Next Article in Journal
Machine-Learning-Based System for the Detection of Entanglement in Dyeing and Finishing Processes
Previous Article in Journal
Auto-Rickshaw Repair, Servicing and Maintenance for Youth-in-Entrepreneurship in Kumasi
Previous Article in Special Issue
Retrieval of Chlorophyll a Concentration in Water Considering High-Concentration Samples and Spectral Absorption Characteristics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Application of Remote-Sensing-Based Hydraulic Model and Hydrological Model in Flood Simulation

College of Urban and Environmental Sciences, Peking University, No. 5 Yiheyuan Road, Haidian District, Beijing 100871, China
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(14), 8576; https://doi.org/10.3390/su14148576
Submission received: 23 May 2022 / Revised: 11 July 2022 / Accepted: 12 July 2022 / Published: 13 July 2022

Abstract

:
Floods are one of the main natural disaster threats to the safety of people’s lives and property. Flood hazards intensify as the global risk of flooding increases. The control of flood disasters on the basin scale has always been an urgent problem to be solved that is firmly associated with the sustainable development of water resources. As important nonengineering measures for flood simulation and flood control, the hydrological and hydraulic models have been widely applied in recent decades. In our study, on the basis of sufficient remote-sensing and hydrological data, a hydrological (Xin’anjiang (XAJ)) and a two-dimensional hydraulic (2D) model were constructed to simulate flood events and provide support for basin flood management. In the Chengcun basin, the two models were applied, and the model parameters were calibrated by the parameter estimation (PEST) automatic calibration algorithm in combination with the measured data of 10 typical flood events from 1990 to 1996. Results show that the two models performed well in the Chengcun basin. The average Nash–Sutcliffe efficiency (NSE), percentage error of peak discharge (PE), and percentage error of flood volume (RE) were 0.79, 16.55%, and 18.27%, respectively, for the XAJ model, and those values were 0.76, 12.83%, and 11.03% for 2D model. These results indicate that the models had high accuracy, and hydrological and hydraulic models both had good application performance in the Chengcun basin. The study can a provide decision-making basis and theoretical support for flood simulation, and the formulation of flood control and disaster mitigation measures in the basin.

1. Introduction

Flooding is an extreme hydrological phenomenon [1,2,3,4]. Floods are one of humanity’s biggest challenges because of their highly destructive attributes [5]. Recently, under the influence of climate change, extreme flood events have occurred frequently, posing a serious threat to the safety of human life and property [6,7]. Due to its special geographical location and climatic conditions, China has suffered from the threat and impact of floods for a long time, which has an adverse effect on the sustainable development of society and the environment [8,9,10]. Clearly, effective flood management methods are urgently needed to minimize flood-related losses, which not only has an important theoretical contribution, but also has strong practical significance.
The most commonly used nonengineering measures in flood simulation and management mainly include hydrological and hydraulic models [11,12]. The hydrological model is a technology for flood process simulation that is generally acknowledged [12,13,14]. Since the 1950s, the hydrological models have been continuously enriched and improved [15,16,17]. Numerous hydrological models have been developed and applied extensively around the world with good results, such as the XAJ [18], TOPMODEL [19], VIC [20], TANK [21], GR3 [22], GR4 J [23], EasyDHM [24], PDM [25], ARNO [26], SAC [27], SWAT [28,29], HBV [30], SWM [31], and G2G [32] models; more details about these models are given in Table 1. These models have been tested and applied through long-term practice, and have high reliability and practicability [33]. Among them, the XAJ model has been widely applied in China and achieved good results [34,35,36].
With the development of numerical simulation technology and the introduction of systems theory, hydraulic methods with higher simulation precision have been gradually developed [43,44,45]. With the rapid development of computer technology, the numerical modelling in Saint-Venant equations has been fully developed and extensively applied in scientific research, which has been a central topic in hydraulic research [46,47,48]. On the basis of these studies, 2D models have been extensively employed in flood simulation and water resource management [44,49,50]. These models can not only provide more hydrological spatial information but also better simulate the hydraulic properties. In recent years, benefiting from the continuous enrichment of remote sensing and 3S (GIS, RS, and GPS) technology, hydrologists and hydraulic researchers are paying increasing attention to 2D models, and hydraulic models were rapidly enhanced [51,52,53,54,55].
In this paper, the XAJ and 2D models were established to compare and explore the performance of the hydrological and hydraulic models. The Chengcun basin was taken as the study area. Results were evaluated by comparing observed and calculated hydrographs for 10 typical flood events. The purpose of this study was to provide a theoretical basis for flood control and disaster reduction. This paper is structured as follows. Section 2 describes the data, study area, and the methods. Section 3 discusses the results of the XAJ and 2D models. In Section 4, we summarize the conclusions.

2. Materials and Methods

2.1. Hydrological Model (XAJ Model)

As a lumped rainfall–runoff model, the XAJ model based on a saturation–excess runoff generation mechanism was developed by Zhao et al. [18]. Figure 1 shows the flowchart of the XAJ model. The inputs of the XAJ model are precipitation (P) and potential evaporation (EM), and the output is the discharge (Figure 1). The model consists of four parts: evapotranspiration, runoff generation, runoff separation, and runoff routing.
Evapotranspiration: A three-layer evapotranspiration approach was adopted to calculate the evaporation, and the soil profile was divided into upper, lower, and deep soil layers. Their evaporation values were EU, EL, and ED, respectively. Total evaporation E could be obtained as follows:
E = EU + EL + ED,
Runoff generation: Runoff generation is one of the most important modules in the XAJ model, and the saturation–excess mechanism was employed, which can be represented by the following equations:
fa F A = 1 - ( 1 - W m W mm ) B ,  
{ i f   P E + A < W mm T h e n :   R = P E W M + W 0 + W M [ 1 ( P E + A ) / W mm ] 1 + B i f   P E + A W mm T h e n : R = P E W M + W 0 ,  
where fa is the partial area where soil moisture storage capacity is less than or equal to W m (km2), F A is the basin area (km2), W m is the storage capacity at a point (mm), W mm is the maximal value of W m , B is the exponent of the spatial distribution curve of tension water storage capacity, P is rainfall (mm),   E is evapotranspiration (mm), A is the initial state of the basin, R is runoff yield (mm), W M is the average tension water capacity (mm), W 0 is the soil water storage at the beginning (mm).
Runoff separation: After runoff generation, the total runoff is divided into surface runoff, interflow, and groundwater runoff with a free water capacity distribution curve:
FS F R = 1 - ( 1 - S M F SMMF ) EX ,
where FS is the area of the basin free water storage capacity that is less than or equal to S M F (km2), F R is the area where runoff yield is more than 0 (km2), S M F is the free water storage capacity (mm), SMMF is the maximal value of S M F (mm), and EX is the exponent of the free water capacity curve influencing the development of the saturated area.
Runoff routing: The surface runoff is routed by the instantaneous unit hydrograph or by the lag-and-route method [56,57], while the interflow and groundwater runoff are routed by linear reservoirs [58]. The outflow of each sub-basin is routed to the basin outlet with the Muskingum method [59].

