0.30 L (Section 3.6) 100 0; 3 0.49 0.26 1.8 1 0.01 80; 100; 120; 140 0.30

za (Section 3.7) 100 0; 3 0.49 0.26 1.8 1 0.01 140 0.30; 1.30; 2.30, 3.30

examine the sensitivity of the flow pathlines and TTD to the archi- tecture of the hillslope permeability, location of hydrological Table 2 Convergence of MTT (Gamma shape parameter) using various numbers of particles boundaries, and infiltration rate in sets of virtual experiment. initially released from the topographic surface for the calibrated model with a ¼ 3

and hypothetical model with a ¼ 0. For this convergence analysis ks0 = 100 [m/d], h Þ a a 3.1. Model calibration, construction of the annual TTD and assessment s0 = 0.49, = 0.26, R=1.8 [mm/d], L = 1 [cm], T = 0.01 [cm], L = 140 m and za = 0.30 [m] were considered. against independent observations

Np =70 Np = 140 Np = 280 Np = 560

Our series solution model was calibrated to define the Ks profile a ¼ 3 145 (0.62) 174 (0.54) 195 (0.52) 199 (0.55) a ¼ (i.e. Ks0 & a) along the S-transect. The calibration objective was to 0 64 (0.94) 65 (0.93) 65 (0.93) 65 (0.93) emulate the observed groundwater depth–discharge relationship at groundwater monitoring wells located at S4, S12, S22 for the average annual runoff rate. Since earlier work on such till hillslopes To compare the model TTD to a tracer-based estimate of TTD, had suggested that a steady state assumption was valid within sev- we assumed the catchment flow system to be a succession of eral tens of meters from the stream, we chose to prioritize the fit of steady states. The semi-analytical model TTD was calculated by the three groundwater tubes nearest the stream and did not use assembling simulated transit times of water particles discharged the groundwater measuring tube located at S140 (sub-catchment into the stream in response to various infiltration rates. These divide) in the calibration process. Note that for this steady state aggregated transit times were weighted according to the frequency model the infiltration and discharge rates were assumed to be of occurrence of the corresponding stream discharge rates (which identical. Thus the calibration was made using the annual average is here equal to infiltration) in the annual discharge frequency dis- discharge of 0.8 mm/d as the infiltration rate. Manually calibrated tribution. A Gamma distribution with a shape parameter of 0.51 values of the parameters were Ks0 = 100 m/d and a = 3 1/m. The was the best fit to the simulated ensemble transit times (Fig. 2b) calibrated exponential function for the saturated hydraulic con- at the S-transect. To simulate transit times, in addition to the cal- ductivity–depth relationship was consistent with the best fit expo- ibrated parameters, observed porosity patterns with Þ = 0.26, nential function fitted to the observed saturated hydraulic hs0 = 0.49 within the hillslope, as well as longitudinal and trans- conductivity–depth data obtained using the permeameter method verse dispersivities of 1 cm and 0.01 cm were considered. The dis-

(the best fit parameters from Bishop (1991) were Ks0 = 86 m/d and persivity values spanned what is typical in till environments. The a=2.49 1/m). annual ensemble TTD using the calibrated semi-analytical model In the calibration phase, the model accurately simulated the yielded a Gamma shape parameter of 0.51. This was comparable groundwater depth–discharge relationship in the hillslope within to the TTD obtained by imposing a convolution approach on the 30 m of the water course (the red triangles in Fig. 2a) for the cali- 10-year time series of isotopic data (d18O and d2H) for precipitation bration condition of a 0.8 mm/d discharge rate. High flow (1.8 mm/ and stream runoff from the catchment which yielded a Gamma d), median flow (0.5 mm/d) and low flow (0.12 mm/d) were used shape parameter of 0.59 (Fig. 2b). As stated earlier the discharge to test the model in the validation phase (the blue triangles in frequency distribution during the study period was similar to the Fig. 2a). The model accurately reproduced the groundwater level long term discharge frequency distribution. This, together with within 30 m of the water course for other discharge rates as well, the fact that the collected isotopic data was taken from the stream while further away from the watercourse at S140 in the vicinity that the S-transect discharged into, suggests that the integrated of the sub-catchment divide the predicted groundwater levels flow and transport model can reasonably emulate the transport remained about 0.5–1 m higher than the observed values. This behavior of the hillslope. was attributed to the fact that in the vicinity of sub-catchment divide, steady state conditions were not expected as shown by manual hydrometric measurements made on a hillslope close to 3.2. Effect of rate of exponential decline in saturated hydraulic the one examined in this study (Seibert et al. (2003). The calibrated conductivity with soil depth subsurface flow model provided a continuous map of the velocity field in the entire domain at a given infiltration rate. This was used The model was then used to explore the impact of changing the within Random Walk Particle Tracking (RWPT) analysis to generate rate of exponential decline in Ks with soil depth on the distribu- flow pathlines and transit times for any given infiltration rate and tions of flow pathlines and transit time. As the rate of exponential the associated stream discharge. The optimum number of particles decline in ks with depth (a) increases, shallow and deep subsurface at which MTT and Gamma shape parameter of the fitted TTD con- flow circulation enhances and lessens, respectively (Fig. 3). This verged was 560 (i.e., at each 25 cm interval along the topographic can be tied to the fact that the water table reaches the superficial surface) (Table 2). This initial condition was used in the original portion of the hillslope with a relatively higher conductivity closer model and the virtual experiments. to the topographic surface. A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30 23

