Compound flooding in coastal regions, that is, the simultaneous or successive occurrence of high sea levels and high river flows, is a significant hazard. By 2100, globally over 600 million people residing on coasts could be exposed to compound floods under RCP8.5 (Kulp & Strauss, 2019). Coastal compound flooding may be a consequence of two mechanisms: First, extreme coastal water levels may affect river flows by backwater effects or by reversing the seaward flow of rivers (Zhang et al., 2019). Rivers located in low‐lying (below 10 m above mean sea level) regions of northwestern Europe (Figure 1 in Hoitink & Jay, 2016) are likely to be impacted by this mechanism. Sea‐level rise (SLR) increases the water levels in the mouth of rivers (Nicholls et al., 2011), thus increasing the probability of extreme coastal water levels. Second, severe storm episodes may lead to extreme storm surges, and at the same time, to heavy precipitation and high river runoff (Kew et al., 2013). Hence, the main drivers of coastal compound flooding are storm surges, SLR and river floods, which in turn, are driven by a range of processes in the atmosphere and oceans, and on the land surface. Quantifying compound flood hazard under climate change poses a particular challenge at the intersection of climate dynamics and statistical analysis, as the estimation of event probabilities is not only affected by climate projection uncertainties in relation to the different drivers but also in relation to the interdependencies between them (Zscheischler & Seneviratne, 2017).
1 Figure. Map of tide (gray triangles) and stream gauge (colored circles) locations. The shades of the circles are proportional to catchment area (in km2) with darker (lighter) hues indicating larger (smaller) catchments.
A number of studies, at local to continental scales across Europe (Bevacqua et al., 2019; Ganguli & Merz, 2019a, 2019b; Kew et al., 2013; Klerk et al., 2015; Petroliagkis, 2018; Svensson & Jones, 2002), the United States (Moftakhari et al., 2017, 2019), Australia (Zheng et al., 2014), China (Lian et al., 2013; Tu et al., 2018), India (Mohanty et al., 2020), Bangladesh (Ikeuchi et al., 2017), and the entire globe (Couasnon et al., 2020; Eilander et al., 2020; Ward et al., 2018) have analyzed compound effects of high coastal water levels (or storm surges and/or tides) and river discharges (or extreme precipitation as proxy for river discharge). Challenges to compound flood hazard assessment at a large scale under projected climate involve precisely estimating the interaction between the various drivers given the uncertainty of Earth system models (Deser et al., 2020). As an initial proof of concept, only four published studies have assessed future changes. These are concentrated around Europe (Bevacqua et al., 2019; Kew et al., 2013; Klerk et al., 2015) and the United States (Moftakhari et al., 2017).
Although a limited number of studies have projected changes in compound floods, these studies (Bevacqua et al., 2019; Kew et al., 2013; Klerk et al., 2015; Moftakhari et al., 2017) have carefully attempted to evaluate the uncertainty. However, earlier studies (Bevacqua et al., 2019; Kew et al., 2013; Klerk et al., 2015) have used different spatial scales, different time periods, variations in the underlying drivers, and different flood sampling schemes (annual maxima or peak over threshold), making it difficult to draw conclusions about future compound flooding for Europe. At the large river basin scale, a few European studies (Kew et al., 2013; Klerk et al., 2015) have analyzed future changes in compound floods from North Sea storm surges and river floods from the Rhine basin considering multiple time lags. While Kew et al. (2013) has used north‐northwesterly winds over the North Sea as a proxy for storm surges and multiple‐day precipitation as a proxy for peak discharge, Klerk et al. (2015) has implemented the Delft Continental Shelf Model to simulate high storm surges at the Hoek van Holland tide gauge and a lumped conceptual hydrological model to simulate daily runoff. Moftakhari et al. (2017) has assessed compound floods resulting from future SLR and river discharge scenarios at eight coastal‐estuarine systems over the Pacific and Atlantic coasts of the United States. Despite their focus on uncertainty quantification, projections of changes in storm surges, one of the drivers, remain elusive. On a continental scale, to our knowledge, only a single study (Bevacqua et al., 2019) covering a large portion of European coasts (including the Mediterranean and Baltic sea regions) has evaluated future trends in compound floods. However, this study investigated the concurrence of high sea levels and precipitation‐induced pluvial flooding.
Second, owing to computational complexity, studies of future compound flood hazards include projection scenarios based on coarse spatial resolutions and simplified assumptions to model individual drivers of compound floods. For instance, in Kew et al. (2013), the forcing variables for simulating extreme North Sea surges were derived from the ESSENCE data set, generated from ECHAM5/MPI‐OM coupled general circulation models (GCM) with a horizontal resolution of 180 km (T63 grids). Likewise, in Bevacqua et al. (2019) the meteorological forcing was directly taken from the output of six GCMs available at the CMIP5 archive, for which neither any downscaling nor bias correction was implemented. However, modeling storm surges in shallow coastal areas requires high‐resolution meshes and meteorological forcing to capture the effects of complex coastal bathymetry and topography (Marsooli et al., 2019; Muis et al., 2016). The same holds for extreme precipitation causing river floods as another driver of coastal compound floods (Prudhomme et al., 2002). The coarse resolution of GCMs possibly introduces large uncertainty in the representation of the drivers of compound floods. In addition, assessing compound flooding based on coarse spatial resolutions of the hydrometeorological forcing may underestimate extremes (Marsooli et al., 2019; Muis et al., 2016). Further, despite strong confidence in SLR (Kopp et al., 2014), only a small number of studies have assessed the contribution of SLR to compound floods over the United States (Moftakhari et al., 2017), Europe (Bevacqua et al., 2019; Klerk et al., 2015), and Bangladesh (Ikeuchi et al., 2015) coasts. A few regional‐scale assessments (Ikeuchi et al., 2015; Klerk et al., 2015) have assumed a spatially homogeneous SLR (i.e., the same increase in sea levels among different locations). However, SLR may stem from various factors, such as nonuniform changes in ocean dynamics, heat and salinity content, glacial isostatic adjustment, and vertical land motion at individual locations (Kopp et al., 2014). Hence, localized SLR projections are required to represent the spatial variability of SLR. Further, Moftakhari et al. (2017) has considered a stationary hydrologic regime (i.e., floods do not change between past and projected periods), which is not a realistic assumption in a changing climate (Stocker et al., 2013).