2.2. Two-Dimensional Hydraulic Model (2D Model)

The 2D model solves the 2D shallow water equations (SWEs) on the basis of the conservation of mass principles and momentum conservation under steady-state flow conditions. The two-dimensional unsteady flow equation is composed of mass balance and the momentum balance, and can be represented as follows [54,60,61]:
q t + f x + g y = st ,
where q represents flow, f is the flux in the x direction, g is the flux in the y direction ,   t denotes the time, and st   stands   for the source terms in which
q = [ d d u d v ] , f = [ d u u 2 d + 1 2 u v d g d 2 ] ,   g = [ d v d v u d v 2 + 1 2 g d 2 ]
st = [ r φ x ρ g d B F x φ y ρ g d B F y ] ,
B F x = s p x , B F y = sp y ,
φ x = ρ C r u u 2 + v 2 , φ y = ρ C r v u 2 + v 2 ,
where d is the depth of the flow, u and v are flow velocities,   g is gravitational acceleration, r is the surface runoff, τ m x and τ m y are the bed friction stresses, ρ is the density of water, B F x and B F y is the bottom friction in the x and y directions, φ x and φ y are the bed friction stresses in the x and y directions, sp is the slope, C r is the bed roughness coefficient ( C r = g n 2 / d 1 3 ), and n is Manning’s coefficient.
The above equations are solved by the Godunov-type scheme and finite volume methods:
Ω   q t dV + Γ   f ( q ) n d Γ + Γ   g ( q ) n d Γ = Ω   s ( q ) dV ,
where Ω is the computational domain, n is the normal vector of the boundary, and   dV and d Γ are the area integration element and line integration element.
Equation (10) can be discretized as follows:
q n + 1 = q n t A i = 1 m [ f n i ( q ) + g n i ( q ) ] L i + t s ( q ) ,
where n is the outer normal vector of the boundary of cell, ∆t is the present time step,   A is the cell area (m2), i is the cell edge’s index, m is the present time level, and L is the corresponding cell edge length.
The HLLC approximate Riemann solver was adopted to solve the numerical fluxes because it is suited to cases involving a wet/dry interface.

2.3. Study Area and Data Description

Chengcun basin is situated in the south of Huangshan city, Anhui province, China (29°32′–29°44′ N, 117°38′–117°53′ E) (Figure 2). The basin area is 290 square kilometers, and the longest river is 36 km. The Chengcun basin has a subtropical monsoon humid climate with four distinctive seasons. Average annual precipitation in the basin is about 2100 mm with uneven rainfall distribution. Annual precipitation is mainly concentrated in the flood season, which is from April to September, accounting for 70% of the total precipitation. The terrain is high-altitude in the south, and low-altitude in the north with a large relative height difference, and the average elevation is about 583 m. The mean slope gradient of the river channel is 0.95%. The vegetation in the basin is good (Figure 3), and the vegetation types mainly include evergreen coniferous, deciduous broadleaved, and mixed forests [62].
The dataset included rainfall, evaporation, discharge, DEM, and land use. The hydrological data (rainfall, evaporation, and discharge) were taken from regional hydrological administrations or the hydrological yearbooks. The DEM dataset was provided by the site of the Geospatial Data Cloud, Computer Network Information Center, Chinese Academy of Sciences (http://www.gscloud.cn) (accessed on 22 April 2022). Land use data come from a NASA website (https://earthexplorer.usgs.gov) (accessed on 22 April 2022).

2.4. Data Preprocessing and Model Settings

XAJ model data preprocessing and model settings mainly included sub-basin division, hydrological data processing, and parameter calibration. According to the modeling idea of the XAJ model, the entire study area was divided into sub-basins according to DEM and GIS. The average areal rainfall was calculated with the Thiessen polygonal method [63]. The parameter estimation (PEST) automatic calibration algorithm was adopted to calibrate the model with a MATLAB environment [64]. This algorithm had the advantages of the inverse Hessian method and the steepest descent method, and can obtain the optimal parameter results through fewer model runs [65]. During parameter optimization, we considered three elements: the parameter boundaries, the objective function, and the termination conditions. The parameter boundaries were identified by the recommended parameter range and their physical meaning (Table 2). The objective function was presented as follows:
O b j = i = 1 n ( Q s i m , i Q o b q , i ) 2 ,
where O b j is the objective function, Q s i m , i is the simulated discharge at time i , and Q o b q , i is the observed discharge at time i .
In our study, 10 flood events were used for model calibration (six events) and validation (four events) at a 1 h time step. The details can be found in Appendix A. The calibration parameters are summarized in Table 2.
The data preprocessing of the 2D hydraulic model mainly included land use and Manning’s coefficient processing, rainfall data processing, and model setting. (1) According to 30 m Landsat image data, considering the influence of land use on roughness [58], the supervised classification method was used to classify the land use types into five categories: water, cultivated land, urban land, forest, and grass land. The Manning’s coefficient of different land uses is shown in Table 3. (2) Rainfall after deducting vertical loss was used as the surface flow of the 2D model. (3) The 2D model adopted the Godunov-type scheme to calculate Equations (5)–(11). On the basis of the Courant–Friedrichs–Lewy (CFL) criterion, the adaptive time step method was implemented. The output time step was 1 h.