Fig. 3. Flow pathline distribution and water table (green line), for different rates in the exponential decline of ks with soil depth (a). (a) a = 3, (b) a = 2, (c) a = 1 and (d) a =0 1 (homogenous case). Only 8 of all particles used in the RWPT analysis were shown. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

shown here). This was because the variation in porosity identically affected pore water velocity in both x and z directions in the entire domain (Eq. (3)). Changes in the strength of exponential decline in porosity with soil depth (Þ) impacted the dimensionless TTD and mean groundwater age; albeit this effect was more pronounced

for homogenous Ks (Fig. 5b) compared to the case with an extreme vertical heterogeneity in Ks (Fig. 5a). As Þ increased, the variability of transit times relative to mean ground water age decreased for

both Ks patterns. This was also supported by the increase in the Gamma shape parameters. Indeed, for the case with an extreme

vertical heterogeneity in Ks with soil depth, as Þ increased, the higher and lower relative porosity at shallow and deep portions, respectively, canceled out higher and lower relative Darcy a Fig. 4. Effect of rate of exponential decline in ks with soil depth ( ) on dimension- (Darcy–Buckingham) fluxes (Eq. (3)) at shallow and deep portions, s less TTD. 0 represents mean groundwater age (day). The Gamma shape parameter which led to a small decrease in MTT and the variability of transit of the TTD varies from 0.92 to 0.50 as a increases from zero to 3. times. On the other hand, for the homogenous Ks, the pore water velocity of deeply penetrating pathlines increased considerably The dimensionless transit time probability density function, with an increase in Þ. This led to a pronounced decrease in MTT qs0, also suggests that as a increased the proportions of (relatively) and the variability of transit times since these deep pathlines early and late particle arrivals discharged into the stream increased formed the tail of the TTD. markedly (Fig. 4), increasing the variability in transit times. This is associated with a considerable decrease in the Gamma shape 3.4. Effect of infiltration rate (R) parameter from 0.92 to 0.50 as a increased from 0 (homogenous case) to 3 (the extreme heterogeneous case). The mean groundwa- s The model was used to assess the effect of infiltration rate on ter age ( 0) also increased, which can be tied to an overall decrease flow pathlines, groundwater table location, TTD and mean ground- a in ks as increased given an identical ks along the topographic water age. High flow (1.8 mm/d), average flow (0.8 mm/d), median surface. flow (0.5 mm/d) and low flow (0.12 mm/d) during the study period were considered here as infiltration rates. As infiltration rate 3.3. Effect of exponential decline in porosity with soil depth (Þ) increased, shallow flow in the vicinity of the watercourse (the dis- charge area) also increased since the water table was raised to The effect of exponential decline in porosity with soil depth on reach the superficial zone of higher conductivity closer to the topo- TTD was assessed for two end members of a (a = 3 and a = 0). As Þ, graphic surface (Fig. 6). In addition, at the S-transect, it was the parameter describing the exponential decline in porosity with observed that the infiltrating rainfall/snowmelt did not percolate soil depth varied, flow pathline distribution did not differ (not deeper than 90 cm at S22 before and after spring flood based on 24 A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30

Fig. 5. Effect of the strength of exponential decline in porosity with depth (Þ) on the dimensionless transit time probability density function. s0 represents mean groundwater age (day). (a) a = 3 (extreme vertical heterogeneity in Ks), the Gamma shape parameter varies from 0.43 to 0.66 as Þ increases from 0 to 0.75. (b) a = 0 (homogenous Ks), the Gamma shape parameter varies from 0.63 to 1.70 as Þ increases from 0 to 0.75.

Fig. 6. Flow pathline distribution and water table (green line) for different values of infiltration rate (R). (a) R = 1.8 mm/d (high flow), (b) R = 0.8 mm/d (average flow), (c) 1 R = 0.5 mm/d (median flow) and (d) R = 0.12 mm/d (low flow). Only 8 of all particles used in the RWPT analysis were shown. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.) the (d18O) analysis. This was comparable to the simulated flow seen in the small variation in the Gamma shape parameter as infil- pathlines with a flow pathline depth of almost 110–140 cm at tration rate varied from low flow to high flow. Indeed, although S22 as infiltration rate decreased from high flow to low flow infiltration rate pronouncedly influenced the water table location (Fig. 6). and the transit time of water particles discharged into the surface There was an overall decrease in transit times of the water par- water course, the impact on the ‘‘variability” of hillslope transit ticles traversed between topographic surface and stream (Fig. 7a) times relative to mean groundwater age was slight. as the infiltration rate increased. The MTT decreased almost expo- nentially (Fig. 7b). Fig. 7 shows that the longest transit time in 3.5. Effect of dispersivity (aL and aT ) response to the high flow rate (almost 10,000 days), was on the order of the MTT for low flow (6396 days). Infiltration rate, how- Dispersivity is a characteristic property of the porous media, ever, had a negligible impact on the shape of the dimensionless which represents the pore scale (micro) heterogeneity. The effect TTD, so the proportions of early and late particle arrivals dis- of different strengths of longitudinal and transverse dispersivity charged into the stream (relative to MTT). This held true regardless (aL and aT ) on the dimensionless TTD and mean groundwater age of the degree of vertical heterogeneity in Ks (Fig. 8). This was also for two end members of Ks vertical pattern was assessed. Four dif- A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30 25

Fig. 7. Effect of infiltration rate (R [mm/d]) on transit times. (a) Transit time probability density function and (b) mean ground water age (s0).