Third, unlike univariate extremes, the joint occurrence of multiple drivers can be measured by upper tail dependence, which takes into account the dependence at the tail of the probability distributions. While the state‐of‐the‐art literature on projected changes in compound floods has mainly assessed the central dependence among the underlying drivers, this measure only considers the probability that one variable exceeds its median given that the other variable also exceeds its median. Ignoring tail dependence among compound flood drivers may underestimate hazards associated with low‐probability but high‐impact events.
Understanding the mechanisms underpinning compound floods, under consideration of the interactions among the drivers, especially at the tail of the distribution, in a high‐resolution climate change projection, is of great importance to manage and minimize risks associated with climate change. To summarize, there is no study available at the continental scale in general and northwestern Europe in particular that estimates possible future changes in compound floods, driven by storm surges, SLR and river peak discharge, in an ensemble of high‐resolution regional climate models (RCMs). Further, the heterogeneity of the few available studies make it impossible to derive solid statements for the European coasts. To fill these gaps, here we quantify the future (2040–2069) changes in compound flood hazard resulting from high coastal water level, SLR, and peak fluvial discharge over northwestern Europe. These projections include an uncertainty assessment and are compared to a reference period (1981–2005). Northwestern Europe comprises one of the densely populated major economic hubs (McMaster, 2016) (Figure 1 and Tables S1 and S2 in the supporting information) and is extremely vulnerable to severe storm‐induced compound flooding (Kew et al., 2013; Ward et al., 2018). We apply a hydrodynamic approach to simulate annual maxima surge and WaterGAP Global Hydrology Model (WGHM) to compute peak river runoff (or streamflow) from individual river basins and analyze the joint frequency of the compound floods (section 2.3). Using high‐resolution (0.11° or 12.5 km), dynamically downscaled meteorological forcing from the World Climate Research Program's CORDEX (COordinated Regional Climate Downscaling EXperiment) framework (Dosio, 2016) for Europe, we simulate projected surge. An analysis of Europe‐wide precipitation climatology (Prein et al., 2016) has shown a clear improvement in 0.11° CORDEX‐RCM simulations in reproducing mean and extreme climatology for all regions and seasons due to improved representation of the orography and surface characteristics, and better resolved physical processes, such as convection. We evaluate the performance of RCM‐forced compound flood climatology against observations; such evaluations were largely ignored in earlier assessments (Bevacqua et al., 2019; Kew et al., 2013). Further, we integrate our results with the physically based, probabilistic SLR projection, labeled here as K14 (Kopp et al., 2014) for the RCP8.5 high emission scenario that represents a possible increase in radiative forcing of 8.5 W m−2 by 2100 (section 2.3). Our study assesses, for the first time, how projected climate change will affect the joint frequency of compound floods over the coastal regions of northwestern Europe involving all three drivers (i.e., localized SLR, storm surges, and river floods) based on high‐resolution, dynamically downscaled climate realizations.
Hourly sea level observations (in meter) for the reference period, 1981–2005 for northwestern Europe (approximately 46–66°N and −12.5°W to 19°E) were obtained from 23 tide gauges (Table S1) archived at Global Extreme Sea Level Analysis version II database (Woodworth et al., 2016). Each tide gauge contains high‐quality sea‐level records with less than 6 months of missing data during the reference period. Relative sea‐level data at Hoek van Holland tide gauge was obtained from Directorate for Public Works and Water Management, Rijkswaterstaat, the Netherlands. Most of the tide gauges contain records of hourly resolution. A few tide gauges have higher sampling frequencies, such as tide gauges from the UK and Norway with sampling frequencies of 15 and 10 min from 1993 and 2001 onward, respectively. For consistency, observations of higher sampling frequencies were transformed to hourly resolution by calculating the median of the subhourly values (Wahl et al., 2017). We obtain daily continuous river discharge data from 39 medium to large‐sized river basins with catchment area between 1,000 and 160,800 km2 (70% of stations with catchment area <10,000 km2; Table S2) from the Global Runoff Data Centre (GRDC; Grabs, 1997). Each of these stream gauges are located within 200 km radius from the tide gauges (Ganguli & Merz, 2019a, 2019b). The use of gauge‐based observations facilitates model validations in the historical (1981–2005) period.
To simulate surge and streamflow in the reference (1981–2005) and projected (2040–2069) periods, we use high‐resolution (0.11° in a rotated‐pole grid, that is, about 12 km) time series from the Euro CORDEX data set archived at
Although temperature and wind speed are typically well simulated by regional climate models, considerable biases exist for precipitation (Schoetter et al., 2012). Further, bias‐corrected meteorological forcing substantially improves hydrologic predictions (Roy et al., 2017; Teutschbein & Seibert, 2013). Hence, to simulate daily river discharge, we use bias‐corrected precipitation and temperature time series (Dosio, 2016), respectively. As downwelling radiation, which is required for the streamflow simulation, is currently not available from the RCM runs, we obtain downwelling longwave and shortwave radiation data from reanalysis, WFDEI‐GPCC (Weedon et al., 2014), available at 0.5°. To force the hydrodynamic model Delft3D, we use the climate model output available at 0.11° resolution. However, since the grid resolution of the hydrological model WGHM is 0.5° (Müller Schmied et al., 2014), the meteorological forcing data are regridded to 0.5° to match the resolution of the hydrologic model.
Following earlier studies (Kopp et al., 2014; Kulp & Strauss, 2019; Lin et al., 2016; Lin & Shullman, 2017; Lowe et al., 2001; Marsooli et al., 2019; Sterl et al., 2009), we assume a negligible influence of mean SLR on surge heights (SLR = 0) during the reference (1981–2005) period. For the future, since SLR projections are not available from CORDEX simulation, we use a downscaled localized SLR scenario, K14 (Kopp et al., 2014) under the RCP8.5 scenario. K14 offers physical model‐based projections based on IPCC AR5 scenarios. Apart from ice sheet components for Greenland, West Antarctic, and the East Antarctic, K14 incorporates all spatially explicit submodels for climatic components of SLR (such as glacier and ice cap surface mass balance, oceanographic processes, land water storage) as well as nonclimatic components (such as global isostatic adjustment, sediment compaction, and tectonics). Leveraging tide gauge observations from the Permanent Service for Mean Sea Level (PMSL), K14 estimates historical SLR rates and generates projections on a 2° global grid that intersects with the world coastline. Further, it offers projections for each RCP, based on 10,000 Monte Carlo samples of SLR at each PMSL tide gauge location.