2.5. Modeling Evaluation Criteria

According to the accuracy standard in China, the Nash–Sutcliffe efficiency (NSE), percentage error of flood volume (RE), percentage error of peak discharge (PE), and the time difference of arrival flood peak ( T ) were employed to estimate the model performance.
NSE = 1 - ( Q o - Q s ) 2 ( Q o - Q o ¯ ) 2   ,
RE = | Q s Q o Q o × 100 % |   ,
PE = | M A X ( Q s ) M A X ( Q o ) M A X ( Q o ) × 100 % | ,
T = | T M a x O b T M a x S i |   ,
where Q o is the observed discharge, Q s   is the simulated discharge, Q o ¯ is the mean of observed discharge, T M a x O b is the arrival time of observed peak discharge, and   T M a x S i is the arrival time of simulated peak discharge.
According to the hydrological modelling accuracy standard and the time scale of this paper [71], if the PE/RE was less than 20%, the simulation results are acceptable. The results were classified into three levels as follows: the first level (PE/RE ≤ 10%; NSE ≥ 0.8), the second level (10% < PE/RE ≤ 15%; 0.7 ≤ NSE < 0.8) and the third level (15% < PE/RE ≤ 20%; 0.6 ≤ NSE < 0.7). Otherwise, the results were unsatisfying.

3. Results and Discussion

3.1. Land Use Change in Chengcun Basin

The stability and consistency of rainfall runoff in the basin greatly impacted the model’s results, which were also strongly related to land use change. Hence, it was critical to analyze the land use change in the basin before analyzing the flood simulation results. The change in land use during the study period is shown in Table 4. There were five land use types, and forest land accounted for the largest proportion (about 99%). The land use type is very suitable for the saturated runoff mechanism. Furthermore, the degree of land-use change was not obvious over the years, so it had little impact on the consistency of the hydrological data. These characteristics were very beneficial for our research, and this is another reason why we chose this study area.

3.2. Comparing the Results of the Hydrological and Hydraulic Models

Considering the availability and consistency of hydrological data, 10 flood events were selected to validate the models. Results were evaluated with the four statistical indices mentioned in Section 2.5. The indices were calculated and are shown in Table 5.
As shown in Table 5: (1) The REs of the XAJ and 2D models were mostly within 20% (XAJ: 7/10, 2D: 9/10), indicating that the two models had achieved satisfactory accuracy in water balance. However, both models had large errors during the first flood event. This phenomenon may have been due to the antecedent precipitation data, and may have needed further improvement. (2) In terms of PE, the 2D model obtained better results with all qualified simulations, while the XAJ model achieved a 60% qualification rate. This demonstrated the strengths of the 2D hydraulic model in confluence. (3) In terms of ΔT, the XAJ model achieved good simulation performance of the peak occurrence time (within 3 h except 19900626, which may have been because of the uneven distribution of rainfall or errors in the meteorological data). (4) From the perspective of NSE, the results of the XAJ model were better. NSE values of XAJ ranged from 0.62 to 0.90 with a mean of 0.81, while they ranged from 0.56 to 0.91 with a mean of 0.75 in the 2D model. These results indicate that the XAJ model was able to reproduce the flood process once the optimal parameters had been obtained. The hydrological model had better overall discharge results. (5) According to the hydrological modeling accuracy standard, in terms of NSE, the percentages of the first, second, and third levels were 60%, 10% and 30%, respectively, for the XAJ model, and those values were 40%, 30%, and 10% for the 2D model. Taken as a whole, both models showed good performance that could reflect the magnitude and trend of the flood process.
The scatter plot of the simulated and observed discharges is shown in Figure 4. The plot indicates that the results of the both models were within acceptable ranges ((R2: 0.85 and 0.77, respectively). Among them, the fitting line of XAJ model was close to the 1:1 line, indicating that the simulated discharge was closest to the observed one. The fitting line of the 2D model was also within a reasonable range. However, the simulated value was too large at low flow, and was low at high flow. These errors were due to both anthropogenic activities (agricultural irrigation, Figure 3 and Table 4) and rainfall intensity (Table 5) [69,72].
The discharge hydrographs are shown in Figure 5. The observed and simulated discharges agreed well, indicating that the determined parameters could basically reflect the characteristics of runoff generation and confluence in the basin. The XAJ model did not fit the regression process well for some flood events due to its generalization in confluence. How to integrate the 2D model into the confluence of the XAJ model is a point worth investigating.
In addition to the discharge at the basin outlet, the 2D model could also provide more information about the water depth and inundation range within the basin. The maximal water depth of the basin was analyzed and presented in Figure 6. The maximal submerged water depth of the flood events (19,930,629) was about 2 m. The flow was distributed in the river channel, and there were also varying degrees of inundation in the surrounding areas of the river channel. The flood spread significantly to both sides across the river, and the residential areas were considered to be at risk of flooding. These results provide additional information for understanding flood movement and flood mitigation.

4. Conclusions

Flood disasters pose a tremendous threat to human life and property. Flood simulation and prediction play a huge role in flood control and disaster mitigation. Hydrological and hydraulic models are major nonengineering methods for flood analysis and management, and their importance is self-evident. In this paper, a hydrological (XAJ) model and a two-dimensional hydraulic (2D) model were established on the basis of measured hydrological, surface-elevation, land-use, and other data in the study area. The simulation performance of the two models for flood volume, flood peak, and arrival time (flood peak) was compared. The results (Table 5) show that our approach is a useful strategy for flood simulation in the Chengcun basin. Compared with the XAJ model, the 2D model achieved much better performance in terms of flood peak (mean PE:11.00% < 16.45%). The XAJ model, on the other hand, performed better in terms of goodness of fit (Figure 4, mean NSE:0.79 > 0.76). These results demonstrate that the XAJ model had more advantages in the overall effect of the flood simulation, while the 2D model was more prominent in flood local feature simulation. The spatial inundation map could also be primarily captured by the 2D model. The residential areas are at risk of being in inundation (Figure 6), which indicates that these areas urgently need to strengthen the river channel.
This study provides theoretical reference for the combined application of hydrological and hydraulic models. However, due to limited conditions, there are still things that need further research or improvement: The Manning’s coefficients of different land use types were directly adopted from the related research. These coefficients may be different from the actual roughness of the basin. In addition, the effect of antecedent precipitation on the results should be further considered. Furthermore, due to the limitations of the data, the flood events that could be used in this paper were limited, so the obtained conclusions may have certain errors. With the future supplementation of data and the improvement of the models, these conclusions should be further tested and revised.