Fig. 8. Effect of infiltration rate (R [mm/d]) on dimensionless transit time distribution for two different rates of exponential decline in Ks with depth (a). (a) a = 3 (extreme heterogeneity), the Gamma shape parameter varies from 0.52 to 0.50 as R increases from 0.12 to 1.8 mm/d. (b) a = 0 (homogenous case), the Gamma shape parameter varies from 0.97 to 0.92 as R increases from 0.12 to 1.8 mm/d.

Fig. 9. Effect of longitudinal dispersivity (aL [cm]) on dimensionless transit time probability density function for two different rates of exponential decline in Ks with soil a a Þ 1 a depth ( ). A transverse dispersivity ( T equals to 100 of the longitudinal dispersivity was considered for each case. (a) = 3 (extreme heterogeneity in Ks), the Gamma shape parameter varies from 0.53 to 0.49 as aL increases from 0 to 10 cm. (b) a = 0 (homogenous Ks), the Gamma shape parameter varies from 1 to 0.85 as aL increases from 0 to 10 cm.

ferent values of aL equal to 10, 1, 0.1 and 0 cm were considered 3.6. Effect of hillslope length (L) a Þ 1 (Fig. 9). A transverse dispersivity ( T equal to 100 of the longitudi- nal dispersivity was also considered for each case (Table 1). An As hillslope length (i.e., the horizontal distance between surface increase in dispersivity decreased the MTT. However, the effect water course and sub-catchment divide) increased, the variability of dispersivity on the dimensionless TTD depended upon the of transit times changed only slightly at both end members of Ks degree of vertical heterogeneity in Ks. For homogenous Ks (a = 0), vertical patterns (Fig. 10). This was reflected in the small decrease an increase in dispersivity values slightly increased the proportions in the Gamma shape parameters. As hillslope length increased, the of young and old waters (relative to mean groundwater age). On water table rose closer to the stream due to a larger amount of the other hand, for extreme vertical heterogeneity in Ks (a = 3), water that must traverse the same soil profile (not shown here). an increase in dispersivity had negligible effects on the variability This led to a shallower unsaturated zone and a slightly increased of transit times. The influence of dispersivity on flow pathlines dis- proportion of young waters for both end members of Ks vertical tribution throughout the hillslope was also negligible (not shown patterns (Fig. 10). For the case with an extreme exponential here). vertical decline in Ks, the elevated water table further enhanced a 26 A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30

Fig. 10. Effect of hillslope length (L [m]) on the dimensionless transit time probability density function for two different rates of exponential decline in Ks with depth (a). (a) a = 3 (extreme heterogeneity in Ks), the Gamma shape parameter varies from 0.55 to 0.50 as L increases from 80 m to 140 m. Even though Gamma shape parameter is almost insensitive to hillslope length, the MTT decreases significantly with an increase in hillslope length. (b) a = 0 (homogenous Ks), the Gamma shape parameter varies from 0.95 to 0.92 as L increases from 80 m to 140 m. Again, Gamma shape parameter is almost insensitive to hillslope length, but the MTT increases slightly with an increase in hillslope length. For the particle tracking simulation in this analysis, particles initially located at each 25 cm interval along the topographic surface were used. Thus, the number of particles was not 560 for all examples, since the length of the hillslope varied.

Fig. 11. Effect of the location of no-flow boundary underlying the hillslope on dimensionless transit time probability density function for two different rates of exponential decline in Ks with depth (a). (a) a = 3 (extreme heterogeneity in Ks), the Gamma shape parameter varies from 0.50 to 0.48 as Za increases. (b) a = 0 (homogenous Ks), the

Gamma shape parameter varies from 0.92 to 1.52 as Za increases. While the gamma shape parameter changes little in the case of extreme Ks heterogeneity, the MTT increases by about 25% (from 199 to 243 days) across the explored range. On a relative scale, the MTT increases more in the case of homogeneous Ks, from 65 to 94 days (45%). shallower flow circulation with a larger pore velocity in the more no-flow boundary at depth promoted deeper flow pathlines and conductive portions of the aquifer closer to the topographic sur- lessened the variability in transit times relative to MTT (Fig. 11b). face. Faster (and shallower) flow circulation together with smaller In addition, MTT for both Ks vertical patterns increased when low- transit times in the unsaturated zone led to a considerable ering the no-flow boundary at the bottom since it increased the decrease in MTT as the hillslope length increased (Fig. 10a). On extent of what the water particles could traverse and also slightly the other hand, for the homogenous Ks (a = 0) with an almost uni- lowered the water table elevation (not shown here). form (with depth) pore water velocity in the entire saturated zone, the increase in the lengths of flow pathlines (as the hillslope length increased) was a dominant control on the slight increase in the 4. Discussion MTT (Fig. 10b). The subsurface vertical permeability architecture and hydrolog- 3.7. Effect of the location of the no-flow boundary underlying the ical boundary locations affect the TTD and flow pathlines (Ali et al., hillslope (Za) 2014; Birkel et al., 2011; Hrachowitz et al., 2010; Kirchner et al., 2001). Few numerical experiments have investigated the impact The original and alternative locations of the no-flow boundary of vertical decline in permeability on TTD either explicitly and underlying the hillslope for this assessment are shown in Fig. 1d implicitly (by using macro-dispersion) (Cardenas and Jiang, 2010; (black and blue lines, respectively) with a minimum vertical dis- Kollet and Maxwell, 2008). However, these numerical experiments tance of Za calculated between the stream bed and the no-flow did not consider the interaction between vertical decline in perme- boundary. Lowering the location of the no-flow boundary (i.e., ability with the other subsurface controls such as location of increasing Za) had a negligible impact on the variability of transit hydrological boundaries and infiltration rate. More importantly, times for the case with an extreme exponential vertical decline these approaches took into account only gradual vertical changes in Ks (Fig. 11a). This was because very few water particles pene- in the permeability pattern. Rapid changes in vertical heterogene- trated the deeper portion of the aquifer. Indeed, the actual depth ity in permeability is a characteristic feature of shallow till soils of the hydrologically active soil layer was almost constant (not where ks and porosity may decline abruptly (e.g., Grip, 2015; varying with the location of the no-flow boundary at the bottom Lundin, 1982; Nyberg, 1995; Seibert et al., 2011). These tills are a of the slope). The deeper parts of the hillslope acted as a low- common soil type in areas with a glaciation history. Recently, flow or even no-flow zone for the case of an extreme exponential Ameli et al. (2016) introduced a new grid-free semi-analytical inte- vertical decline in Ks. In the case of homogenous Ks, lowering the grated flow and transport solution with an ability to explicitly and A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30 27