The SLR projections are based on single realizations of 29 GCMs and represent a 19‐year running average value available for each decade since the year 2010 (i.e., average over 2001–2019) until 2300. These SLR projections assume the year 2000 as the baseline year (i.e., SLR = 0), which we treat as a part of the reference period with respect to sea level for hazard assessment. The coastal flood height should be estimated by combining distributions of projected surge height and SLR, projected by the same climate model since a change in storm climatology and SLR both are affected by the large‐scale environment (Lin & Shullman, 2017). However, neither probabilistic SLR projections based on individual RCM nor GCM realizations are currently available in the literature. Hence, following the earlier studies (Lin et al., 2016, 2019; Lin & Shullman, 2017; Marsooli et al., 2019), we combine projected surge height with composite SLR distributions for individual GCM‐RCM realizations and their multimodel ensemble (discussed later). For the future period, we compute median values of decadal localized SLR projections between 2040 and 2069 (i.e., 2040 ≤ s < 2069, where s indicates a random variable, SLR) at individual tide gauge locations before adding it to the projected surge.
While we force dynamically downscaled (as a preprocessor) climate model simulations to reduce meteorological input uncertainty and to minimize a possible error propagation through hydrodynamic and hydrologic models, we incorporate a statistical postprocessor, which effectively reduces the uncertainty in ensemble climate projections (Kang et al., 2010). Following earlier studies (Lin et al., 2016, 2019; Roy et al., 2017), we bias correct the simulated surge and discharge climatology during the reference period using a simple quantile mapping (or CDF matching) approach, in which a quantile of the present‐day simulated distribution is replaced by the same quantile of the present‐day observed distribution preserving its statistical moments (Maraun, 2016). Given a historical surge or discharge time series x, the method is formulated as [Image Omitted. See PDF]where is the bias‐corrected surge or discharge time series, F(·) is the Cumulative Distribution Function (CDF) and F−1(·) is the inverse CDF of either observation (“o”) or model (“m”) for the historical period or current climate condition (“c”). We use the Gaussian kernel density function with the bandwidth according to Silverman's rule of thumb to estimate the Probability Density Function (PDF) and its inverse PDF (MATLAB “ksdensity” function). To prevent negative values for discharge, we restrict the kernel density estimator to positive values by setting the ksdensity function parameter to “positive.”
Equation 1 assumes only changes in the mean but stationarity in the variance and skew of the distribution. This, however, may not hold since climate model projection may lead to changes in higher order moments as well. In order to incorporate information from the distribution of the model projection, the difference between CDFs of the future and the reference period is taken into account. To bias correct the hourly surge simulated by the Delft3D hydrodynamic model, in the projected (“p”) period, we apply an equidistant quantile mapping by adding an additive factor (Yang et al., 2018) as below: [Image Omitted. See PDF]
To avoid the problem of negative values in the projected discharge time series simulated by WGHM, we use an equiratio quantile mapping by applying a multiplicative factor (Yang et al., 2018) as below: [Image Omitted. See PDF]Equation 2 is analogous to the bias correction of temperature, where an additive adjustment is usually applied to correct the mean biases of the model, whereas Equation 3, that is, scaling adjustment by a multiplicative factor, is often utilized for precipitation to avoid negative values. Figures S1 and S2 compare the performance of RCM‐forced annual maxima surge and peak discharge time series relative to observations, with and without bias correction for selected tide and stream gauges. The bias correction clearly improves the reproduction of the variability of the compound flood drivers. Figure S3 compares the variability of reference (1981–2005) versus projected (2040–2069) annual maxima surge and corresponding peak fluvial discharge after bias correction across all GCM‐RCM realizations for selected tide and stream gauges. There is no clear change between the reference and future periods for the annual maxima surge heights (Figure S3a). In contrast, future peak discharge (Figure S3b) tends to be higher than the observed streamflow, which agrees with earlier assessments over northwestern Europe (Bosshard et al., 2014; Lobanova et al., 2018).
To account for the nonstationarity resulting from SLR and interannual variability in hourly sea level observations, we subtract the annual average sea level on a year‐to‐year basis (Muis et al., 2016; Wahl et al., 2017). We use the MATLAB UTide package (Codiga, 2011) for a year‐to‐year harmonic tidal analysis with all 67 tidal constituents (Wahl et al., 2015, 2017). The advantage of this tool is its ability to extract tidal constituents from sea level records with uneven temporal sampling and data gaps. Then we compute hourly skew surge as the difference between the maximum observed high water (i.e., detrended time series) and the maximum predicted tidal water level within a tidal cycle for each tide gauge record (Wahl et al., 2017).
To simulate surge with meteorological forcing from GCM‐RCM runs, we run the hydrodynamic model Delft3D‐FLOW (Delft3D‐FLOW, 2014), which uses the depth‐averaged shallow water equations. Since we are mostly interested in the meteorological phenomena that drive compound floods, we focus on modeling surges and do not simulate tides. Further, the simulated storm surge does not include the surge component due to wave setup. The review of the literature (Lowe et al., 2001; Sterl et al., 2009) shows a negligible influence of mean sea level rise on surge height; hence, following these studies, the mean sea level rise was not included during the calibration of the hydrodynamic model. The hydrodynamic model was previously calibrated (Paprotny et al., 2019, 2020) with a computational grid of 0.11° with the same spatial extent as the EURO‐CORDEX domain using observed skew surges. The model was validated using reanalysis data ERA‐Interim for the period 1979–2014 and showed a satisfactory model performance (Paprotny et al., 2019, 2020; Figure 2, right panel of Paprotny et al., 2019 shows the calibrated hydrodynamic model performance, which should have been placed in Figure 3, left panel of the paper. Further, in Paprotny et al., 2019, the relative error of surge output was found to vary between −25% to 3% for different tide gauges, which was obtained without applying any statistical postprocessor. We argue that we have bias corrected the surge to ensure better model performance). Details of the model parameters (such as wind drag coefficients and channel roughness) and boundary conditions are discussed in Paprotny et al. (2019, 2020). The simulation is driven by 6‐hourly sea level pressure and winds. To accurately simulate annual maxima storm surges, the time step of the calculation is kept as 30‐min. The modeled surge output is outputted at every 6‐hr interval.
