Open access
Technical Papers
Mar 27, 2017

Hydrodynamic Modeling of the St. Lawrence Fluvial Estuary. II: Reproduction of Spatial and Temporal Patterns

Publication: Journal of Waterway, Port, Coastal, and Ocean Engineering
Volume 143, Issue 5

Abstract

This is the second part of an investigation aimed at documenting the tidal hydrodynamics of the St. Lawrence fluvial estuary (SLFE). In Part I, the calibration and validation of a high-resolution, two-dimensional (2D), time-dependent hydrodynamic model of the SLFE was presented. Herein, the process-based (structural) validation procedure used to quantitatively assess the ability of the model to reproduce spatial and temporal patterns observed in the field data is presented. Tidal and flow features were found to be reproduced satisfactorily in terms of their lateral and longitudinal variability at both the intratidal and fortnightly scales. These properties can be used to describe the real system dynamics, including flooding–drying processes, tidal propagation and modulation, and transient momentum balance, providing insights into the general physical processes of the SLFE and of large tidal rivers globally.

Introduction

In the study of tidal rivers, the use of high-resolution hydrodynamic models is essential to understanding the complex interaction between tides, river flow, and geometry. Such tools enable reliable and sustainable ecosystem monitoring and are a prerequisite for predicting the likely consequences of management practices, anthropogenic changes, and natural events on a system [e.g., Araújo et al. (2008); Babu et al. (2005); Horsburgh and Wilson (2007); Morin and Champoux (2006); Picado et al. (2010)]. Furthermore, hydrodynamic models can be coupled or used as input to morphodynamic models (Nabi et al. 2012; Rinaldi et al. 2008), hydrologic models (de Paiva et al. 2013), advection–diffusion models (Bárcena et al. 2012; Simons et al. 2006), wave–current interaction models (Liu et al. 2007), and aquatic habitat models (Morin et al. 2003), among others.
In Part I (Matte et al. 2017), a high-resolution, two-dimensional (2D), time-dependent hydrodynamic model of the St. Lawrence fluvial estuary (SLFE) was presented, which includes a drying–wetting component that allows water in intertidal areas to be cyclically stored and evacuated. The numerical terrain model (NTM) is based on high-density topographic data stemming from light detection and ranging (LIDAR) surveys and multibeam bathymetric soundings. Friction fields are defined according to bottom substrate composition and macrophytes, whose coefficients were adjusted during calibration. A finite-element mesh with an average spatial resolution of 50 m was designed, far denser than that of previous/existing models, with refinements for regions of complex topography. The model was proven to be replicatively valid in that it matches the field data used during the calibration phase with high accuracy. Demonstration of model replicative validity was performed based on a comparison of modeled and observed water levels, mean water levels (MWLs), tidal ranges, and harmonic properties at 29 tide gauges and discharges at 13 transects during two simulation periods in June and August 2009 for which detailed data were available.
In this second part, a process-based validation is presented that exploits tide gauge observations and cross-sectional water-level and velocity data collected in the SLFE (Matte et al. 2014a). The focus of this process-based validation is oriented toward the reproduction of temporal and spatial patterns observed in the field data, along with quantities extracted from the reconstructed fields. These properties can be used to describe the real system dynamics, in terms of its lateral and longitudinal variability, at both the intratidal and fortnightly scales, providing new insights into the tidal hydrodynamics and circulation complexities of the SLFE.

Data and Methods

Data

Detailed hydrodynamic data were collected in the SLFE during a field campaign conducted in the summer of 2009, as summarized in Part I. Data from an extended tide gauge network composed of 29 stations were obtained, along with water-level and velocity data simultaneously acquired along 13 transects [cf. Matte et al. (2017), Fig. 1], each repeatedly surveyed by boat during approximately one tidal cycle. At each measurement transect, the irregular mixed space–time data series were interpolated in space and in time, following a procedure developed by Matte et al. (2014b), to allow reconstruction of continuous and synoptic fields and facilitate comparison with model results. Details of the data processing and interpolation errors have been provided by Matte et al. (2014b, c).
Fig. 1. (Color) Water levels (in meters) at Grondines on June 19, 2009, as a function of cross-sectional distance (or bathymetry) and time: (a) observed; (b) modeled (Note: Black dots represent the gridded data points along the boat track; time reference is EDT)

Model Validation

In Part I, a replicative validation was performed, focusing on a comparison of observed and modeled water levels at the tide gauges and discharges at the transects. In this paper, a process-based validation is presented as a demonstration of the model’s capability to reproduce the real system dynamics. The same 15-day simulation periods starting on June 14, 2009, 00:00:00 EDT, and August 19, 2009, 00:00:00 EDT, where transect data were acquired were used for this purpose. Measured and simulated data were used for direct comparisons, whereas quantities representative of the tidal dynamics were extracted from the observed and modeled fields and compared both qualitatively and quantitatively. The validation process involved the following steps:
1.
Lateral water level and velocity fields reconstructed over a tidal cycle were compared with model results at a selected transect to validate their general spatiotemporal patterns.
2.
A series of variables extracted from the water level and velocity fields was used to quantify lateral and intratidal differences at the surveyed transect [namely, timing and height of low water (LW) and high water (HW), tidal range (TR), maximum velocity, timing of slack water, and inclination of velocity vectors].
3.
Observed and modeled water-level time series were compared at the tide gauges for the two simulation periods and used to investigate tidal damping and distortion.
4.
Tidal range and tidal datum levels [i.e., higher high water (HHW), MWL, and lower low water (LLW)] were extracted for conditions of neap, mean, and spring tides both at tide stations and along the thalweg. They were used to validate the longitudinal displacement of the tidal amplitude maximum and the fortnightly modulation of tidal datum levels, including the neap–spring reversal of LLWs.
5.
Longitudinal and fortnightly variations in ebb and flood velocities and in the position of the upstream limit of current reversal were assessed based on modeled tidal velocities, extracted for conditions of neap, mean, and spring tides along the thalweg.
6.
The flow distribution between the two arms of Orleans Island was validated with tidal discharge measurements at two transects. The velocity patterns at the island junction were analyzed based on model results over a tidal cycle.
7.
Transient momentum balance was characterized based on a snapshot taken during spring tide, encompassing the various features observed during a tidal cycle (i.e., slack currents, peak ebb and flood currents).