Author Contributions

Conceptualization, C.X. and J.Y.; methodology, C.X.; software, C.X.; validation, C.X., J.Y. and L.W.; formal analysis, C.X.; investigation, J.Y.; resources, J.Y.; data curation, L.W.; writing—original draft preparation, C.X. and J.Y.; writing—review and editing, L.W.; visualization, J.Y.; supervision, L.W.; project administration, C.X.; funding acquisition, C.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Major Science and Technology Program for Water Pollution Control and Treatment, grant number 2014ZX07203-008.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

The authors would like to acknowledge Xiaoliu Yang of Peking University, China for his theoretical support and guidance.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

The PEST calibration method used in this paper adopted a single objective function (Equation (12)). The automatic calibration process is the process of interaction and iteration between the input and output files of the XAJ model and PEST. PEST read the output file of the XAJ model through the TSPROC program, optimized and obtained the calibration parameters through the GML algorithm, and then wrote it into the XAJ model for resimulation until the iteration termination condition had been met. The main structure consisted of three parts: the input files, the time series processing program (Tsproc program), and the PEST operation and interpretation. The input files of the calibration process included the initial parameter simulation result of the XAJ model, parameter file (parameter name, initial value, upper and lower limits (Table 2), and parameter group), and template file (.tpl format, one for each model input file with parameters from calibrating the model that managed the calibration application). Tsproc was mainly used to automatically generate the control files and instruction files, and deal with the simulation value sequence output by the XAJ model. The PEST operation and interpretation were mainly used for the interaction between the main PEST program and the XAJ model. The whole procedure process is shown in Figure A1, and the required input files are shown in Table A1.
Figure A1. The workflow of the calibration process.
Figure A1. The workflow of the calibration process.
Sustainability 14 08576 g0a1
Table A1. The required input files for calibration.
Table A1. The required input files for calibration.
NumberNameDescription
ISdischarge.txtSimulation discharge results by initial parameter values
Format: identifier (estimate/test); date (YYYYMMDD HHMMSS); simulated discharge (m3/s)
IIOdischarge.txtObserved discharge (m3/s)
Format: same as Sdischarge.txt
IIIParameter.txtParameter file
Format: initial parameter values; parameter’s name; parameter ranges
IVParameter.tplFiles with parameter templates
Format: identifier (ptf #); #
Parameter name; parameter ranges
VGroup.txtParameter group file
Format: parameter’s name; type; initial values; upper bounds; lower bounds; parameter group
VITsproc.txtTime series control file
VIIModel.pstThe PEST control file provides the PEST program with the names of all the files with templates and instructions, the names of the appropriate input and output files of the model, the values of the control variables, the values of the initial parameters, the measurement value, and the weights
(* parameter data)
VIIIModel.insOne for each model output file from which to read numbers, and files with parameter templates (.tpl);
one for each model input file with parameters from calibrating the model that manages the calibration application.
(pif $)
IXTsproc.txtTime series control file
Set keywords and date formats
Read Sdischarge.txt and Odischarge.txt and Write output file (Simulated discharge, model.pst and model.ins)
XXAJ.recThe results of the calculation process for each iteration