Assuming a negligible nonlinear interaction between surge, S and SLR, R for the RCP8.5 scenario, following the earlier literature (Lin et al., 2016, 2019; Marsooli et al., 2019), we estimate the future coastal flood elevation (Hf) as the sum of the future surge elevation and the SLR scenario. As both variables are given as PDFs, we employ a convolution integral operation to obtain the PDF associated with surge and SLR. Assuming surge and SLR be two independent continuous random variables with PDFs fS(s) and fR(r), and Hf = S + R, yields the PDF of Hf: [Image Omitted. See PDF]
Our choice of neglecting nonlinear interactions between water level components is justified as follows: (1) Besides a few regional exceptions (Devlin et al., 2017; Zijl et al., 2019), the assumption of nonlinear interactions between the water level components are relaxed in continental‐scale assessments (Losada et al., 2013; Paprotny et al., 2020) and climate change projections (Lowe et al., 2001; Sterl et al., 2009; Vousdoukas et al., 2018), since the resulting error is negligible as compared to the other sources of uncertainty, such as the cascading uncertainty in numerical modeling chains and climate model experiments (Vousdoukas et al., 2018). (2) Simulating water level resulting from nonlinear interaction of streamflow and coastal water level integrating tide, surge, and wave setup would require a fully coupled modeling approach (Santiago‐Collazo et al., 2019), which goes beyond the computational capabilities for a continental‐scale assessment.
We use the global hydrological model, WGHM version 2.2d (Döll et al., 2003; Müller Schmied et al., 2020) to simulate current and future streamflow at daily time steps. WGHM simulates storage and flux components of the continental water cycle with a spatial resolution of 0.5° (~55 km). WGHM is run with daily reanalysis‐based WFDEI‐GPCC (Watch Forcing Data based on ERA‐Interim) (Weedon et al., 2014) meteorological forcing. The model is calibrated (Müller Schmied et al., 2014) based on observed river discharge at stations around the world by tuning the runoff coefficient parameter. Figure S4 compares the performance of WFDEI‐forced WGHM runs with and without bias correction using an overall score, OS, which is the median of runoff signatures and temporal variability (Text S1.1). 0.8 < OS < 1 indicates an excellent performance; 0.6 < OS < 0.8 represents good performance, and 0.4 < OS < 0.6 shows moderate performance. Without bias correction, we note (from Figure S4a) a moderate to good performance especially for stream gauges located in southern UK. Bias correction clearly improves the overall performance for the gauges across the southern and western coasts of the UK, Germany, Denmark, and the Swedish coast. Figure S5 shows the heat map of the performance of bias‐corrected daily discharge obtained from RCM‐driven WGHM runs for each stream gauge location during the reference period, indicating that most of the sites show a good to excellent model performance with OS larger than 0.6.
Following the earlier literature (Ganguli & Merz, 2019a, 2019b), we identify compound floods when hourly annual maxima surge events coincide with d day (where d = 1,.., n, which is proportional to the catchment area) lagged peak river discharge within ±7 days of the surge event. To account for the delay between storm surges and river peak flows, which are caused by the same meteorological event, we shift the discharge time series by a lag time based on catchment‐specific watershed response time. Following the earlier literature (Berghuijs et al., 2019; Merz et al., 2018), we consider two events, that is, a coastal surge and a river peak flow, as a compound event when they occur within a window of ±7 days. This window is motivated by the fact that two events do not need to occur at exactly the same day to enhance their impacts. For instance, the first event could weaken flood defense systems or stress disaster management capacities, posing a problem when the second event occurs with a few days delay.
While other statistical methods, such as r‐largest methods or peak‐over‐threshold approach (Wahl et al., 2017), could be applied to sample surge by choosing more than one event per year, we prefer annual maxima since this typically ensures the iid (independent, identically distributed) assumption of extreme value statistics. The annual maxima method is also easier to apply without the need of choosing an appropriate threshold from surge time series over a large number of tide gauges and across a suite of climate model run (Muis et al., 2016).
The inclusion of basin lag time is motivated by the delay between storm surges and river floods when both are caused by the same large‐scale meteorological event. A lag time of a few days has been suggested for establishing a moderate to a strong correlation between surge and river discharge (Petroliagkis, 2018). We adopt the watershed response time (d), representing the time needed for rainfall to propagate through the catchment to the downstream streamflow gauge (Ganguli & Merz, 2019a, 2019b): [Image Omitted. See PDF]
Where Ad denotes the catchment area (in km2). For each tide gauge‐stream gauge combination, we search for the highest peak discharge of the d‐shifted streamflow time series within an interval of ±7 days of the occurrence of the annual maxima surge event.
The upper‐tail‐dependence coefficient (UTDC) between surge and peak discharge is determined using a nonparametric tail‐dependence metrics, namely, Capéraá‐Fougéres‐Genest (CFG) estimator, λCFG. We estimate as follows (Frahm et al., 2005) [Image Omitted. See PDF]
Where,ui and vi denote the empirical cumulative distributions of the compound flood drivers and n indicates the number of samples. We determine the empirical UTDC by calculating the empirical cumulative distribution of each of the variables and then substituting the corresponding values in Equation 6. We determine statistical significance (p value < 0.10) of UTDC estimates by drawing 10,000 random bootstrap samples and then calculating the p value of the test using a standard percentile‐based approach.
For the marginal distributions of annual maxima surge and peak discharge the Gamma, Log‐normal, Log‐logistic, and Generalized Extreme Value (GEV) are used. While the parameters of the first three distributions are determined using the method of maximum likelihood, the GEV parameters are estimated via Bayesian inference combined with Differential Evaluation Markov Chain (DE‐MC) Monte Carlo (MC) simulation (Ganguli & Merz, 2019b) owing to the fact that the application of maximum likelihood estimator in small samples (n ≤ 50) often leads to an absurd value of the GEV shape parameter. The use of a Bayesian prior distribution restricts the shape parameter to a physically consistent range (Martins & Stedinger, 2000). The DE‐MC simulations are run for 3,000 iterations of multiple (n = 5) parallel chains and the convergence of MC simulation is checked by a scale reduction factor that suggests the value of should remain below the threshold value of 1.1. The parameters of the GEV distribution are derived by computing the 50th percentile of postburn in (2001st–3000th) random draw.
We select the marginal distribution fit based on the corrected Akaike Information Criterion (AICc) (Lee & Ghosh, 2009). The validity of the best fitted distribution is evaluated using Kolmogorov‐Smirnov and Cramer‐von‐Mises goodness‐of‐fit statistics at a 10% significance level. Figures S6 and S7 show the probability‐probability plots of empirical versus simulated compound flood drivers for the historical and projected periods, which indicates good performance of the modeled distribution. Since SLR variates are available in a discrete form as an output from a Monte Carlo simulations, following the previous studies (Lin et al., 2016, 2019; Lin & Shullman, 2017; Marsooli et al., 2019), the marginal distribution of SLR is estimated using nonparametric kernel density estimators. We use the Gaussian kernel density function to fit PDF of SLR whereas the bandwidth is obtained using Silverman's rule of thumb.