exactly account for smooth to rapid exponential decline in ks and but was closely related to catchment landscape organization (e.g., porosity with soil depth. This, together with the ability to effi- Hrachowitz et al., 2010). Indeed, this paper is the first modeling ciently generate subsurface flow pathlines and transit times along experiment that we are aware of that explicitly links the shape the pathlines using continuous particle tracking, offers a frame- of the TTD and variability in transit times with subsurface vertical work to systematically assess how the rate of vertical decline in macro heterogeneity by considering a relatively rapid decline in Ks ks and porosity interact with other hydrological controls (location and porosity with soil depth as is common in till mantled catch- of hillslope hydrological boundaries, dispersivity and infiltration ments. The Gamma shape parameters reported in this paper also rate) to influence the groundwater table location, flow pathline showed that as vertical heterogeneity in permeability increased, distribution, TTD and MTT. the behavior of the hillslope approached the fractal behavior. The The current study was designed to first determine the efficacy conclusions made here can also provide a guideline for the of this semi-analytical approach in simulating observed groundwa- required level of complexity in subsurface structure to explicitly ter flow and transport dynamics along a well-studied, till-mantled and efficiently model flow pathlines and TTD. hillslope on the Västrabäcken sub-catchment in Sweden. Our mod- While these results are interesting hydrologically, they may eling results showed that the steady-state semi-analytical solution also open the door for greater insights into how systematic can accurately predict the observed groundwater depth for various changes in runoff regimes related to climate and land use will flow rates within tens of meters from the stream, as well as the influence hydrogeochemistry. Such changes in the MTT and TTD long-term TTD obtained from the convolution approach applied and the prevalence of specific flow paths can greatly impact reac- to the isotopic tracers. The simulated pathlines for the calibrated tive and non-reactive solute concentrations (e.g., Peters et al., model were also consistent with the observation of little annual 2014). Short term variation in flow pathlines have already been change in an isotopic hydrological tracer at 90 cm depth some shown to be a first-order control on the concentration of dissolved 20 m from the stream on the hillslope. organic carbon (Bishop et al., 2004) and landscape organization has The model was then used to examine systematically the influ- been implicated in the more complex patterns of dissolved organic ence of changing the (ks) and porosity patterns with soil depth, carbon (Troch et al., 2013) and nutrient processing (Pinay et al., mechanical dispersion, depth of the no-flow boundary underlying 2015) in the stream network. With the insights from this current the hillslope, hillslope length and infiltration rate for computed modeling study, and long-term predictions of the hydrological MTT, TTD, and flow pathline distribution. Our results showed that regime at a highly resolved spatial and temporal scale an exponential decline in ks with soil depth can significantly affect (Teutschbein et al., 2015), it should be possible to pose stronger the flow pathlines and the variability (relative to MTT) of transit hypotheses which can be tested using the decades of high- times. The results also showed that changes in porosity pattern resolution data that are being developed at long term hydroecolog- can influence the shape of TTD and MTT, but these impacts are ical research sites (like the Krycklan Catchment Study). highly dependent upon the rate of Ks change with soil depth within the hillslope. While the impacts of systematic ‘‘macro scale” sub- 4.1. Need for future research surface vertical heterogeneity (i.e., ks and porosity changes with soil depth) on the structure of TTD was considerable, pore scale The semi-analytical solution used in this paper was a useful and heterogeneity (represented in this paper by mechanical dispersiv- appropriate tool for simulating flow and transport in a till environ- ity) had only a slight impact on subsurface flow pathlines, TTD and ments as well as for systematically assessing the impact of vertical MTT, particularly as the rate of vertical decline in Ks increased. This permeability architecture on time-invariant TTD and flow path- is consistent with the findings by Cardenas and Jiang (2010) and lines. However, the steady state condition is still a necessary Fiori and Russo (2008). assumption for most (semi)analytical solutions, including the one Although both the infiltration rate and the pattern of vertical used in this paper. TTDs are by nature time variant (Harman, decline in Ks pronouncedly impact transit times, water table loca- 2015; Klaus et al., 2015), and vary with precipitation regime tion, MTT and the shape of the regular (non-dimensionless) TTD, (Sayama and McDonnell, 2009) as well as wetness conditions the infiltration rate has a negligible impact on the ‘‘variability” of (Heidbüchel et al., 2013; Tetzlaff et al., 2014). While the time transit times relative to MTT (and the shape of the dimensionless invariant TTD may still be valid for wet conditions with little sea- TTD). The results also suggest that the location of hydrological sonality and/or if one focuses on long term behavior of the catch- boundaries including water divide and no-flow boundary underly- ment (see also Botter et al., 2010), most systems have time ing the hillslope have negligible effects on the structure of dimen- varying TTD. This can underestimate the proportion of early arrival sionless TTD and the variability of transit times as the rate of waters in humid catchments (Botter, 2012; Hrachowitz et al., vertical decline in ks increases. This may imply that accurate deter- 2010). A time variant integrated subsurface flow and particle mination of the location of the no-flow boundary underlying the movement approach would be desirable for an explicit simulation hillslope in subsurface flow and transport models is not necessary of transient subsurface flow pathlines and transit times. This can with regard to TTD delineation for the cases with extreme vertical be accomplished through application of robust numerical transient exponential decline in Ks. This may become less true as the degree subsurface flow and transport solutions, or by adding a Laplace of vertical exponential decline in Ks decreases or other Ks-depth Transform simulator (in a manner similar to Bakker, 2013) to the functions govern subsurface structure. For example, for the present semi-analytical solution (something we plan to explore homogenous case, the location of hydrological boundaries has a in the future). Notwithstanding, the steady-state assumption still larger effect on the TTD particularly the changes in the location seems to be valid for modeling flow pathlines and transit times of the no-flow boundary underlying the hillslope. A longer hillslope in the riparian portion of the hillslope studied in this paper; also increases the MTT for homogenous ks but decreases the MTT although the validity of the steady state assumption weakens fur- for the extreme vertical decline in ks. ther away from the water course as one moves closer to the sub- Our model results obtained through virtual experiments are catchment divide. consistent with observations from experimental studies that Another factor that should be examined in future work is the showed that the shape parameter of Gamma distribution, which lateral variation in the pattern of ks exponential decline. In this describes the variability of time-invariant transit times relative paper we ignored this feature and only considered vertical perme- to MTT, had no relationship with precipitation intensity variations ability patterns. The consistent underestimation of groundwater 28 A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30