References

  1. Willner, S.N.; Otto, C.; Levermann, A. Global economic response to river floods. Nat. Clim Change 2018, 8, 594–598. [Google Scholar] [CrossRef]
  2. Costache, R.; Bui, D.T. Spatial prediction of flood potential using new ensembles of bivariate statistics and artificial intelligence: A case study at the Putna river catchment of Romania. Sci. Total Environ. 2019, 691, 1098–1118. [Google Scholar] [CrossRef]
  3. Toosi, A.S.; Calbimonte, G.H.; Nouri, H.; Alaghmand, S. River basin-scale flood hazard assessment using a modified multi-criteria decision analysis approach: A case study. J. Hydrol. 2019, 574, 660–671. [Google Scholar] [CrossRef]
  4. Tsakiris, G. Flood risk assessment: Concepts, modelling, applications. Nat. Hazards Earth Syst. Sci. 2014, 14, 1361–1369. [Google Scholar] [CrossRef] [Green Version]
  5. Dottori, F.; Szewczyk, W.; Ciscar, J.C.; Zhao, F.; Alfieri, L.; Hirabayashi, Y.; Bianchi, A.; Mongelli, I.; Frieler, K.; Betts, R.A.; et al. Increased human and economic losses from river flooding with anthropogenic warming. Nat. Clim. Change 2018, 8, 1021. [Google Scholar] [CrossRef] [Green Version]
  6. Rehman, S.; Sahana, M.; Hong, H.Y.; Sajjad, H.; Ahmed, B.B. A systematic review on approaches and methods used for flood vulnerability assessment: Framework for future research. Nat. Hazards 2019, 96, 975–998. [Google Scholar] [CrossRef]
  7. Mateo-Garcia, G.; Veitch-Michaelis, J.; Smith, L.; Oprea, S.V.; Schumann, G.; Gal, Y.; Baydin, A.G.; Backes, D. Towards global flood mapping onboard low cost satellites with machine learning. Sci. Rep. 2021, 11, 7249. [Google Scholar] [CrossRef]
  8. Luo, P.P.; He, B.; Takara, K.; Xiong, Y.E.; Nouer, D.; Duan, W.L.; Fukushi, K. Historical assessment of Chinese and Japanese flood management policies and implications for managing future floods. Environ. Sci. Policy 2015, 48, 265–277. [Google Scholar] [CrossRef] [Green Version]
  9. Song, S.; Xu, Y.P.; Wu, Z.F.; Deng, X.J.; Wang, Q. The relative impact of urbanization and precipitation on long-term water level variations in the Yangtze River Delta. Sci. Total Environ. 2019, 648, 460–471. [Google Scholar] [CrossRef]
  10. Shi, J.; Cui, L.L.; Tian, Z. Spatial and temporal distribution and trend in flood and drought disasters in East China. Environ. Res. 2020, 185, 109406. [Google Scholar] [CrossRef]
  11. Icyimpaye, G.; Abdelbaki, C.; Mourad, K.A. Hydrological and hydraulic model for flood forecasting in Rwanda. Modeling Earth Syst. Environ. 2021, 8, 1179–1189. [Google Scholar] [CrossRef]
  12. Segura-Beltran, F.; Sanchis-Ibor, C.; Morales-Hernandez, M.; Gonzalez-Sanchis, M.; Bussi, G.; Ortiz, E. Using post-flood surveys and geomorphologic mapping to evaluate hydrological and hydraulic models: The flash flood of the Girona River (Spain) in 2007. J. Hydrol. 2016, 541, 310–329. [Google Scholar] [CrossRef] [Green Version]
  13. Sood, A.; Smakhtin, V. Global hydrological models: A review. Hydrol. Sci. J. 2015, 60, 549–565. [Google Scholar] [CrossRef]
  14. Bierkens, M.F.P. Global hydrology 2015: State, trends, and directions. Water Resour. Res. 2015, 51, 4923–4947. [Google Scholar] [CrossRef]
  15. Fatichi, S.; Vivoni, E.R.; Ogden, F.L.; Ivanov, V.Y.; Mirus, B.; Gochis, D.; Downer, C.W.; Camporese, M.; Davison, J.H.; Ebel, B.A.; et al. An overview of current applications, challenges, and future trends in distributed process-based models in hydrology. J. Hydrol. 2016, 537, 45–60. [Google Scholar] [CrossRef] [Green Version]
  16. Guo, Y.; Zhang, Y.; Zhang, L.; Wang, Z. Regionalization of hydrological modeling for predicting streamflow in ungauged catchments: A comprehensive review. WIREs Water 2020, 8, e1487. [Google Scholar] [CrossRef]
  17. Johnston, R.; Smakhtin, V. Hydrological Modeling of Large river Basins: How Much is Enough? Water Resour. Manag. 2014, 28, 2695–2730. [Google Scholar] [CrossRef] [Green Version]
  18. Zhao, R.J. The Xinanjiang Model Applied in China. J. Hydrol. 1992, 135, 371–381. [Google Scholar] [CrossRef]
  19. Beven, K. TOPMODEL: A critique. Hydrol. Process. 1997, 11, 1069–1085. [Google Scholar] [CrossRef]
  20. Liang, X.; Lettenmaier, D.P.; Wood, E.F. One-dimensional statistical dynamic representation of subgrid spatial variability of precipitation in the two-layer variable infiltration capacity model. J. Geophys Res. Atmos. 1996, 101, 21403–21422. [Google Scholar] [CrossRef]
  21. Zotz, G.; Thomas, V. How much water is in the tank? Model calculations for two epiphytic bromeliads. Ann. Bot. 1999, 83, 183–192. [Google Scholar] [CrossRef] [Green Version]
  22. Nascimento, N.D.; Yang, X.I.; Makhlouf, Z.; Michel, C. GR3J: A daily watershed model with three free parameters. Hydrol. Sci. J. 1999, 44, 263–277. [Google Scholar] [CrossRef]
  23. Yang, X.L.; Michel, C. Flood forecasting with a watershed model: A new method of parameter updating. Hydrol. Sci. J. 2000, 45, 537–546. [Google Scholar] [CrossRef]
  24. Lei, X.H.; Liao, W.H.; Wang, Y.H.; Jiang, Y.Z.; Wang, H.; Tian, Y. Development and Application of a Distributed Hydrological Model: EasyDHM. J. Hydrol. Eng. 2014, 19, 44–59. [Google Scholar] [CrossRef]
  25. Zhang, R.; Cuartas, L.A.; Carvalho, L.V.D.; Leal, K.R.D.; Mendiondo, E.M.; Abe, N.; Birkinshaw, S.; Mohor, G.S.; Seluchi, M.E.; Nobre, C.A. Season-based rainfall-runoff modelling using the probability-distributed model (PDM) for large basins in southeastern Brazil. Hydrol. Process. 2018, 32, 2217–2230. [Google Scholar] [CrossRef]