Results

Lateral and Intratidal Variability

The model was validated by comparison with water-level and velocity measurements at the surveyed transects. Results are presented in Figs. 14 for a selected river cross section located at Grondines [river kilometer (rkm) 179.5].
Fig. 2. (Color) Comparison between observed and modeled variables extracted from the water-level fields at Grondines on June 19, 2009, as a function of cross-sectional distance (or bathymetry): times and heights of LW, times and heights of HW, and tidal range
Fig. 3. (Color) Depth-averaged velocity components (in meters per second) at Grondines on June 19, 2009, as a function of cross-sectional distance (or bathymetry) and time: (a) observed u velocities; (b) modeled u velocities; (c) observed v velocities; (d) modeled v velocities (Note: Black dots represent the gridded data points along the boat track; white contours identify the slack waters; time reference is EDT)
Fig. 4. (Color) Comparison between observed and modeled variables extracted from the velocity fields at Grondines on June 19, 2009, as a function of cross-sectional distance (or bathymetry): maximum absolute u and v velocities, times of slack water, and inclination of the tidal ellipse
Observed and modeled water levels are shown in Fig. 1. The agreement between the two was found to be good, both in terms of synchronicity of the signals and reproducibility of the lateral patterns. Of particular interest is the lateral gradient that formed during the falling tide between 1.0 and 1.7 km from the south shore, which is responsible for the emptying of the tidal flats into the channel. Variables extracted from the water-level fields are shown in Fig. 2, allowing a quantitative assessment of model performance at the transect. The timing and heights of HW and LW and tidal range are plotted as a function of cross-sectional distance. LW occurred slightly (∼1–5 min) earlier in the model data than in the observed data, with a later arrival and a higher level on the southern tidal flat (first kilometer). Modeled LW was also ∼10 cm higher than expected, but it exhibited lateral gradients similar to those in the observations. The timing of HW in the model was relatively similar to that in the observations but showed less lateral variability. The modeled heights of HW were also higher than the observed heights, with the differences being smaller than those for LW (<10 cm); they remained almost unchanged across the section. Overall, the LW is more sensitive than HW to lateral gradients in bathymetry. As a result, the tidal ranges were larger in the channel than on the shoal at Grondines. They were lower in the model compared with observations, but they followed a similar trend. Slightly decreasing friction near Grondines would potentially reduce the local predicted tidal heights and increase tidal range. However, the observed differences mostly remain within measurements errors (Matte et al. 2014b). They can also be partly explained by differences between the bathymetry measured along the section and that obtained from the Canadian Hydrographic Service (CHS) soundings; they are referred to as observed and modeled bathymetries in bottom panel of Fig. 2, the latter being the one actually incorporated in the NTM.
Depth-averaged u and v velocity components are shown in Fig. 3, corresponding to the along- and cross-channel velocities, respectively. Current reversals occurred only on the north shore during the surveyed period (identified as a white contour in Fig. 3). The highest u velocities were concentrated in the channel, whereas v velocities presented two distinct regions characterized by large positive values (oriented to the north). The southernmost region is the result of the tidal flats emptying into the channel, in concordance with the observed lateral gradients in water levels (cf. Fig. 1). The second region is in the channel and likely corresponds to the effects of local channel curvature on the velocity directions. These general flow features were adequately reproduced by the model, suggesting that the lateral variations in topography and friction were well captured. Variables extracted from the velocity fields are presented in Fig. 4. The modeled maximum u and v velocities were found to have patterns very similar to those of the observations, although they were slightly higher in the tidal flat, between 1.0 and 1.7 km. Part of these discrepancies may be explained by interpolation errors in the observed data, which amount to approximately 0.1 m/s at Grondines (Matte et al. 2014b). Conversely, the observed times of slack water in both the ebb-to-flood and flood-to-ebb transitions [i.e. the low-water slack (LWS) and high-water slack (HWS)] were found to be relatively well synchronized with the model. Current reversals, however, occurred within a shorter distance from the north shore in the model. Finally, the inclination of the tidal ellipse formed by the velocity vector over a tidal cycle (a measure of the relative strength of the lateral and longitudinal velocity components) presented generally small differences between the model and observations. The highest inclinations were found in the southern tidal flat, where lateral exchanges are maximal.
Although comparisons are herein limited to one cross section, detailed results for all surveyed transects have been given by Matte (2014, Appendix 3).