depths at the water divide might be related in part to a different ks Appendix A. Semi-analytical subsurface flow solution profiles there compared to that closer to the stream where we had better data to parameterize the ks depth profile. More importantly, Ameli et al. (2016) have shown that the series solution to the the impact of large-scale variation in lateral permeability on TTD saturated flow governing equation with no-flow conditions along and flow pathlines may be significant and should be assessed sys- the sides of the saturated domain and exponentially depth tematically. This assessment can be accomplished in the future by decaying saturated hydraulic conductivity with soil depth

aðzZt Þ the implementation of statistically driven large-scale lateral (Ks ¼ Ks0e ) can be obtained in terms of discharge potential heterogeneity in the permeability. / ð ; Þ function ( s x z ) as: Finally, topographic controls such as topographic convergences hi hi XN np np can also influence TTD and flow pathlines. A 2-D representation, as / ðx; zÞ¼A þ A cos x expðc zÞ þ B cos x expðc zÞ s 0 n L n n L n adopted in this study seems appropriate for the slope studied here n¼1 since it was characterized by a fairly uniform geometry (width and ðA:1Þ slope). However, a study of other hillslopes with more variable sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi geometries showed links between terrain convergence and flow   a 1 2np 2 a 1 2np 2 paths in the riparian zone (Grabs et al., 2012). To apply the semi- c ¼ þ a2 þ ; c ¼ a2 þ n 2 2 L n 2 2 L analytical model to such cases one would thus have to overcome some of the limitations inherent in a 2-D hillslope representation. / ð ; Þ¼ ð ; Þ where s x z Ks0hs x z . This could be achieved potentially by using 3-D semi-analytical The series solution for the unsaturated moisture movement model (e.g., Ameli and Craig, 2014), or a numerical subsurface flow with exponentially depth decaying saturated hydraulic conductiv- and transport model, as well as by combining the semi-analytical ity with soil depth can be calculated in terms of Kirchhoff potential model with mathematical approximations of hillslope geometries / ( u) as: (Troch et al., 2002).  hi XM mp £ / ðx; zÞ¼C ½expðbzÞ C cos x expð£ zÞ m u 0 m L m mp 5. Conclusion m¼1 L hip  þ m ð Þ £m ð : Þ Dm cos x exp £mz mp A 2 A semi-analytical series solution model was developed to simu- L L late subsurface flow and particle movement in the well-studied S- sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  transect hillslope on the Västrabäcken sub-catchment in Sweden. ða þ bÞ p 2 1 2 2n The steady state model emulated the observed groundwater £ ¼ þ ða þ bÞ 4ab þ ; £ m 2 2 L m depth–discharge relationship within tens of meters of the stream. sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  The integrated model also reasonably simulated the hillslope TTD. 2 ða þ bÞ 1 2 2np The results also suggested that the macro scale vertical hetero- ¼ ða þ bÞ 4ab þ 2 2 L geneity in subsurface permeability including exponential decline ðbðuueÞÞ in saturated hydraulic conductivity and porosity with soil depth / ðuÞ¼Ks0 exp where u b . significantly impact the structure of TTD and the variability of tran- In the above equations hs [L] is the total hydraulic head, u is sit times (relative to MTT) within the hillslope. The exponential 1 suction pressure head [L], L is aquifer length, Ks0 [L T ] is the sat- decline in saturated hydraulic conductivity with depth also urated hydraulic conductivity along the topographic surface ZtðxÞ; impacts the flow path distribution with shallower flow circulation a is the parameter of the exponential relationship between satu- as the strength of exponential decline in K increases. In contrast, s rated hydraulic conductivity with soil depth, b and ue are the sorp- subsurface pore scale micro heterogeneity (mechanical dispersion) tive number and air entry pressure of the Gardner’s constitutive only slightly influenced the variability of transit times and pathline function (Gardner, 1958) used to characterize the suction- distribution. The impact of infiltration rate was also negligible on hydraulic conductivity relationship in the vadose zone. In Eqs. the ‘‘variability” of transit times relative to MTT, while an increase (A.1) and (A.2), additionally, n and m denote the coefficient index, in this rate significantly decreased the transit times and MTT in the and An, Bn, Cm, Dm are the unknown series coefficients associated hillslope. The location of the hydrological boundaries at the water with the nth and mth coefficient index, respectively. N and M also divide and the no-flow boundary underlying the hillslope also refer to the total number of terms in the series solutions to the sat- influenced the TTD but only slightly when the rate of exponential urated and unsaturated flow governing equations, respectively. vertical decline in Ks was large. Location of these boundaries had The unknown series solution coefficients (An, Bn and Cm, Dm) for a somewhat larger effect in more homogeneous soils on the TTD. Eqs. (A.1) and (A.2) were calculated by enforcing the boundary con- Our findings provide useful guidelines for understanding the ditions at the top and the bottom of saturated and unsaturated required level of complexity in subsurface structure to explicitly zones. These boundary conditions were imposed using a simple model time invariant flow pathlines and TTD. Future work is least square scheme at control points located along each interface. needed to explicitly simulate ‘‘time variant” subsurface flow path- These boundary conditions include Neumann boundary conditions lines and transit time distributions on hillslopes with more com- (infiltration rate) along the top of the unsaturated domain (i.e. land plex topography where the soil properties can change with surface), constant head (air entry pressure) along the unsaturated distance from the stream. domain bottom (i.e. the a priori unknown top of the capillary fringe zone interface), continuity of flux between saturated and unsatu- Acknowledgement rated zones along the top of the saturated domain (again the a pri- ori unknown top of the capillary fringe zone interface), no-flow We thank James Craig and Jan Seibert for their support through- condition along the bottom boundary of the computational domain out the process. This research was funded by NSERC Discovery and constant head at the stream. The a priori unknown location of Grant and NSERC Accelerator to J.J.M, NSERC Discovery Grant to top of capillary fringe was also obtained using an iterative scheme I.F.C. The Krycklan catchment study is supported by the Swedish between the saturated and unsaturated solutions. The water table

Science Foundation (VR) SITES, ForWater (Formas), Future Forest, location (Zwt(x)inFig. 1d) was then calculated as the boundary Kempe Foundation, SLU FOMA and SKB. with a zero pressure head. We refer the readers Ameli et al. A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30 29