  26. Todini, E. The ARNO rainfall-runoff model. J. Hydrol. 1996, 175, 339–382. [Google Scholar] [CrossRef]
  27. Wren, S.A.C.; Rowe, R.C. Theoretical Aspects of Chiral Separation in Capillary Electrophoresis.1. Initial Evaluation of a Model. J. Chromatogr. 1992, 603, 235–241. [Google Scholar] [CrossRef]
  28. Arnold, J.G.; Srinivasan, R.; Muttiah, R.S.; Williams, J.R. Large area hydrologic modeling and assessment-Part 1: Model development. J. Am. Water Resour. Assoc. 1998, 34, 73–89. [Google Scholar] [CrossRef]
  29. Xu, C.; Wang, Y.; Fu, H.; Yang, J. Comprehensive Analysis for Long-Term Hydrological Simulation by Deep Learning Techniques and Remote Sensing. Front. Earth Sci. 2022, 10, 875145. [Google Scholar] [CrossRef]
  30. Geris, J.; Tetzlaff, D.; Seibert, J.; Vis, M.; Soulsby, C. Conceptual Modelling to Assess Hydrological Impacts and Evaluate Environmental Flow Scenarios in Montane River Systems Regulated for Hydropower. River Res. Appl. 2015, 31, 1066–1081. [Google Scholar] [CrossRef]
  31. Crawford, N.H.; Linsley, R.K. Digital Simulation in Hydrology: Stanford Watershed Model IV; Evapotranspiration; Department of Civil Engineering Stanford University: Stanford, CA, USA, 1966. [Google Scholar]
  32. Moore, R.J. The PDM rainfall-runoff model. Hydrol. Earth Syst. Sci. 2007, 11, 483–499. [Google Scholar] [CrossRef]
  33. Jiang, T.; Chen, Y.Q.D.; Xu, C.Y.Y.; Chen, X.H.; Chen, X.; Singh, V.P. Comparison of hydrological impacts of climate change simulated by six hydrological models in the Dongjiang Basin, South China. J. Hydrol. 2007, 336, 316–333. [Google Scholar] [CrossRef]
  34. Zhang, Y.Q.; Chiew, F.H.S. Relative merits of different methods for runoff predictions in ungauged catchments. Water Resour. Res. 2009, 45, W07412. [Google Scholar] [CrossRef]
  35. Gong, J.F.; Yao, C.; Li, Z.J.; Chen, Y.F.; Huang, Y.C.; Tong, B.X. Improving the flood forecasting capability of the Xinanjiang model for small- and medium-sized ungauged catchments in South China. Nat. Hazards 2021, 106, 2077–2109. [Google Scholar] [CrossRef]
  36. Bai, P.; Liu, X.M.; Liang, K.; Liu, X.J.; Liu, C.M. A comparison of simple and complex versions of the Xinanjiang hydrological model in predicting runoff in ungauged basins. Hydrol. Res. 2017, 48, 1282–1295. [Google Scholar] [CrossRef]
  37. Zhao, L.L.; Xia, J.; Xu, C.Y.; Wang, Z.G.; Sobkowiak, L.; Long, C.R. Evapotranspiration estimation methods in hydrological models. J. Geogr. Sci. 2013, 23, 359–369. [Google Scholar] [CrossRef]
  38. Song, X.M.; Zhang, J.Y.; Zhan, C.S.; Xuan, Y.Q.; Ye, M.; Xu, C.G. Global sensitivity analysis in hydrological modeling: Review of concepts, methods, theoretical framework, and applications. J. Hydrol. 2015, 523, 739–757. [Google Scholar] [CrossRef] [Green Version]
  39. Dong, C.Y. Remote sensing, hydrological modeling and in situ observations in snow cover research: A review. J. Hydrol. 2018, 561, 573–583. [Google Scholar] [CrossRef]
  40. Praskievicz, S.; Chang, H.J. A review of hydrological modelling of basin-scale climate change and urban development impacts. Prog. Phys. Geogr. 2009, 33, 650–671. [Google Scholar] [CrossRef]
  41. Gao, H.K.; Sabo, J.L.; Chen, X.H.; Liu, Z.Y.; Yang, Z.J.; Ren, Z.; Liu, M. Landscape heterogeneity and hydrological processes: A review of landscape-based hydrological models. Landsc. Ecol. 2018, 33, 1461–1480. [Google Scholar] [CrossRef]
  42. Dong, Q.Q.; Zhan, C.S.; Wang, H.X.; Wang, F.Y.; Zhu, M.C. A review on evapotranspiration data assimilation based on hydrological models. J. Geogr. Sci. 2016, 26, 230–242. [Google Scholar] [CrossRef] [Green Version]
  43. Fernández-Pato, J.; Caviedes-Voullième, D.; García-Navarro, P. Rainfall/runoff simulation with 2D full shallow water equations: Sensitivity analysis and calibration of infiltration parameters. J. Hydrol. 2016, 536, 496–513. [Google Scholar] [CrossRef]
  44. Costabile, P.; Costanzo, C. A 2D-SWEs framework for efficient catchment-scale simulations: Hydrodynamic scaling properties of river networks and implications for non-uniform grids generation. J. Hydrol. 2021, 599, 126306. [Google Scholar] [CrossRef]
  45. Hien, L.T.T.; Chien, N.V. Investigate Impact Force of Dam-Break Flow against Structures by Both 2D and 3D Numerical Simulations. Water 2021, 13, 344. [Google Scholar] [CrossRef]
  46. Wang, G.T.; Chen, S.; Boll, J. A semianalytical solution of the Saint-Venant equations for channel flood routing. Water Resour. Res. 2003, 39, 1076. [Google Scholar] [CrossRef]
  47. Zarmehi, F.; Tavakoli, A. A Simple Scheme to Solve Saint-Venant Equations by Finite Element Method. Int. J. Comput. Methods 2016, 13, 1650001. [Google Scholar] [CrossRef] [Green Version]
  48. Yu, C.W.; Hodges, B.; Liu, F. A new form of the Saint-Venant equations for variable topography. Hydrol. Earth Syst. Sci. 2020, 24, 4001–4024. [Google Scholar] [CrossRef]
  49. Yang, X.; An, W.; Li, W.; Zhang, S. Implementation of a Local Time Stepping Algorithm and Its Acceleration Effect on Two-Dimensional Hydrodynamic Models. Water 2020, 12, 1148. [Google Scholar] [CrossRef] [Green Version]
  50. Kim, B.; Sanders, B.F.; Schubert, J.E.; Famiglietti, J.S. Mesh type tradeoffs in 2D hydrodynamic modeling of flooding with a Godunov-based flow solver. Adv. Water Resour. 2014, 68, 42–61. [Google Scholar] [CrossRef] [Green Version]