Longitudinal and Neap–Spring Variability

An appreciation of the longitudinal variations in tidal amplitude, signal distortion, and neap–spring variability in the system can be gained by comparing the observed and modeled water levels for the two simulation periods of June and August 2009 (Fig. 5). From downstream to upstream, the tidal wave is progressively damped and distorted, with significantly shorter rising than falling tides. Generally, the observed water levels were well reproduced by the model at the stations during both neap and spring tides. Fortnightly variations in MWL were also well captured at upstream stations [e.g., Trois-Rivières (rkm 231) and Bécancour (rkm 217)]. However, the differences between observed and modeled water levels at Bécancour were larger in August than June, with MWL being lower than expected in August (cf. Matte et al. 2017, Table 3). This points toward a gradual underestimation of friction in the Gentilly shoal, located 2 km downstream of Bécancour, following the growth of macrophytes during summer. The same friction field (cf. Matte et al. 2017, Fig. 3) was used during both simulation periods, but time-varying friction based on density and growth phase factors could be implemented to take this effect into account (Morin et al. 2000).
Fig. 5. (Color) Observed (solid lines) and modeled (dotted lines) water levels from the global model at six permanent tide gauges for the two simulation periods: (a) June 14–29, 2009; (b) August 19–September 3, 2009
The observed and modeled signals at the stations were used to extract information on tidal range, HHW, MWL, and LLW. They are presented in Fig. 6 and compared with modeled longitudinal profiles aligned with the thalweg for the simulation of August 2009. Tidal ranges were significantly reduced landward, exceeding 6 m downstream during spring tides and almost vanishing upstream during neap tides. Most of the damping occurred between Portneuf (rkm 163.5) and Cap-à-la-Roche (rkm 186) as a result of the presence of rapids associated with a sharp increase of the bottom slope. Over the neap–spring cycle, modulations of the tidal ranges were significantly more pronounced at downstream stations, with differences in tidal range between spring and neap tides exceeding 3 m at Saint-Joseph-de-la-Rive (rkm 0). More specifically, during the analyzed period, observed tidal ranges were the largest in the St. Lawrence at Saint-Joseph-de-la-Rive (rkm 0) during spring tides, whereas during neap tides, they reached their maximum at Saint-François, about 66 rkm upstream. In other words, tidal amplification occurring in the estuary persists further landward during neap tides than spring tides. Overall, these features were well captured by the model, as shown by the proximity of the observed and modeled data (represented by circles and dots in Fig. 6, respectively). However, the model tended to push the amplitude maximum upstream toward Lauzon (rkm 100) during neap tides, possibly as a result of underestimated friction between the Orleans Island and the downstream boundary. Differences between the station data and the longitudinal profiles are an indication of the lateral variability in tidal ranges throughout the system. Smaller tidal ranges near the shore than in the channel were generally observed as the extent of tidal flats increased, for example, between Deschambault (rkm 168) and Cap-à-la-Roche (rkm 186), as also shown in Fig. 2. These smaller tidal ranges correlated with larger MWL on the shore.
Fig. 6. (Color) Longitudinal and neap–spring variability in water levels—modeled longitudinal profiles aligned with the thalweg (solid lines), along with observed (circles) and modeled (dots) values at the stations under conditions of spring (blue), mean (red), and neap (yellow) tides for the simulation period of August 19 through September 3, 2009: (a) TR; (b) HHW; (c) MWL; (d) LLW
Both HHW and MWL were systematically higher on spring tides than on neap or mean tides. Longitudinal variations of HHW were, however, smoother (almost linear landward of the Quebec Bridge; rkm 115) than MWL and LLW because they are less affected by longitudinal changes in bottom slope. In contrast, at the neap–spring scale, HHWs were quite variable throughout the system, with a range of variability decreasing landward. Downstream, differences in HHW between spring and neap tides were similar to those in LLW, as they mostly arise from the amplitude difference between M2 and S2. LLWs were higher on neap tides than on spring tides, but they presented almost identical heights over roughly 115 rkm during neap tides, from Saint-Joseph-de-la-Rive (rkm 0) to the Quebec Bridge (rkm 115). This follows the amplification of the tidal ranges only occurring between these stations during neap tides. The neap–spring modulations of MWL gradually increased from Saint-Joseph-de-la-Rive (rkm 0) landward, reaching a maximum between Portneuf (rkm 163.5) and Cap-à-la-Roche (rkm 186). Upstream, even though the tidal range became weak, the MWL still varied by more than 0.5 m over the neap–spring cycle as a result of the nonlinear growth of low-frequency compound tides (e.g., MSf), thus exceeding the semidiurnal tide in amplitude. As a result, both HHW and LLW presented larger variations than what tidal range alone could explain. Furthermore, because the MWL was much lower on neap tides, the LLW also occurs on neap tides at upstream locations. This shift in the relative levels of LLW between neap and spring tides typically occurs around Deschambault (rkm 168), where the amplitude of the fortnightly tidal component (MSf) is equal to that of S2 (Hoitink and Jay 2016). However, it could appear sooner or later (in the rkm sense) depending on tidal and river flow conditions. In fact, in the present example, this transition seemed to occur earlier (around Neuville; rkm 138), although this may be a result of the rather short simulation period used to extract the tidal datum levels, comprising only one neap–spring cycle.
Lateral differences between the near-shore stations and the channel profiles were the most significant with HHW, especially during neap tides. Similarly, the discrepancies between observed and modeled HHW were the largest among all tidal datum levels. These differences could be attributable to lateral gradients in topography or friction, but further investigation is needed to identify the exact source of variability.
Tidal velocities were also characterized by strong longitudinal and neap–spring variability (Fig. 7). Maximum ebb velocities generally occur during the falling tide, slightly before low tide. They were the strongest around Íle aux Coudres (cf. Matte et al. 2017, Fig. 1) and under the Québec Bridge (rkm 115) and Richelieu rapids near Deschambault (rkm 168), where they exceeded 3 m/s under spring-tide conditions. Downstream, ebb velocities during spring tides were approximately twice as strong as during neap tides. This neap–spring variability was reduced moving upstream, up to Deschambault, landward of which the tidal range had practically no effect on peak ebb velocities, as they were mostly controlled by the river discharge. In contrast, peak flood velocities exhibited marked variations between neap- and spring-tide conditions throughout the system. They were the highest at the Québec Bridge constriction, the narrowest and deepest section of the St. Lawrence, but not so strong at Deschambault compared with its respective ebb velocities. This is attributable to the rapid widening of the Deschambault cross section on the rising tide, which is associated with large intertidal flats, proportionally reducing the flood velocities as they get flooded.
Fig. 7. (Color) Longitudinal and neap–spring variability in velocities—modeled longitudinal profiles under conditions of spring (blue), mean (red), and neap (yellow) tides for the simulation period of August 19 through September 3, 2009: (a) maximum ebb velocity; (b) maximum flood velocity (Note: Negative flood velocities are oriented in the landward direction)
Another manifestation of the neap–spring variability appears in the location of the upstream limit of current reversal. This is illustrated in Fig. 7 by the flood velocities crossing zero (i.e., reaching slack current) before becoming positive. In the present case, the upstream limit of current reversals moved between rkm 154 (neap tide) and rkm 200 (spring tide), but it is expected to vary even further in each direction under more extreme conditions of tidal range and river flow. Upstream of this limit, the tidal current was decelerated during the rising tide and accelerated during the falling tide. Lateral patterns in flow reversal were also observed, as illustrated in Fig. 3 by the white contours identifying the slack waters. Overall, these lateral patterns were characterized by current reversals being lagged and occurring earlier near the shore than in the channel [see also Matte (2014), Appendix 3].
Fig. 8 provides a detailed picture of the longitudinal and temporal variability of water levels and velocities, separately for neap and spring tides, from which the longitudinal profiles of Figs. 6 and 7 were extracted. Although Fig. 8 summarizes some of the information noted previously, the following additional features are worthy of note:
Fig. 8. (Color) Longitudinal and temporal variability during neap and spring tides for the simulation in August, 2009: (a and b) water levels; (c and d) tangential velocities (Note: Thick black contours indicate the time and location of slack currents; black dotted lines show the time and location of LW and HW)
1.
Tides are progressive in nature and become increasingly asymmetric moving upstream. The LWs propagate slower than the HWs, and the rising tide is considerably shortened compared with the falling tide, as shown by the diminishing distance separating the black dotted lines in Fig. 8. Landward of Deschambault (rkm 168), this asymmetric behavior was accentuated as the tide entered the tidal-fluvial reach (see Discussion for details on the reach classification). It was also exacerbated during spring tides compared with neap tides.
2.
The duration of flood currents, circumscribed by the thick black contours in Fig. 8, was approximately equivalent to the duration of ebb currents downstream but became relatively reduced upstream, up to the point of no current reversal.
3.
Variations in the location of the upstream limit of current reversals were observed between neap and spring tides but also between successive semidiurnal tides as a result of the diurnal inequality in tidal range.
4.
The LWS systematically occurred after LW, and the HWS came after HW, everywhere except near the upstream limit of current reversal, where the HWS sometimes occurred slightly before HW. Past Portneuf (rkm 163.5), the LWS got progressively closer to HW than LW during the rising tide.
5.
Landward of Saint-François (rkm 66), the HWS occurred nearly simultaneously over ∼75–100 rkm, especially during neap tides.