(2016) for a detailed discussion of the mathematical formulation of Kirchner, J.W., Feng, X., Neal, C., 2000. Fractal stream chemistry and its implications series solution method, boundary conditions, least square scheme for contaminant transport in catchments. Nature 403 (6769), 524–527. Kirchner, J.W., Feng, X., Neal, C., 2001. Catchment-scale advection and dispersion as used to enforce boundary conditions and iterative algorithm used a mechanism for fractal scaling in stream tracer concentrations. J. Hydrol. 254 to determine the a priori unknown location of water table. (1), 82–101. Kirchner, J., 2016. Aggregation in environmental systems–Part 1: Seasonal tracer cycles quantify young water fractions, but not mean transit times, in spatially References heterogeneous catchments. Hydrol. Earth Syst. Sci 20, 279–297. Klaminder, J. et al., 2006. Flux rates of atmospheric lead pollution within soils of a Ali, M., Fiori, A., Russo, D., 2014. A comparison of travel-time based catchment small catchment in northern Sweden and their implications for future stream transport models, with application to numerical experiments. J. Hydrol. 511, water quality. Environ. Sci. Technol. 40 (15), 4639–4645. 605–618. Klaus, J., Chun, K.P., McGuire, K.J., McDonnell, J.J., 2015. Temporal dynamics of Ameli, A.A., Craig, J.R., 2014. Semianalytical series solutions for three-dimensional catchment transit times from stable isotope data. Water Resour. Res. 51 (6). groundwater-surface water interaction. Water Resour. Res. 50 (5). Kollet, S.J., Maxwell, R.M., 2008. Demonstrating fractal scaling of baseflow residence Ameli, A.A., Craig, J.R., Wong, S., 2013. Series solutions for saturated–unsaturated time distributions using a fully-coupled groundwater and land surface model. flow in multi-layer unconfined aquifers. Adv. Water Resour. 60, 24–33. Geophys. Res. Lett. 35 (7). Ameli, A.A., McDonnell, J.J., Bishop, K., 2016. The exponential decline in saturated Laudon, H., Seibert, J., Köhler, S., Bishop, K., 2004. Hydrological flow paths during hydraulic conductivity with depth and its effect on water flow paths and transit snowmelt: congruence between hydrometric measurements and oxygen 18 in time distribution. Hydrol. Process. http://dx.doi.org/10.1002/hyp.10777. meltwater, soil water, and runoff. Water Resour. Res. 40 (3). Bakker, M., 2013. Semi-analytic modeling of transient multi-layer flow with TTim. Laudon, H. et al., 2013. The Krycklan Catchment Study—a flagship infrastructure for Hydrogeol. J. 21 (4), 935–943. http://dx.doi.org/10.1007/s10040-013-0975-2. hydrology, biogeochemistry, and climate research in the boreal landscape. Basu, N.B., Jindal, P., Schilling, K.E., Wolter, C.F., Takle, E.S., 2012. Evaluation of Water Resour. Res. 49 (10), 7154–7158. analytical and numerical approaches for the estimation of groundwater travel Leith, F. et al., 2014. Carbon dioxide transport across the hillslope-riparian-stream time distribution. J. Hydrol. 475, 65–73. continuum in a boreal headwater catchment. Biogeosci. Discuss. 11 (11), Birkel, C., Tetzlaff, D., Dunn, S.M., Soulsby, C., 2011. Using lumped conceptual 15585–15619. rainfall–runoff models to simulate daily isotope variability with fractionation in Lundin, L., 1982. Soil moisture and ground water in till soil and the significance of a nested mesoscale catchment. Adv. Water Resour. 34 (3), 383–394. soil type for runoff.—UNGI Rapport Nr 56, Uppsala, 216 pp. Swedish, English Bishop, K., Lee, Y.-H., Pettersson, C., Allard, B., 1995. Terrestrial sources of summary. methylmercury in surface waters: the importance of the riparian zone on the Maxwell, R.M. et al., 2015. The imprint of climate and geology on the residence Svartberget catchment. Water Air Soil Pollut. 80 (1–4), 435–444. times of groundwater. Geophys. Res. Lett 43 (2). Bishop, K., Seibert, J., Köhler, S., Laudon, H., 2004. Resolving the double paradox of McDonnell, J.J., Beven, K., 2014. Debates—the future of hydrological sciences: a rapidly mobilized old water with highly variable responses in runoff chemistry. (common) path forward? A call to action aimed at understanding velocities, Hydrol. Process. 