To model the joint distribution, we select the tide gauge‐stream gauge pairs for which empirical upper tail dependence coefficient, UTDC between surge and peak discharge, λemp > 0 and absolute signal (simulated change) to noise (simulated variability) ratio, |SNR| > 1 for both historical and projected periods. We relax the statistical significance level to include all sites that show positive upper tail dependency, even when the total correlation value is weakly negative or insignificant. This also allows us to include gauges that show a large climate change signal over its variability (or noise).
Based on analytical tractability and ability to model upper tail dependence, we use three copula families: Gumbel‐Hougaard, flipped Clayton and Student's t. We estimate the copula parameters using the maximum pseudo‐likelihood method. We evaluate the adequacy of the copula model in capturing the upper tail dependence using Mean Error to Standard Error (MESE) statistics (Frahm et al., 2005) between empirical UTDC estimate, and simulated UTDC from the parametric family of the copula, (see Equation 7). For comparison, synthetic bivariate samples are simulated from the copula family, Cθ with the same length as the observation, Xi,d = {Xi,1, Xi,2}. The computation of is repeated for N = 500 random runs. MESE statistics is then computed as the absolute deviations between and the mean of , for all 500 runs normalized by its standard deviation, [Image Omitted. See PDF]
Out of these three copula families, we select the one that results in the lowest value of MESE statistics for a given location. Further, we assess the goodness‐of‐fit of the selected copula model using parametric bootstrap‐based Cramér‐von Mises test statistics at a 10% significance level (Genest et al., 2009). Figure S8 shows the performance of empirical versus the best selected copula models in terms of tail dependency for historical and projected periods. The scatter close to the 1:1 line indicates a good performance of the selected copula.
Considering the dependency between annual maxima surge and associated peak river discharge, we select the 95th percentile value of each of the compound flood drivers to compute the joint return period. For a given compound flood event pair, the bivariate joint return periods (JRP) of the cooccurrence or successive occurrence (within ±7 days' time window) of annual maxima surge and peak discharge (i.e., both) exceeding 95th percentile threshold using “AND” operator can be expressed as [Image Omitted. See PDF]
Where FS (s) and FQ (q) are the marginal distributions of the surge and discharge, respectively, and CSQ (s,q) is the copula‐based joint distribution.
Accounting for the 50th percentile (median) SLR condition, the joint return period of coastal flood elevation and river discharge for the projected period using “AND” operator [Image Omitted. See PDF]
Where = s95 + r50; s95 and r50 correspond to 95th percentile projected surge and 50th percentile SLR values for a tide gauge. is the CDF value corresponding to and obtained from interpolation of coastal flood elevation with its CDF, . is obtained using numerical integration of as obtained from Equation 4.
A comparison of the univariate return level plots between observed and modeled surge heights during the reference period, 1981–2005 (when SLR = 0; Figures 2 and S9), shows a good agreement for most of the selected tide gauges, except in Dover, where return levels of surge heights are overestimated by four out of five models. This is due to the complex topography of the Dover Straits that limits the ability of hydrodynamic model to adequately simulate the flow of water through such narrow straits. Except for a few models, the return level plots show a marginal increase in surge heights for the projected time period. Considering the contribution of SLR (Figures 2 and S9) along with surge, the return levels of projected coastal flood elevation are significantly higher across all gauges. This suggests that SLR plays an important role in future compound floods in northwestern Europe and should not be neglected.
2 Figure. Comparison of historical (1981–2005) versus projected (2040–2069) extreme coastal water levels with and without sea level rise at selected tide gauges in northwestern Europe. (a) Cuxhaven, (b) Dunkirk, (c) Hoek van Holland, (d) Oslo, (e) Gothenburg, (f) Dover. Observations of annual maxima are shown as circles, simulations from individual climate model simulations (ALL) are shown as thin solid lines (historical in gray; projected in light brown). The median (MED) of the multimodel RCM ensemble is shown as thick solid lines (historical in black; projected in dark brown). It is the median of the single GCM‐RCM realizations for each return period. The return levels of projected surge from the multimodel RCM ensemble considering the effect of SLR is shown using thick dashed red lines (surge and SLRMED).
The univariate return level plots of river floods (Figure 3) demonstrate a good agreement between the median of the multimodel ensemble and the observed peak discharge for most of the selected stream gauges. The observed peak discharge is underestimated at the gauge Gothenburg (River Lagan, Figure 3e) and overestimated at the gauge Dover (River Thames, Figure 3f). GHM often tends to underestimate the peak discharge of river basins at high latitudes (Gudmundsson et al., 2012). The underestimation at Gothenburg and other high‐altitude gauges is likely a consequence of biases in the forcing data, for instance, the local orographic effects on precipitation for a region with complex topography cannot be resolved well with the large grid cells of RCM simulations (Gudmundsson et al., 2012). For the future, a marked increase in peak discharge is seen at most stream gauges, especially for the larger rivers Rhine (gauge Hoek van Holland) and Elbe (gauge Cuxhaven). Overall, we find that future peak discharge (Figure 3) tends to be higher than the observed streamflow, which agrees well with earlier assessments over northwestern Europe (Lobanova et al., 2018; Schneider et al., 2013).
3 Figure. Comparison of historical (1981–2005) versus projected (2040–2069) peak river discharge at selected streamflow gauges in northwestern Europe. (a) Cuxhaven, (b) Dunkirk, (c) Hoek van Holland, (d) Oslo, (e) Gothenburg, (f) Dover. The observed return levels are shown using circles (in black), simulated return levels forced with individual RCMs are shown using thin solid lines (historical in gray; projected in light brown). The return levels of the multimodel ensemble is shown using thick solid lines (historical in black; projected in dark brown). The multimodel return level curve is obtained as the median of each return level curve of the different GCM‐RCM realizations.