Flow Division at a Tidal Junction

Computed tidal discharges in the north and south arms of Orleans Island are presented in Fig. 9 to illustrate flow distribution around the island. Boat measurements in the two arms were taken 1 day apart from each other (Matte et al. 2014b) so that the river discharge conditions during each survey would be comparable. Similarly, the tidal range only slightly decreased between the 2 days (cf. Fig. 5). At both locations, observed and modeled discharges were found to compare very well over the tidal cycle, meaning that flow was well distributed between the channels. The average discharge at Québec was 11,600 m3/s during the measurement period, but the peak discharge exceeded 60,000 m3/s during ebb tide and reached almost the same (negative) value during flood tide in the south arm of Orleans Island. In the north arm, another 10,000 m3/s was discharged during ebb tide at its peak. These differences in discharge, measured during spring tide, follow the geometry of the channels, with the higher depths and wider sections being encountered in the south arm. However, to see how flow is partitioned in an average sense, the residual (tidally averaged) discharges should be computed at the two branches; such a comparison could not be performed here because of the length of the available discharge records, which did not exceed one tidal cycle.
Fig. 9. (Color) Observed and modeled flow division around Orleans Island: (a) north arm; (b) south arm
Fig. 10 presents simulation results at the junction of Orleans Island at different stages of the tidal cycle. Results are presented for a spring tide, measured at Lauzon (rkm 100) on June 24, 2009, under approximately average discharge conditions (11,100 m3/s). Arrows indicate the mean direction of currents and recirculations. At high tide, currents were reversed in both arms of Orleans Island, with velocities reaching approximately 1.5 m/s in the deepest regions. One hour after HW, currents were weakened, and recirculation appeared in shallow regions where water was redirected downstream. Two hours after HW, currents were oriented downstream and increased in importance with the falling tide; tidal flats were also progressively dried. Currents were at their maximum approximately 1 hour before low tide, with velocities reaching 2.3 m/s. At low tide, currents started decreasing; they rapidly changed in the following hours as a result of the more abrupt rising tide. One hour after LW, slack water had reached Lauzon, but ebb currents were still strong in the north arm of Orleans Island. Two hours after LW, currents were completely reversed in the south arm; they were partly diverted into the north arm and partly directed upstream. Slack water finally arrived in the north arm 1 hour before the next HW, and currents were completely reversed thereafter for approximately the next 3 hours.
Fig. 10. (Color) Modeled depth-averaged velocity modulus (in meters per second) at the junction of Orleans Island at different stages of the tidal cycle, identified by red squares over a tidal signal measured at Lauzon on June 24, 2009 (spring tide), with corresponding HW and LW values (Note: Arrows indicate the direction of currents and recirculations)
The flow patterns just described are in concordance with continuous water-level and velocity measurements taken at the junction of Orleans Island during the same tidal cycle as the one presented in Fig. 10 [cf. Matte (2014), Appendix 3].