18 (1), 185–189. http://dx.doi.org/10.1002/hyp.5209. celerities and residence time distributions of the headwater hydrograph. Water Bishop, K.H., 1991. Episodic Increases in Stream Acidity, Catchment Flow Pathways Resour. Res. 50 (6), 5342–5350. and Hydrograph Separation. University of Cambridge. McGuire, K.J., McDonnell, J.J., 2010. Hydrological connectivity of hillslopes and Botter, G., 2012. Catchment mixing processes and travel time distributions. Water : characteristic time scales and nonlinearities. Water Resour. Res. 46 Resour. Res. 48 (5). (10). http://dx.doi.org/10.1029/2010wr009341. Botter, G., Bertuzzo, E., Rinaldo, A., 2010. Transport in the hydrologic response: Molénat, J., Gascuel-Odoux, C., Aquilina, L., Ruiz, L., 2013. Use of gaseous tracers travel time distributions, soil moisture dynamics, and the old water paradox. (CFCs and SF 6) and transit-time distribution spectrum to validate a shallow Water Resour. Res. 46 (3). groundwater transport model. J. Hydrol. 480, 1–9. Botter, G., Bertuzzo, E., Rinaldo, A., 2011. Catchment residence and travel time Nyberg, L., 1995. Water flow path interactions with soil hydraulic properties in till distributions: the master equation. Geophys. Res. Lett. 38 (11). soil at Gårdsjön, Sweden. J. Hydrol. 170 (1), 255–275. Cardenas, M.B., Jiang, X.-W., 2010. Groundwater flow, transport, and residence Nyberg, L., Stähli, M., Mellander, P.E., Bishop, K.H., 2001. Soil frost effects on soil times through topography-driven basins with exponentially decreasing water and runoff dynamics along a boreal forest transect: 1. Field permeability and porosity. Water Resour. Res. 46 (11). investigations. Hydrol. Process. 15 (6), 909–926. Cory, N., Laudon, H., Köhler, S., Seibert, J., Bishop, K., 2007. Evolution of soil solution Peralta-Tapia, A. et al., 2016. Hydroclimatic influences on non-stationary transit aluminum during transport along a forested boreal hillslope. J. Geophys. Res.: time distributions in a boreal headwater catchment. J. Hydrol. 543 (PA), 7–16. Biogeosci. (2005–2012) 112 (G3). Peralta-Tapia, A., Sponseller, R., Tetzlaff, D., Soulsby, C., Laudon, H., 2014. Eklöf, K. et al., 2014. Impact of forestry on total and methyl-mercury in surface Connecting precipitation inputs and soil flow pathways to stream water in waters: distinguishing effects of logging and site preparation. Environ. Sci. contrasting boreal catchments. Hydrol. Process. 29 (16). Technol. 48 (9), 4690–4698. Peters, N.E., Burns, D.A., Aulenbach, B.T., 2014. Evaluation of high-frequency mean Fiori, A., Russo, D., 2008. Travel time distribution in a hillslope: insight from streamwater transit-time estimates using groundwater age and dissolved silica numerical simulations. Water Resour. Res. 44 (12). concentrations in a small forested watershed. Aquat. Geochem. 20 (2–3), 183– Fiori, A., Russo, D., Di Lazzaro, M., 2009. Stochastic analysis of transport in 202. hillslopes: travel time distribution and source zone dispersion. Water Resour. Pinay, G. et al., 2015. Upscaling nitrogen removal capacity from local hotspots to Res. 45 (8). low stream orders’ drainage basins. Ecosystems, 1–20. Gardner, W., 1958. Some steady-state solutions of the unsaturated moisture flow Rinaldo, A. et al., 2015. Storage selection functions: a coherent framework for equation with application to evaporation from a water table. Soil Sci. 85 (4), quantifying how catchments store and release water and solutes. Water Resour. 228–232. Res. 51 (6), 4840–4847. Godsey, S.E. et al., 2010. Generality of fractal 1/f scaling in catchment tracer time Rinaldo, A. et al., 2011. Catchment travel time distributions and water flow in soils. series, and its implications for catchment travel time distributions. Hydrol. Water Resour. Res. 47 (7). Process. 