Next, we analyze the dependence between compound flood drivers. Since central dependence measures (such as rank correlation coefficients) are more suitable to reflect dependence at the centre of the distribution (Ganguli & Merz, 2019a), we use an upper tail dependence coefficient (UTDC; section 2.3) to assess the dependence between annual maxima surge and peak river discharge (Figure 4). In the observations (Figure 4a), we find stronger positive UTDC values along the north shore of Scotland; the dependence pattern weakens gradually toward the south. The strong dependence along the north shore of Scotland has been attributed to orographically enhanced precipitation owing to the presence of hills on the northern side and cyclone traveling from northeastward to the north (Svensson & Jones, 2004). Statistically significant, rather strong positive UTDC values (i.e., λ ≥ 0.4) are also observed for the German tide gauge Cuxhaven, Danish tide gauge Esbjerg, Dutch tide gauge Delfzijl, and Swedish tide gauge Stenungsund. Weakly negative to insignificant positive dependence values are mostly concentrated in southeast England, which could be due to recurrent dry winter and persistent groundwater droughts leading to multiyear hydrological droughts (Kendon et al., 2013). The values of observed UTDC vary between −0.12 and 0.61 (Figure 4a), whereas the values of RCM‐modeled UTDC during the reference period (1981–2005) vary between 0.15 and 0.70 (Figures 4b and 4d). Overall, the strength of dependence weakens in the projected period (2040–2069) among all models (Figure 4c) and ranges between −0.17 and 0.51. The spatial pattern of the mean UTDC values, that is, averaged across climate realizations, for the reference period (Figures 4a and 4b) shows mostly an overestimation relative to the observations, along the western coast of the United Kingdom (UK), France, Germany, and the Netherlands. However, there are also a few locations where the dependence is underestimated, in particular along the north shore of Scotland and a few gauges in Scandinavian coasts (Figures 4a and 4b). These regions are characterized by hills and high plateau in the north of Scotland and by deep valleys and fjords around the Scandinavian coasts. The underestimation in dependence patterns is likely caused by this pronounced orography, which is not well simulated by many climate models (Gudmundsson et al., 2012).
4 Figure. Dependence between annual maxima surge and peak river discharge for observation and GCM‐RCM ensembles. Spatial distribution of upper tail dependence coefficient, λCFG is shown for (a) observation, (b) simulation for the reference period 1981–2005, and (c) the projected period 2040–2069. Circles show stream gauge locations. Shaded circles, ranging from yellow to red, indicate the fraction of models that show significant correlation at 0.10 level if this fraction is 60% or larger (i.e., ≥3/5 models agree in dependence pattern). The circles in white indicate insignificant positive correlation. The circles in cyan show a weak negative dependence, indicating an inverse relation between them that does not lead to a compound event. Locations with |SNR| > 1 (signal‐to‐noise ratio) are marked with a “plus” sign. (d) Kullback‐Leibler (KL) divergence metric between observed and historical upper tail dependence for sites with SNR > 1; the lower the KL divergence measure, the better the match to the empirical distribution. (e) probability density functions (PDF) of upper tail dependence coefficients λCFG for observations (dashed line), individual GCM‐RCM runs during the reference (shaded solid lines in gray) and projected (shaded solid lines in red) periods for sites where SNR > 1.
The spatial pattern of the mean UTDC for the projected period indicates a weakening of the dependence strength with a decrease in UTDC for 93% of the stream gauges (Figures 4c and 4d). The decrease in mean UTDC ranges between −3.5% and −84.6%. A few gauges in the UK and Denmark show an increase in the mean UTDC in the range from 0.3% to 34%. Among the proposed mechanisms for a weakening of dependence strength is a decrease in wind speed in large parts of northwestern Europe (the North Atlantic region) associated with a poleward shift of the North Atlantic jet (Gonzalez et al., 2019; Kjellström et al., 2018). Pooling RCM realizations, 92% (66 out of 72) and 78% (56 out of 72) of the tide gauge–stream gauge pairs show an absolute signal (simulated change) to noise (simulated variability) ratio (SNR) >1 during the historical and projected periods, respectively, indicating that the simulated change is larger than the simulated variability among models (Figures 4a and 4c). This implies a higher model agreement in the reference period compared to the projected period. This decrease in SNR (Figure S10) from reference to projected period is seen over large portions of northwestern Europe, except for gauges across the southeast UK and Scotland (Figure 1), where SNR value tends to be greater than 3.
Next, we compare the UTDC values between observations and simulations for the historical (Figure 4d) and projected period (Figure 4e), considering only those stream gauges for which SNR > 1. In Figure 4d, we show the performance of individual RCM realizations in simulating the observed dependence using the Kullback–Leibler (KL) divergence metric (Thorarinsdottir et al., 2013), which quantifies the divergence between the probability distributions of observation and simulation. While the PDFs of the models tend to encompass the observed dependence (Figure 4e), they differ in their shape and width relative to observations, which may result from model structural differences and relative roles of internal variability (Deser et al., 2020). All models overestimate the dependence strength. RCA4 forced with HadGEM2‐ES shows the largest discrepancy between observed and simulated dependence with the largest divergence value as well as the strongest overestimation, whereas IPSL‐CM5A performs the best. The changes in simulated standard deviations of UTDC, forced with the RCM ensemble, are up to ±23% of the observed. The biases of the different GCM‐RCM realizations show that the driving GCMs have some impact on the dependence pattern. The model structural differences emerge because of limitations in the process representation, such as the parametrization of convection (Prein et al., 2015; Púčik et al., 2017; Rajczak & Schär, 2017). Finally, except HadGEM2‐ES and MPI‐ESM‐LR, three out of five models show statistically significant (p value < 0.05 in two‐sample Kolmogorov‐Smirnov test) leftward shifts in projected PDFs relative to their historical simulation, indicating a substantial decrease in the strength of dependence in the projected period (Figure 4e).
An overestimation in the dependence strength by the models highlights the need to be careful when deriving the interdependence between compound flood drivers from climate model realizations. This cautionary note seems particularly valuable, as most of the earlier assessments have not validated their simulated dependence strength (Bevacqua et al., 2019; Kew et al., 2013), which is a robust statistic affecting the occurrence frequency of compound floods.
We model the dependence between extreme surge and peak river discharge using a copula‐based multivariate distribution (section 2.3) for gauges that show positive UTDC and SNR > 1. Next, we focus on the 95th percentile (occurring on average once in every 20 years) of the compound flood drivers, storm surge, and river peak. The joint return periods of such events, that is, storm surge and river peak flow are both higher than the 95th percentile, for most gauges between 20 and 50 years in the reference period (Figure S11a). This implies a considerable dependence between surge and peak discharge.