Momentum Balance

Instantaneous momentum balances were computed for the spring tide of June 24, 2009, 08:00:00 EDT. Each term of the balance appears in the momentum conservation of the shallow-water equations given by Matte et al. [2017, Eq. (4)]. Their moduli are reported in Fig. 11, calculated from the x and y components of the balance. This snapshot encompasses the various features that can be observed during a tidal cycle (e.g., slack currents, peak ebb and flood currents), in contrast to tidally averaged momentum balances (not shown); the latter would be slightly less contrasted spatially but would highlight the regions where each term is the most influential in an average sense. The hydrodynamic conditions prevailing at the time of the snapshot are shown in Figs. 11(a and b). Upstream, the flow was unidirectional down to approximately Batiscan (rkm 199). A slack water before flood (or LWS) occurred at Deschambault (rkm 168), as shown by near-zero velocities and by a minimum in water levels close by [Figs. 11(a and b)]. This was then followed by a flood tide where currents were reversed and water-level gradients were positive (in the seaward direction). A slack before ebb (or HWS) was observed at the eastern end of Orleans Island, followed by increasing ebb currents in the seaward direction that were accompanied by a sharp decrease in water levels. Note that neither the LWS nor the HWS was exactly synchronized with the water-level extrema. They were both located downstream (or occurred later, in a time-reference frame) of the minimum and maximum in water levels, respectively.
Fig. 11. (Color) Terms of the momentum balance (modulus in meters squared per second squared) on June 24, 2009, 08:00:00 EDT (spring tide): (a) depth-averaged velocities with arrows indicating the direction of currents; (b) water levels; (c) local acceleration; (d) advective acceleration; (e) pressure gradient; (f) bottom friction; (g) Coriolis acceleration; (h) turbulent viscosity, whose scale is 1 order of magnitude smaller than the other terms [Note: River kilometer marks are provided in (a)]
Figs. 11(c–h) show the contribution of local acceleration Hu/t, advective acceleration ·Huu, pressure gradient gHh, bottom friction τb/ρ, Coriolis acceleration fcHu, and turbulent viscosity ·νtHu to the momentum balance, respectively [cf. Matte et al. 2017, Eqs. (4a) and (4b)]. Clearly, the dynamic balance was dominated by the local acceleration and pressure gradients, which are the major driving force of the flow. At both full ebb and full flood, local and advective accelerations, bottom friction, and, to a lesser extent, Coriolis acceleration balanced the pressure gradient. In the two upstream regions of high velocities, the local and advective accelerations were comparable to the pressure gradient because of the relatively low-water-level slopes. In contrast, in the downstream ebb, the flow was primarily driven by gravity because of the much-steeper gradients of water levels. Ratios calculated between the terms (not presented) show that bottom friction dominated over the effects of pressure gradient almost exclusively in shallow areas (e.g., intertidal flats, shoals). The Coriolis acceleration, for its part, typically exceeded the advective acceleration in regions where velocities were lower than 1 m/s. As for turbulence, it was 1 order of magnitude smaller than the other terms. The highest values were located in the deepest portions of the river, in front of Québec and Íle aux Coudres, and around engineering structures and islands. Turbulence was also higher in the channel than in the intertidal regions. Near slack, only the pressure gradient remained significant, which was balanced by local acceleration.
Hench and Luettich (2003) analyzed the transient momentum balances at shallow barotropic tidal inlets and found that they oscillated between two dynamical states, depending on whether the phase of the tide was near maximum ebb or flood or near slack. Here, similar analyzes were carried out with complete coverage of the tidal cycle in space rather than in time. Similar conclusions can be drawn from the analyzed fields with respect to the differing dynamic states observed at high velocity versus near slack. Near maximum ebb or flood, the pressure gradients were balanced by the acceleration and bottom friction terms. As the dynamic state approached slack waters, the velocities were reduced, and the balance shifted between the pressure gradient and local acceleration only. Lateral variations in the respective contribution of each force could also be observed and are associated with cross-channel gradients in bathymetry and friction.

Discussion

Flooding–Drying Processes

In the SLFE, large tidal ranges are responsible for rapid variations in flow conditions and in the wetted areas. At downstream locations, where tidal ranges exceed 6 m, increases in water levels of more than 1 m/h can be observed during the rising tide. These variations are the main driving force behind the flooding–drying processes, which play a key role not only in the transport of water over shallow areas but also in the strength of tidal currents in the main river channel. In fact, the convergence of lateral flow toward the channel tends to add mass and increase the magnitude of the along-estuary flow in the channel during ebb (Valle-Levinson et al. 2000). In contrast, the filling of intertidal flats during the rising tide reduces the flood velocities as part of the flow is redirected toward the shores. Ignoring these lateral exchanges could lead to significant under- and/or overestimation of the amplitude of tidal currents (Zheng et al. 2003).
Observations in the SLFE and other large tidal rivers [e.g., Jay et al. (2015)] have shown that LWs are higher on the tidal flats with respect to the channel (cf. Fig. 2) and may even be truncated when the floodplain is dried, depending on river flow and tidal conditions. Conversely, HWs remain relatively constant across the river section. As a result, tidal ranges are overall larger in the main channel than in the floodplain. Moreover, as observed by Valle-Levinson et al. (2000), lateral convergences are produced by phase lags of the tidal flow between the channel and the shoals, with magnitudes that are proportional to the along-estuary bathymetry gradients and to the tidal range. This is corroborated by results in the SLFE, where HW and LW timings were found to occur later over shallow topography than in the deeper channel, with the time lags of LWs being more significant than those of HWs (cf. Fig. 2). This yields a more pronounced tidal asymmetry in shallow areas. Laterally, these differences in amplitude and timing create gradients in velocity directing water toward the channel during ebb and toward the shores during flood.
Accurate representation of the floodplain topography, channel bathymetry, and bottom friction is therefore essential from a modeling perspective because they directly affect the magnitudes of these exchanges. In the SLFE, the availability of high-resolution LIDAR data and bathymetric soundings and detailed cross-sectional hydrodynamic data (i.e., simultaneous water levels and velocities) combined with a high-resolution finite-element mesh allowed for a precise description and validation of the flooding–drying processes occurring over shallow topography.