24 (12), 1660–1671. Salamon, P., Fernàndez-Garcia, D., Gómez-Hernández, J.J., 2006. A review and Goode, D.J., 1996. Direct simulation of groundwater age. Water Resour. Res. 32 (2), numerical assessment of the random walk particle tracking method. J. Contam. 289–296. Hydrol. 87 (3), 277–305. Grabs, T., Bishop, K., Laudon, H., Lyon, S.W., Seibert, J., 2012. Riparian zone Sayama, T., McDonnell, J.J., 2009. A new time-space accounting scheme to predict hydrology and soil water total organic carbon (TOC): implications for spatial stream water residence time and hydrograph source components at the variability and upscaling of lateral riparian TOC exports. Biogeosciences 9 (10), watershed scale. Water Resour. Res. 45 (7). 3901–3916. Seibert, J., 2000. Multi-criteria calibration of a conceptual runoff model using a Grip, H., 2015. Sweden’s first forest hydrology field study 1905–1926: genetic algorithm. Hydrol. Earth Syst. Sci. Discuss. 4 (2), 215–224. contemporary relevance of inherited conclusions and data from the Rokliden Seibert, J., Bishop, K., Nyberg, L., Rodhe, A., 2011. Water storage in a till catchment. I: Hillslope. Hydrol. Process 29 (16). Distributed modelling and relationship to runoff. Hydrol. Process. 25 (25), Harman, C.J., 2015. Time-variable transit time distributions and transport: theory 3937–3949. and application to storage-dependent transport of chloride in a watershed. Seibert, J., Bishop, K., Rodhe, A., McDonnell, J.J., 2003. Groundwater dynamics along Water Resour. Res. 51 (1), 1–30. a hillslope: a test of the steady state hypothesis. Water Resour. Res. 39 (1). Harr, R.D., 1977. Water flux in soil and subsoil on a steep forested slope. J. Hydrol. http://dx.doi.org/10.1029/2002wr001404. 33 (1), 37–58. Seibert, J. et al., 2009. Linking soil-and stream-water chemistry based on a Riparian Heidbüchel, I., Troch, P.A., Lyon, S.W., 2013. Separating physical and meteorological flow-concentration integration model. Hydrol. Earth Syst. Sci. 13 (12), 2287– controls of variable transit times in zero-order catchments. Water Resour. Res. 2297. 49 (11), 7644–7657. Stähli, M., Nyberg, L., Mellander, P.E., Jansson, P.E., Bishop, K.H., 2001. Soil frost Hrachowitz, M. et al., 2009. Using long-term data sets to understand transit times in effects on soil water and runoff dynamics along a boreal transect: 2. contrasting headwater catchments. J. Hydrol. 367 (3), 237–248. Simulations. Hydrol. Process. 15 (6), 927–941. Hrachowitz, M., Soulsby, C., Tetzlaff, D., Malcolm, I., Schoups, G., 2010. Gamma Tetzlaff, D., Birkel, C., Dick, J., Geris, J., Soulsby, C., 2014. Storage dynamics in distribution models for transit time estimation in catchments: physical hydropedological units control hillslope connectivity, runoff generation, and interpretation of parameters and implications for time-variant transit time the evolution of catchment transit time distributions. Water Resour. Res. 50 (2), assessment. Water Resour. Res. 46 (10). 969–985. 30 A.A. Ameli et al. / Journal of Hydrology 543 (2016) 17–30

Teutschbein, C., Grabs, T., Karlsen, R., Laudon, H., Bishop, K., 2015. Hydrological Vaché, K.B., McDonnell, J.J., 2006. A process-based rejectionist framework for response to changing climate conditions: spatial streamflow variability in the evaluating catchment runoff model structure. Water Resour. Res. 42 (2). boreal region. Water Resour. Res. 51 (12), 9425–9446. Weiler, M., McDonnell, J., 2004. Virtual experiments: a new approach for improving Troch, P., Van Loon, E., Hilberts, A., 2002. Analytical solutions to a hillslope-storage process conceptualization in hillslope hydrology. J. Hydrol. 285 (1–4), 3–18. kinematic wave equation for subsurface flow. Adv. Water Resour. 25 (6), 637– http://dx.doi.org/10.1016/s0022-1694(03)00271-3. 649. Woolfenden, L.R., Ginn, T.R., 2009. Modeled ground water age distributions. Troch, P.A. et al., 2013. The importance of hydraulic groundwater theory in Groundwater 47 (4), 547–557. : the legacy of Wilfried Brutsaert and Jean-Yves Parlange. Water Resour. Res. 49 (9), 5099–5116.