Projected changes in compound flood hazard, that is, changes in the mean of the multimodel ensemble relative to the reference period, are shown in Figure 5 for the different streamflow locations and in Figure 6b (indicated by the density plot in black) as PDF across all locations. These changes vary between −41% and 291% across the locations. For instance, selecting those locations with a Joint Return Period (JRP) of 60 years for the 95th percentile of the compound flood drivers in the reference period, this event would occur on average every 35 and 76 years, respectively, in the future period. Neglecting SLR (Figures 5a and 6b), anthropogenic climate forcing decreases the flood hazard for the large majority of locations. 87% of the gauges show positive changes in JRPs, out of which 66% show a more than 50% increase in JRP. This increase in JRPs is mainly a consequence of the weakening in dependence strength between storm surges and river peaks (Zscheischler & Seneviratne, 2017) by the 2050s (Figure 4c). On the other hand, only 13% of the gauges show more severe compound flood hazard. Their projected decreases in JRPs range from −0.5% to −41%, with a comparatively more robust decrease for gauges in southeast England. These sites show |SNR| > 1 in RCM simulations (Figures 4b and 4c), indicating a strong agreement among the climate models.
5 Figure. Projected changes in joint return periods. Changes (2040–2069 minus 1981–2005) in joint return periods (JRPs) for the 95th percentile of storm surge and peak river flow without (a) and with (b) the effect of probabilistic sea level rise (SLR) under K14 scenario. The changes in joint return periods are mapped for gauges with |SNR| > 1 owing to the highest agreement among climate models at these locations. Values show the changes in joint return periods (in %) derived from multimodel median climate realization with positive (negative) values showing a decrease (increase) in hazard, relative to the reference period (1981–2005).
6 Figure. Interquartile range and density functions for the projected changes in joint return periods. (a) Interquartile range of changes in joint return periods (in %) with (red) and without (gray) SLR effect for individual climate models and the multimodel ensemble. The uncertainty, expressed as interquartile range, of changes in joint return periods is larger for individual GCM‐RCM realization than for the multimodel ensemble. The median changes are shown using the vertical line within the horizontal boxplot. (b) Kernel density functions of the median of the multimodel RCM ensemble. Incorporating localized SLR, shifts the PDF substantially to the left, indicating the important role of SLR in future compound flood hazard estimation.
When the effect of SLR is incorporated (Figures 5b and 6b), we find again that anthropogenic climate change tends to decrease compound flood hazard over northwestern Europe. The majority of locations (66%) shows positive changes in JRPs, that is, the given combination of storm surge and river peak flow occurs less often. However, the PDF of changes in JRPs is clearly shifted toward a less favorable situation compared to the situation without SLR (Figure 6b). Of the locations, 34% show negative changes in JRPs, indicating more severe hazard, compared to only 13% when we do not consider the effect of SLR. Hence, SLR is an important driver of compound flood hazard in northwestern Europe and should not be neglected. Stream gauges with more severe hazard by 2050s are mostly concentrated along the coasts of Great Britain (including southeast England, Wales, and Scotland), and a few of them are located along the Scandinavian coast. We find a similar trend in JRPs, when forced with individual RCM realizations (Figures S11b and S11c). When the RCP8.5 RCM ensembles are pooled together (Figure 6a), we observe a high intermodel spread of projected changes in JRPs without considering SLR. This spread tends to decrease when considering the SLR. Further, the uncertainty of the multimodel ensemble is smaller as compared to the single RCM realizations for both with and without the SLR. The smaller spread of multimodel ensembles results from error cancelation (i.e., the inclusion of an apparent poor model can still improve the skill of the multimodel projection), and leads to greater consistency and increased reliability as shown in earlier studies (Deser et al., 2020; Hagedorn et al., 2005; Tebaldi & Knutti, 2007).
With SLR, the future change in JRPs lies between −7% and 48% (interquartile spread) with a median increase of 22% (Figure 6a) for locations with |SNR| > 1. Only considering the compounding effects of storm surge and peak discharge, the median increase in JRP is 75% (with 36% to 112% interquartile spread).
The PDFs of the projected changes in JRPs of the median of the multimodel ensemble show an elongated right tail (Figure 6b). This distortion arises from the large differences in JRP values between the future and reference periods in the ICHEC‐EC‐EARTH model. Considering SLR, the PDF shows a significant leftward shift (two‐sample Kolmogorov‐Smirnov test comparing the two PDFs shows p value ≤0.05) with an extended right tail and a little change on the lower tail. This shift in PDF is an indication of “changed symmetry” in combination with the “shifted mean/and variance” (changes in median and variance between the two distributions are statistically significant at 5% level using the Wilcoxon rank‐sum test and the F test for the homogeneity of variance), as reported in the IPCC AR5 report (page 134, Figure 1.8 in the IPCC Fifth Assessment Report: physical science basis, Stocker et al., 2013), suggesting considering a median SLR projection (section 2.3) favors more frequent occurrence of compound floods than neglecting the SLR.
We assess future compound floods due to high sea levels, resulting from storm surge together with probabilistic SLR projection, and peak fluvial discharge under the highly energy‐intensive RCP8.5 scenario and identify potentially vulnerable sites across northwestern Europe. While past studies relied on GCMs (Bevacqua et al., 2019; Bloemendaal et al., 2019; Lin et al., 2019; Marsooli et al., 2019) and were, thus, limited due to the very coarse spatial model resolution, this study for the first time uses a suite of dynamically downscaled regional climate model simulations archived at EURO‐CORDEX domain to analyze future compound flood hazards and associated uncertainty. While most of the earlier assessments on future trends in compound floods were limited to only one (Kew et al., 2013; Klerk et al., 2015) or a few gauges (Moftakhari et al., 2017), we analyze 72 tide gauge‐stream gauge pairs representing the entire northwestern Europe. Further, unlike previous assessments that considered either a spatially homogeneous SLR (Ikeuchi et al., 2015; Klerk et al., 2015) or stationary fluvial flood regime (Moftakhari et al., 2017), we consider physical‐model‐based, probabilistic and localized SLR projections (Kopp et al., 2014) and simulate fluvial peak discharge by forcing a global hydrological model with downscaled climate model simulations (Dosio, 2016).
Our results contrast with a previous study (Bevacqua et al., 2019) that suggests an increased probability of compound floods, in particular for northern Europe, due to higher sea levels and heavier precipitation in a warming climate. As they included precipitation as one of their compound flood drivers, it is not clear whether their result has implications for river flooding or is limited to pluvial floods. Further, they have used meteorological forcing directly from GCMs to simulate coastal water levels and did not bias correct the forcing variables. While the limited ability of GCMs to resolve fine‐scale processes may lead to underestimation of extremes, such as storm surge heights (Bloemendaal et al., 2019; Muis et al., 2016) and fluvial peak discharges (Prudhomme et al., 2002), RCMs are expected to provide an added value in the frequency distribution of local weather anomalies (Laprise, 2008). Unlike earlier assessments (Bevacqua et al., 2019; Kew et al., 2013), in which no validation was performed, our validation, comparing dynamically downscaled regional climate model simulations with observations, reveals an overestimation in the strength of dependence between the compound flood drivers. This could be a consequence of the models' structural uncertainties and limitations in their parameterizations (Fowler et al., 2005). This overestimation should motivate careful evaluations of future assessments on how well model simulations agree with observations.
A few caveats should be considered so that our results are not overinterpreted. Although a recent study (Zeppetello et al., 2019) has shown that the downwelling longwave radiation response to greenhouse forcing is controlled primarily by an increase in surface temperature, due to a lack of available RCM projections, we assume that the future downwelling longwave and shortwave radiation is the same as in the historical period. Second, due to the limited number of available ensemble members, we limit our analysis to a 30‐year window in the future using a single RCM, RCA4 driven by a limited number of host‐GCMs barring us to assess the full spectrum of multimodel response variability and the internal climate variability (Deser et al., 2020), which contributes significantly to projection uncertainty, especially at regional scales. Finally, the spatial resolution of WGHM is limited to 0.5°; however, this has been widely used to explore hydrologic response to climate change at a continental scale (Gosling et al., 2017; Lobanova et al., 2018; Prudhomme et al., 2014).
Global warming is expected to affect compound flood hazard in different ways: through rising sea levels, changes in precipitation and river discharge, and changes in storm activity, affecting the magnitude and timing of these flood drivers and the interactions between them. Our climate projection analysis demonstrates the intricate nature of changes in compound floods: Changes in surge elevations are expected to be small with no clear trend in most of the locations, whereas trends in river peaks tend to increase for northwestern Europe. The dependence strength between surges and river peaks shows a clear downward change, decreasing the compound flood hazard, whereas SLR is a strong driver of increasing compound floods. When these drivers and their interactions are acting together, we find mostly a decrease in coastal compound flood hazard in northwestern Europe. The substantial role of SLR in future compound flood hazard is in agreement with an earlier study for New York City (Garner et al., 2017), where the projected SLR dominated the overall increase in coastal flood levels, hence increasing coastal flood hazards, whereas the storm intensity and shifts in storm tracks showed minimal changes. However, this study was solely focused on coastal flood hazards and did not investigate trends in compound floods and the role of SLR in modulating compound flood hazard. In future research, the proposed modeling framework could be combined with hydrodynamic inundation modeling to assess the inundation extent and depth in the low‐lying delta regions. Another potential research avenue is to integrate projections of compound flood hazards with expected damages and possible adaptation costs to inform risk reduction decisions (Kreibich et al., 2017; Rosner et al., 2014). Although the present work is applied to the locations of tide and stream gauges in northwestern Europe, the methodology can be extended to any geographical region across the globe and over finer‐resolution data grids. The findings are based on state‐of‐the‐art regional climate models and available greenhouse gas emission scenarios but may be subjected to change with new scenarios and socioeconomic pathways in the future. Given the cascade of uncertainty across model chains, the results presented herein should be interpreted with caution. Nevertheless, the obtained insights will assist modelers to identify model deficiencies to reduce uncertainty associated with projected changes in coastal compound floods. A close coordination between regional climate modelers and hydrologists is required to further improve model parameterizations, so that the coupled modeling framework is capable of capturing hydrologically relevant metrics for end‐user applications.
Dr. P. Ganguli, is a recipient of the Humboldt Research Fellowship for early career researchers from Alexander von Humboldt Foundation, Germany. Support from the foundation and the host institute—GFZ German Research Centre for Geosciences, Potsdam, Germany is deeply acknowledged. The first author of the manuscript would like to thank Dr. Thomas Wahl at the University of Central Florida and Dr. Ivan D Haigh at the University of Southampton for their fruitful suggestions on the computation of skew surge from tide gauge records. The first author would like to thank Dr. Alessandro Dosio at the European Commission Joint Research Centre for his suggestions on implementation of dynamically downscaled bias‐adjusted precipitation and temperature data at a WGHM scale as well as Dr. Robert Kopp of Rutgers University and Dr. Ning Lin of Princeton University for helpful correspondence on implementation of localized SLR. Open access funding enabled and organized by Projekt DEAL.
Observed daily streamflow and hourly sea level data are obtained from GRDC data center, (
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2020. This work is published under http://creativecommons.org/licenses/by/4.0/ (the “License”). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
Compound flooding in coastal regions, that is, the simultaneous or successive occurrence of high sea levels and high river flows, is expected to increase in a warmer world. To date, however, there is no robust evidence on projected changes in compound flooding for northwestern Europe. We combine projected storm surges and river floods with probabilistic, localized relative sea‐level rise (SLR) scenarios to assess the future compound flood hazard over northwestern coastal Europe in the high (RCP8.5) emission scenario. We use high‐resolution, dynamically downscaled regional climate models (RCM) to drive a storm surge model and a hydrological model, and analyze the joint occurrence of high coastal water levels and associated river peaks in a multivariate copula‐based approach. The RCM‐forced multimodel mean reasonably represents the observed spatial pattern of the dependence strength between annual maxima surge and peak river discharge, although substantial discrepancies exist between observed and simulated dependence strength. All models overestimate the dependence strength, possibly due to limitations in model parameterizations. This bias affects compound flood hazard estimates and requires further investigation. While our results suggest decreasing compound flood hazard over the majority of sites by 2050s (2040–2069) compared to the reference period (1985–2005), an increase in projected compound flood hazard is limited to around 34% of the sites. Further, we show the substantial role of SLR, a driver of compound floods, which has frequently been neglected. Our findings highlight the need to be aware of the limitations of the current generation of Earth system models in simulating coastal compound floods.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details




1 Agricultural and Food Engineering Department, Indian Institute of Technology Kharagpur, Kharagpur, India; German Research Centre for Geosciences (GFZ), Potsdam, Germany
2 German Research Centre for Geosciences (GFZ), Potsdam, Germany
3 German Research Centre for Geosciences (GFZ), Potsdam, Germany; Institute of Environmental Sciences and Geography, University of Potsdam, Potsdam, Germany