Tidal Propagation and Modulation

Factors influencing the strength of the tidal forcing over the tidal month include the neap–spring effect, the apogee–perigee effect, and the lunar declination, which is responsible for the diurnal inequality (Jay et al. 1990). Neap–spring variability in tidal systems is a well-known phenomenon that arises from the interaction of the dominant lunar and solar semidiurnal tides (M2 and S2, respectively), whereas apogee-perigee modulations and the diurnal inequality involve interactions between M2 and N2, and K1 and O1, respectively. In tidal rivers, the effects of these interactions are numerous, three of which are the longitudinal displacement of the tidal amplitude maximum [e.g., Jay et al. (1990)]; the fortnightly modulation of tidal ranges, HW, MWL, and LW [e.g., LeBlond (1979)]; and the longitudinal displacement of the upstream limit of current reversal [e.g., Jay (1984)].
As a result of topographic funneling effects, tides in the St. Lawrence are amplified as they enter the gulf and estuary. They reach their maximum amplitude between Íle aux Coudres and Orleans Island (Matte et al. 2017; Fig. 1), depending on the phase on the neap–spring tidal cycle and river discharge. The longitudinal position of the tidal amplitude maximum can be explained in terms of the tidal energy budget, where the tidal energy flux divergence is balanced against dissipation (Jay et al. 1990). Whereas the former is proportional to the square of the velocity, the latter is cubic in velocity, meaning that dissipation increases more than the incoming energy flux divergence as tidal range increases. This tends to damp the tide earlier and push any amplitude maximum seaward on spring tides. As a result, less energy is transmitted upriver, and the differences in tidal range between the neap and spring tides are smaller upstream than at the mouth (cf. Fig. 6).
In the SLFE, tides are increasingly distorted and damped as they propagate upriver as a result of nonlinear interactions of the incoming tide with river flow, friction, and topography. The respective contribution of these factors to the system’s internal variability is a function of both time and space, thus yielding distinct spatiotemporal patterns in water levels and velocities. Among these patterns are, for example, the flooding–drying processes occurring over shallow intertidal flats, marked changes in the tidal behavior associated with breaks in morphology [e.g., Matte et al. (2014a); Sassi et al. (2012)], and flow division at tidal junctions, for which topography and friction are known to be critical components (Buschman et al. 2010). In general, the timing of current reversals is strongly dependent on the geometry of the channel, and in the case of multiple channels (e.g., around the Orleans Island), their respective geometries will affect both tidal propagation and flow distribution. Cyclic exchanges occurring between the south and north arms of the Orleans Island are one example of the effects of asynchronous tidal propagation.
From downstream to upstream, tidal energy is progressively transferred from major astronomical constituents to overtides and subtidal frequencies. Overtides contribute to tidal asymmetry (i.e., wave steepening) by exerting an influence on the amplitude and timing of HW and LW. Low-frequency compound tides, for their part, contribute to the lowering of MWLs and LWs during neap tides with respect to spring tides (Aubrey and Speer 1985; Godin 1984, 1999; Jay et al. 2015; LeBlond 1978, 1979; Nidzieko 2010; Speer and Aubrey 1985). These effects are modulated by the river discharge, which adds to frictional damping and alters constituent amplitudes by stimulating energy transfers between frequency bands (Godin 1985, 1999; Jay and Flinchem 1997). In the St. Lawrence, the neap–spring reversal of MWLs begins seaward of the limit of salinity intrusion (which moves between Orleans Island and Íle aux Coudres), whereas the neap–spring reversal of LLWs more or less begins at Deschambault (rkm 168), that is, ∼100 rkm upstream of the salinity intrusion limit. Roughly, this corresponds to the point where the amplitude of MSf equals that of S2 (Hoitink and Jay 2016). It also coincides with the location where the amplitude of MSf is maximal in the SLFE and where a change in tidal asymmetry occurs (Matte et al. 2014a). The amplitude of the fortnightly tide eventually exceeds that of the semidiurnal tide near Trois-Rivières (rkm 231), approximately 160 rkm beyond the salinity intrusion limit. In sum, the SLFE can be divided into four regions (Matte et al. 2014a): (1) a tide-dominated reach downstream of Portneuf (rkm 0–163.5); (2) a transitional reach between the tidal and tidal-fluvial regimes (rkm 163.5–186), characterized by a rapid increase in bottom slope at the Richelieu Rapid near Deschambault (rkm 168); (3) a tidal-fluvial reach from Cap-à-la-Roche to Laviolette Bridge near Trois-Rivières (rkm 186–235), acting as a major restriction to the flow; and (4) a river-dominated reach upstream (rkm 235–302), where most of the short-period tide gets extinguished but where long-period oscillations persist.
Downstream, where tidal flows are bidirectional, the tidal storage volume can be estimated from the volume differences between HWS and LWS (Zhang et al. 2015). The limit where currents cease to reverse (i.e., where the HWS and LWS coincide) moves between Regions (1) and (3) in the SLFE as a function of tidal range and river discharge. Although the tidal wave propagates beyond this limit, the water is no longer transported upstream by the tidal current during flood tide (cf. Fig. 7). Instead, the currents experience a tidal backwater effect, where they decelerate during the rising tide and accelerate during the falling tide as water is periodically stored and released. Similarly, at the neap–spring scale, water is temporarily stored in the system during spring tide and released as the neap tide approaches, with a direct consequence on flood currents but little effect on ebb currents in the tidal-fluvial reach (cf. Fig. 7). The main reason for this is that during spring tide, a higher surface-level gradient is needed to realize the same discharge than during neap tide (Hoitink and Jay 2016), causing a fortnightly oscillation in water levels and leaving the outgoing flow almost unaffected. On the long term, river–tide interaction creates an oscillatory gradient of the MWL and steepens the surface-level profile as a result of the enhanced subtidal friction, up to the point of tidal extinction or even beyond (Buschman et al. 2009, 2010; Sassi and Hoitink 2013).

Conclusion

To gain insight into the tidal hydrodynamics of the SLFE, a 2D time-dependent hydrodynamic model was developed. In Part I (Matte et al. 2017), model calibration and validation were performed based on statistical assessments of modeled water levels at 29 tide gauges and discharges at 13 transects. In Part II, spatial and temporal patterns observed in the transect data were quantitatively assessed as a demonstration of model structural validity. This step, often neglected in the development of numerical models, is a necessary requirement for establishing confidence in model results, especially from an environmental modeling perspective. In fact, environmental processes are known to be strongly linked to physical variables, most specifically to their spatial and temporal distribution and to their gradients and connectivity (Morin et al. 2003).
The present research provides insights into the general physical processes of the SLFE, and of large tidal rivers globally, in relation to flooding–drying processes, tidal propagation and modulation, and transient momentum balance. These tidal and flow features manifest themselves at various spatial and temporal scales and are sensitive to topography and friction parameterizations. Future work should focus on the validation of the model under varying discharge, wind, ice, and macrophyte-distribution conditions. Efforts are currently under way to extend the model upstream and to set up an operational model of the St. Lawrence River and fluvial estuary (Matte et al. 2015).

Acknowledgments

Work by Pascal Matte was supported by scholarships from the Natural Sciences and Engineering Research Council of Canada and Fonds de recherche du Québec—Nature et technologies. The authors thank Environment Canada (Meteorological Service of Canada) for financial support; the Ministère du Développement Durable, de l’Environnement et de la Lutte contre les Changements Climatiques (MDDELCC) for financing the LIDAR campaign; and the Canadian Hydrographic Service (CHS) and Ministère des Transports du Québec (MTQ) for providing bathymetric and topographic data. The authors also thank Daniel Bourgault for his valued comments on a previous version of the manuscript and David A. Jay for discussions regarding the propagation of tides in rivers. Finally, the authors thank the two anonymous reviewers for their constructive comments on the manuscript.