  51. Xia, X.L.; Liang, Q.H.; Ming, X.D. A full-scale fluvial flood modelling framework based on a high-performance integrated hydrodynamic modelling system (HiPIMS). Adv. Water Resour. 2019, 132, 103392. [Google Scholar] [CrossRef]
  52. Liang, Q.H.; Smith, L.S. A high-performance integrated hydrodynamic modelling system for urban flood simulations. J. Hydroinform. 2015, 17, 518–533. [Google Scholar] [CrossRef]
  53. Fernandez, A.; Najafi, M.R.; Durand, M.; Mark, B.G.; Moritz, M.; Jung, H.C.; Neal, J.; Shastry, A.; Laborde, S.; Phang, S.C.; et al. Testing the skill of numerical hydraulic modeling to simulate spatiotemporal flooding patterns in the Logone floodplain, Cameroon. J. Hydrol. 2016, 539, 265–280. [Google Scholar] [CrossRef] [Green Version]
  54. Xu, C.W.; Han, Z.Y.; Fu, H. Remote Sensing and Hydrologic-Hydrodynamic Modeling Integrated Approach for Rainfall-Runoff Simulation in Farm Dam Dominated Basin. Front. Environ. Sci. 2022, 9, 672. [Google Scholar] [CrossRef]
  55. Sauvagnargues, S.; Salze, D.; AyralP, A.; Laganier, O. A coupling of hydrologic and hydraulic models appropriate for the fast floods of the Gardon River basin (France). Nat. Hazards Earth Syst. Sci. 2014, 14, 2899–2920. [Google Scholar]
  56. Sherman, L. Streamflow from rainfall by the unit-graph method. Eng. News Rec. 1932, 108, 501–505. [Google Scholar]
  57. Luo, B. Theory and application of lag-and-route method and comparison with other flood routing methods. J. Hydraul. Eng. 1987, 32, 485–496. (In Chinese) [Google Scholar]
  58. Monajemi, P.; Khaleghi, S.; Maleki, S. Derivation of instantaneous unit hydrographs using linear reservoir models. Hydrol. Res. 2021, 52, 339–355. [Google Scholar] [CrossRef]
  59. Gill, M.A. Flood Routing by the Muskingum Method-Reply. J. Hydrol. 1979, 41, 169–170. [Google Scholar] [CrossRef]
  60. Liang, Q.H.; Smith, L.; Xia, X.L. New prospects for computational hydraulics by leveraging high-performance heterogeneous computing techniques. J. Hydrodyn. 2016, 28, 977–985. [Google Scholar] [CrossRef]
  61. Wang, W.Q.; Chen, W.J.; Huang, G.R. Urban Stormwater Modeling with Local Inertial Approximation Form of Shallow Water Equations: A Comparative Study. Int. J. Disaster Risk Sci. 2021, 12, 745–763. [Google Scholar] [CrossRef]
  62. Chao, L.J.; Zhang, K.; Li, Z.J.; Wang, J.F.; Yao, C.; Li, Q.L. Applicability assessment of the CASCade Two Dimensional SEDiment (CASC2D-SED) distributed hydrological model for flood forecasting across four typical medium and small watersheds in China. J. Flood Risk Manag. 2019, 12, e12518. [Google Scholar] [CrossRef] [Green Version]
  63. Wagner, P.D.; Fiener, P.; Wilken, F.; Kumar, S.; Schneider, K. Comparison and evaluation of spatial interpolation schemes for daily rainfall in data scarce regions. J. Hydrol. 2012, 464, 388–400. [Google Scholar] [CrossRef]
  64. Goegebeur, M.; Pauwels, V.R.N. Improvement of the PEST parameter estimation algorithm through Extended Kalman Filtering. J. Hydrol. 2007, 337, 436–451. [Google Scholar] [CrossRef]
  65. Li, X.H.; Zhang, Q.; Shao, M.; Li, Y.L. A comparison of parameter estimation for distributed hydrological modelling using automatic and manual methods. Adv. Mater. Res. 2012, 356–360, 2372–2375. [Google Scholar] [CrossRef]
  66. Li, H.X.; Zhang, Y.Q.; Chiew, F.H.S.; Xu, S.G. Predicting runoff in ungauged catchments by using Xinanjiang model with MODIS leaf area index. J. Hydrol. 2009, 370, 155–162. [Google Scholar] [CrossRef]
  67. Yao, C.; Zhang, K.; Yu, Z.B.; Li, Z.J.; Li, Q.L. Improving the flood prediction capability of the Xinanjiang model in ungauged nested catchments by coupling it with the geomorphologic instantaneous unit hydrograph. J. Hydrol. 2014, 517, 1035–1048. [Google Scholar] [CrossRef]
  68. Lin, K.R.; Lv, F.S.; Chen, L.; Singh, V.P.; Zhang, Q.; Chen, X.H. Xinanjiang model combined with Curve Number to simulate the effect of land use change on environmental flow. J. Hydrol. 2014, 519, 3142–3152. [Google Scholar] [CrossRef]
  69. Li, Z.J.; Kan, G.Y.; Yao, C.; Liu, Z.Y.; Li, Q.L.; Yu, S. Improved Neural Network Model and Its Application in Hydrological Simulation. J. Hydrol. Eng. 2014, 19, 04014019. [Google Scholar] [CrossRef]
  70. Wang, Y.L.; Yang, X.L. A Coupled Hydrologic-Hydraulic Model (XAJ-HiPIMS) for Flood Simulation. Water 2020, 12, 1288. [Google Scholar] [CrossRef]
  71. Moriasi, D.N.; Gitau, M.W.; Pai, N.; Daggupati, P. Hydrologic and Water Quality Models: Performance Measures and Evaluation Criteria. Trans. ASABE 2015, 58, 1763–1785. [Google Scholar]
  72. Amo-Boateng, M.; Li, Z.J.; Guan, Y.Q. Inter-annual variation of streamflow, precipitation and evaporation in a small humid watershed (Chengcun Basin, China). Chin. J. Oceanol. Limnol. 2014, 32, 455–468. [Google Scholar] [CrossRef]
Figure 1. Framework of XAJ model (red fonts represent the parameters that are introduced in Section 2.4).
Figure 1. Framework of XAJ model (red fonts represent the parameters that are introduced in Section 2.4).
Sustainability 14 08576 g001
Figure 2. Location map for the study area.
Figure 2. Location map for the study area.
Sustainability 14 08576 g002
Figure 3. Land use map (1992).
Figure 3. Land use map (1992).
Sustainability 14 08576 g003
Figure 4. Comparison between observed and simulated discharges.