References

Araújo, I. B., Dias, J. M., and Pugh, D. T. (2008). “Model simulations of tidal changes in a coastal lagoon, the Ria de Aveiro (Portugal).” Cont. Shelf Res., 28(8), 1010–1025.
Aubrey, D. G., and Speer, P. E. (1985). “A study of non-linear tidal propagation in shallow inlet/estuarine systems. Part I: Observations.” Estuarine Coastal Shelf Sci., 21(2), 185–205.
Babu, M. T., Vethamony, P., and Desa, E. (2005). “Modelling tide-driven currents and residual eddies in the Gulf of Kachchh and their seasonal variability: A marine environmental planning perspective.” Ecol. Modell., 184(2–4), 299–312.
Bárcena, J. F., García, A., Gómez, A. G., Álvarez, C., Juanes, J. A., and Revilla, J. A. (2012). “Spatial and temporal flushing time approach in estuaries influenced by river and tide. An application in Suances Estuary (Northern Spain).” Estuarine Coastal Shelf Sci., 112, 40–51.
Buschman, F. A., Hoitink, A. J. F., van der Vegt, M., and Hoekstra, P. (2009). “Subtidal water level variation controlled by river flow and tides.” Water Resour. Res., 45(10).
Buschman, F. A., Hoitink, A. J. F., van der Vegt, M., and Hoekstra, P. (2010). “Subtidal flow division at a shallow tidal junction.” Water Resour. Res., 46(12).
de Paiva, R. C. D., et al. (2013). “Large-scale hydrologic and hydrodynamic modeling of the Amazon River basin.” Water Resour. Res., 49(3), 1226–1243.
Godin, G. (1984). “The tide in rivers.” Int. Hydrogr. Rev., 61, 159–170.
Godin, G. (1985). “Modification of river tides by the discharge.” J. Waterway, Port, Coastal, Ocean Eng., 257–274.
Godin, G. (1999). “The propagation of tides up rivers with special considerations on the Upper Saint Lawrence River.” Estuarine Coastal Shelf Sci., 48(3), 307–324.
Hench, J. L., and Luettich, R. A. (2003). “Transient tidal circulation and momentum balances at a shallow inlet.” J. Phys. Oceanogr., 33(4), 913–932.
Hoitink, A. J. F., and Jay, D. A. (2016). “Tidal river dynamics: Implications for deltas.” Rev. Geophys., 54(1), 240–272.
Horsburgh, K. J., and Wilson, C. (2007). “Tide-surge interaction and its role in the distribution of surge residuals in the North Sea.” J. Geophys. Res., 112(C8).
Jay, D. A. (1984). Circulatory processes in the Columbia River estuary, Univ. of Washington, Seattle.
Jay, D. A., and Flinchem, E. P. (1997). “Interaction of fluctuating river flow with a barotropic tide: A demonstration of wavelet tidal analysis methods.” J. Geophys. Res, 102(C3), 5705–5720.
Jay, D. A., Giese, B. S., and Sherwood, C. R. (1990). “Energetics and sedimentary processes in the Columbia River estuary.” Prog. Oceanogr., 25(1–4), 157–174.
Jay, D. A., Leffler, K., Diefenderfer, H. L., and Borde, A. B. (2015). “Tidal-fluvial and estuarine processes in the lower Columbia River: I. Along-channel water level variations, Pacific Ocean to Bonneville Dam.” Estuaries Coasts, 38(2), 415–433.
LeBlond, P. H. (1978). “On tidal propagation in shallow rivers.” J. Geophys. Res., 83(C9), 4717–4721.
LeBlond, P. H. (1979). “Forced fortnightly tides in shallow rivers.” Atmos. Ocean, 17(3), 253–264.
Liu, Y. Z., Shi, J. Z., and Perrie, W. (2007). “A theoretical formulation for modeling 3D wave and current interactions in estuaries.” Adv. Water Resour., 30(8), 1737–1745.
Matte, P. (2014). “Modélisation hydrodynamique de l'estuaire fluvial du Saint-Laurent.” Ph.D. thesis, INRS—Centre Eau Terre Environnement, Québec.
Matte, P., Champoux, O., Secretan, Y., Morin, J., Smith, G. C., and Pellerin, P. (2015). “Towards an operational 2D non-stationary hydrodynamic model of the St. Lawrence River and fluvial estuary.” Proc., 49th CMOS Congress and 13th AMS Conf. on Polar Meteorology and Oceanography, Canadian Meteorological and Oceanographic Society, Ottawa.
Matte, P., Secretan, Y., and Morin, J. (2014a). “Temporal and spatial variability of tidal-fluvial dynamics in the St. Lawrence fluvial estuary: An application of nonstationary tidal harmonic analysis.” J. Geophys. Res., 119(9), 5724–5744.
Matte, P., Secretan, Y., and Morin, J. (2014b). “Quantifying lateral and intratidal variability in water level and velocity in a tide-dominated river using combined RTK GPS and ADCP measurements.” Limnol. Oceanogr. Methods, 12(5), 281–302.
Matte, P., Secretan, Y., and Morin, J. (2014c). “A robust estimation method for correcting dynamic draft error in PPK GPS elevation using ADCP tilt data.” J. Atmos. Oceanic Technol., 31(3), 729–738.
Matte, P., Secretan, Y., and Morin, J. (2017). “Hydrodynamic modeling of the St. Lawrence fluvial estuary. I: Model setup, calibration, and validation.” J. Waterway, Port, Coastal, Ocean Eng., 04017010.
Morin, J., and Champoux, O. (2006). “Integrated modelling of the physical processes and habitats of the St. Lawrence River.” Water availability issues for the St. Lawrence River: An environmental synthesis, A. Talbot, ed., Environment Canada, Montréal, 24–39.
Morin, J., Leclerc, M., Secretan, Y., and Boudreau, P. (2000). “Integrated two-dimensional macrophytes-hydrodynamic modeling.” J. Hydraul. Res., 38(3), 163–172.
Morin, J., et al. (2003). “Emergence of new explanatory variables for 2D habitat modelling in large rivers: The St. Lawrence experience.” Can. Water Resour. J., 28(2), 249–272.
Nabi, M., de Vriend, H. J., Mosselman, E., Sloff, C. J., and Shimizu, Y. (2012). “Detailed simulation of morphodynamics: 1. Hydrodynamic model.” Water Resour. Res., 48(12).
Nidzieko, N. J. (2010). “Tidal asymmetry in estuaries with mixed semidiurnal/diurnal tides.” J. Geophys. Res., 115(C8).
Picado, A., Dias, J. M., and Fortunato, A. B. (2010). “Tidal changes in estuarine systems induced by local geomorphologic modifications.” Cont. Shelf. Res, 30(17), 1854–1864.
Rinaldi, M., Mengoni, B., Luppi, L., Darby, S. E., and Mosselman, E. (2008). “Numerical simulation of hydrodynamics and bank erosion in a river bend.” Water Resour. Res., 44(9).
Sassi, M. G., and Hoitink, A. J. F. (2013). “River flow controls on tides and tide-mean water level profiles in a tidal freshwater river.” J. Geophys. Res., 118(9), 4139–4151.
Sassi, M. G., Hoitink, A. J. F., de Brye, B., and Deleersnijder, E. (2012). “Downstream hydraulic geometry of a tidally influenced river delta.” J. Geophys. Res. Earth Surf., 117(F4).
Simons, R. D., Monismith, S. G., Johnson, L. E., Winkler, G., and Saucier, F. J. (2006). “Zooplankton retention in the estuarine transition zone of the St. Lawrence Estuary.” Limnol. Oceanogr., 51(6), 2621–2631.
Speer, P. E., and Aubrey, D. G. (1985). “A study of non-linear tidal propagation in shallow inlet/estuarine systems Part II: Theory.” Estuarine Coastal Shelf Sci., 21(2), 207–224.
Valle-Levinson, A., Li, C., Wong, K. C., and Lwiza, K. M. M. (2000). “Convergence of lateral flow along a coastal plain estuary.” J. Geophys. Res., 105(C7), 17045–17061.
Zhang, M., Townend, I. H., Cai, H., and Zhou, Y. (2015). “Seasonal variation of tidal prism and energy in the Changjiang River estuary: A numerical study.” Chin. J. Oceanol. Limnol., 1–12.
Zheng, L., Chen, C., and Liu, H. (2003). “A modeling study of the Satilla River estuary, Georgia. i: flooding-drying process and water exchange over the salt marsh-estuary-shelf complex.” Estuaries, 26(3), 651–669.

Information & Authors

Information

Published In

Go to Journal of Waterway, Port, Coastal, and Ocean Engineering
Journal of Waterway, Port, Coastal, and Ocean Engineering
Volume 143Issue 5September 2017

History

Received: Jul 5, 2016
Accepted: Dec 27, 2016
Published online: Mar 27, 2017
Discussion open until: Aug 27, 2017
Published in print: Sep 1, 2017

Authors

Affiliations

Research Scientist, Environmental Numerical Prediction Research Section (RPN-E), Meteorological Research Division, Environment and Climate Change Canada, Government of Canada, 801-1550 avenue d’Estimauville, Québec, QC, Canada G1J 0C3 (corresponding author). ORCID: https://orcid.org/0000-0003-0968-507X. E-mail: [email protected]
Yves Secretan [email protected]
Professor, Centre Eau Terre Environnement, Institut National de la Recherche Scientifique (INRS-ETE), Univ. du Québec, 490 rue de la Couronne, Québec, QC, Canada G1K 9A9. E-mail: [email protected]
Chief, Hydrology and Ecohydraulic Section, National Hydrological Service, Environment and Climate Change Canada, Government of Canada, 801-1550 avenue d’Estimauville, Québec, QC, Canada G1J 0C3. E-mail: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share