Figure 4. Comparison between observed and simulated discharges.
Sustainability 14 08576 g004
Figure 5. Hydrographs of models XAJ and 2D in the study area (Ob: observed discharge).
Figure 5. Hydrographs of models XAJ and 2D in the study area (Ob: observed discharge).
Sustainability 14 08576 g005
Figure 6. The map of maximal water depth (19930629, unit: m).
Figure 6. The map of maximal water depth (19930629, unit: m).
Sustainability 14 08576 g006
Table 1. Model characteristics.
Table 1. Model characteristics.
Hydrological ModelDeveloped/Maintained byTypeRunoff
Characteristic
Applications
XAJZhao, 1984LumpedSaturation—excess runoffHumid/semihumid
Hourly/daily
TOPMODELBeven and Kirkby, 1979SemidistributedSaturation—excess
runoff
Humid/arid
Daily
VICLiang et al., 1994DistributedInfiltration—excess
runoff
Daily/monthly
River basin
TANKSugawara, 1961LumpedInfiltration—excess
runoff
Single rainfall simulation
Humid/arid
GR3Edijaton, 1999LumpedSaturation—excess
runoff
Hourly/daily
Small/medium watershed
GR4 JPerrin et al., 1993LumpedSaturation—excess
runoff
Hourly/daily
Small/middle watershed
EasyDHMLei, X., et al., 2010DistributedSaturation—excess/infiltration—excessWide applicability
PDMUniversity of Reading, UK, 1998LumpedSaturation—excess
runoff
Daily
Continuous simulation
A macroscale globe model
ARNOCiarapica, Todini, 1996, 2002Lumped/
distributed
Saturation-excess
runoff
Large range of spatial scales
Continuous simulation
SACBurnash, Ferral, Ferral, 1970sLumpedSaturation—excess
runoff
Long-time continuous simulation
Daily/6 h
Large/medium watershed
Humid/arid
SWATArnold et al., 1995DistributedInfiltration—excess runoffUp to large river basin
Daily/monthly
HBVGardelin, 1997Lumped/
Distributed
Saturation—excess runoffLarge/medium watershed
Daily/donthly
SWMLinsley, Crawford,1959LumpedSaturation—excess
runoff
Continuous simulation
Hourly
Reference: [13,24,37,38,39,40,41,42].
Table 2. Optimized model parameters based on PEST.
Table 2. Optimized model parameters based on PEST.
ParameterPhysical MeaningRange and Units [66,67,68]Value
(Hourly Model) [69]
EvapotranspirationKEvaporation coefficient0.5–1.1 (-)0.8
CEvaporation coefficient of the deep layer0.1–0.3 (-)0.1
WUMTension water capacity of upper layer5–100 (mm)17
WLMTension water capacity of lower layer50–300 (mm)87
WDMTension water capacity of deep layer5–100 (mm)33
Runoff generationBRepresentation of the nonuniformity of the spatial distribution of the tension water capacity0.1–2 (-)0.3
IMPProportion of impervious surface0.01–0.1 (%)0.01
Runoff separationSMMean free water storage capacity5–100 (mm)97
EXThe distribution of free water storage capacity1–1.5 (-)1.5
Runoff routingKSSOutflow coefficient of the free water storage reservoir to interflow0.01–0.7 (-)0.4
KGOutflow coefficient of the free water storage reservoir to groundwater0.01–0.7 (-)0.3
KKIRecession coefficient of the interflow0.05–0.95 (-)0.9
KKGRecession coefficient of the groundwater0.9–0.999 (-)0.98
Sum KSS + KG was kept as 0.7.
Table 3. Manning’s coefficients corresponding to different land use types [54,70].
Table 3. Manning’s coefficients corresponding to different land use types [54,70].
Land UseManning’s Coefficient (s/[m1/3])
Urban land0.013
Water0.021
Grassland0.031
Cultivated land0.041
Forest land0.139
Table 4. Comparison of land use areas over the study period.
Table 4. Comparison of land use areas over the study period.
Land Use1990199119921993199419951996
Area ( k m 2 ) Ratio(%)Area ( k m 2 ) Ratio(%)Area ( k m 2 ) Ratio(%)Area ( k m 2 ) Ratio(%)Area ( k m 2 ) Ratio(%)Area ( k m 2 ) Ratio(%)Area ( k m 2 ) Ratio(%)
Cultivated land1.440.511.620.581.700.611.800.641.800.641.750.621.680.60
Forest279.3399.43279.1499.36279.0699.33278.9599.30278.9699.30279.0299.32279.0999.35
Grass land0.110.040.110.040.090.030.090.030.090.030.080.030.080.03
Water0.010.000.020.010.030.010.020.010.010.000.020.010.010.00
Urban land0.040.010.050.020.060.020.070.020.070.020.070.020.070.02
Table 5. Performance of the XAJ and 2D hydraulic models.
Table 5. Performance of the XAJ and 2D hydraulic models.
Flood CodeRainfall (mm)RE (%)PE (%)NSE T   ( h )
XAJ2DXAJ2DXAJ2DXAJ2D
19900614112.8831.6136.1310.5813.540.850.7700
19900626246.0918.234.6818.866.020.900.8573
19910518158.978.2426.1713.807.210.860.6413
1992051681.5819.4519.7525.4517.270.830.5603
19920701225.6231.451.166.2515.030.870.5706
19930629566.6917.162.5114.6211.610.900.8314
19940608574.1514.9610.385.1419.530.830.8027
19950519500.7817.109.0224.1712.560.690.6413
19950701160.9826.766.4118.7011.130.790.8013
19960619169.212.9919.1220.841.810.620.9104
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xu, C.; Yang, J.; Wang, L. Application of Remote-Sensing-Based Hydraulic Model and Hydrological Model in Flood Simulation. Sustainability 2022, 14, 8576. https://doi.org/10.3390/su14148576

AMA Style

Xu C, Yang J, Wang L. Application of Remote-Sensing-Based Hydraulic Model and Hydrological Model in Flood Simulation. Sustainability. 2022; 14(14):8576. https://doi.org/10.3390/su14148576

Chicago/Turabian Style

Xu, Chaowei, Jiashuai Yang, and Lingyue Wang. 2022. "Application of Remote-Sensing-Based Hydraulic Model and Hydrological Model in Flood Simulation" Sustainability 14, no. 14: 8576. https://doi.org/10.3390/su14148576

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop